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

Density estimate from below
in relation to a conjecture of A. Zygmund
on Lipschitz differentiation

Thierry De Pauw School of Mathematical Sciences
Shanghai Key Laboratory of PMMP
East China Normal University
500 Dongchuang Road
Shanghai 200062
P.R. of China
and NYU-ECNU Institute of Mathematical Sciences at NYU Shanghai
3663 Zhongshan Road North
Shanghai 200062
China
Université Paris Diderot
Sorbonne Université
CNRS
Institut de Mathématiques de Jussieu – Paris Rive Gauche, IMJ-PRG
F-75013, Paris
France [email protected],[email protected]
Abstract.

Letting AnA\subseteq\mathbb{R}^{n} be Borel measurable and 𝐖0:A𝔾(n,m)\mathbf{W}_{0}:A\to\mathbb{G}(n,m) Lipschitzian, we establish that

lim supr0+m[A𝐁(x,r)(x+𝐖0(x))]𝜶(m)rm12n,\limsup_{r\to 0^{+}}\frac{\mathscr{H}^{m}\left[A\cap\mathbf{B}(x,r)\cap(x+\mathbf{W}_{0}(x))\right]}{\boldsymbol{\alpha}(m)r^{m}}\geqslant\frac{1}{2^{n}},

for n\mathscr{L}^{n}-almost every xAx\in A. In particular, it follows that AA is n\mathscr{L}^{n}-negligible if and only if m(A(x+𝐖0(x))=0\mathscr{H}^{m}(A\cap(x+\mathbf{W}_{0}(x))=0, for n\mathscr{L}^{n}-almost every xAx\in A.

Key words and phrases:
Lebesgue measure, Nikodým set, Negligible set, Derivation basis, Zygmund conjecture, Lipschitz differentiation
2010 Mathematics Subject Classification:
Primary 28A75,26B15
The author was partially supported by the Science and Technology Commission of Shanghai (No. 18dz2271000).

1. Foreword

Let 𝒮x\mathscr{S}_{x} be the set of squares centered at x2x\in\mathbb{R}^{2}. The Lebesgue density theorem states that if f:2f:\mathbb{R}^{2}\to\mathbb{R} is Lebesgue summable, then

f(x)=limS𝒮xdiamS0+Sf𝑑2,f(x)=\lim_{\begin{subarray}{c}S\in\mathscr{S}_{x}\\ \operatorname{\mathrm{diam}}S\to 0^{+}\end{subarray}}\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int_{S}fd\mathscr{L}^{2}, (*)

for 2\mathscr{L}^{2}-almost every x2x\in\mathbb{R}^{2}. This consequence of the Vitali covering theorem fails if 𝒮x\mathscr{S}_{x} is replaced with x\mathscr{R}_{x}, the set of rectangles centered at xx of arbitrary direction and eccentricity. It fails even for some indicator function f=𝟙Af=\mathbbm{1}_{A}. Indeed, Nikodým [12] defined a set A[0,1]×[0,1]A\subseteq[0,1]\times[0,1] of full measure with the following property: For every xAx\in A, there exists a line L(x)L(x) such that AL(x)={x}A\cap L(x)=\{x\}. In fact, Nikodým’s example shows that the Lebesgue density theorem fails if we replace 𝒮x\mathscr{S}_{x} with xL\mathscr{R}_{x}^{L}, the set of rectangles of arbitrary eccentricity, centered at xx, and one side of which is parallel to L(x)L(x). Furthermore [4, Chap. IV Theorem 3.5], replacing AA by a nonnegligible subset of AA, one may choose LL to be continuous with respect to xx. Yet, if LL is constant, then the Lebesgue density theorem with respect to xL\mathscr{R}_{x}^{L} holds, for pp-summable functions ff, 1<p1<p\leqslant\infty, by virtue of a theorem of Zygmund, though the corresponding version of the Vitali covering theorem fails, according to an example of H. Bohr [4, Chap. IV Theorem 1.1]. This raises the question: What regularity condition of xL(x)x\mapsto L(x) guarantees that the Lebesgue density theorem holds with respect to xL\mathscr{R}_{x}^{L} for, say, functions that are square summable? The following version is a conjecture reportedly [11] attributed to Zygmund. Assume that xL(x)x\mapsto L(x) is a Lipschitzian field of lines, with xL(x)x\in L(x), and f:2f:\mathbb{R}^{2}\to\mathbb{R} is square summable. Is it true that

f(x)=limr0+L(x)𝐁(x,r)f𝑑1,f(x)=\lim_{r\to 0^{+}}\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int_{L(x)\cap\mathbf{B}(x,r)}fd\mathscr{H}^{1},

for 2\mathscr{L}^{2}-almost every x2x\in\mathbb{R}^{2}? Here, 1\mathscr{H}^{1} is the 1-dimensional Hausdorff measure.

E.M. Stein raised the singular integral variant of this conjecture. Both have received much attention from the harmonic analysis community. To the author’s knowledge, most results recorded so far in the literature, via the maximal function approach, assume some extra regularity property of LL – namely, that LL be C1C^{1} together with a hypothesis on the variation of the derivative – see, for instance, [1] and [11].

In this paper we offer a novel approach – a kind of change of variable based on an appropriate fibration and the coarea formula. This allows for treating the case when LL is Lipschitzian, without further restriction. We obtain the following lower density bound for indicator functions. Let xL(x)x\mapsto L(x) be a Lipschitzian field of lines, with xL(x)x\in L(x), and let A2A\subseteq\mathbb{R}^{2} be Lebesgue measurable. Then,

lim supr0+1(AB(x,r)L(x))2r14,\limsup_{r\to 0^{+}}\frac{\mathscr{H}^{1}(A\cap B(x,r)\cap L(x))}{2r}\geqslant\frac{1}{4},

for 2\mathscr{L}^{2}-almost every xAx\in A. Incidentally, the following corollary – a nonparallel version of Fubini theorem – seems to be new as well. The set AA is Lebesgue negligible if and only if 1(AL(x))=0\mathscr{H}^{1}(A\cap L(x))=0, for 2\mathscr{L}^{2}-almost every x2x\in\mathbb{R}^{2}.

Our results hold in any dimension and codimension.

2. Sketch of proof

Let AA be a subset of Euclidean space n\mathbb{R}^{n}, n2n\geqslant 2, and let n\mathscr{L}^{n} be the Lebesgue outer measure. We start by considering the following weak question: Can one tell whether AA is Lebesgue negligible from the knowledge only of its trace on each member of some given collection of “lower dimensional” subsets Γin\Gamma_{i}\subseteq\mathbb{R}^{n}, iIi\in I. One expects that if AΓiA\cap\Gamma_{i} is “negligible in the dimension of Γi\Gamma_{i}”, for each iIi\in I, then n(A)=0\mathscr{L}^{n}(A)=0. Of course, a necessary condition is that the sets Γi\Gamma_{i} cover almost all of AA, i.e., n(AiIΓi)=0\mathscr{L}^{n}(A\setminus\cup_{i\in I}\Gamma_{i})=0. Consider, for instance, n=2n=2, I=I=\mathbb{R}, and Γt={t}×\Gamma_{t}=\{t\}\times\mathbb{R}, for tt\in\mathbb{R}, the collection of all vertical lines in the plane. It is not true in general that if A2A\subseteq\mathbb{R}^{2} and AΓtA\cap\Gamma_{t} is a singleton, for each tt\in\mathbb{R}, then 2(A)=0\mathscr{L}^{2}(A)=0. There exist, indeed, functions f:f:\mathbb{R}\to\mathbb{R} whose graph AA satisfies 2(A)>0\mathscr{L}^{2}(A)>0 – see, e.g., [10, Chapter 2 Theorem 4] for an example due to Sierpiński. In order to rule out such examples, we will henceforth assume that AnA\subseteq\mathbb{R}^{n} is Borel measurable. In that case, the theorem of Fubini, together with the invariance of the Lebesgue measure under orthogonal transformations, imply the following. Given an integer 1mn11\leqslant m\leqslant n-1, if (Γi)iI(\Gamma_{i})_{i\in I} is the collection of all mm-dimensional affine subspaces of n\mathbb{R}^{n} of some fixed direction, and if m(AΓi)=0\mathscr{H}^{m}(A\cap\Gamma_{i})=0 for all iIi\in I, then n(A)=0\mathscr{L}^{n}(A)=0. Here, m\mathscr{H}^{m} denotes the mm-dimensional Hausdorff measure. A special feature of this collection (Γi)iI(\Gamma_{i})_{i\in I} is that it partitions n\mathbb{R}^{n}, its members being the level sets f1{y}f^{-1}\{y\}, ynmy\in\mathbb{R}^{n-m}, of a “nice map” f:nnmf:\mathbb{R}^{n}\to\mathbb{R}^{n-m}, indeed, an orthogonal projection. This is an occurrence of the following more general situation when ff and its leaves f1{y}f^{-1}\{y\} are allowed to be nonlinear. The coarea formula due to Federer [8] asserts that if f:nnmf:\mathbb{R}^{n}\to\mathbb{R}^{n-m} is Lipschitzian and AnA\subseteq\mathbb{R}^{n} is Borel measurable, then

AJf(x)𝑑n(x)=nmm(Af1{y})𝑑nm(y).\int_{A}Jf(x)d\mathscr{L}^{n}(x)=\int_{\mathbb{R}^{n-m}}\mathscr{H}^{m}\left(A\cap f^{-1}\{y\}\right)d\mathscr{L}^{n-m}(y).

Thus, if the Jacobian coarea factor JfJf is positive, n\mathscr{L}^{n}-almost everywhere in AA, then the collection (f1{y})ynm\left(f^{-1}\{y\}\right)_{y\in\mathbb{R}^{n-m}} is suitable for detecting whether or not AA is Lebesgue null. At n\mathscr{L}^{n}-almost all xnx\in\mathbb{R}^{n}, the map ff is differentiable, by virtue of Rademacher’s theorem, and

Jf(x)=|det(Df(x)Df(x))|=nmDf(x),Jf(x)=\sqrt{\left|\det\left(Df(x)\circ Df(x)^{*}\right)\right|}=\left\|\wedge_{n-m}Df(x)\right\|,

see [6, Chapter 3 §4] and [9, 3.2.1 and 3.2.11].

In this paper, we focus on the case when Γi\Gamma_{i}, iIi\in I, are affine subspaces of n\mathbb{R}^{n}, but not necessarily members of a partition of the ambient space. Specifically, we assume that with each xnx\in\mathbb{R}^{n} is associated an mm-dimensional affine subspace 𝐖(x)\mathbf{W}(x) of n\mathbb{R}^{n} containing xx. Given a Borel set AnA\subseteq\mathbb{R}^{n}, the question whether

If m(A𝐖(x))=0, for all xA, then n(A)=0,\textit{If }\mathscr{H}^{m}(A\cap\mathbf{W}(x))=0\text{, for all }x\in A\textit{, then }\mathscr{L}^{n}(A)=0\,, (1)

has a negative answer, in view of Nikodým’s set A2A\subseteq\mathbb{R}^{2} evoked in the previous section. Indeed, corresponding to this set AA, there exists a field of lines x𝐖(x)x\mapsto\mathbf{W}(x) such that A𝐖(x)={x}A\cap\mathbf{W}(x)=\{x\}, for all xAx\in A. In this context, a selection theorem due to von Neumann implies that (possibly considering a smaller, non Lebesgue null, Borel subset of AA) the correspondence x𝐖(x)x\mapsto\mathbf{W}(x) can be chosen to be Borel measurable (see 3.19) and, in turn, it can be chosen to be continuous, according to Lusin’s theorem. Nonetheless, when 𝐖\mathbf{W} is Lipschitzian, the situation improves, as illustrated in our theorem below; 𝔾(n,m)\mathbb{G}(n,m) is the Grassmannian manifold consisting of mm-dimensional vector subspaces of n\mathbb{R}^{n}.

Theorem. —

Assume 𝐖0:n𝔾(n,m)\mathbf{W}_{0}:\mathbb{R}^{n}\to\mathbb{G}(n,m) is Lipschitzian and AnA\subseteq\mathbb{R}^{n} is Borel measurable. The following are equivalent.

  1. (A)

    n(A)=0\mathscr{L}^{n}(A)=0.

  2. (B)

    For n\mathscr{L}^{n} almost every xAx\in A, m(A(x+𝐖0(x)))=0\mathscr{H}^{m}\left(A\cap(x+\mathbf{W}_{0}(x))\right)=0.

  3. (C)

    For n\mathscr{L}^{n} almost every xnx\in\mathbb{R}^{n}, m(A(x+𝐖0(x)))=0\mathscr{H}^{m}\left(A\cap(x+\mathbf{W}_{0}(x))\right)=0.

As should be apparent from the discussion above, one difficulty stands with the fact that the affine mm-planes 𝐖(x)=x+𝐖0(x)\mathbf{W}(x)=x+\mathbf{W}_{0}(x) may not be disjointed. Nevertheless, they locally are, in the following sense. Given x0Ax_{0}\in A there exist a neighborhood UU of x0x_{0} and Lipschitzian maps 𝐰1,,𝐰m:Un\mathbf{w}_{1},\ldots,\mathbf{w}_{m}:U\to\mathbb{R}^{n} so that 𝐰1(x),,𝐰m(x)\mathbf{w}_{1}(x),\ldots,\mathbf{w}_{m}(x) is an orthonomal frame spanning 𝐖0(x)\mathbf{W}_{0}(x), for xUx\in U, and the map Φ:(Vx0U)×mn:(ξ,t)ξ+i=1mti𝐰i(ξ)\Phi:(V_{x_{0}}\cap U)\times\mathbb{R}^{m}\to\mathbb{R}^{n}:(\xi,t)\mapsto\xi+\sum_{i=1}^{m}t_{i}\mathbf{w}_{i}(\xi) is a lipeomorphism of a neighborhood of x0x_{0} onto its image – here, Vx0=x0+VV_{x_{0}}=x_{0}+V and V𝔾(n,nm)V\in\mathbb{G}(n,n-m) is close to 𝐖0(x0)\mathbf{W}_{0}(x_{0})^{\perp}. This, and applications of Fubini’s theorem, yield (B)(A)(B)\Rightarrow(A) in the theorem above (see 8.2).

However, we aim at obtaining a quantative version of this result, that the change of variable just described does not seem to provide. A natural route is to reduce the problem to applying the coarea formula, by spreading out the 𝐖(x)\mathbf{W}(x)’s in a disjointed way, in a higher dimensional space – i.e., adding a variable u𝐖(x)u\in\mathbf{W}(x) to the given xnx\in\mathbb{R}^{n} and considering 𝐖(x)\mathbf{W}(x) as a fiber above the base space n\mathbb{R}^{n}. We thus define

Σ=n×n{(x,u):xE and u𝐖(x)},\Sigma=\mathbb{R}^{n}\times\mathbb{R}^{n}\cap\{(x,u):x\in E\text{ and }u\in\mathbf{W}(x)\},

where EnE\subseteq\mathbb{R}^{n} is Borel measurable. This set is (n+m)(n+m)-rectifiable, owing to the Lipschitz continuity of 𝐖\mathbf{W}. It is convenient to assume that n(E)<\mathscr{L}^{n}(E)<\infty, so that

ϕE(B)=Em(B𝐖(x))𝑑n(x),\phi_{E}(B)=\int_{E}\mathscr{H}^{m}\left(B\cap\mathbf{W}(x)\right)d\mathscr{L}^{n}(x), (2)

BnB\subseteq\mathbb{R}^{n}, is a locally finite Borel measure 3.16. Now, Σ\Sigma was precisely set up so that, for each xEx\in E,

m(Σπ21(B)π11{x})=m(B𝐖(x)),\mathscr{H}^{m}\left(\Sigma\cap\pi_{2}^{-1}(B)\cap\pi_{1}^{-1}\{x\}\right)=\mathscr{H}^{m}\left(B\cap\mathbf{W}(x)\right),

where π1\pi_{1} and π2\pi_{2} denote the projections of n×n\mathbb{R}^{n}\times\mathbb{R}^{n} to the xx and uu variable, respectively. Abbreviating ΣB=Σπ21(B)\Sigma_{B}=\Sigma\cap\pi_{2}^{-1}(B), the coarea formula yields

ϕE(B)=ΣBJΣπ1𝑑n+m.\phi_{E}(B)=\int_{\Sigma_{B}}J_{\Sigma}\pi_{1}d\mathscr{H}^{n+m}. (3)

A simple calculation 5.4 shows that JΣπ2>0J_{\Sigma}\pi_{2}>0, n+m\mathscr{H}^{n+m}-almost everywhere on ΣB\Sigma_{B}. Since also

ΣBJΣπ2𝑑n+m=Bm(ΣBπ21{u})𝑑n(u),\int_{\Sigma_{B}}J_{\Sigma}\pi_{2}d\mathscr{H}^{n+m}=\int_{B}\mathscr{H}^{m}\left(\Sigma_{B}\cap\pi_{2}^{-1}\{u\}\right)d\mathscr{L}^{n}(u), (4)

the implication (A)(C)(A)\Rightarrow(C) above should now be clear: Letting B=AB=A and E=𝐁(0,R)E=\mathbf{B}(0,R), one infers from hypothesis (A) and (4) that n+m(ΣB)=0\mathscr{H}^{n+m}(\Sigma_{B})=0, whence, ϕ𝐁(0,R)(A)=0\phi_{\mathbf{B}(0,R)}(A)=0, by (3), and, in turn, conclusion (C) ensues from (2).

In order to establish that (B)(A)(B)\Rightarrow(A) by using this new change of variable, we need to observe 5.2 that JΣπ1>0J_{\Sigma}\pi_{1}>0, n+m\mathscr{H}^{n+m}-almost everywhere, and, ideally, to show that m(ΣBπ21{u})>0\mathscr{H}^{m}\left(\Sigma_{B}\cap\pi_{2}^{-1}\{u\}\right)>0, for almost every uBu\in B. This last part involves some difficulty. To understand this, we let m=n1m=n-1, in order to keep the notations short. Now u𝐖(x)u\in\mathbf{W}(x) iff ux𝐖0(x)u-x\in\mathbf{W}_{0}(x) iff 𝐯0(x),xu=0\langle\mathbf{v}_{0}(x),x-u\rangle=0, where 𝐯0(x)𝐖0(x)\mathbf{v}_{0}(x)\in\mathbf{W}_{0}(x)^{\perp} is, say, a unit vector. Abbreviating gu(x)=𝐯0(x),xug_{u}(x)=\langle\mathbf{v}_{0}(x),x-u\rangle, we infer that

m(ΣBπ21{u})=m(Egu1{0}).\mathscr{H}^{m}\left(\Sigma_{B}\cap\pi_{2}^{-1}\{u\}\right)=\mathscr{H}^{m}\left(E\cap g_{u}^{-1}\{0\}\right).

The problem remains that two of the nonlinear mm sets Egu1{0}E\cap g_{u}^{-1}\{0\} and Egu1{0}E\cap g_{u^{\prime}}^{-1}\{0\} may intersect, thereby preventing another application of the coarea formula when looking out for their lower bound. Yet, we already know that

ϕE(B)=B𝒵E𝐖𝑑n,\phi_{E}(B)=\int_{B}\mathscr{Z}_{E}\mathbf{W}d\mathscr{L}^{n},

where 𝒵E𝐖\mathscr{Z}_{E}\mathbf{W} is a Radon-Nikodým derivative (see 5.6) and also that (𝒵E𝐖)(u)(\mathscr{Z}_{E}\mathbf{W})(u) is comparable to m(Egu1{0})\mathscr{H}^{m}\left(E\cap g_{u}^{-1}\{0\}\right) (see 5.8). Adding an extra variable yy to the fibered space Σ\Sigma, 5.10, we improve on this by showing that

(𝒵E𝐖)(u)𝜼(n,m)lim infjj1j1m(Egu1{y})𝑑1(y)=𝜼(n,m)(𝒴E𝐖)(u),(\mathscr{Z}_{E}\mathbf{W})(u)\geqslant\boldsymbol{\eta}(n,m)\liminf_{j}\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int_{-j^{-1}}^{j^{-1}}\mathscr{H}^{m}\left(E\cap g_{u}^{-1}\{y\}\right)d\mathscr{L}^{1}(y)=\boldsymbol{\eta}(n,m)(\mathscr{Y}_{E}\mathbf{W})(u),

where the last equality defines 𝒴E𝐖\mathscr{Y}_{E}\mathbf{W}, and 𝜼(n,m)>0\boldsymbol{\eta}(n,m)>0 is a lower bound for the coarea Jacobian factor of the restriction to the fibered space of the projection (x,u,y)(x,y)(x,u,y)\mapsto(x,y) (see 5.12 and 5.16). We are reduced to showing that 𝒴E𝐖>0\mathscr{Y}_{E}\mathbf{W}>0, almost everywhere. The reason why this holds is the following. Fix a Borel set ZnZ\subseteq\mathbb{R}^{n}, x0nx_{0}\in\mathbb{R}^{n} and r>0r>0. Let 𝐂𝐖(x0,r)\mathbf{C}_{\mathbf{W}}(x_{0},r) be the cylindrical box consisting of those xnx\in\mathbb{R}^{n} such that |P𝐖0(x0)(xx0)|r\left|P_{\mathbf{W}_{0}(x_{0})}(x-x_{0})\right|\leqslant r and |P𝐖0(x0)(xx0)|r\left|P_{\mathbf{W}_{0}(x_{0})^{\perp}}(x-x_{0})\right|\leqslant r. We want to find a lower bound for

Z𝐂𝐖(x0,r)(𝒴Z𝐂𝐖(x0,r))(u)𝑑n(u).\int_{Z\cap\mathbf{C}_{\mathbf{W}}(x_{0},r)}\left(\mathscr{Y}_{Z\cap\mathbf{C}_{\mathbf{W}}(x_{0},r)}\right)(u)d\mathscr{L}^{n}(u).

To this end, we fix z𝐖0(x0)𝐁(0,r)z\in\mathbf{W}_{0}(x_{0})\cap\mathbf{B}(0,r) and we let Vz=n{x0+z+s𝐯0(x0):rsr}V_{z}=\mathbb{R}^{n}\cap\{x_{0}+z+s\mathbf{v}_{0}(x_{0}):-r\leqslant s\leqslant r\} denote the corresponding vertical line segment. According to Fubini’s theorem we are reduced to estimating

Vz(𝒴Z𝐂𝐖(x0,r))(u)𝑑1(u).\int_{V_{z}}\left(\mathscr{Y}_{Z\cap\mathbf{C}_{\mathbf{W}}(x_{0},r)}\right)(u)d\mathscr{H}^{1}(u).

By virtue of Vitali’s covering theorem, we can find a disjointed family of line segments I1,I2,I_{1},I_{2},\ldots, covering almost all VzV_{z}, such that the above integral nearly equals

k1(Ik)Ikm(Z𝐂𝐖(x0,r)guk1{y})𝑑1(y)kZ𝐂𝐖(x0,r)guk1(Ik)|guk(x)|𝑑n(x)n(Z𝐂𝐖(x0,r)),\sum_{k}\mathscr{H}^{1}(I_{k})\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int_{I_{k}}\mathscr{H}^{m}\left(Z\cap\mathbf{C}_{\mathbf{W}}(x_{0},r)\cap g_{u_{k}}^{-1}\{y\}\right)d\mathscr{L}^{1}(y)\\ \cong\sum_{k}\int_{Z\cap\mathbf{C}_{\mathbf{W}}(x_{0},r)\cap g_{u_{k}}^{-1}(I_{k})}\left|\nabla g_{u_{k}}(x)\right|d\mathscr{L}^{n}(x)\cong\mathscr{L}^{n}\left(Z\cap\mathbf{C}_{\mathbf{W}}(x_{0},r)\right),

where the first near equality follows from the coarea formula, and the second one because |guk|1\left|\nabla g_{u_{k}}\right|\cong 1 at small scales (see 3.12) and the “nonlinear horizontal stripes” guk1(Ik)g_{u_{k}}^{-1}(I_{k}) are nearly pairwise disjoint. Verification of these claims takes up sections 6 and 7. Now, we reach a contradiction if Z=n{𝒴E𝐖=0}Z=\mathbb{R}^{n}\cap\{\mathscr{Y}_{E}\mathbf{W}=0\} is assumed to have n(Z)>0\mathscr{L}^{n}(Z)>0, and x0x_{0} is a point of density of ZZ.

Along this path, we have, in fact, achieved a quantitative version of these observations. Indeed, assuming that lim supr0+m(A𝐁(x,r)𝐖(x))𝜶(m)rm<𝜼(n,m)2m\limsup_{r\to 0^{+}}\frac{\mathscr{H}^{m}(A\cap\mathbf{B}(x,r)\cap\mathbf{W}(x))}{\boldsymbol{\alpha}(m)r^{m}}<\frac{\boldsymbol{\eta}(n,m)}{2^{m}}, for n\mathscr{L}^{n} positively many xAx\in A, we start by choosing EAE\subseteq A with the same property, whose diameter is small enough for the various estimates above to hold, and we refer to Egoroff’s theorem to select ε>0\varepsilon>0 so that the set

Z=E{lim supr0+m(A𝐁(x,r)𝐖(x))𝜶(m)rm<(1ε)𝜼(n,m)(2+ε)m}Z=E\cap\left\{\limsup_{r\to 0^{+}}\frac{\mathscr{H}^{m}(A\cap\mathbf{B}(x,r)\cap\mathbf{W}(x))}{\boldsymbol{\alpha}(m)r^{m}}<(1-\varepsilon)\frac{\boldsymbol{\eta}(n,m)}{(2+\varepsilon)^{m}}\right\}

has positive n\mathscr{L}^{n} measure. Thus, for 0<ε^ε0<\hat{\varepsilon}\ll\varepsilon, there exists x0Zx_{0}\in Z and there exists r>0r>0 as small as we please so that

(1ε^)m(𝐂𝐖(x0,r))n(Z𝐂𝐖(x0,r))Z𝐂𝐖(x0,r)𝒴Z𝐂𝐖(x0,r)𝐖(u)𝜶(m)rm𝑑n(u)1𝜼(n,m)Z𝐂𝐖(x0,r)m(Z𝐂𝐖(x0,r)𝐖(x))𝜶(m)rm𝑑n(x)1𝜼(n,m)Z𝐂𝐖(x0,r)m(Z𝐁(x,2r+o(r))𝐖(x))𝜶(m)rm𝑑n(x)1𝜼(n,m)Z𝐂𝐖(x0,r)(1ε)𝜼(n,m)(2+ε)m𝑑n(x)(1Cε)n(𝐂𝐖(x0,r)),\left(1-\hat{\varepsilon}\right)\mathscr{L}^{m}\left(\mathbf{C}_{\mathbf{W}}(x_{0},r)\right)\leqslant\mathscr{L}^{n}\left(Z\cap\mathbf{C}_{\mathbf{W}}(x_{0},r)\right)\lesssim\int_{Z\cap\mathbf{C}_{\mathbf{W}}(x_{0},r)}\frac{\mathscr{Y}_{Z\cap\mathbf{C}_{\mathbf{W}}(x_{0},r)}\mathbf{W}(u)}{\boldsymbol{\alpha}(m)r^{m}}d\mathscr{L}^{n}(u)\\ \lesssim\frac{1}{\boldsymbol{\eta}(n,m)}\int_{Z\cap\mathbf{C}_{\mathbf{W}}(x_{0},r)}\frac{\mathscr{H}^{m}(Z\cap\mathbf{C}_{\mathbf{W}}(x_{0},r)\cap\mathbf{W}(x))}{\boldsymbol{\alpha}(m)r^{m}}d\mathscr{L}^{n}(x)\\ \leqslant\frac{1}{\boldsymbol{\eta}(n,m)}\int_{Z\cap\mathbf{C}_{\mathbf{W}}(x_{0},r)}\frac{\mathscr{H}^{m}(Z\cap\mathbf{B}(x,2r+o(r))\cap\mathbf{W}(x))}{\boldsymbol{\alpha}(m)r^{m}}d\mathscr{L}^{n}(x)\\ \lesssim\frac{1}{\boldsymbol{\eta}(n,m)}\int_{Z\cap\mathbf{C}_{\mathbf{W}}(x_{0},r)}(1-\varepsilon)\frac{\boldsymbol{\eta}(n,m)}{(2+\varepsilon)^{m}}d\mathscr{L}^{n}(x)\leqslant(1-C\varepsilon)\mathscr{L}^{n}\left(\mathbf{C}_{\mathbf{W}}(x_{0},r)\right),

where the symbol \lesssim means that we loose a multiplicative factor λ>1\lambda>1 as close as we wish to 1, according to the scale rr. The above contradiction proves our main result 8.4, namely,

Theorem. —

Assume 𝐖0:n𝔾(n,m)\mathbf{W}_{0}:\mathbb{R}^{n}\to\mathbb{G}(n,m) is Lipschitzian and AnA\subseteq\mathbb{R}^{n} is Borel measurable. Abbreviating 𝐖(x)=x+𝐖0(x)\mathbf{W}(x)=x+\mathbf{W}_{0}(x), it follows that

lim supr0+m(A𝐁(x,r)𝐖(x))𝜶(m)rm12n,\limsup_{r\to 0^{+}}\frac{\mathscr{H}^{m}\big{(}A\cap\mathbf{B}(x,r)\cap\mathbf{W}(x)\big{)}}{\boldsymbol{\alpha}(m)r^{m}}\geqslant\frac{1}{2^{n}},

for n\mathscr{L}^{n}-almost every xAx\in A.

I extend my warm thanks to Jean-Christophe Léger for carefully reading an early version of the manuscript.

3. Preliminaries

3.1. —

In this paper, 1mn11\leqslant m\leqslant n-1 are integers. The ambient space is n\mathbb{R}^{n}. The canonical inner product of x,xnx,x^{\prime}\in\mathbb{R}^{n} is denoted x,x\langle x,x^{\prime}\rangle and the corresponding Euclidean norm of xx is |x||x|. If SnS\subseteq\mathbb{R}^{n}, we let (S)\mathscr{B}(S) be the σ\sigma-algebra of Borel measurable subsets of SS.

3.2Hausdorff measure. —

We let n\mathscr{L}^{n} be the Lebesgue outer measure in n\mathbb{R}^{n} and 𝜶(n)=n(𝐁(0,1))\boldsymbol{\alpha}(n)=\mathscr{L}^{n}(\mathbf{B}(0,1)). For SnS\subseteq\mathbb{R}^{n}, we abbreviate ζm(S)=𝜶(m)2m(diamS)m\zeta^{m}(S)=\boldsymbol{\alpha}(m)2^{-m}(\operatorname{\mathrm{diam}}S)^{m}. Given 0<δ0<\delta\leqslant\infty, we call δ\delta-cover of AnA\subseteq\mathbb{R}^{n} a finite or countably infinite family (Sj)jJ(S_{j})_{j\in J} of subsets of n\mathbb{R}^{n} such that AjJSjA\subseteq\cup_{j\in J}S_{j} and diamSjδ\operatorname{\mathrm{diam}}S_{j}\leqslant\delta, for every jJj\in J. We define

δm(A)=inf{jJζm(Sj):(Sj)jJ is a δ-cover of A }\mathscr{H}^{m}_{\delta}(A)=\inf\left\{\sum_{j\in J}\zeta^{m}(S_{j}):(S_{j})_{j\in J}\text{ is a $\delta$-cover of $A$ }\right\}

and m(A)=limδ0+δm(A)=supδ>0δm(A)\mathscr{H}^{m}(A)=\lim_{\delta\to 0^{+}}\mathscr{H}^{m}_{\delta}(A)=\sup_{\delta>0}\mathscr{H}^{m}_{\delta}(A). Thus, m\mathscr{H}^{m} is the mm-dimensional Hausdorff outer measure in n\mathbb{R}^{n}.

  1. (1)

    If (Kk)k(K_{k})_{k} is a sequence of nonempty compact subsets of n\mathbb{R}^{n} converging in Hausdorff distance to KK, then m(K)lim supkm(Kk)\mathscr{H}^{m}_{\infty}(K)\geqslant\limsup_{k}\mathscr{H}^{m}_{\infty}(K_{k}).

Given ε>0\varepsilon>0, choose an \infty-cover (Sj)j(S_{j})_{j\in\mathbb{N}} of KK such that m(K)+εjζm(Sj)\mathscr{H}^{m}_{\infty}(K)+\varepsilon\geqslant\sum_{j}\zeta^{m}(S_{j}). Since limr0+ζm(𝐔(Sj,r))=ζm(Sj)\lim_{r\to 0^{+}}\zeta^{m}(\mathbf{U}(S_{j},r))=\zeta^{m}(S_{j}), for each jj\in\mathbb{N}, we can choose an open set UjU_{j} containing SjS_{j} such that ζm(Uj)ε2j+ζm(Sj)\zeta^{m}(U_{j})\leqslant\varepsilon 2^{-j}+\zeta^{m}(S_{j}). Since U=jUjU=\cup_{j}U_{j} is open, there exists a positive integer k0k_{0} such that KkUK_{k}\subseteq U, whenever kk0k\geqslant k_{0}. Thus, in that case (Uj)j(U_{j})_{j} is an \infty-cover of KkK_{k} and, therefore, m(Kk)jζm(Uj)2ε+m(K)\mathscr{H}^{m}_{\infty}(K_{k})\leqslant\sum_{j}\zeta^{m}(U_{j})\leqslant 2\varepsilon+\mathscr{H}^{m}_{\infty}(K). Take the limit superior of the left hand side, as kk\to\infty, and then let ε0\varepsilon\to 0.

  1. (2)

    For all AmA\subseteq\mathbb{R}^{m}, one has m(A)=m(A)=m(A)\mathscr{L}^{m}(A)=\mathscr{H}^{m}(A)=\mathscr{H}^{m}_{\infty}(A).

It suffices to note that m(A)m(A)m(A)m(A)\mathscr{H}^{m}(A)\geqslant\mathscr{H}^{m}_{\infty}(A)\geqslant\mathscr{L}^{m}(A)\geqslant\mathscr{H}^{m}(A). The first inequality is trivial; the second one follows from the isodiametric inequality [9, 2.10.33]; the last one is a consequence of the Vitali covering theorem [6, Chapter 2 §2 Theorem 2].

  1. (3)

    If WnW\subseteq\mathbb{R}^{n} is an mm-dimensional affine subspace and AWA\subseteq W, then m(A)=m(A)\mathscr{H}^{m}(A)=\mathscr{H}^{m}_{\infty}(A).

Let m\mathfrak{H}^{m} denote the mm-dimensional Hausdorff outer measure in the metric space WW. In other words,

δm(A)=inf{jJζm(Sj):(Sj)jJ is a δ- cover of A and SjW, for all jJ},\mathfrak{H}^{m}_{\delta}(A)=\inf\left\{\sum_{j\in J}\zeta^{m}(S_{j}):(S_{j})_{j\in J}\text{ is a $\delta$- cover of $A$ and $S_{j}\subseteq W$, for all $j\in J$}\right\},

and m(A)=supδ>0δm(A)\mathfrak{H}^{m}(A)=\sup_{\delta>0}\mathfrak{H}^{m}_{\delta}(A). It is elementary to observe that m(A)=m(A)\mathfrak{H}^{m}(A)=\mathscr{H}^{m}(A) and that m(A)=m(A)\mathfrak{H}^{m}_{\infty}(A)=\mathscr{H}^{m}_{\infty}(A). Now, if f:Wmf:W\to\mathbb{R}^{m} is an isometry, then m(A)=m(A)=m(f(A))=m(f(A))=m(A)=m(A)\mathscr{H}^{m}(A)=\mathfrak{H}^{m}(A)=\mathscr{H}^{m}(f(A))=\mathscr{H}^{m}_{\infty}(f(A))=\mathfrak{H}^{m}_{\infty}(A)=\mathscr{H}^{m}_{\infty}(A), where the third equality follows from claim (2) above.

3.3Coarea formula. —

We recall two versions of the coarea formula. First, if AnA\subseteq\mathbb{R}^{n} is n\mathscr{L}^{n}-measurable and f:Anmf:A\to\mathbb{R}^{n-m} is Lipschitzian, then nm[0,]:ym(Af1{y})\mathbb{R}^{n-m}\to[0,\infty]:y\mapsto\mathscr{H}^{m}\left(A\cap f^{-1}\{y\}\right) is nm\mathscr{L}^{n-m}- measurable and

AJf(x)𝑑n(x)=nmm(Af1{y})𝑑nm(y).\int_{A}Jf(x)d\mathscr{L}^{n}(x)=\int_{\mathbb{R}^{n-m}}\mathscr{H}^{m}\left(A\cap f^{-1}\{y\}\right)d\mathscr{L}^{n-m}(y).

Here, the coarea Jacobian factor is well-defined n\mathscr{L}^{n}-almost everywhere, according to Rademacher’s theorem, and equals

Jf(x)=|det(Df(x)Df(x))|=nmDf(x);Jf(x)=\sqrt{\left|\det\left(Df(x)\circ Df(x)^{*}\right)\right|}=\left\|\wedge_{n-m}Df(x)\right\|;

see, for instance, [6, Chapter 3 §4].

Secondly, if ApA\subseteq\mathbb{R}^{p} is n\mathscr{H}^{n}-measurable and countably (n,n)(\mathscr{H}^{n},n)-rectifiable and if f:Anmf:A\to\mathbb{R}^{n-m} is Lipschitzian, then nm[0,]:ym(Af1{y})\mathbb{R}^{n-m}\to[0,\infty]:y\mapsto\mathscr{H}^{m}\left(A\cap f^{-1}\{y\}\right) is nm\mathscr{L}^{n-m}-measurable and

AJAf(x)𝑑n(x)=nmm(Af1{y})𝑑nm(y).\int_{A}J_{A}f(x)d\mathscr{H}^{n}(x)=\int_{\mathbb{R}^{n-m}}\mathscr{H}^{m}\left(A\cap f^{-1}\{y\}\right)d\mathscr{L}^{n-m}(y).

In order to state a formula for the coarea Jacobian factor JAf(x)J_{A}f(x) of ff relative to AA, we consider a point xAx\in A, where AA admits an approximate nn-dimensional tangent space TxAT_{x}A and ff is differentiable at xx along AA. Letting L:TxAnmL:T_{x}A\to\mathbb{R}^{n-m} denote the derivative of ff at xx, we have

JAf(x)=|det(LL)|=nmL;J_{A}f(x)=\sqrt{\left|\det\left(L\circ L^{*}\right)\right|}=\left\|\wedge_{n-m}L\right\|;

see, for instance, [9, 3.2.22].

In both cases, it is useful to recall the following. If L:VVL:V\to V^{\prime} is a linear map between two inner product spaces VV and VV^{\prime}, then

kL=sup{kL,ξ:ξkV and |ξ|=1}.\|\wedge_{k}L\|=\sup\left\{\langle\wedge_{k}L,\xi\rangle:\xi\in\wedge_{k}V\text{ and }|\xi|=1\right\}. (5)

On the one hand, kLLk\|\wedge_{k}L\|\leqslant\|L\|^{k} [9, 1.7.6] and LLipf\|L\|\leqslant\operatorname{\mathrm{Lip}}f, with LL as above. On the other hand, if v1,,vkv_{1},\ldots,v_{k} are linearly independent vectors of VV, then

kL|L(v1)L(vk)||v1vk|.\|\wedge_{k}L\|\geqslant\frac{|L(v_{1})\wedge\ldots\wedge L(v_{k})|}{|v_{1}\wedge\ldots\wedge v_{k}|}. (6)

Finally, we observe that both coarea formulæ hold true when ff is merely locally Lipschitzian, according to the monotone convergence theorem.

3.4Grassmannian. —

We let 𝔾(n,m)\mathbb{G}(n,m) be the set whose members are the mm-dimensional vector subspaces of n\mathbb{R}^{n}. With W𝔾(n,m)W\in\mathbb{G}(n,m), we associate PW:nnP_{W}:\mathbb{R}^{n}\to\mathbb{R}^{n}, the orthogonal projection on WW. We give 𝔾(n,m)\mathbb{G}(n,m) the structure of a compact metric space by letting d(W1,W2)=PW1PW2d(W_{1},W_{2})=\|P_{W_{1}}-P_{W_{2}}\|, where \|\cdot\| is the operator norm. If W𝔾(n,m)W\in\mathbb{G}(n,m), then W𝔾(n,nm)W^{\perp}\in\mathbb{G}(n,n-m) satisfies PW+PW=idnP_{W}+P_{W^{\perp}}=\operatorname{\mathrm{id}}_{\mathbb{R}^{n}}, therefore, 𝔾(n,m)𝔾(n,nm):WW\mathbb{G}(n,m)\to\mathbb{G}(n,n-m):W\mapsto W^{\perp} is an isometry. The bijective correspondence φ:𝔾(n,m)Hom(n,n):WPW\varphi:\mathbb{G}(n,m)\to\operatorname{\mathrm{Hom}}(\mathbb{R}^{n},\mathbb{R}^{n}):W\mapsto P_{W} identifies 𝔾(n,m)\mathbb{G}(n,m) with the submanifold Mn,m=Hom(n,n){L:LL=L,L=L, and traceL=m}M_{n,m}=\operatorname{\mathrm{Hom}}(\mathbb{R}^{n},\mathbb{R}^{n})\cap\{L:L\circ L=L\,,\,L^{*}=L\text{, and }\operatorname{\mathrm{trace}}L=m\}. There exists an open neighborhood VV of Mn,mM_{n,m} in Hom(n,n)\operatorname{\mathrm{Hom}}(\mathbb{R}^{n},\mathbb{R}^{n}) and a Lipschitzian retraction ρ:VMn,m\rho:V\to M_{n,m}, according, for instance, to [9, 3.1.20]. Therefore, if SnS\subseteq\mathbb{R}^{n} and 𝐖0:S𝔾(n,m)\mathbf{W}_{0}:S\to\mathbb{G}(n,m) is Lipschitzian, then there exist an open neighborhood UU of SS in n\mathbb{R}^{n} and a Lipschitzian extension 𝐖^0:U𝔾(n,m)\widehat{\mathbf{W}}_{0}:U\to\mathbb{G}(n,m) of 𝐖0\mathbf{W}_{0}. Indeed, φ𝐖0\varphi\circ\mathbf{W}_{0} admits a Lipschitzian extension Y:nHom(n,n)Y:\mathbb{R}^{n}\to\operatorname{\mathrm{Hom}}(\mathbb{R}^{n},\mathbb{R}^{n}) – see, e.g., [9, 2.10.43] – and it suffices to let U=Y1(V)U=Y^{-1}(V) and 𝐖^0=ρ(Y|U)\widehat{\mathbf{W}}_{0}=\rho\circ\left(Y|_{U}\right).

3.5Orthonormal frames. —

We let 𝕍(n,m)\mathbb{V}(n,m) be the set of orthonormal mm-frames in n\mathbb{R}^{n}, i.e., 𝕍(n,m)=(n)m{(w1,,wm): the family w1,,wm is orthonormal}\mathbb{V}(n,m)=(\mathbb{R}^{n})^{m}\cap\{(w_{1},\ldots,w_{m}):\text{ the family }w_{1},\ldots,w_{m}\text{ is orthonormal}\}. We will consider it as a metric space, with its structure inherited from (n)m(\mathbb{R}^{n})^{m}.

3.6. —

Let 𝒱𝔾(n,m)\mathscr{V}\subseteq\mathbb{G}(n,m) be a nonempty closed set such that diam𝒱<1\operatorname{\mathrm{diam}}\mathscr{V}<1. There exists a Lipschitzian map Ξ:𝒱𝕍(n,m)\Xi:\mathscr{V}\to\mathbb{V}(n,m) such that W=span{Ξ1(W),,Ξm(W)}W=\operatorname{\mathrm{span}}\{\Xi_{1}(W),\ldots,\Xi_{m}(W)\}, for every W𝒱W\in\mathscr{V}.

Proof.

Pick arbitrarily W0𝒱W_{0}\in\mathscr{V}. If W𝒱W\in\mathscr{V}, then the map W0W:wPW(w)W_{0}\to W:w\mapsto P_{W}(w) is bijective: If wW0{0}w\in W_{0}\setminus\{0\}, then |PW(w)w|=|PW(w)PW0(w)|<|w|\left|P_{W}(w)-w\right|=\left|P_{W}(w)-P_{W_{0}}(w)\right|<|w|, thus, PW(w)0P_{W}(w)\neq 0. Letting w1,,wmw_{1},\ldots,w_{m} be an arbitrary basis of W0W_{0}, it follows that, for each W𝒱W\in\mathscr{V}, the vectors wi(W)=PW(wi)w_{i}(W)=P_{W}(w_{i}), i=1,,mi=1,\ldots,m, constitute a basis of WW. Furthermore, the maps wi:𝒱nw_{i}:\mathscr{V}\to\mathbb{R}^{n} are Lipschitzian: |wi(W)wi(W)|=|PW(wi)PW(wi)|d(W,W)|wi|\left|w_{i}(W)-w_{i}(W^{\prime})\right|=\left|P_{W}(w_{i})-P_{W^{\prime}}(w_{i})\right|\leqslant d(W,W^{\prime})|w_{i}|. We apply the Gram-Schmidt process:

w¯1(W)=w1(W) and w¯i(W)=wi(W)j=1i1wi(W),w¯j(W)w¯j(W),i=2,,m,\overline{w}_{1}(W)=w_{1}(W)\quad\text{ and }\quad\overline{w}_{i}(W)=w_{i}(W)-\sum_{j=1}^{i-1}\langle w_{i}(W),\overline{w}_{j}(W)\rangle\overline{w}_{j}(W)\,,\,\,i=2,\ldots,m,

so that w¯1(W),,w¯m(W)\overline{w}_{1}(W),\ldots,\overline{w}_{m}(W) is readily an orthogonal basis of WW depending upon WW in a Lipschitzian way. Since each |w¯i|\left|\overline{w}_{i}\right| is bounded away from zero on 𝒱\mathscr{V}, the formula Ξi(W)=|w¯i(W)|1w¯i(W)\Xi_{i}(W)=\left|\overline{w}_{i}(W)\right|^{-1}\overline{w}_{i}(W), i=1,,mi=1,\ldots,m, defines Ξ\Xi with the required property. ∎

3.7. —

There exists a Borel measurable map Ξ:𝔾(n,m)𝕍(n,m)\Xi:\mathbb{G}(n,m)\to\mathbb{V}(n,m) with the property that W=span{Ξ1(W),,Ξm(W)}W=\operatorname{\mathrm{span}}\{\Xi_{1}(W),\ldots,\Xi_{m}(W)\}, for every W𝔾(n,m)W\in\mathbb{G}(n,m).

Proof.

Since 𝔾(n,m)\mathbb{G}(n,m) is compact, it can partitioned into finitely many Borel measurable sets 𝒱1,,𝒱J\mathscr{V}_{1},\ldots,\mathscr{V}_{J}, each having diameter bounded by 1/21/2. Define Ξ\Xi, piecewise, to coincide on 𝒱j\mathscr{V}_{j} with Ξj\Xi_{j} associated to Clos𝒱j\operatorname{\mathrm{Clos}}\mathscr{V}_{j} in 3.6, j=1,,Jj=1,\ldots,J. ∎

3.8. —

Assume SnS\subseteq\mathbb{R}^{n}, x0Sx_{0}\in S, and 𝐖0:S𝔾(n,m)\mathbf{W}_{0}:S\to\mathbb{G}(n,m) is Lipschitzian. There exist an open neighborhood UU of x0x_{0} in n\mathbb{R}^{n} and Lipschitzian maps 𝐰1,,𝐰m,𝐯1,,𝐯nm:Un\mathbf{w}_{1},\ldots,\mathbf{w}_{m},\mathbf{v}_{1},\ldots,\mathbf{v}_{n-m}:U\to\mathbb{R}^{n} with the following properties.

  1. (1)

    For every xUx\in U, the family 𝐰1(x),,𝐰m(x),𝐯1(x),,𝐯nm(x)\mathbf{w}_{1}(x),\ldots,\mathbf{w}_{m}(x),\mathbf{v}_{1}(x),\ldots,\mathbf{v}_{n-m}(x) is an orthonormal basis of n\mathbb{R}^{n}.

  2. (2)

    For every xSUx\in S\cap U, one has

    𝐖0(x)=span{𝐰1(x),,𝐰m(x)}\mathbf{W}_{0}(x)=\operatorname{\mathrm{span}}\{\mathbf{w}_{1}(x),\ldots,\mathbf{w}_{m}(x)\}

    and

    𝐖0(x)=span{𝐯1(x),,𝐯nm(x)}.\mathbf{W}_{0}(x)^{\perp}=\operatorname{\mathrm{span}}\{\mathbf{v}_{1}(x),\ldots,\mathbf{v}_{n-m}(x)\}.
Proof.

We let 𝐖^0:U^𝔾(n,m)\widehat{\mathbf{W}}_{0}:\widehat{U}\to\mathbb{G}(n,m) be a Lipschitzian extension of 𝐖0\mathbf{W}_{0}, where U^\widehat{U} is an open neighborhood of SS in n\mathbb{R}^{n} (recall 3.4). Abbreviate W0:=𝐖0(x0)W_{0}:=\mathbf{W}_{0}(x_{0}). Define 𝒱=𝔾(n,m){W:d(W,W0)<1/4}\mathscr{V}=\mathbb{G}(n,m)\cap\left\{W:d(W,W_{0})<1/4\right\} and U=U^𝐖^01(𝒱)U=\widehat{U}\cap\widehat{\mathbf{W}}_{0}^{-1}(\mathscr{V}). Apply 3.6 to Clos𝒱\operatorname{\mathrm{Clos}}\mathscr{V} and denote by Ξ\Xi the resulting Lipschitzian map 𝒱(n)m\mathscr{V}\to(\mathbb{R}^{n})^{m}. Next, define 𝒱=𝔾(n,nm){W:W𝒱}\mathscr{V}^{\perp}=\mathbb{G}(n,n-m)\cap\left\{W^{\perp}:W\in\mathscr{V}\right\}, apply 3.6 to Clos𝒱\operatorname{\mathrm{Clos}}\mathscr{V}^{\perp}, and denote by Ξ\Xi^{\perp} the resulting Lipschitzian map 𝒱(n)nm\mathscr{V}^{\perp}\to(\mathbb{R}^{n})^{n-m}. Letting 𝐰i(x)=(Ξi𝐖^0)(x)\mathbf{w}_{i}(x)=\left(\Xi_{i}\circ\widehat{\mathbf{W}}_{0}\right)(x), i=1,,mi=1,\ldots,m, and 𝐯i(x)=(Ξi𝐖^0)(x)\mathbf{v}_{i}(x)=\left(\Xi^{\perp}_{i}\circ\widehat{\mathbf{W}}_{0}\right)(x), i=1,,nmi=1,\ldots,n-m, completes the proof. ∎

3.9. —

Assume 𝐖0:n𝔾(n,m)\mathbf{W}_{0}:\mathbb{R}^{n}\to\mathbb{G}(n,m) is Borel measurable. There exist Borel measurable maps 𝐰1,,𝐰m,𝐯1,,𝐯nm:nn\mathbf{w}_{1},\ldots,\mathbf{w}_{m},\mathbf{v}_{1},\ldots,\mathbf{v}_{n-m}:\mathbb{R}^{n}\to\mathbb{R}^{n} with the following properties.

  1. (1)

    For every xnx\in\mathbb{R}^{n}, the family 𝐰1(x),,𝐰m(x),𝐯1(x),,𝐯nm(x)\mathbf{w}_{1}(x),\ldots,\mathbf{w}_{m}(x),\mathbf{v}_{1}(x),\ldots,\mathbf{v}_{n-m}(x) is an orthonormal basis of n\mathbb{R}^{n}.

  2. (2)

    For every xnx\in\mathbb{R}^{n}, one has

    𝐖0(x)=span{𝐰1(x),,𝐰m(x)}\mathbf{W}_{0}(x)=\operatorname{\mathrm{span}}\{\mathbf{w}_{1}(x),\ldots,\mathbf{w}_{m}(x)\}

    and

    𝐖0(x)=span{𝐯1(x),,𝐯nm(x)}.\mathbf{W}_{0}(x)^{\perp}=\operatorname{\mathrm{span}}\{\mathbf{v}_{1}(x),\ldots,\mathbf{v}_{n-m}(x)\}.
Proof.

Choose Ξ:𝔾(n,m)𝕍(n,m)\Xi:\mathbb{G}(n,m)\to\mathbb{V}(n,m) and Ξ:𝔾(n,nm)𝕍(n,nm)\Xi^{\perp}:\mathbb{G}(n,n-m)\to\mathbb{V}(n,n-m) as in 3.7. Letting (𝐰1(x),,𝐰m(x))=(Ξ𝐖0)(x)\left(\mathbf{w}_{1}(x),\ldots,\mathbf{w}_{m}(x)\right)=\left(\Xi\circ\mathbf{W}_{0}\right)(x) and (𝐯1(x),,𝐯nm(x))=(Ξ𝐖0)(x)\left(\mathbf{v}_{1}(x),\ldots,\mathbf{v}_{n-m}(x)\right)=\left(\Xi^{\perp}\circ\mathbf{W}_{0}^{\perp}\right)(x), for xnx\in\mathbb{R}^{n}, completes the proof. ∎

3.10Definition of 𝐖(x)\mathbf{W}(x). —

The typical situation that arises in the remaining part of this paper is that we are given a set SnS\subseteq\mathbb{R}^{n}, a Lipschitzian map 𝐖0:S𝔾(n,m)\mathbf{W}_{0}:S\to\mathbb{G}(n,m), and x0Sx_{0}\in S. We will represent 𝐖0(x)\mathbf{W}_{0}(x) and 𝐖0(x)\mathbf{W}_{0}^{\perp}(x) in a neighborhood UU of x0x_{0} as in 3.8. We will then further reduce the size of UU several times, in order that various conditions be met. With no exception, we will denote as 𝐖(x)=x+𝐖0(x)\mathbf{W}(x)=x+\mathbf{W}_{0}(x) the affine subspace containing xx, of direction 𝐖0(x)\mathbf{W}_{0}(x), whenever 𝐖0(x)\mathbf{W}_{0}(x) is defined.

3.11Definition of g𝐯1,,𝐯nm,ug_{\mathbf{v}_{1},\ldots,\mathbf{v}_{n-m},u} and lower bound of its coarea factor. —

Given an open set UnU\subseteq\mathbb{R}^{n}, a Lipschitzian map 𝐯:Un\mathbf{v}:U\to\mathbb{R}^{n}, and unu\in\mathbb{R}^{n}, we define g𝐯,u:Ug_{\mathbf{v},u}:U\to\mathbb{R} by the formula

g𝐯,u(x)=𝐯(x),xu.g_{\mathbf{v},u}(x)=\langle\mathbf{v}(x),x-u\rangle.

Clearly, g𝐯,ug_{\mathbf{v},u} is locally Lipschitzian. If 𝐯\mathbf{v} is differentiable at xUx\in U, then so is g𝐯,ug_{\mathbf{v},u} and, for every hnh\in\mathbb{R}^{n}, one has

Dg𝐯,u(x)(h)=g𝐯,u(x),h=D𝐯(x)(h),xu+𝐯(x),h.Dg_{\mathbf{v},u}(x)(h)=\langle\nabla g_{\mathbf{v},u}(x),h\rangle=\langle D\mathbf{v}(x)(h),x-u\rangle+\langle\mathbf{v}(x),h\rangle. (7)

Next, we assume that we are given Lipschitzian maps 𝐯1,,𝐯nm:Un\mathbf{v}_{1},\ldots,\mathbf{v}_{n-m}:U\to\mathbb{R}^{n}. We define g𝐯1,,𝐯nm,u:Unmg_{\mathbf{v}_{1},\ldots,\mathbf{v}_{n-m},u}:U\to\mathbb{R}^{n-m} by the formula

g𝐯1,,𝐯nm,u(x)=(g𝐯1,u(x),,g𝐯nm,u(x)).g_{\mathbf{v}_{1},\ldots,\mathbf{v}_{n-m},u}(x)=\left(g_{\mathbf{v}_{1},u}(x),\ldots,g_{\mathbf{v}_{n-m},u}(x)\right).

It is locally Lipschitzian as well. The relevance of g𝐯1,,𝐯nm,ug_{\mathbf{v}_{1},\ldots,\mathbf{v}_{n-m},u} stems from the following observation, assuming that 𝐯1,,𝐯nm\mathbf{v}_{1},\ldots,\mathbf{v}_{n-m} are associated with 𝐖0\mathbf{W}_{0} and 𝐖\mathbf{W} as in 3.8 and 3.10:

u𝐖(x)ux𝐖0(x)𝐯i(x),ux=0, for all i=1,,nmg𝐯1,,𝐯nm,u(x)=0xg𝐯1,,𝐯nm,u1{0}.\begin{split}u\in\mathbf{W}(x)&\iff u-x\in\mathbf{W}_{0}(x)\\ &\iff\langle\mathbf{v}_{i}(x),u-x\rangle=0\text{, for all }i=1,\ldots,n-m\\ &\iff g_{\mathbf{v}_{1},\ldots,\mathbf{v}_{n-m},u}(x)=0\\ &\iff x\in g_{\mathbf{v}_{1},\ldots,\mathbf{v}_{n-m},u}^{-1}\{0\}.\end{split} (8)

In fact, |g𝐯1,,𝐯nm,u(x)|=|P𝐖0(x)(xu)||g_{\mathbf{v}_{1},\ldots,\mathbf{v}_{n-m},u}(x)|=\left|P_{\mathbf{W}_{0}(x)^{\perp}}(x-u)\right|.

Abbreviate g=g𝐯1,,𝐯nm,ug=g_{\mathbf{v}_{1},\ldots,\mathbf{v}_{n-m},u}. If each 𝐯i\mathbf{v}_{i} is differentiable at xUx\in U, and hnh\in\mathbb{R}^{n}, then

Dg(x)(h)=i=1nmDg𝐯i,u(x)(h)ei,Dg(x)(h)=\sum_{i=1}^{n-m}Dg_{\mathbf{v}_{i},u}(x)(h)e_{i},

where, here and elsewhere, e1,,enme_{1},\ldots,e_{n-m} denotes the canonical basis of nm\mathbb{R}^{n-m}. Thus, if 𝐯1(x),,𝐯nm(x)\mathbf{v}_{1}(x),\ldots,\mathbf{v}_{n-m}(x) constitute an orthonormal family in n\mathbb{R}^{n}, then

Dg𝐯i,u(x)(𝐯j(x))=δi,j+εi,j(x,u),Dg_{\mathbf{v}_{i},u}(x)(\mathbf{v}_{j}(x))=\delta_{i,j}+\varepsilon_{i,j}(x,u),

where

|εi,j(x,u)|=|D𝐯i(x)(𝐯j(x)),xu|(Lip𝐯i)|xu|,|\varepsilon_{i,j}(x,u)|=\left|\langle D\mathbf{v}_{i}(x)(\mathbf{v}_{j}(x)),x-u\rangle\right|\leqslant\left(\operatorname{\mathrm{Lip}}\mathbf{v}_{i}\right)|x-u|, (9)

according to (7), and, in turn,

Dg(x)(𝐯j(x))=i=1nm(δi,j+εi,j(x,u))ei.Dg(x)(\mathbf{v}_{j}(x))=\sum_{i=1}^{n-m}\left(\delta_{i,j}+\varepsilon_{i,j}(x,u)\right)e_{i}.

This allows for a lower bound of the coarea factor of gg at xx as follows.

nmDg(x)|Dg(x)(𝐯1(x))Dg(x)(𝐯nm(x))|=|(i=1nm(δi,1+εi,1(x,u))ei)(i=1nm(δi,nm+εi,nm(x,u))ei)|=|det(δi,j+εi,j(x,u))i,j=1,,nm|.\begin{split}\left\|\wedge_{n-m}Dg(x)\right\|&\geqslant\left|Dg(x)(\mathbf{v}_{1}(x))\wedge\ldots\wedge Dg(x)(\mathbf{v}_{n-m}(x))\right|\\ &=\left|\left(\sum_{i=1}^{n-m}\left(\delta_{i,1}+\varepsilon_{i,1}(x,u)\right)e_{i}\right)\wedge\ldots\wedge\left(\sum_{i=1}^{n-m}\left(\delta_{i,n-m}+\varepsilon_{i,n-m}(x,u)\right)e_{i}\right)\right|\\ &=\left|\det\left(\delta_{i,j}+\varepsilon_{i,j}(x,u)\right)_{i,j=1,\ldots,n-m}\right|.\end{split}

In view of (9), we obtain the next lemma.

3.12. —

Given Λ>0\Lambda>0 and 0<ε<10<\varepsilon<1, there exists 𝛅3.12(n,Λ,ε)>0\boldsymbol{\delta}_{3.12}(n,\Lambda,\varepsilon)>0 with the following property. Assume that

  1. (1)

    UnU\subseteq\mathbb{R}^{n} is open and unu\in\mathbb{R}^{n};

  2. (2)

    𝐯1,,𝐯nm:Un\mathbf{v}_{1},\ldots,\mathbf{v}_{n-m}:U\to\mathbb{R}^{n} are Lipschitzian;

  3. (3)

    𝐯1(x),,𝐯nm(x)\mathbf{v}_{1}(x),\ldots,\mathbf{v}_{n-m}(x) is an orthonormal family, for every xUx\in U.

If

  1. (4)

    Lip𝐯iΛ\operatorname{\mathrm{Lip}}\mathbf{v}_{i}\leqslant\Lambda, for each i=1,,nmi=1,\ldots,n-m;

  2. (5)

    diam(U{u})𝜹3.12(n,Λ,ε)\operatorname{\mathrm{diam}}\left(U\cup\{u\}\right)\leqslant\boldsymbol{\delta}_{3.12}(n,\Lambda,\varepsilon);

then

Jg𝐯1,,𝐯nm,u(x)1ε,Jg_{\mathbf{v}_{1},\ldots,\mathbf{v}_{n-m},u}(x)\geqslant 1-\varepsilon,

at n\mathscr{L}^{n}-almost every xUx\in U.

3.13Definition of πu\pi_{u} and its relation with g𝐯1,,𝐯nm,ug_{\mathbf{v}_{1},\ldots,\mathbf{v}_{n-m},u}. —

With unu\in\mathbb{R}^{n}, we associate

πu:𝕍(n,nm)×nnm:(ξ1,,ξnm,x)(ξ1,xu,,ξnm,xu).\pi_{u}:\mathbb{V}(n,n-m)\times\mathbb{R}^{n}\to\mathbb{R}^{n-m}:(\xi_{1},\ldots,\xi_{n-m},x)\mapsto\left(\langle\xi_{1},x-u\rangle,\ldots,\langle\xi_{n-m},x-u\rangle\right).

When (ξ1,,ξnm)𝕍(n,nm)(\xi_{1},\ldots,\xi_{n-m})\in\mathbb{V}(n,n-m) is fixed, we also abbreviate as πξ1,,ξnm,u\pi_{\xi_{1},\ldots,\xi_{n-m},u} the map nnm\mathbb{R}^{n}\to\mathbb{R}^{n-m} defined by πξ1,,ξnm,u(x)=πu(ξ1,,ξnm,x)\pi_{\xi_{1},\ldots,\xi_{n-m},u}(x)=\pi_{u}(\xi_{1},\ldots,\xi_{n-m},x). It is then rather useful to observe that, in the context described in 3.8 and 3.10, the following holds:

π𝐯1(x),,𝐯nm(x),u1{g𝐯1,,𝐯nm,u(x)}=𝐖(x).\pi_{\mathbf{v}_{1}(x),\ldots,\mathbf{v}_{n-m}(x),u}^{-1}\left\{g_{\mathbf{v}_{1},\ldots,\mathbf{v}_{n-m},u}(x)\right\}=\mathbf{W}(x). (10)

Indeed,

h𝐖(x)hx𝐖0(x)𝐯i(x),hx=0, for all i=1,,nm𝐯i(x),hu=𝐯i(x),xu, for all i=1,,nmπ𝐯1(x),,𝐯nm(x),u(h)=g𝐯1,,𝐯nm,u(x).\begin{split}h\in\mathbf{W}(x)&\iff h-x\in\mathbf{W}_{0}(x)\\ &\iff\langle\mathbf{v}_{i}(x),h-x\rangle=0\text{, for all }i=1,\ldots,n-m\\ &\iff\langle\mathbf{v}_{i}(x),h-u\rangle=\langle\mathbf{v}_{i}(x),x-u\rangle\text{, for all }i=1,\ldots,n-m\\ &\iff\pi_{\mathbf{v}_{1}(x),\ldots,\mathbf{v}_{n-m}(x),u}(h)=g_{\mathbf{v}_{1},\ldots,\mathbf{v}_{n-m},u}(x).\end{split}

In the sequel, we will sometimes abbreviate ξ=(ξ1,,ξnm)𝕍(n,nm)\xi=(\xi_{1},\ldots,\xi_{n-m})\in\mathbb{V}(n,n-m). It also helps to notice that, for given ξ𝕍(n,nm)\xi\in\mathbb{V}(n,n-m) and ynmy\in\mathbb{R}^{n-m}, the set πξ,u1{y}\pi_{\xi,u}^{-1}\{y\} is an mm-dimensional affine subspace of n\mathbb{R}^{n}.

3.14. —

If B(n)B\in\mathscr{B}(\mathbb{R}^{n}) and unu\in\mathbb{R}^{n}, then

hB:𝕍(n,nm)×nm[0,]:(ξ,y)m(Bπξ,u1{y})h_{B}:\mathbb{V}(n,n-m)\times\mathbb{R}^{n-m}\to[0,\infty]:(\xi,y)\mapsto\mathscr{H}^{m}\left(B\cap\pi_{\xi,u}^{-1}\{y\}\right)

is Borel measurable.

Proof.

We start by showing that, when BB is compact, hBh_{B} is upper semicontinuous. Thus, if ((ξk,yk))k((\xi_{k},y_{k}))_{k} is a sequence in 𝕍(n,nm)×nm\mathbb{V}(n,n-m)\times\mathbb{R}^{n-m} that converges to (ξ,y)(\xi,y), we ought to show that

m(K)lim supkm(Kk),\mathscr{H}^{m}_{\infty}(K)\geqslant\limsup_{k}\mathscr{H}^{m}_{\infty}(K_{k}), (11)

where K=Bπξ,u1{y}K=B\cap\pi_{\xi,u}^{-1}\{y\} and Kk=Bπξk,u1{yk}K_{k}=B\cap\pi_{\xi_{k},u}^{-1}\{y_{k}\}. This is indeed equivalent to the same inequality with m\mathscr{H}^{m}_{\infty} replaced by m\mathscr{H}^{m}, according to 3.2(3) and the last sentence of 3.13. Considering, if necessary, a subsequence of (Kk)k(K_{k})_{k} we may assume that none of the compact sets KkK_{k} is empty and that the limit superior in (11) is a limit. Since the set of nonempty compact subsets of the compact set BB, equipped with the Hausdorff metric, is compact, the sequence (Kk)k(K_{k})_{k} admits a subsequence (denoted the same way) converging to a compact set LBL\subseteq B. Given zLz\in L, there are zkKkz_{k}\in K_{k} converging to zz. Thus, πu(ξ,z)=limkπu(ξk,zk)=limkyk=y\pi_{u}(\xi,z)=\lim_{k}\pi_{u}(\xi_{k},z_{k})=\lim_{k}y_{k}=y. In other words, zKz\in K. Therefore, m(K)m(L)\mathscr{H}^{m}_{\infty}(K)\geqslant\mathscr{H}^{m}_{\infty}(L), so that (11) follows from 3.2(1).

Next, we fix r>0r>0 and we abbreviate 𝒜=(n){B:hB𝐁(0,r) is Borel measurable}\mathscr{A}=\mathscr{B}(\mathbb{R}^{n})\cap\{B:h_{B\cap\mathbf{B}(0,r)}\text{ is Borel measurable}\}. Thus, we have just shown that 𝒜\mathscr{A} contains the collection (n)\mathscr{F}(\mathbb{R}^{n}) of all closed subsets of n\mathbb{R}^{n}. Observe that if (Bj)j(B_{j})_{j} is an nondecreasing sequence in 𝒜\mathscr{A} and B=jBjB=\cup_{j}B_{j}, then hB𝐁(0,r)=limjhBj𝐁(0,r)h_{B\cap\mathbf{B}(0,r)}=\lim_{j}h_{B_{j}\cap\mathbf{B}(0,r)} pointwise, thus, B𝒜B\in\mathscr{A}. In particular, n𝒜\mathbb{R}^{n}\in\mathscr{A}. Furthermore, if B,B𝒜B,B^{\prime}\in\mathscr{A} and BBB^{\prime}\subseteq B, then hBB=hBhBh_{B\setminus B^{\prime}}=h_{B}-h_{B^{\prime}}, because all measures involved are finite; indeed, hB𝐁(0,r)(ξ,y)𝜶(m)rmh_{B\cap\mathbf{B}(0,r)}(\xi,y)\leqslant\boldsymbol{\alpha}(m)r^{m}, for all (ξ,y)(\xi,y). Accordingly, BB𝒜B\setminus B^{\prime}\in\mathscr{A}. This means that 𝒜\mathscr{A} is a Dynkin class. Since (n)\mathscr{F}(\mathbb{R}^{n}) is a π\pi-system, 𝒜\mathscr{A} contains the σ\sigma-algebra generated by (n)\mathscr{F}(\mathbb{R}^{n}), i.e., (n)\mathscr{B}(\mathbb{R}^{n}) [3, Theorem 1.6.2]. Finally, if B(n)B\in\mathscr{B}(\mathbb{R}^{n}), then hB=limjhB𝐁(0,j)h_{B}=\lim_{j}h_{B\cap\mathbf{B}(0,j)} pointwise, whence hBh_{B} is Borel measurable. ∎

3.15. —

Assume B(n)B\in\mathscr{B}(\mathbb{R}^{n}) and 𝐖0:n𝔾(n,m)\mathbf{W}_{0}:\mathbb{R}^{n}\to\mathbb{G}(n,m) is Borel measurable. The following function is Borel measurable.

n[0,]:xm(B𝐖(x)).\mathbb{R}^{n}\to[0,\infty]:x\mapsto\mathscr{H}^{m}\left(B\cap\mathbf{W}(x)\right).
Proof.

Let h𝐖,Bh_{\mathbf{W},B} denote this function. Let 𝐯1,,𝐯nm:nn\mathbf{v}_{1},\ldots,\mathbf{v}_{n-m}:\mathbb{R}^{n}\to\mathbb{R}^{n} be Borel measurable maps associated with 𝐖0\mathbf{W}_{0} in 3.9. Fix unu\in\mathbb{R}^{n} arbitrarily. Define

Υ:n𝕍(n,nm)×nm:x(𝐯1(x),,𝐯nm(x),g𝐯1(x),,𝐯nm(x),u(x))\Upsilon:\mathbb{R}^{n}\to\mathbb{V}(n,n-m)\times\mathbb{R}^{n-m}:x\mapsto\left(\mathbf{v}_{1}(x),\ldots,\mathbf{v}_{n-m}(x),g_{\mathbf{v}_{1}(x),\ldots,\mathbf{v}_{n-m}(x),u}(x)\right)

and notice that

h𝐖,B=hBΥh_{\mathbf{W},B}=h_{B}\circ\Upsilon

(where hBh_{B} is the function associated with BB and uu in 3.14), according to (10). One notes that Υ\Upsilon is Borel measurable; whence, the conclusion ensues from 3.14. ∎

3.16Definition of ϕE,𝐖\phi_{E,\mathbf{W}}. —

Let 𝐖0:n𝔾(n,m)\mathbf{W}_{0}:\mathbb{R}^{n}\to\mathbb{G}(n,m) be Borel measurable and E(n)E\in\mathscr{B}(\mathbb{R}^{n}) be such that n(E)<\mathscr{L}^{n}(E)<\infty. For each B(n)B\in\mathscr{B}(\mathbb{R}^{n}), we define

ϕE,𝐖(B)=Em(B𝐖(x))𝑑n(x).\phi_{E,\mathbf{W}}(B)=\int_{E}\mathscr{H}^{m}\left(B\cap\mathbf{W}(x)\right)d\mathscr{L}^{n}(x).

This is well-defined, according to 3.15. It is easy to check that ϕE,𝐖\phi_{E,\mathbf{W}} is a locally finite – hence, σ\sigma-finite – Borel measure on n\mathbb{R}^{n}; indeed, ϕE,𝐖(B)𝜶(m)(diamB)mn(E)\phi_{E,\mathbf{W}}(B)\leqslant\boldsymbol{\alpha}(m)(\operatorname{\mathrm{diam}}B)^{m}\mathscr{L}^{n}(E).

To close this section, we discuss the relevance of ϕE,𝐖\phi_{E,\mathbf{W}} to the problem of existence of “nearly Nikodým sets”.

3.17Definition of Nearly Nikodým set. —

Let E(n)E\in\mathscr{B}(\mathbb{R}^{n}). We say that B(E)B\in\mathscr{B}(E) is nearly mm-Nikodým in EE if

  1. (1)

    n(B)>0\mathscr{L}^{n}(B)>0;

  2. (2)

    For n\mathscr{L}^{n}-almost each xEx\in E, there is W𝔾(n,m)W\in\mathbb{G}(n,m) such that m(B(x+W))=0\mathscr{H}^{m}\left(B\cap(x+W)\right)=0.

In case n=2n=2, m=1m=1, and E=[0,1]×[0,1]E=[0,1]\times[0,1], the existence of such a BB (with 2(B)=1\mathscr{L}^{2}(B)=1) was established by Nikodým [12], see also [4, Chapter 8]. For arbitrary n2n\geqslant 2 and m=n1m=n-1, the existence of such a BB was established by Falconer [7]. In fact, in both cases, these authors established the stronger condition that, for every xBx\in B, m(B(x+W))=0\mathscr{H}^{m}(B\cap(x+W))=0 can be replaced by B(x+W)={x}B\cap(x+W)=\{x\}. Thus, in case 1m<n11\leqslant m<n-1, letting BB be a set exhibited by Falconer, if xBx\in B and W𝔾(n,n1)W\subseteq\mathbb{G}(n,n-1) is such that B(x+W)={x}B\cap(x+W)=\{x\}, then picking arbitrarily V𝔾(n,m)V\in\mathbb{G}(n,m) such that VWV\subseteq W, we see that B(x+V)={x}B\cap(x+V)=\{x\}. Whence, BB is also nearly mm-Nikodým in BB.

Assuming that 𝐖0:E𝔾(n,m)\mathbf{W}_{0}:E\to\mathbb{G}(n,m) is Borel measurable, we say that B(E)B\in\mathscr{B}(E) is nearly mm-Nikodým in EE relative to 𝐖\mathbf{W} if

  1. (1)

    n(B)>0\mathscr{L}^{n}(B)>0;

  2. (2)

    For n\mathscr{L}^{n}-almost each xEx\in E, one has m(B𝐖(x))=0\mathscr{H}^{m}\left(B\cap\mathbf{W}(x)\right)=0.

3.18. —

Let E(n)E\in\mathscr{B}(\mathbb{R}^{n}) have finite n\mathscr{L}^{n} measure and 𝐖0:n𝔾(n,m)\mathbf{W}_{0}:\mathbb{R}^{n}\to\mathbb{G}(n,m) be Borel measurable. The following are equivalent.

  1. (1)

    n|(E)\mathscr{L}^{n}|_{\mathscr{B}(E)} is absolutely continuous with respect to ϕE,𝐖|(E)\phi_{E,\mathbf{W}}|_{\mathscr{B}(E)}.

  2. (2)

    There does not exist a nearly mm-Nikodým set in EE relative to 𝐖\mathbf{W}.

Proof.

A set B(E)B\in\mathscr{B}(E) such that ϕE,𝐖(B)=0\phi_{E,\mathbf{W}}(B)=0 and n(B)>0\mathscr{L}^{n}(B)>0 is, by definition, a nearly mm-Nikodým set relative to 𝐖\mathbf{W}. Condition (1) is equivalent to their nonexistence. ∎

3.19. —

Assume that E(n)E\in\mathscr{B}(\mathbb{R}^{n}) and that B(E)B\in\mathscr{B}(E) is nearly mm-Nikodým. The following hold.

  1. (1)

    There exists 𝐖0:n𝔾(n,m)\mathbf{W}_{0}:\mathbb{R}^{n}\to\mathbb{G}(n,m) Borel measurable such that BB is nearly mm- Nikodým in EE relative to 𝐖\mathbf{W}.

  2. (2)

    There exists CBC\subseteq B compact and 𝐖¯0:n𝔾(n,m)\overline{\mathbf{W}}_{0}:\mathbb{R}^{n}\to\mathbb{G}(n,m) continuous such that CC is nearly mm-Nikodým in CC relative to 𝐖¯\overline{\mathbf{W}}.

Proof.

Define a Borel measurable map 𝝃:𝔾(n,m)𝕍(n,nm)\boldsymbol{\xi}:\mathbb{G}(n,m)\to\mathbb{V}(n,n-m) by 𝝃(W)=Ξ(W)\boldsymbol{\xi}(W)=\Xi\left(W^{\perp}\right), where Ξ:𝔾(n,nm)𝕍(n,nm)\Xi:\mathbb{G}(n,n-m)\to\mathbb{V}(n,n-m) is as in 3.7. Choose arbitrarily unu\in\mathbb{R}^{n} and define a Borel measurable map

Υ:E×𝔾(n,m)𝕍(n,nm)×nm(x,W)(𝝃(W),𝝃1(W),xu,,𝝃nm(W),xu).\Upsilon:E\times\mathbb{G}(n,m)\to\mathbb{V}(n,n-m)\times\mathbb{R}^{n-m}\\ (x,W)\mapsto\left(\boldsymbol{\xi}(W),\langle\boldsymbol{\xi}_{1}(W),x-u\rangle,\ldots,\langle\boldsymbol{\xi}_{n-m}(W),x-u\rangle\right).

Similarly to (10), observe that

x+W=π𝝃(W),u1{(𝝃1(W),xu,,𝝃nm(W),xu)},x+W=\pi_{\boldsymbol{\xi}(W),u}^{-1}\left\{\left(\langle\boldsymbol{\xi}_{1}(W),x-u\rangle,\ldots,\langle\boldsymbol{\xi}_{n-m}(W),x-u\rangle\right)\right\},

for every (x,W)E×𝔾(n,m)(x,W)\in E\times\mathbb{G}(n,m). We infer from 3.14 that

hBΥ:E×𝔾(n,m)[0,]:(x,W)m(B(x+W))h_{B}\circ\Upsilon:E\times\mathbb{G}(n,m)\to[0,\infty]:(x,W)\mapsto\mathscr{H}^{m}\left(B\cap(x+W)\right)

is Borel measurable. Thus, the set

=E×𝔾(n,m){(x,W):m(B(x+W))=0}\mathscr{E}=E\times\mathbb{G}(n,m)\cap\left\{(x,W):\mathscr{H}^{m}\left(B\cap(x+W)\right)=0\right\}

is Borel measurable as well. The set N=E{x:x=}N=E\cap\{x:\mathscr{E}_{x}=\emptyset\} is coanalytic and n(N)=0\mathscr{L}^{n}(N)=0, by assumption. By virtue of von Neumann’s selection theorem [13, 5.5.3], there exists a universally measurable map 𝐖~0:EN𝔾(n,m)\tilde{\mathbf{W}}_{0}:E\setminus N\to\mathbb{G}(n,m) such that 𝐖~0(x)x\tilde{\mathbf{W}}_{0}(x)\in\mathscr{E}_{x}, for every xENx\in E\setminus N, i.e., m(B(x+𝐖~0(x)))=0\mathscr{H}^{m}\left(B\cap\left(x+\tilde{\mathbf{W}}_{0}(x)\right)\right)=0. We extend 𝐖~0\tilde{\mathbf{W}}_{0} to be an arbitrary constant on N(nE)N\cup(\mathbb{R}^{n}\setminus E). This makes 𝐖~0\tilde{\mathbf{W}}_{0} an n\mathscr{L}^{n}-measurable map defined on n\mathbb{R}^{n}. Therefore, it is equal n\mathscr{L}^{n}-almost everywhere to a Borel measurable map 𝐖0:n𝔾(n,m)\mathbf{W}_{0}:\mathbb{R}^{n}\to\mathbb{G}(n,m). This proves (1).

In order to prove (2), we recall 3.4, specifically, the retraction ρ:VMn,m\rho:V\to M_{n,m} and the homeomorphic identification φ:𝔾(n,m)Mn,m\varphi:\mathbb{G}(n,m)\to M_{n,m}. Owing to the compactness of Mn,mM_{n,m}, there are finitely many open balls UjU_{j}, j=1,,Jj=1,\ldots,J, whose closures are contained in VV and whose union contains Mn,mM_{n,m}. Since n(B)>0\mathscr{L}^{n}(B)>0, there exists j{1,,J}j\in\{1,\ldots,J\} such that n(BEj)>0\mathscr{L}^{n}\left(B\cap E_{j}\right)>0, where Ej=(φ𝐖0)1(Uj)E_{j}=\left(\varphi\circ\mathbf{W}_{0}\right)^{-1}(U_{j}). It follows from Lusin’s theorem [9, 2.5.3] that there exists a compact set CBEjC\subseteq B\cap E_{j} such that n(C)>0\mathscr{L}^{n}(C)>0 and the restriction 𝐖0|C\mathbf{W}_{0}|_{C} is continuous. The map φ𝐖0|C\varphi\circ\mathbf{W}_{0}|_{C} takes its values in the closed ball ClosUj\operatorname{\mathrm{Clos}}U_{j}, therefore, it admits a continuous extension Y0:nClosUjVY_{0}:\mathbb{R}^{n}\to\operatorname{\mathrm{Clos}}U_{j}\subseteq V. Letting 𝐖¯0=φ1ρY0\overline{\mathbf{W}}_{0}=\varphi^{-1}\circ\rho\circ Y_{0} completes the proof. ∎

4. Common setting

4.1Setting for the next three sections. —

In the next three sections, we shall assume the following.

  1. (1)

    EnE\subseteq\mathbb{R}^{n} is Borel measurable and n(E)<\mathscr{L}^{n}(E)<\infty.

  2. (2)

    UnU\subseteq\mathbb{R}^{n} is open and EUE\subseteq U.

  3. (3)

    BnB\subseteq\mathbb{R}^{n} is Borel measurable.

  4. (4)

    𝐖0:U𝔾(n,m)\mathbf{W}_{0}:U\to\mathbb{G}(n,m) is Lipschitzian.

  5. (5)

    𝐖(x)=x+𝐖0(x)\mathbf{W}(x)=x+\mathbf{W}_{0}(x), for each xUx\in U.

  6. (6)

    Λ>0\Lambda>0.

  7. (7)

    𝐰1,,𝐰m:Un\mathbf{w}_{1},\ldots,\mathbf{w}_{m}:U\to\mathbb{R}^{n} and Lip𝐰iΛ\operatorname{\mathrm{Lip}}\mathbf{w}_{i}\leqslant\Lambda, i=1,,mi=1,\ldots,m.

  8. (8)

    𝐯1,,𝐯nm:Un\mathbf{v}_{1},\ldots,\mathbf{v}_{n-m}:U\to\mathbb{R}^{n} and Lip𝐯iΛ\operatorname{\mathrm{Lip}}\mathbf{v}_{i}\leqslant\Lambda, i=1,,nmi=1,\ldots,n-m.

  9. (9)

    𝐖0(x)=span{𝐰1(x),,𝐰m(x)}\mathbf{W}_{0}(x)=\operatorname{\mathrm{span}}\{\mathbf{w}_{1}(x),\ldots,\mathbf{w}_{m}(x)\}, for every xUx\in U.

  10. (10)

    𝐖0(x)=span{𝐯1(x),,𝐯nm(x)}\mathbf{W}_{0}(x)^{\perp}=\operatorname{\mathrm{span}}\{\mathbf{v}_{1}(x),\ldots,\mathbf{v}_{n-m}(x)\}, for every xUx\in U.

  11. (11)

    𝐰1(x),,𝐰m(x),𝐯1(x),,𝐯nm(x)\mathbf{w}_{1}(x),\ldots,\mathbf{w}_{m}(x),\mathbf{v}_{1}(x),\ldots,\mathbf{v}_{n-m}(x) constitute an orthonormal basis of n\mathbb{R}^{n}, for every xUx\in U.

5. Two fibrations

5.1A fibered space associated with E,B,𝐰1,,𝐰mE,B,\mathbf{w}_{1},\ldots,\mathbf{w}_{m}. —

We define

F:E×mn×n:(x,t1,,tm)(x,x+i=1mti𝐰i(x))F:E\times\mathbb{R}^{m}\to\mathbb{R}^{n}\times\mathbb{R}^{n}:(x,t_{1},\ldots,t_{m})\mapsto\left(x,x+\sum_{i=1}^{m}t_{i}\mathbf{w}_{i}(x)\right)

as well as

Σ=F(E×m)=n×n{(x,u):xE and u𝐖(x)}.\Sigma=F(E\times\mathbb{R}^{m})=\mathbb{R}^{n}\times\mathbb{R}^{n}\cap\left\{(x,u):x\in E\text{ and }u\in\mathbf{W}(x)\right\}.

We will oftentimes abbreviate (x,t)=(x,t1,,tm)(x,t)=(x,t_{1},\ldots,t_{m}). It is obvious that FF is locally Lipschitzian and, therefore, Σ\Sigma is countably (n+m)(n+m)-rectifiable and n+m\mathscr{H}^{n+m}-measurable. We also consider the two canonical projections

π1:n×nn:(x,u)x and π2:n×nn:(x,u)u,\pi_{1}:\mathbb{R}^{n}\times\mathbb{R}^{n}\to\mathbb{R}^{n}:(x,u)\mapsto x\quad\text{ and }\quad\pi_{2}:\mathbb{R}^{n}\times\mathbb{R}^{n}\to\mathbb{R}^{n}:(x,u)\mapsto u,

as well as the set

ΣB=Σπ21(B)=n×n{(x,u):xE and uB𝐖(x)},\Sigma_{B}=\Sigma\cap\pi_{2}^{-1}(B)=\mathbb{R}^{n}\times\mathbb{R}^{n}\cap\left\{(x,u):x\in E\text{ and }u\in B\cap\mathbf{W}(x)\right\},

which also is, clearly, countably (n+m)(n+m)-rectifiable and n+m\mathscr{H}^{n+m}-measurable. With the prospect of applying the coarea formula to ΣB\Sigma_{B} and π1\pi_{1}, and to ΣB\Sigma_{B} and π2\pi_{2}, respectively, we observe that, for each fixed xEx\in E,

ΣBπ11{x}={x}×(n{u:uB𝐖(x)}),\Sigma_{B}\cap\pi_{1}^{-1}\{x\}=\{x\}\times\big{(}\mathbb{R}^{n}\cap\{u:u\in B\cap\mathbf{W}(x)\}\big{)},

so that

m(ΣBπ11{x})=m(B𝐖(x)),\mathscr{H}^{m}\left(\Sigma_{B}\cap\pi_{1}^{-1}\{x\}\right)=\mathscr{H}^{m}\left(B\cap\mathbf{W}(x)\right), (12)

and that, for each fixed uBu\in B,

ΣBπ21{u}=(n{x:xE and u𝐖(x)})×{u}=(n{x:xEg𝐯1,,𝐯nm,u1{0}})×{u},\begin{split}\Sigma_{B}\cap\pi_{2}^{-1}\{u\}&=\big{(}\mathbb{R}^{n}\cap\{x:x\in E\text{ and }u\in\mathbf{W}(x)\}\big{)}\times\{u\}\\ &=\left(\mathbb{R}^{n}\cap\left\{x:x\in E\cap g_{\mathbf{v}_{1},\ldots,\mathbf{v}_{n-m},u}^{-1}\{0\}\right\}\right)\times\{u\},\end{split}

according to (8), so that

m(ΣBπ21{u})=m(Eg𝐯1,,𝐯nm,u1{0}),\mathscr{H}^{m}\left(\Sigma_{B}\cap\pi_{2}^{-1}\{u\}\right)=\mathscr{H}^{m}\left(E\cap g_{\mathbf{v}_{1},\ldots,\mathbf{v}_{n-m},u}^{-1}\{0\}\right), (13)

whenever uBu\in B. It now follows from the coarea formula that

ΣBJΣπ1𝑑n+m=Em(B𝐖(x))𝑑n(x)=ϕE,𝐖(B)\int_{\Sigma_{B}}J_{\Sigma}\pi_{1}d\mathscr{H}^{n+m}=\int_{E}\mathscr{H}^{m}\left(B\cap\mathbf{W}(x)\right)d\mathscr{L}^{n}(x)=\phi_{E,\mathbf{W}}(B) (14)

and

ΣBJΣπ2𝑑n+m=Bm(Eg𝐯1,,𝐯nm,u1{0})𝑑n(u).\int_{\Sigma_{B}}J_{\Sigma}\pi_{2}d\mathscr{H}^{n+m}=\int_{B}\mathscr{H}^{m}\left(E\cap g_{\mathbf{v}_{1},\ldots,\mathbf{v}_{n-m},u}^{-1}\{0\}\right)d\mathscr{L}^{n}(u). (15)

For these formulæ to be useful, we need to establish bounds for the coarea Jacobian factors JΣπ1J_{\Sigma}\pi_{1} and JΣπ2J_{\Sigma}\pi_{2}. In order to do so, we notice that if Σ(x,u)=F(x,t)\Sigma\ni(x,u)=F(x,t), FF is differentiable at (x,t)(x,t), i.e., each 𝐰i\mathbf{w}_{i} is differentiable at xx, i=1,,mi=1,\ldots,m, then the approximate tangent space T(x,u)ΣT_{(x,u)}\Sigma exists and is generated by the following n+mn+m vectors of n×n\mathbb{R}^{n}\times\mathbb{R}^{n}:

p=1nwj,epFxp(x,t)=(wj,wj+i=1mtiD𝐰i(x)(wj)),j=1,,n,Ftk(x,t)=(0,𝐰k(x)),k=1,,m,\begin{split}\sum_{p=1}^{n}\langle w_{j},e_{p}\rangle\frac{\partial F}{\partial x_{p}}(x,t)&=\left(w_{j},w_{j}+\sum_{i=1}^{m}t_{i}D\mathbf{w}_{i}(x)(w_{j})\right)\,,\,j=1,\ldots,n,\\ \frac{\partial F}{\partial t_{k}}(x,t)&=\left(0,\mathbf{w}_{k}(x)\right)\,,\,k=1,\ldots,m,\end{split}

where w1,,wnw_{1},\ldots,w_{n} is an arbitrary basis of n\mathbb{R}^{n}. As usual, e1,,ene_{1},\ldots,e_{n} denotes the canonical basis of n\mathbb{R}^{n}.

5.2Coarea Jacobian factor of π1\pi_{1}. —

For n+m\mathscr{H}^{n+m}-almost every (x,u)Σ(x,u)\in\Sigma, one has

(1+m2Λ2|xu|2)m2(2+2mΛ|xu|+m2Λ2|xu|2)nm2JΣπ1(x,u)1.\left(1+m^{2}\Lambda^{2}|x-u|^{2}\right)^{-\frac{m}{2}}\left(2+2m\Lambda|x-u|+m^{2}\Lambda^{2}|x-u|^{2}\right)^{-\frac{n-m}{2}}\leqslant J_{\Sigma}\pi_{1}(x,u)\leqslant 1.
Proof.

We recall 3.3. That the right hand inequality be valid follows from Lipπ1=1\operatorname{\mathrm{Lip}}\pi_{1}=1. Regarding the left hand inequality, fix (x,u)=F(x,t)(x,u)=F(x,t) such that FF is differentiable at (x,t)(x,t) and let L:T(x,u)ΣnL:T_{(x,u)}\Sigma\to\mathbb{R}^{n} denote the restriction of π1\pi_{1} to T(x,u)ΣT_{(x,u)}\Sigma. Define wj=𝐰j(x)w_{j}=\mathbf{w}_{j}(x), j=1,,mj=1,\ldots,m, and wj=𝐯jm(x)w_{j}=\mathbf{v}_{j-m}(x), j=m+1,,nj=m+1,\ldots,n. Put

vj=p=1nwj,epFxp(x,t)Ftj(x,t)=(wj,i=1mtiD𝐰i(x)(wj)),v_{j}=\sum_{p=1}^{n}\langle w_{j},e_{p}\rangle\frac{\partial F}{\partial x_{p}}(x,t)-\frac{\partial F}{\partial t_{j}}(x,t)=\left(w_{j},\sum_{i=1}^{m}t_{i}D\mathbf{w}_{i}(x)(w_{j})\right),

j=1,,mj=1,\ldots,m, and

vj=p=1nwj,epFxp(x,t)=(wj,wj+i=1mtiD𝐰i(x)(wj)),v_{j}=\sum_{p=1}^{n}\langle w_{j},e_{p}\rangle\frac{\partial F}{\partial x_{p}}(x,t)=\left(w_{j},w_{j}+\sum_{i=1}^{m}t_{i}D\mathbf{w}_{i}(x)(w_{j})\right),

j=m+1,,nj=m+1,\ldots,n. Recall (6) that

JΣπ1(x,u)=nL1|v1vn|,J_{\Sigma}\pi_{1}(x,u)=\|\wedge_{n}L\|\geqslant\frac{1}{|v_{1}\wedge\ldots\wedge v_{n}|},

since L(vj)=wjL(v_{j})=w_{j}, j=1,nj=1\ldots,n. Now, notice that, for j=1,,mj=1,\ldots,m,

|vj|2=|wj|2+|i=1mtiD𝐰i(x)(wj)|21+m2Λ2|t|2,\left|v_{j}\right|^{2}=|w_{j}|^{2}+\left|\sum_{i=1}^{m}t_{i}D\mathbf{w}_{i}(x)(w_{j})\right|^{2}\leqslant 1+m^{2}\Lambda^{2}|t|^{2},

whereas, for j=m+1,,nj=m+1,\ldots,n,

|vj|2=|wj|2+|wj+i=1mtiD𝐰i(x)(wj)|22+2|i=1mtiD𝐰i(x)(wj)|+|i=1mtiD𝐰i(x)(wj)|22+2mΛ|t|+m2Λ2|t|2.\left|v_{j}\right|^{2}=|w_{j}|^{2}+\left|w_{j}+\sum_{i=1}^{m}t_{i}D\mathbf{w}_{i}(x)(w_{j})\right|^{2}\leqslant 2+2\left|\sum_{i=1}^{m}t_{i}D\mathbf{w}_{i}(x)(w_{j})\right|+\left|\sum_{i=1}^{m}t_{i}D\mathbf{w}_{i}(x)(w_{j})\right|^{2}\\ \leqslant 2+2m\Lambda|t|+m^{2}\Lambda^{2}|t|^{2}.

Since u=x+i=1mti𝐰i(x)u=x+\sum_{i=1}^{m}t_{i}\mathbf{w}_{i}(x), one also has

|ux|2=|i=1mti𝐰i(x)|2=|t|2.|u-x|^{2}=\left|\sum_{i=1}^{m}t_{i}\mathbf{w}_{i}(x)\right|^{2}=|t|^{2}.

Finally,

|v1vn|(1+m2Λ2|xu|2)m2(2+2mΛ|xu|+m2Λ2|xu|2)nm2,|v_{1}\wedge\ldots\wedge v_{n}|\leqslant\left(1+m^{2}\Lambda^{2}|x-u|^{2}\right)^{\frac{m}{2}}\left(2+2m\Lambda|x-u|+m^{2}\Lambda^{2}|x-u|^{2}\right)^{\frac{n-m}{2}},

and the conclusion follows. ∎

5.3. —

Let 1qn11\leqslant q\leqslant n-1 be an integer and let v1,,vqv_{1},\ldots,v_{q} be an orthonormal family in n\mathbb{R}^{n}. There exists λΛ(n,q)\lambda\in\Lambda(n,q) such that

|det(vk,eλ(j))j,k=1,,q|(nq)12.\left|\det\left(\langle v_{k},e_{\lambda(j)}\rangle\right)_{j,k=1,\ldots,q}\right|\geqslant\begin{pmatrix}n\\ q\end{pmatrix}^{-\frac{1}{2}}\,.

Here, Λ(n,q)\Lambda(n,q) denotes the set of increasing maps {1,,q}{1,,n}\{1,\ldots,q\}\to\{1,\ldots,n\}.

Proof.

We define a linear map L:qn:(s1,,sq)k=1qskvkL:\mathbb{R}^{q}\to\mathbb{R}^{n}:(s_{1},\ldots,s_{q})\mapsto\sum_{k=1}^{q}s_{k}v_{k} and we observe that LL is an isometry. Therefore, its area Jacobian factor JL=1JL=1, by definition. Now, also

(JL)2=λΛ(n,q)|det(vk,eλ(j))j,k=1,,q|2,(JL)^{2}=\sum_{\lambda\in\Lambda(n,q)}\left|\det\left(\langle v_{k},e_{\lambda(j)}\rangle\right)_{j,k=1,\ldots,q}\right|^{2},

according to the Binet-Cauchy formula [6, Chapter 3 §2 Theorem 4]. The conclusion easily follows. ∎

5.4Coarea Jacobian factor of π2\pi_{2}. —

The following hold.

  1. (1)

    For n+m\mathscr{H}^{n+m}-almost every (x,u)Σ(x,u)\in\Sigma, one has

    (nnm)12m(nm)Λ|xu|(1+mΛ|xu|)nm1(2+2mΛ|xu|+m2Λ2|xu|2)nm2JΣπ2(x,u)1.\frac{\begin{pmatrix}n\\ n-m\end{pmatrix}^{-\frac{1}{2}}-m(n-m)\Lambda|x-u|\left(1+m\Lambda|x-u|\right)^{n-m-1}}{\left(2+2m\Lambda|x-u|+m^{2}\Lambda^{2}|x-u|^{2}\right)^{\frac{n-m}{2}}}\leqslant J_{\Sigma}\pi_{2}(x,u)\leqslant 1.
  2. (2)

    For n+m\mathscr{H}^{n+m}-almost every (x,u)Σ(x,u)\in\Sigma, one has JΣπ2(x,u)>0J_{\Sigma}\pi_{2}(x,u)>0.

Proof.

Clearly, JΣπ2(x,u)(Lipπ2)n1J_{\Sigma}\pi_{2}(x,u)\leqslant(\operatorname{\mathrm{Lip}}\pi_{2})^{n}\leqslant 1. Regarding the left hand inequality, fix (x,u)=F(x,t)(x,u)=F(x,t) such that FF is differentiable at (x,t)(x,t) and, this time, let L:T(x,u)ΣnL:T_{(x,u)}\Sigma\to\mathbb{R}^{n} denote the restriction of π2\pi_{2} to T(x,u)ΣT_{(x,u)}\Sigma. We will now define a family of nn vectors v1,,vnv_{1},\ldots,v_{n} belonging to T(x,u)ΣT_{(x,u)}\Sigma. We choose vk=Ftk(x,t)=(0,𝐰k(x))v_{k}=\frac{\partial F}{\partial t_{k}}(x,t)=(0,\mathbf{w}_{k}(x)), for k=1,,mk=1,\ldots,m. For choosing the nmn-m remaining vectors, we proceed as follows. We select λΛ(n,nm)\lambda\in\Lambda(n,n-m) as in 5.3 applied with q=nmq=n-m to 𝐯1(x),,𝐯nm(x)\mathbf{v}_{1}(x),\ldots,\mathbf{v}_{n-m}(x) and we let vm+j=Fxλ(j)(x,t)v_{m+j}=\frac{\partial F}{\partial x_{\lambda(j)}}(x,t), j=1,,nmj=1,\ldots,n-m. Recalling (6), we have

JΣπ1(x,u)=nL|L(v1)L(vn)||v1vn|.J_{\Sigma}\pi_{1}(x,u)=\|\wedge_{n}L\|\geqslant\frac{|L(v_{1})\wedge\ldots\wedge L(v_{n})|}{|v_{1}\wedge\ldots\wedge v_{n}|}.

As in the proof of 5.2, we find that

|v1vn|(2+2mΛ|xu|+m2Λ2|xu|2)nm2|v_{1}\wedge\ldots\wedge v_{n}|\leqslant\left(2+2m\Lambda|x-u|+m^{2}\Lambda^{2}|x-u|^{2}\right)^{\frac{n-m}{2}}

and it remains only to find a lower bound for |L(v1)L(vn)||L(v_{1})\wedge\ldots\wedge L(v_{n})|. The latter equals the absolute value of the determinant of the matrix of coefficients of L(vi)L(v_{i}), i=1,,ni=1,\ldots,n, with respect to any orthonormal basis of n\mathbb{R}^{n}. We choose the basis 𝐰1(x),,𝐰m(x),𝐯1(x),,𝐯nm(x)\mathbf{w}_{1}(x),\ldots,\mathbf{w}_{m}(x),\mathbf{v}_{1}(x),\ldots,\mathbf{v}_{n-m}(x). Thus,

|L(v1)L(vn)|=|det(100100eλ(j)+i=1mti𝐰ixλ(j)(x),𝐯k(x))|=|det(eλ(j)+i=1mti𝐰ixλ(j)(x),𝐯k(x))j,k=1,,nm|.\begin{split}|L(v_{1})\wedge\ldots\wedge L(v_{n})|&=\left|\det\left(\begin{array}[]{c c c | c}1&\cdots&0&*\\ \vdots&\ddots&\vdots&\vdots\\ 0&\cdots&1&*\\ \hline\cr 0&\cdots&0&\left\langle e_{\lambda(j)}+\sum_{i=1}^{m}t_{i}\frac{\partial\mathbf{w}_{i}}{\partial x_{\lambda(j)}}(x),\mathbf{v}_{k}(x)\right\rangle\\ \end{array}\right)\right|\\ &=\left|\det\left(\left\langle e_{\lambda(j)}+\sum_{i=1}^{m}t_{i}\frac{\partial\mathbf{w}_{i}}{\partial x_{\lambda(j)}}(x),\mathbf{v}_{k}(x)\right\rangle\right)_{j,k=1,\ldots,n-m}\right|.\end{split} (16)

Abbreviate

hλ(j)=i=1mti𝐰ixλ(j)(x)h_{\lambda(j)}=\sum_{i=1}^{m}t_{i}\frac{\partial\mathbf{w}_{i}}{\partial x_{\lambda(j)}}(x)

and observe that |hλ(j)|mΛ|t|=mΛ|xu|\left|h_{\lambda(j)}\right|\leqslant m\Lambda|t|=m\Lambda|x-u|, j=1,,nmj=1,\ldots,n-m (recall the proof of 5.2). It remains only to remember that λ\lambda has been selected in order that

|det(eλ(j),𝐯k(x))j,k=1,,nm|(nnm)12\left|\det\left(\left\langle e_{\lambda(j)},\mathbf{v}_{k}(x)\right\rangle\right)_{j,k=1,\ldots,n-m}\right|\geqslant\begin{pmatrix}n\\ n-m\end{pmatrix}^{-\frac{1}{2}}

and to infer from the multilinearity of the determinant that

|det(eλ(j),𝐯k(x)+hλ(j),𝐯k(x))j,kdet(eλ(j),𝐯k(x))j,k|(nm)(maxj=1,,nm|hλ(j)|)(maxj=1,,nm|hλ(j)|+|eλ(j)|)nm1(nm)mΛ|xu|(1+mΛ|xu|)nm1.\left|\det\left(\left\langle e_{\lambda(j)},\mathbf{v}_{k}(x)\right\rangle+\left\langle h_{\lambda(j)},\mathbf{v}_{k}(x)\right\rangle\right)_{j,k}-\det\left(\left\langle e_{\lambda(j)},\mathbf{v}_{k}(x)\right\rangle\right)_{j,k}\right|\\ \leqslant(n-m)\left(\max_{j=1,\ldots,n-m}\left|h_{\lambda(j)}\right|\right)\left(\max_{j=1,\ldots,n-m}\left|h_{\lambda(j)}\right|+\left|e_{\lambda(j)}\right|\right)^{n-m-1}\\ \leqslant(n-m)m\Lambda|x-u|\left(1+m\Lambda|x-u|\right)^{n-m-1}.

This completes the proof of conclusion (1).

Let E0E_{0} denote the subset of EE consisting of those xx such that each 𝐰i\mathbf{w}_{i}, i=1,,mi=1,\ldots,m, is differentiable at xx. Thus, E0E_{0} is Borel measurable and so is

A=E0×m{(x,t):rank(𝐰1(x)𝐰m(x)e1+i=1mti𝐰ix1(x)en+i=1mti𝐰ixn(x))<n}.A=E_{0}\times\mathbb{R}^{m}\cap\bigg{\{}(x,t):\\ \operatorname{\mathrm{rank}}\left(\begin{array}[]{c|c|c|c|c|c}\mathbf{w}_{1}(x)&\cdots&\mathbf{w}_{m}(x)&e_{1}+\sum_{i=1}^{m}t_{i}\frac{\partial\mathbf{w}_{i}}{\partial x_{1}}(x)&\ldots&e_{n}+\sum_{i=1}^{m}t_{i}\frac{\partial\mathbf{w}_{i}}{\partial x_{n}}(x)\end{array}\right)<n\bigg{\}}.

If (x,u)ΣF(A)(x,u)\in\Sigma\setminus F(A), then the restriction of π2\pi_{2} to T(x,u)ΣT_{(x,u)}\Sigma is surjective and, therefore, JΣπ2(x,u)>0J_{\Sigma}\pi_{2}(x,u)>0. Thus, we ought to show that n+m(F(A))=0\mathscr{H}^{n+m}(F(A))=0. Since FF is Lipschitzian, it suffices to establish that n+m(A)=0\mathscr{L}^{n+m}(A)=0. As AA is Borel measurable, it is enough to prove that m(Ax)=0\mathscr{L}^{m}(A_{x})=0, for every xE0x\in E_{0}, according to Fubini’s theorem. Fix xE0x\in E_{0}. As in the proof of conclusion (1), choose λΛ(n,nm)\lambda\in\Lambda(n,n-m) associated with 𝐯1(x),,𝐯nm(x)\mathbf{v}_{1}(x),\ldots,\mathbf{v}_{n-m}(x), according to 5.3. Based on (16), we see that

Axm{t:det(eλ(j)+i=1mti𝐰ixλ(j)(x),𝐯k(x))j,k=1,,nm=0}.A_{x}\subseteq\mathbb{R}^{m}\cap\left\{t:\det\left(\left\langle e_{\lambda(j)}+\sum_{i=1}^{m}t_{i}\frac{\partial\mathbf{w}_{i}}{\partial x_{\lambda(j)}}(x),\mathbf{v}_{k}(x)\right\rangle\right)_{j,k=1,\ldots,n-m}=0\right\}.

The set on the right is of the form Sx=m{(t1,,tm):Px(t1,,tm)=0}S_{x}=\mathbb{R}^{m}\cap\{(t_{1},\ldots,t_{m}):P_{x}(t_{1},\ldots,t_{m})=0\}, for some polynomial Px[T1,,Tm]P_{x}\in\mathbb{R}[T_{1},\ldots,T_{m}], and Px(0,,0)=det(eλ(j),𝐯k(x))j,k=1,,nm0P_{x}(0,\ldots,0)=\det\left(\langle e_{\lambda(j)},\mathbf{v}_{k}(x)\rangle\right)_{j,k=1,\ldots,n-m}\neq 0. It follows that m(Sx)=0\mathscr{L}^{m}(S_{x})=0 – see, e.g., [9, 2.6.5] – and the proof of (2) is complete. ∎

5.5 Proposition. —

The measure ϕE,𝐖\phi_{E,\mathbf{W}} is absolutely continuous with respect to n\mathscr{L}^{n}.

Proof.

Let B(n)B\in\mathscr{B}(\mathbb{R}^{n}) be such that n(B)=0\mathscr{L}^{n}(B)=0. It follows from (15) that

ΣBJΣπ2𝑑n+m=0.\int_{\Sigma_{B}}J_{\Sigma}\pi_{2}d\mathscr{H}^{n+m}=0.

It next follows from 5.4(2) that n+m(ΣB)=0\mathscr{H}^{n+m}(\Sigma_{B})=0. In turn, (14) implies that

ϕE,𝐖(B)=ΣBJΣπ1𝑑n+m=0.\phi_{E,\mathbf{W}}(B)=\int_{\Sigma_{B}}J_{\Sigma}\pi_{1}d\mathscr{H}^{n+m}=0.

5.6Definition of 𝒵E𝐖\mathscr{Z}_{E}\mathbf{W}. —

Note that ϕE,𝐖\phi_{E,\mathbf{W}} is a σ\sigma-finite Borel measure on n\mathbb{R}^{n} (see 3.16) and it is absolutely continuous with respect to n\mathscr{L}^{n} (see 5.5). It then ensues from the Radon-Nikodým theorem that there exists a Borel measurable function

𝒵E𝐖:n\mathscr{Z}_{E}\mathbf{W}:\mathbb{R}^{n}\to\mathbb{R}

such that, for every B(n)B\in\mathscr{B}(\mathbb{R}^{n}), one has

Em(B𝐖(x))𝑑n(x)=ϕE,𝐖(B)=B𝒵E𝐖(u)𝑑n(u).\int_{E}\mathscr{H}^{m}\left(B\cap\mathbf{W}(x)\right)d\mathscr{L}^{n}(x)=\phi_{E,\mathbf{W}}(B)=\int_{B}\mathscr{Z}_{E}\mathbf{W}(u)d\mathscr{L}^{n}(u).

Furthermore, 𝒵E𝐖\mathscr{Z}_{E}\mathbf{W} is univoquely defined (only) up to an n\mathscr{L}^{n} null set. This will not affect the reasoning in this paper. Each time we write 𝒵E𝐖\mathscr{Z}_{E}\mathbf{W}, we mean one particular Borel measurable function satisfying the above equality, for every B(n)B\in\mathscr{B}(\mathbb{R}^{n}).

5.7Definition of 𝒴E0𝐖\mathscr{Y}_{E}^{0}\mathbf{W}. —

We define 𝒴E0𝐖:n[0,]\mathscr{Y}^{0}_{E}\mathbf{W}:\mathbb{R}^{n}\to[0,\infty] by the formula

𝒴E0𝐖(u)=m(Eg𝐯1,,𝐯nm,u1{0}),\mathscr{Y}_{E}^{0}\mathbf{W}(u)=\mathscr{H}^{m}\left(E\cap g_{\mathbf{v}_{1},\ldots,\mathbf{v}_{n-m},u}^{-1}\{0\}\right), (17)

unu\in\mathbb{R}^{n}. Letting B=nB=\mathbb{R}^{n} in (13), one infers from 3.3 that 𝒴E0𝐖\mathscr{Y}^{0}_{E}\mathbf{W} is n\mathscr{L}^{n}-measurable. Using the estimates we have established so far regarding coarea Jacobian factors, we now show that 𝒵E𝐖\mathscr{Z}_{E}\mathbf{W} and 𝒴E0𝐖\mathscr{Y}_{E}^{0}\mathbf{W} are comparable, when the diameter of EE is not too large.

5.8 Proposition. —

Given 0<ε<10<\varepsilon<1, there exists 𝛅5.8(n,Λ,ε)>0\boldsymbol{\delta}_{5.8}(n,\Lambda,\varepsilon)>0 with the following property. If diamE𝛅5.8(n,Λ,ε)\operatorname{\mathrm{diam}}E\leqslant\boldsymbol{\delta}_{5.8}(n,\Lambda,\varepsilon), then

(1ε)2nm2𝒴E0𝐖(u)𝒵E𝐖(u)(1+ε)2nm2(nnm)12𝒴E0𝐖(u),(1-\varepsilon)2^{-\frac{n-m}{2}}\mathscr{Y}^{0}_{E}\mathbf{W}(u)\leqslant\mathscr{Z}_{E}\mathbf{W}(u)\leqslant(1+\varepsilon)2^{\frac{n-m}{2}}\begin{pmatrix}n\\ n-m\end{pmatrix}^{\frac{1}{2}}\mathscr{Y}^{0}_{E}\mathbf{W}(u),

for n\mathscr{L}^{n}-almost every uEu\in E.

Proof.

We readily infer from 5.2 and 5.4(1) that there exists 𝜹(n,Λ,ε)>0\boldsymbol{\delta}(n,\Lambda,\varepsilon)>0 such that, for n+m\mathscr{H}^{n+m}-almost all (x,u)Σ(x,u)\in\Sigma, if |xu|𝜹(n,Λ,ε)|x-u|\leqslant\boldsymbol{\delta}(n,\Lambda,\varepsilon), then

α:=(1ε)2nm2JΣπ1(x,u)\alpha:=(1-\varepsilon)2^{-\frac{n-m}{2}}\leqslant J_{\Sigma}\pi_{1}(x,u) (18)

and

β:=(1+ε)12nm2(nnm)12JΣπ2(x,u),\beta:=(1+\varepsilon)^{-1}2^{-\frac{n-m}{2}}\begin{pmatrix}n\\ n-m\end{pmatrix}^{-\frac{1}{2}}\leqslant J_{\Sigma}\pi_{2}(x,u), (19)

where the above define α\alpha and β\beta.

Assume now that diamE𝜹(n,Λ,ε)\operatorname{\mathrm{diam}}E\leqslant\boldsymbol{\delta}(n,\Lambda,\varepsilon). Given B(E)B\in\mathscr{B}(E), we infer from (14), 5.2, 5.4(1), (15), and the above lower bounds, that

ϕE,𝐖(B)=ΣBJΣπ1𝑑n+mαn+m(ΣB)αΣBJΣπ2𝑑n+m=αB𝒴E0𝐖𝑑n\phi_{E,\mathbf{W}}(B)=\int_{\Sigma_{B}}J_{\Sigma}\pi_{1}d\mathscr{H}^{n+m}\geqslant\alpha\mathscr{H}^{n+m}(\Sigma_{B})\geqslant\alpha\int_{\Sigma_{B}}J_{\Sigma}\pi_{2}d\mathscr{H}^{n+m}\\ =\alpha\int_{B}\mathscr{Y}^{0}_{E}\mathbf{W}d\mathscr{L}^{n}

and

ϕE,𝐖(B)=ΣBJΣπ1𝑑n+mn+m(ΣB)β1ΣBJΣπ2𝑑n+m=β1B𝒴E0𝐖𝑑n.\phi_{E,\mathbf{W}}(B)=\int_{\Sigma_{B}}J_{\Sigma}\pi_{1}d\mathscr{H}^{n+m}\leqslant\mathscr{H}^{n+m}(\Sigma_{B})\leqslant\beta^{-1}\int_{\Sigma_{B}}J_{\Sigma}\pi_{2}d\mathscr{H}^{n+m}\\ =\beta^{-1}\int_{B}\mathscr{Y}^{0}_{E}\mathbf{W}d\mathscr{L}^{n}.

Thus,

Bα𝒴E0𝐖𝑑nB𝒵E𝐖𝑑nBβ1𝒴E0𝐖𝑑n,\int_{B}\alpha\mathscr{Y}^{0}_{E}\mathbf{W}d\mathscr{L}^{n}\leqslant\int_{B}\mathscr{Z}_{E}\mathbf{W}d\mathscr{L}^{n}\leqslant\int_{B}\beta^{-1}\mathscr{Y}^{0}_{E}\mathbf{W}d\mathscr{L}^{n},

for every B(E)B\in\mathscr{B}(E). The conclusion follows from the n\mathscr{L}^{n}-measurability of both 𝒵E𝐖\mathscr{Z}_{E}\mathbf{W} and 𝒴E0𝐖\mathscr{Y}^{0}_{E}\mathbf{W}. ∎

5.9Rest stop. —

The above upper bound for 𝒵E𝐖\mathscr{Z}_{E}\mathbf{W} is already enough to bound it from above, in turn, by a constant times (diamE)m(\operatorname{\mathrm{diam}}E)^{m} – see 6.4. We next want to establish that 𝒵E𝐖>0\mathscr{Z}_{E}\mathbf{W}>0, almost everywhere in EE. Yet, in the definition (17) of 𝒴E0𝐖(u)\mathscr{Y}^{0}_{E}\mathbf{W}(u), uu does not appear as the covariable of a function whose level sets we are measuring, thereby preventing the use of the coarea formula in an attempt to estimate 𝒴E0𝐖(u)\mathscr{Y}^{0}_{E}\mathbf{W}(u). This naturally leads to adding a variable ynmy\in\mathbb{R}^{n-m} to the fibered space Σ\Sigma, a covariable for g𝐯1,,𝐯nm,ug_{\mathbf{v}_{1},\ldots,\mathbf{v}_{n-m},u}.

5.10A fibered space associated with E,B,𝐰1,,𝐰m,𝐯1,,𝐯nmE,B,\mathbf{w}_{1},\ldots,\mathbf{w}_{m},\mathbf{v}_{1},\ldots,\mathbf{v}_{n-m}. —

Let r>0r>0. Abbreviate Cr=nm{y:|y|r}C_{r}=\mathbb{R}^{n-m}\cap\left\{y:|y|\leqslant r\right\}, the Euclidean ball centered at the origin, of radius rr, in nm\mathbb{R}^{n-m}. We define

F^r:E×m×Crn×n×nm:(x,t,y)(x,x+i=1mti𝐰i(x)+i=1nmyi𝐯i(x),y)\hat{F}_{r}:E\times\mathbb{R}^{m}\times C_{r}\to\mathbb{R}^{n}\times\mathbb{R}^{n}\times\mathbb{R}^{n-m}:(x,t,y)\mapsto\left(x,x+\sum_{i=1}^{m}t_{i}\mathbf{w}_{i}(x)+\sum_{i=1}^{n-m}y_{i}\mathbf{v}_{i}(x),y\right)

and

Σ^r=F^r(E×m×Cr)=n×n×Cr{(x,u,y):xE and u𝐖(x)+i=1nmyi𝐯i(x)}.\hat{\Sigma}_{r}=\hat{F}_{r}\left(E\times\mathbb{R}^{m}\times C_{r}\right)=\mathbb{R}^{n}\times\mathbb{R}^{n}\times C_{r}\cap\left\{(x,u,y):x\in E\text{ and }u\in\mathbf{W}(x)+\sum_{i=1}^{n-m}y_{i}\mathbf{v}_{i}(x)\right\}.

We note that F^r\hat{F}_{r} is locally Lipschitzian and Σ^r\hat{\Sigma}_{r} is countably 2n2n-rectifiable and 2n\mathscr{H}^{2n}-measurable. Similarly to 5.1, we define

Σ^r,B=Σ^rπ21(B)\hat{\Sigma}_{r,B}=\hat{\Sigma}_{r}\cap\pi_{2}^{-1}(B)

which, clearly, is also countably 2n2n-rectifiable and 2n\mathscr{H}^{2n}-measurable. We aim to apply the coarea formula to Σ^r,B\hat{\Sigma}_{r,B} and to the two projections

π1×π3:n×n×nmn×nm:(x,u,y)(x,y)\pi_{1}\times\pi_{3}:\mathbb{R}^{n}\times\mathbb{R}^{n}\times\mathbb{R}^{n-m}\to\mathbb{R}^{n}\times\mathbb{R}^{n-m}:(x,u,y)\mapsto(x,y)

and

π2×π3:n×n×nmn×nm:(x,u,y)(u,y).\pi_{2}\times\pi_{3}:\mathbb{R}^{n}\times\mathbb{R}^{n}\times\mathbb{R}^{n-m}\to\mathbb{R}^{n}\times\mathbb{R}^{n-m}:(x,u,y)\mapsto(u,y).

To this end, we notice that

Σ^r,B(π1×π3)1{(x,y)}=n×n×nm{(x,u,y):uB(𝐖(x)+i=1nmyi𝐯i(x))}\hat{\Sigma}_{r,B}\cap\left(\pi_{1}\times\pi_{3}\right)^{-1}\{(x,y)\}\\ =\mathbb{R}^{n}\times\mathbb{R}^{n}\times\mathbb{R}^{n-m}\cap\left\{(x,u,y):u\in B\cap\left(\mathbf{W}(x)+\sum_{i=1}^{n-m}y_{i}\mathbf{v}_{i}(x)\right)\right\}

and, thus,

m(Σ^r,B(π1×π3)1{(x,y)})=m(B(𝐖(x)+i=1nmyi𝐯i(x))),\mathscr{H}^{m}\left(\hat{\Sigma}_{r,B}\cap\left(\pi_{1}\times\pi_{3}\right)^{-1}\{(x,y)\}\right)=\mathscr{H}^{m}\left(B\cap\left(\mathbf{W}(x)+\sum_{i=1}^{n-m}y_{i}\mathbf{v}_{i}(x)\right)\right),

for every (x,y)E×Cr(x,y)\in E\times C_{r}. We further notice that

Σ^r,B(π2×π3)1{(u,y)}=n×n×nm{(x,u,y):xE and u𝐖(x)+i=1nmyi𝐯i(x)}=n×n×nm{(x,u,y):xEg𝐯1,,𝐯nm,u1{y}},\begin{split}\hat{\Sigma}_{r,B}\cap(\pi_{2}&\times\pi_{3})^{-1}\{(u,y)\}\\ &=\mathbb{R}^{n}\times\mathbb{R}^{n}\times\mathbb{R}^{n-m}\cap\left\{(x,u,y):x\in E\text{ and }u\in\mathbf{W}(x)+\sum_{i=1}^{n-m}y_{i}\mathbf{v}_{i}(x)\right\}\\ &=\mathbb{R}^{n}\times\mathbb{R}^{n}\times\mathbb{R}^{n-m}\cap\left\{(x,u,y):x\in E\cap g_{\mathbf{v}_{1},\ldots,\mathbf{v}_{n-m},u}^{-1}\{y\}\right\},\end{split}

because

u𝐖(x)+i=1nmyi𝐯i(x)uxi=1nmyi𝐯i(x)𝐖0(x)𝐯j(x),uxi=1nmyi𝐯i(x)=0, for all j=1,,nmg𝐯1,,𝐯nm,u(x)=y\begin{split}u\in\mathbf{W}(x)+\sum_{i=1}^{n-m}y_{i}\mathbf{v}_{i}(x)&\iff u-x-\sum_{i=1}^{n-m}y_{i}\mathbf{v}_{i}(x)\in\mathbf{W}_{0}(x)\\ &\iff\left\langle\mathbf{v}_{j}(x),u-x-\sum_{i=1}^{n-m}y_{i}\mathbf{v}_{i}(x)\right\rangle=0\text{, for all }j=1,\ldots,n-m\\ &\iff g_{\mathbf{v}_{1},\ldots,\mathbf{v}_{n-m},u}(x)=y\end{split}

and, therefore,

m(Σ^r,B(π2×π3)1{(u,y)})=m(Eg𝐯1,,𝐯nm,u1{y}),\mathscr{H}^{m}\left(\hat{\Sigma}_{r,B}\cap(\pi_{2}\times\pi_{3})^{-1}\{(u,y)\}\right)=\mathscr{H}^{m}\left(E\cap g_{\mathbf{v}_{1},\ldots,\mathbf{v}_{n-m},u}^{-1}\{y\}\right),

whenever uBu\in B and yCry\in C_{r}.

It now follows from the coarea formula and Fubini’s theorem that

Σ^r,BJΣ^r(π1×π3)𝑑2n=E𝑑n(x)Crm(B(𝐖(x)+i=1nmyi𝐯i(x)))𝑑nm(y)\int_{\hat{\Sigma}_{r,B}}J_{\hat{\Sigma}_{r}}(\pi_{1}\times\pi_{3})d\mathscr{H}^{2n}=\int_{E}d\mathscr{L}^{n}(x)\int_{C_{r}}\mathscr{H}^{m}\left(B\cap\left(\mathbf{W}(x)+\sum_{i=1}^{n-m}y_{i}\mathbf{v}_{i}(x)\right)\right)d\mathscr{L}^{n-m}(y) (20)

and

Σ^r,BJΣ^r(π2×π3)𝑑2n=B𝑑n(u)Crm(Eg𝐯1,,𝐯nm,u1{y})𝑑nm(y).\int_{\hat{\Sigma}_{r,B}}J_{\hat{\Sigma}_{r}}(\pi_{2}\times\pi_{3})d\mathscr{H}^{2n}=\int_{B}d\mathscr{L}^{n}(u)\int_{C_{r}}\mathscr{H}^{m}\left(E\cap g_{\mathbf{v}_{1},\ldots,\mathbf{v}_{n-m},u}^{-1}\{y\}\right)d\mathscr{L}^{n-m}(y). (21)
5.11Coarea Jacobian factors of π1×π3\pi_{1}\times\pi_{3} and π2×π3\pi_{2}\times\pi_{3}. —

The following inequalities hold, for 2n\mathscr{H}^{2n}-almost every (x,u,y)Σ^r(x,u,y)\in\hat{\Sigma}_{r}.

2(nm)(1+2nΛ|ux|+2n2Λ2|ux|2)n2JΣ^r(π1×π3)(x,u,y)2^{-(n-m)}\left(1+2n\Lambda|u-x|+2n^{2}\Lambda^{2}|u-x|^{2}\right)^{-\frac{n}{2}}\leqslant J_{\hat{\Sigma}_{r}}(\pi_{1}\times\pi_{3})(x,u,y)

and

JΣ^r(π2×π3)(x,u,y)1.J_{\hat{\Sigma}_{r}}(\pi_{2}\times\pi_{3})(x,u,y)\leqslant 1.
Proof.

The second conclusion is obvious, since Lipπ2×π3=1\operatorname{\mathrm{Lip}}\pi_{2}\times\pi_{3}=1. Regarding the first conclusion, we reason similarly to the proof of 5.2. Fix (x,u,y)=F^r(x,t,y)(x,u,y)=\hat{F}_{r}(x,t,y) such that F^r\hat{F}_{r} is differentiable at (x,t,y)(x,t,y) and denote by LL the restriction of π1×π3\pi_{1}\times\pi_{3} to T(x,u,y)Σ^rT_{(x,u,y)}\hat{\Sigma}_{r}. This tangent space is generated by the following 2n2n vectors of n×n×nm\mathbb{R}^{n}\times\mathbb{R}^{n}\times\mathbb{R}^{n-m}:

F^rxj(x,t,y)=(ej,ej+i=1mti𝐰i(x)xj(x)+i=1nmyi𝐯ixj(x),0),j=1,,n,F^rtk(x,t,y)=(0,𝐰k(x),0),k=1,,m,F^ry(x,t,y)=(0,𝐯(x),e),=1,,nm.\begin{split}\frac{\partial\hat{F}_{r}}{\partial x_{j}}(x,t,y)&=\left(e_{j},e_{j}+\sum_{i=1}^{m}t_{i}\frac{\partial\mathbf{w}_{i}(x)}{\partial x_{j}}(x)+\sum_{i=1}^{n-m}y_{i}\frac{\partial\mathbf{v}_{i}}{\partial x_{j}}(x),0\right),\,j=1,\ldots,n,\\ \frac{\partial\hat{F}_{r}}{\partial t_{k}}(x,t,y)&=\left(0,\mathbf{w}_{k}(x),0\right),\,k=1,\ldots,m,\\ \frac{\partial\hat{F}_{r}}{\partial y_{\ell}}(x,t,y)&=\left(0,\mathbf{v}_{\ell}(x),e_{\ell}\right),\,\ell=1,\ldots,n-m.\end{split}

The range of π1×π3\pi_{1}\times\pi_{3} being (2nm)(2n-m)-dimensional, we need to select 2nm2n-m vectors v1,,v2nmv_{1},\ldots,v_{2n-m} in T(x,u,y)Σ^rT_{(x,u,y)}\hat{\Sigma}_{r} in view of obtaining a lower bound

JΣ^r(π1×π3)(x,u,y)=2nmL|L(v1)L(v2nm)||v1v2nm|.J_{\hat{\Sigma}_{r}}(\pi_{1}\times\pi_{3})(x,u,y)=\|\wedge_{2n-m}L\|\geqslant\frac{|L(v_{1})\wedge\ldots\wedge L(v_{2n-m})|}{|v_{1}\wedge\ldots\wedge v_{2n-m}|}. (22)

Our choice of v1,,v2nmv_{1},\ldots,v_{2n-m} is as follows. As in the proof of 5.2, we let wj=𝐰j(x)w_{j}=\mathbf{w}_{j}(x), for j=1,,mj=1,\ldots,m, and wj=𝐯jm(x)w_{j}=\mathbf{v}_{j-m}(x), for j=m+1,,nj=m+1,\ldots,n. For j=1,,mj=1,\ldots,m, we define

vj=(p=1nwj,epF^rxp(x,t,y))F^rtj(x,t,y)=(wj,i=1mtiD𝐰i(x)(wj)+i=1nmyiD𝐯i(x)(wj),0),v_{j}=\left(\sum_{p=1}^{n}\langle w_{j},e_{p}\rangle\frac{\partial\hat{F}_{r}}{\partial x_{p}}(x,t,y)\right)-\frac{\partial\hat{F}_{r}}{\partial t_{j}}(x,t,y)\\ =\left(w_{j},\sum_{i=1}^{m}t_{i}D\mathbf{w}_{i}(x)(w_{j})+\sum_{i=1}^{n-m}y_{i}D\mathbf{v}_{i}(x)(w_{j}),0\right),

for j=m+1,,nj=m+1,\ldots,n, we define

vj=p=1mwj,epF^rxp(x,t,y)=(wj,wj+i=1mtiD𝐰i(x)(wj)+i=1nmyiD𝐯i(x)(wj),0),v_{j}=\sum_{p=1}^{m}\langle w_{j},e_{p}\rangle\frac{\partial\hat{F}_{r}}{\partial x_{p}}(x,t,y)=\left(w_{j},w_{j}+\sum_{i=1}^{m}t_{i}D\mathbf{w}_{i}(x)(w_{j})+\sum_{i=1}^{n-m}y_{i}D\mathbf{v}_{i}(x)(w_{j}),0\right),

and, for j=n+1,,2nmj=n+1,\ldots,2n-m, we define

vj=F^ryjn(x,t,y)=(0,𝐯jn(x),ejn),v_{j}=\frac{\partial\hat{F}_{r}}{\partial y_{j-n}}(x,t,y)=(0,\mathbf{v}_{j-n}(x),e_{j-n}),

so that L(v1),,L(v2nm)L(v_{1}),\ldots,L(v_{2n-m}) is an orthonormal basis of n×nm\mathbb{R}^{n}\times\mathbb{R}^{n-m} and, therefore, the numerator in (22) equals 1. In order to determine an upper bound for its denominator, we start by fixing j=1,,nj=1,\ldots,n, we abbreviate aj(x,t,y)=i=1mtiD𝐰i(x)(x)(wj)a_{j}(x,t,y)=\sum_{i=1}^{m}t_{i}D\mathbf{w}_{i}(x)(x)(w_{j}) and bj(x,t,y)=i=1nmyiD𝐯i(x)(wj)b_{j}(x,t,y)=\sum_{i=1}^{n-m}y_{i}D\mathbf{v}_{i}(x)(w_{j}), and we notice that |aj(x,t,y)|mΛ|t|nΛ|t||a_{j}(x,t,y)|\leqslant m\Lambda|t|\leqslant n\Lambda|t|, |bj(x,t,y)|(nm)Λ|y|nΛ|y||b_{j}(x,t,y)|\leqslant(n-m)\Lambda|y|\leqslant n\Lambda|y|. Furthermore, since ux=i=1mti𝐰i(x)+i=1nmyi𝐯i(x)u-x=\sum_{i=1}^{m}t_{i}\mathbf{w}_{i}(x)+\sum_{i=1}^{n-m}y_{i}\mathbf{v}_{i}(x), one has |ux|2=|t|2+|y|2max{|t|2,|y|2}|u-x|^{2}=|t|^{2}+|y|^{2}\geqslant\max\{|t|^{2},|y|^{2}\}. Therefore, if j=1,,mj=1,\ldots,m, then

|vj|2=|wj|2+|aj(x,t,y)+bj(x,t,y)|21+|aj(x,t,y)|2+|bj(x,t,y)|2+2|aj(x,t,y)||bj(x,t,y)|1+n2Λ2(|t|2+|y|2+2|t||y|)1+2n2Λ2|ux|2,\begin{split}|v_{j}|^{2}&=|w_{j}|^{2}+\left|a_{j}(x,t,y)+b_{j}(x,t,y)\right|^{2}\\ &\leqslant 1+\left|a_{j}(x,t,y)\right|^{2}+\left|b_{j}(x,t,y)\right|^{2}+2\left|a_{j}(x,t,y)\right|\left|b_{j}(x,t,y)\right|\\ &\leqslant 1+n^{2}\Lambda^{2}\left(|t|^{2}+|y|^{2}+2|t||y|\right)\\ &\leqslant 1+2n^{2}\Lambda^{2}|u-x|^{2},\end{split}

whereas, if j=m+1,,nj=m+1,\ldots,n, then

|vj|2=|wj|2+|wj+aj(x,t,y)+bj(x,t,y)|21+1+|aj(x,t,y)|2+|bj(x,t,y)|2+2|aj(x,t,y)|+2|bj(x,t,y)|+2|aj(x,t,y)||bj(x,t,y)|2(1+n2Λ2|ux|2+2nΛ|ux|),\begin{split}\left|v_{j}\right|^{2}&=|w_{j}|^{2}+|w_{j}+a_{j}(x,t,y)+b_{j}(x,t,y)|^{2}\\ &\leqslant 1+1+|a_{j}(x,t,y)|^{2}+|b_{j}(x,t,y)|^{2}\\ &\quad\quad\quad+2|a_{j}(x,t,y)|+2|b_{j}(x,t,y)|+2|a_{j}(x,t,y)||b_{j}(x,t,y)|\\ &\leqslant 2\left(1+n^{2}\Lambda^{2}|u-x|^{2}+2n\Lambda|u-x|\right),\end{split}

and, if j=n+1,,2nmj=n+1,\ldots,2n-m, then

|vj|=2.\left|v_{j}\right|=\sqrt{2}.

We conclude that

|v1v2nm|2nm(1+2n2Λ2|ux|2)m2(1+n2Λ2|ux|2+2nΛ|ux|)nm22nm(1+2n2Λ2|ux|2+2nΛ|ux|)n2|v_{1}\wedge\ldots\wedge v_{2n-m}|\leqslant 2^{n-m}\left(1+2n^{2}\Lambda^{2}|u-x|^{2}\right)^{\frac{m}{2}}\left(1+n^{2}\Lambda^{2}|u-x|^{2}+2n\Lambda|u-x|\right)^{\frac{n-m}{2}}\\ \leqslant 2^{n-m}\left(1+2n^{2}\Lambda^{2}|u-x|^{2}+2n\Lambda|u-x|\right)^{\frac{n}{2}}

and the proof is complete. ∎

5.12Definition of 𝒴E𝐖\mathscr{Y}_{E}\mathbf{W}. —

It follows from the coarea theorem that the function

n×nm[0,]:(u,y)m(Eg𝐯1,,𝐯nm,u1{y})\mathbb{R}^{n}\times\mathbb{R}^{n-m}\to[0,\infty]:(u,y)\to\mathscr{H}^{m}\left(E\cap g_{\mathbf{v}_{1},\ldots,\mathbf{v}_{n-m},u}^{-1}\{y\}\right)

is nnm\mathscr{L}^{n}\otimes\mathscr{L}^{n-m}-measurable (recall 5.10 applied with B=nB=\mathbb{R}^{n}). It now follows from Fubini’s theorem that, for each r>0r>0, the function

n[0,]:uCrm(Eg𝐯1,,𝐯nm,u1{y})𝑑nm(y)\mathbb{R}^{n}\to[0,\infty]:u\mapsto\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int_{C_{r}}\mathscr{H}^{m}\left(E\cap g_{\mathbf{v}_{1},\ldots,\mathbf{v}_{n-m},u}^{-1}\{y\}\right)d\mathscr{L}^{n-m}(y)

is n\mathscr{L}^{n}-measurable. In turn, the function

𝒴E𝐖:n[0,]:ulim infjCj1m(Eg𝐯1,,𝐯nm,u1{y})𝑑nm(y)\mathscr{Y}_{E}\mathbf{W}:\mathbb{R}^{n}\to[0,\infty]:u\mapsto\liminf_{j}\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int_{C_{j^{-1}}}\mathscr{H}^{m}\left(E\cap g_{\mathbf{v}_{1},\ldots,\mathbf{v}_{n-m},u}^{-1}\{y\}\right)d\mathscr{L}^{n-m}(y)

is n\mathscr{L}^{n}-measurable. It is a replacement for 𝒴E0𝐖\mathscr{Y}^{0}_{E}\mathbf{W} defined in 5.7. We shall establish, for 𝒵E𝐖\mathscr{Z}_{E}\mathbf{W}, a similar lower bound to that in 5.8, this time involving 𝒴E𝐖\mathscr{Y}_{E}\mathbf{W}. Before doing so, we notice the rather trivial fact that if FEF\subseteq E, then

𝒴F𝐖(u)𝒴E𝐖(u),\mathscr{Y}_{F}\mathbf{W}(u)\leqslant\mathscr{Y}_{E}\mathbf{W}(u),

for all unu\in\mathbb{R}^{n}.

5.13Preparatory remark for the proof of 5.15. —

It follows from the coarea theorem that the function

n×nm[0,]:(x,y)m(B(𝐖(x)+i=1nmyi𝐯i(x)))\mathbb{R}^{n}\times\mathbb{R}^{n-m}\to[0,\infty]:(x,y)\mapsto\mathscr{H}^{m}\left(B\cap\left(\mathbf{W}(x)+\sum_{i=1}^{n-m}y_{i}\mathbf{v}_{i}(x)\right)\right)

is nnm\mathscr{L}^{n}\otimes\mathscr{L}^{n-m}-measurable (recall 5.10 applied with B=nB=\mathbb{R}^{n}). It therefore follows from Fubini’s theorem as in 5.12 that

fj:n[0,]:xCj1m(B(𝐖(x)+i=1nmyi𝐯i(x)))𝑑nm(y)f_{j}:\mathbb{R}^{n}\to[0,\infty]:x\mapsto\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int_{C_{j^{-1}}}\mathscr{H}^{m}\left(B\cap\left(\mathbf{W}(x)+\sum_{i=1}^{n-m}y_{i}\mathbf{v}_{i}(x)\right)\right)d\mathscr{L}^{n-m}(y)

is n\mathscr{L}^{n}-measurable. Furthermore, if BB is bounded, then |fj(x)|𝜶(m)(diamB)m|f_{j}(x)|\leqslant\boldsymbol{\alpha}(m)(\operatorname{\mathrm{diam}}B)^{m}, for every xnx\in\mathbb{R}^{n}.

5.14. —

If BB is compact, then, for every xE,x\in E, the function

nm+:ym(B(𝐖(x)+i=1nmyi𝐯i(x)))\mathbb{R}^{n-m}\to\mathbb{R}_{+}:y\mapsto\mathscr{H}^{m}\left(B\cap\left(\mathbf{W}(x)+\sum_{i=1}^{n-m}y_{i}\mathbf{v}_{i}(x)\right)\right)

is upper semicontinuous.

Proof.

The proof is analogous to that of 3.14. For each ynmy\in\mathbb{R}^{n-m}, define the compact set Ky=B(𝐖(x)+i=1nmyi𝐯i(x))K_{y}=B\cap\left(\mathbf{W}(x)+\sum_{i=1}^{n-m}y_{i}\mathbf{v}_{i}(x)\right). If (yk)k(y_{k})_{k} is a sequence converging to yy, we ought to show that

m(Ky)lim supkm(Kyk).\mathscr{H}^{m}_{\infty}\left(K_{y}\right)\geqslant\limsup_{k}\mathscr{H}^{m}_{\infty}\left(K_{y_{k}}\right).

Since each KyK_{y} is a subset of an mm-dimensional affine subspace of n\mathbb{R}^{n}, this is indeed equivalent to the same inequality with m\mathscr{H}^{m}_{\infty} replaced by m\mathscr{H}^{m}, according to 3.2(3). Considering, if necessary, a subsequence of (yk)k(y_{k})_{k}, we may assume that none of the compact sets KykK_{y_{k}} is empty and that the above limit superior is a limit. Considering yet a further subsequence, we may now assume that (Kyk)k(K_{y_{k}})_{k} converges in Hausdorff distance to some compact set LBL\subseteq B. One checks that LKyL\subseteq K_{y}. It then follows from 3.2(1) that m(Ky)m(L)lim supkm(Kyk)\mathscr{H}^{m}_{\infty}\left(K_{y}\right)\geqslant\mathscr{H}^{m}_{\infty}(L)\geqslant\limsup_{k}\mathscr{H}^{m}_{\infty}\left(K_{y_{k}}\right). ∎

5.15 Proposition. —

Given 0<ε<10<\varepsilon<1, there exists 𝛅5.15(n,Λ,ε)>0\boldsymbol{\delta}_{5.15}(n,\Lambda,\varepsilon)>0 with the following property. If diam(EB)𝛅5.15(n,Λ,ε)\operatorname{\mathrm{diam}}(E\cup B)\leqslant\boldsymbol{\delta}_{5.15}(n,\Lambda,\varepsilon) and BB is compact, then

Em(B𝐖(x))𝑑n(x)(1ε)2(nm)B𝒴E𝐖(u)𝑑n(u).\int_{E}\mathscr{H}^{m}\left(B\cap\mathbf{W}(x)\right)d\mathscr{L}^{n}(x)\geqslant(1-\varepsilon)2^{-(n-m)}\int_{B}\mathscr{Y}_{E}\mathbf{W}(u)d\mathscr{L}^{n}(u).
Proof.

We first observe that we can choose 𝜹5.15(n,Λ,ε)>0\boldsymbol{\delta}_{\ref{lb.1}}(n,\Lambda,\varepsilon)>0 small enough so that

JΣ^r(π1×π3)(x,u,y)(1ε)2(nm),J_{\hat{\Sigma}_{r}}(\pi_{1}\times\pi_{3})(x,u,y)\geqslant(1-\varepsilon)2^{-(n-m)}, (23)

for 2n\mathscr{H}^{2n}-almost every (x,u,y)Σ^r(x,u,y)\in\hat{\Sigma}_{r}, provided |ux|𝜹5.15(n,Λ,ε)|u-x|\leqslant\boldsymbol{\delta}_{\ref{lb.1}}(n,\Lambda,\varepsilon), according to 5.11. Thus, (23) holds, for 2n\mathscr{H}^{2n}-almost every (x,u,y)Σ^r,B(x,u,y)\in\hat{\Sigma}_{r,B}, under the assumption that diam(EB)𝜹5.15(n,Λ,ε)\operatorname{\mathrm{diam}}(E\cup B)\leqslant\boldsymbol{\delta}_{\ref{lb.1}}(n,\Lambda,\varepsilon). In that case, (20), (21), and 5.11 imply that

E𝑑n(x)Crm(B(𝐖(x)+i=1nmyi𝐯i(x)))𝑑nm(y)=Σ^r,BJΣ^r(π1×π3)𝑑2n(1ε)2(nm)2n(Σ^r,B)(1ε)2(nm)Σ^r,BJΣ^r(π2×π3)𝑑2n=(1ε)2(nm)B𝑑n(u)Crm(Eg𝐯1,,𝐯nm,u1{y})𝑑nm(y).\begin{split}\int_{E}d\mathscr{L}^{n}(x)&\int_{C_{r}}\mathscr{H}^{m}\left(B\cap\left(\mathbf{W}(x)+\sum_{i=1}^{n-m}y_{i}\mathbf{v}_{i}(x)\right)\right)d\mathscr{L}^{n-m}(y)\\ &=\int_{\hat{\Sigma}_{r,B}}J_{\hat{\Sigma}_{r}}(\pi_{1}\times\pi_{3})d\mathscr{H}^{2n}\\ &\geqslant(1-\varepsilon)2^{-(n-m)}\mathscr{H}^{2n}\left(\hat{\Sigma}_{r,B}\right)\\ &\geqslant(1-\varepsilon)2^{-(n-m)}\int_{\hat{\Sigma}_{r,B}}J_{\hat{\Sigma}_{r}}(\pi_{2}\times\pi_{3})d\mathscr{H}^{2n}\\ &=(1-\varepsilon)2^{-(n-m)}\int_{B}d\mathscr{L}^{n}(u)\int_{C_{r}}\mathscr{H}^{m}\left(E\cap g_{\mathbf{v}_{1},\ldots,\mathbf{v}_{n-m},u}^{-1}\{y\}\right)d\mathscr{L}^{n-m}(y).\end{split} (24)

Fix xEx\in E and β>0\beta>0. According to 5.14, there exists a positive integer j(x,β)j(x,\beta) such that if jj(x,β)j\geqslant j(x,\beta), then

m(B𝐖(x))+βsupyCj1m(B(𝐖(x)+i=1nmyi𝐯i(x)))Cj1m(B(𝐖(x)+i=1nmyi𝐯i(x)))𝑑nm(y).\begin{split}\mathscr{H}^{m}\left(B\cap\mathbf{W}(x)\right)+\beta&\geqslant\sup_{y\in C_{j^{-1}}}\mathscr{H}^{m}\left(B\cap\left(\mathbf{W}(x)+\sum_{i=1}^{n-m}y_{i}\mathbf{v}_{i}(x)\right)\right)\\ &\geqslant\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int_{C_{j^{-1}}}\mathscr{H}^{m}\left(B\cap\left(\mathbf{W}(x)+\sum_{i=1}^{n-m}y_{i}\mathbf{v}_{i}(x)\right)\right)d\mathscr{L}^{n-m}(y).\end{split}

Taking the limit superior as jj\to\infty, on the right hand side, and letting β0\beta\to 0, we obtain

m(B𝐖(x))lim supjCj1m(B(𝐖(x)+i=1nmyi𝐯i(x)))𝑑nm(y).\mathscr{H}^{m}\left(B\cap\mathbf{W}(x)\right)\geqslant\limsup_{j}\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int_{C_{j^{-1}}}\mathscr{H}^{m}\left(B\cap\left(\mathbf{W}(x)+\sum_{i=1}^{n-m}y_{i}\mathbf{v}_{i}(x)\right)\right)d\mathscr{L}^{n-m}(y). (25)

As this holds for all xEx\in E, we may integrate over EE with respect to n\mathscr{L}^{n}. We notice that for every j=1,2,j=1,2,\ldots, |fj|𝜶(m)(diamB)m𝟙n|f_{j}|\leqslant\boldsymbol{\alpha}(m)(\operatorname{\mathrm{diam}}B)^{m}\mathbbm{1}_{\mathbb{R}^{n}} (recall the notation of 5.13); the latter being n    E\mathscr{L}^{n}\hskip 2.5pt{\vrule height=7.0pt,width=0.5pt,depth=0.0pt}\hskip-0.2pt\vbox{\hrule height=0.5pt,width=7.0pt,depth=0.0pt}\,E-summable, this justifies the application of the reverse Fatou lemma below. Thus, the following ensues from (25), the reverse Fatou lemma, (24), and the Fatou lemma:

Em(B𝐖(x))dn(x)E𝑑n(x)lim supjCj1m(B(𝐖(x)+i=1nmyi𝐯i(x)))𝑑nm(y)lim supjE𝑑n(x)Cj1m(B(𝐖(x)+i=1nmyi𝐯i(x)))𝑑nm(y)(1ε)2(nm)lim supjB𝑑n(u)Cj1m(Eg𝐯1,,𝐯nm,u1{y})𝑑nm(y)(1ε)2(nm)lim infjB𝑑n(u)Cj1m(Eg𝐯1,,𝐯nm,u1{y})𝑑nm(y)(1ε)2(nm)B𝑑n(u)lim infjCj1m(Eg𝐯1,,𝐯nm,u1{y})𝑑nm(y)=(1ε)2(nm)B𝒴E𝐖(u)𝑑n(u).\begin{split}\int_{E}\mathscr{H}^{m}&\left(B\cap\mathbf{W}(x)\right)d\mathscr{L}^{n}(x)\\ &\geqslant\int_{E}d\mathscr{L}^{n}(x)\limsup_{j}\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int_{C_{j^{-1}}}\mathscr{H}^{m}\left(B\cap\left(\mathbf{W}(x)+\sum_{i=1}^{n-m}y_{i}\mathbf{v}_{i}(x)\right)\right)d\mathscr{L}^{n-m}(y)\\ &\geqslant\limsup_{j}\int_{E}d\mathscr{L}^{n}(x)\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int_{C_{j^{-1}}}\mathscr{H}^{m}\left(B\cap\left(\mathbf{W}(x)+\sum_{i=1}^{n-m}y_{i}\mathbf{v}_{i}(x)\right)\right)d\mathscr{L}^{n-m}(y)\\ &\geqslant(1-\varepsilon)2^{-(n-m)}\limsup_{j}\int_{B}d\mathscr{L}^{n}(u)\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int_{C_{j^{-1}}}\mathscr{H}^{m}\left(E\cap g_{\mathbf{v}_{1},\ldots,\mathbf{v}_{n-m},u}^{-1}\{y\}\right)d\mathscr{L}^{n-m}(y)\\ &\geqslant(1-\varepsilon)2^{-(n-m)}\liminf_{j}\int_{B}d\mathscr{L}^{n}(u)\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int_{C_{j^{-1}}}\mathscr{H}^{m}\left(E\cap g_{\mathbf{v}_{1},\ldots,\mathbf{v}_{n-m},u}^{-1}\{y\}\right)d\mathscr{L}^{n-m}(y)\\ &\geqslant(1-\varepsilon)2^{-(n-m)}\int_{B}d\mathscr{L}^{n}(u)\liminf_{j}\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int_{C_{j^{-1}}}\mathscr{H}^{m}\left(E\cap g_{\mathbf{v}_{1},\ldots,\mathbf{v}_{n-m},u}^{-1}\{y\}\right)d\mathscr{L}^{n-m}(y)\\ &=(1-\varepsilon)2^{-(n-m)}\int_{B}\mathscr{Y}_{E}\mathbf{W}(u)d\mathscr{L}^{n}(u).\end{split}

5.16 Corollary. —

If 0<ε<10<\varepsilon<1 and diamE𝛅5.15(n,Λ,ε)\operatorname{\mathrm{diam}}E\leqslant\boldsymbol{\delta}_{\ref{lb.1}}(n,\Lambda,\varepsilon), then

𝒵E𝐖(u)(1ε)2(nm)𝒴E𝐖(u),\mathscr{Z}_{E}\mathbf{W}(u)\geqslant(1-\varepsilon)2^{-(n-m)}\mathscr{Y}_{E}\mathbf{W}(u),

for n\mathscr{L}^{n}-almost every uEu\in E.

6. Upper bound for 𝒴E𝐖\mathscr{Y}_{E}\mathbf{W} and 𝒵E𝐖\mathscr{Z}_{E}\mathbf{W}

6.1Bow tie lemma. —

Let SnS\subseteq\mathbb{R}^{n}, W𝔾(n,m)W\in\mathbb{G}(n,m), and 0<τ<10<\tau<1. Assume that

(xS)( 0<ρdiamS):S𝐁(x,ρ)𝐁(x+W,τρ).(\forall\,x\in S)(\forall\,0<\rho\leqslant\operatorname{\mathrm{diam}}S):S\cap\mathbf{B}(x,\rho)\subseteq\mathbf{B}(x+W,\tau\rho).

Then there exists F:PW(S)nF:P_{W}(S)\to\mathbb{R}^{n} such that S=imFS=\operatorname{\mathrm{im}}F and LipF11τ2\operatorname{\mathrm{Lip}}F\leqslant\frac{1}{\sqrt{1-\tau^{2}}}. In particular,

m(S)(11τ2)m𝜶(m)(diamS)m.\mathscr{H}^{m}(S)\leqslant\left(\frac{1}{\sqrt{1-\tau^{2}}}\right)^{m}\boldsymbol{\alpha}(m)(\operatorname{\mathrm{diam}}S)^{m}.
Proof.

Let x,xSx,x^{\prime}\in S and define ρ=|xx|diamS\rho=|x-x^{\prime}|\leqslant\operatorname{\mathrm{diam}}S. Thus, xS𝐁(x,ρ)x^{\prime}\in S\cap\mathbf{B}(x,\rho) and, therefore, |PW(xx)|τρ=τ|xx|\left|P_{W^{\perp}}(x-x^{\prime})\right|\leqslant\tau\rho=\tau|x-x^{\prime}|. Since |xx|2=|PW(xx)|2+|PW(xx)|2|x-x^{\prime}|^{2}=\left|P_{W}(x-x^{\prime})\right|^{2}+\left|P_{W^{\perp}}(x-x^{\prime})\right|^{2}, we infer that

(1τ2)|xx|2|PW(xx)|2.(1-\tau^{2})\left|x-x^{\prime}\right|^{2}\leqslant\left|P_{W}(x-x^{\prime})\right|^{2}.

Therefore, PW|SP_{W}|_{S} is injective; and the Lipschitzian bound on F=(PW|S)1F=\left(P_{W}|_{S}\right)^{-1} clearly follows from the above inequality. Regarding the second conclusion, we note that

m(S)=m(F(PW(S)))(LipF)mm(PW(S))\mathscr{H}^{m}(S)=\mathscr{H}^{m}\left(F(P_{W}(S))\right)\leqslant\left(\operatorname{\mathrm{Lip}}F\right)^{m}\mathscr{H}^{m}\left(P_{W}(S)\right)

and PW(S)P_{W}(S) is contained in a ball of radius diamPW(S)diamS\operatorname{\mathrm{diam}}P_{W}(S)\leqslant\operatorname{\mathrm{diam}}S. ∎

6.2. —

Given 0<τ<10<\tau<1, there exists 𝛅6.2(n,Λ,τ)>0\boldsymbol{\delta}_{6.2}(n,\Lambda,\tau)>0 with the following property. If

  1. (1)

    x0Ux_{0}\in U and unu\in\mathbb{R}^{n};

  2. (2)

    diam(E{x0}{u})𝜹6.2(n,Λ,τ)\operatorname{\mathrm{diam}}\left(E\cup\{x_{0}\}\cup\{u\}\right)\leqslant\boldsymbol{\delta}_{6.2}(n,\Lambda,\tau);

then: For each ynmy\in\mathbb{R}^{n-m}, each xEg𝐯1,,𝐯nm,u1{y}x\in E\cap g_{\mathbf{v}_{1},\ldots,\mathbf{v}_{n-m},u}^{-1}\{y\}, and each 0<ρ<0<\rho<\infty, one has

Eg𝐯1,,𝐯nm,u1{y}𝐁(x,ρ)𝐁(x+𝐖0(x0),τρ).E\cap g_{\mathbf{v}_{1},\ldots,\mathbf{v}_{n-m},u}^{-1}\{y\}\cap\mathbf{B}(x,\rho)\subseteq\mathbf{B}\left(x+\mathbf{W}_{0}(x_{0}),\tau\rho\right).
Proof.

We show that 𝜹6.2(n,Λ,τ)=τ2Λn\boldsymbol{\delta}_{\ref{ub.1}}(n,\Lambda,\tau)=\frac{\tau}{2\Lambda\sqrt{n}} will do. Let x,xEg𝐯1,,𝐯nm,u1{y}x,x^{\prime}\in E\cap g_{\mathbf{v}_{1},\ldots,\mathbf{v}_{n-m},u}^{-1}\{y\}, for some ynmy\in\mathbb{R}^{n-m}. Thus, g𝐯1,,𝐯nm,u(x)=g𝐯1,,𝐯nm,u(x)g_{\mathbf{v}_{1},\ldots,\mathbf{v}_{n-m},u}(x)=g_{\mathbf{v}_{1},\ldots,\mathbf{v}_{n-m},u}(x^{\prime}) and, hence,

0=|g𝐯1,,𝐯nm,u(x)g𝐯1,,𝐯nm,u(x)|=i=1nm|𝐯i(x),xu𝐯i(x),xu|2=i=1nm|𝐯i(x),xx𝐯i(x)𝐯i(x),xu|2i=1nm|𝐯i(x),xx|2i=1nm|𝐯i(x)𝐯i(x),xu|2,0=\left|g_{\mathbf{v}_{1},\ldots,\mathbf{v}_{n-m},u}(x)-g_{\mathbf{v}_{1},\ldots,\mathbf{v}_{n-m},u}(x^{\prime})\right|=\sqrt{\sum_{i=1}^{n-m}\left|\langle\mathbf{v}_{i}(x),x-u\rangle-\langle\mathbf{v}_{i}(x^{\prime}),x^{\prime}-u\rangle\right|^{2}}\\ =\sqrt{\sum_{i=1}^{n-m}\left|\langle\mathbf{v}_{i}(x),x-x^{\prime}\rangle-\langle\mathbf{v}_{i}(x^{\prime})-\mathbf{v}_{i}(x),x^{\prime}-u\rangle\right|^{2}}\\ \geqslant\sqrt{\sum_{i=1}^{n-m}\left|\langle\mathbf{v}_{i}(x),x-x^{\prime}\rangle\right|^{2}}-\sqrt{\sum_{i=1}^{n-m}\left|\langle\mathbf{v}_{i}(x^{\prime})-\mathbf{v}_{i}(x),x^{\prime}-u\rangle\right|^{2}},

thus,

i=1nm|𝐯i(x),xx|2i=1nm|𝐯i(x)𝐯i(x),xu|2nmΛ|xx||xu|τ2|xx|.\sqrt{\sum_{i=1}^{n-m}\left|\langle\mathbf{v}_{i}(x),x-x^{\prime}\rangle\right|^{2}}\leqslant\sqrt{\sum_{i=1}^{n-m}\left|\langle\mathbf{v}_{i}(x^{\prime})-\mathbf{v}_{i}(x),x^{\prime}-u\rangle\right|^{2}}\\ \leqslant\sqrt{n-m}\Lambda|x-x^{\prime}||x^{\prime}-u|\leqslant\frac{\tau}{2}|x-x^{\prime}|.

In turn,

|P𝐖0(x0)(xx)|=i=1nm|𝐯i(x0),xx|2i=1nm|𝐯i(x),xx|2+i=1nm|𝐯i(x)𝐯i(x0),xx|2τ2|xx|+nmΛ|xx0||xx|τ|xx|.\left|P_{\mathbf{W}_{0}(x_{0})^{\perp}}(x-x^{\prime})\right|=\sqrt{\sum_{i=1}^{n-m}\left|\langle\mathbf{v}_{i}(x_{0}),x-x^{\prime}\rangle\right|^{2}}\\ \leqslant\sqrt{\sum_{i=1}^{n-m}\left|\langle\mathbf{v}_{i}(x^{\prime}),x-x^{\prime}\rangle\right|^{2}}+\sqrt{\sum_{i=1}^{n-m}\left|\langle\mathbf{v}_{i}(x^{\prime})-\mathbf{v}_{i}(x_{0}),x-x^{\prime}\rangle\right|^{2}}\\ \leqslant\frac{\tau}{2}|x-x^{\prime}|+\sqrt{n-m}\Lambda|x^{\prime}-x_{0}||x-x^{\prime}|\leqslant\tau|x-x^{\prime}|.

6.3 Proposition. —

There are 𝛅6.3(n,Λ)>0\boldsymbol{\delta}_{6.3}(n,\Lambda)>0 and 𝐜6.3(m)1\mathbf{c}_{6.3}(m)\geqslant 1 with the following property. If unu\in\mathbb{R}^{n} and diam(E{u})𝛅6.3(n,Λ)\operatorname{\mathrm{diam}}(E\cup\{u\})\leqslant\boldsymbol{\delta}_{6.3}(n,\Lambda), then

max{𝒴E0𝐖(u),𝒴E𝐖(u)}𝐜6.3(m)(diamE)m.\max\left\{\mathscr{Y}^{0}_{E}\mathbf{W}(u),\mathscr{Y}_{E}\mathbf{W}(u)\right\}\leqslant\mathbf{c}_{6.3}(m)(\operatorname{\mathrm{diam}}E)^{m}.
Proof.

Let 𝜹6.3(n,Λ)=𝜹6.2(n,Λ,1/2)\boldsymbol{\delta}_{\ref{upper.bound}}(n,\Lambda)=\boldsymbol{\delta}_{\ref{ub.1}}(n,\Lambda,1/2). Recall the definitions of 𝒴E0𝐖\mathscr{Y}^{0}_{E}\mathbf{W} and 𝒴E𝐖\mathscr{Y}_{E}\mathbf{W}, 5.7 and 5.12, respectively. If E=E=\emptyset, the conclusion is obvious. If not, pick x0Ex_{0}\in E arbitrarily. Given any ynmy\in\mathbb{R}^{n-m}, we see that 6.2 applies with τ=1/2\tau=1/2 and, in turn, the bow tie lemma 6.1 applies to S=Eg𝐯1,,𝐯nm,u1{y}S=E\cap g_{\mathbf{v}_{1},\ldots,\mathbf{v}_{n-m},u}^{-1}\{y\} and W=𝐖0(x0)W=\mathbf{W}_{0}(x_{0}). Thus,

m(Eg𝐯1,,𝐯nm,u1{y})(23)m𝜶(m)(diamE)m.\mathscr{H}^{m}\left(E\cap g_{\mathbf{v}_{1},\ldots,\mathbf{v}_{n-m},u}^{-1}\{y\}\right)\leqslant\left(\frac{2}{\sqrt{3}}\right)^{m}\boldsymbol{\alpha}(m)(\operatorname{\mathrm{diam}}E)^{m}.

The proposition is proved. ∎

6.4 Corollary. —

There are 𝛅6.4(n,Λ)>0\boldsymbol{\delta}_{6.4}(n,\Lambda)>0 and 𝐜6.4(n)1\mathbf{c}_{6.4}(n)\geqslant 1 with the following property. If diamE𝛅6.4(n,Λ)\operatorname{\mathrm{diam}}E\leqslant\boldsymbol{\delta}_{6.4}(n,\Lambda), then

𝒵E𝐖(u)𝐜6.4(n)(diamE)m,\mathscr{Z}_{E}\mathbf{W}(u)\leqslant\mathbf{c}_{6.4}(n)(\operatorname{\mathrm{diam}}E)^{m},

for n\mathscr{L}^{n}-almost every uEu\in E.

Proof.

Let 𝜹6.4(n,Λ)=min{𝜹6.3(n,Λ),𝜹5.8(n,Λ,1/2)}\boldsymbol{\delta}_{\ref{cor.ub}}(n,\Lambda)=\min\{\boldsymbol{\delta}_{\ref{upper.bound}}(n,\Lambda),\boldsymbol{\delta}_{\ref{Z.1}}(n,\Lambda,1/2)\}. ∎

7. Lower bound for 𝒴E𝐖\mathscr{Y}_{E}\mathbf{W} and 𝒵E𝐖\mathscr{Z}_{E}\mathbf{W}

7.1Setting for this section. —

We enforce again the exact same assumptions as in 4.1; and as in 5.10, we let Cr=nm{y:|y|r}C_{r}=\mathbb{R}^{n-m}\cap\{y:|y|\leqslant r\}.

7.2Polyballs. —

Given x0nx_{0}\in\mathbb{R}^{n} and r>0r>0, we define

𝐂𝐖(x0,r)=n{x:|P𝐖0(x0)(xx0)|r and |P𝐖0(x0)(xx0)|r}.\mathbf{C}_{\mathbf{W}}(x_{0},r)=\mathbb{R}^{n}\cap\left\{x:\left|P_{\mathbf{W}_{0}(x_{0})}(x-x_{0})\right|\leqslant r\text{ and }\left|P_{\mathbf{W}_{0}(x_{0})^{\perp}}(x-x_{0})\right|\leqslant r\right\}.

We notice that, if x𝐂𝐖(x0,r)x\in\mathbf{C}_{\mathbf{W}}(x_{0},r), then |xx0|r2|x-x_{0}|\leqslant r\sqrt{2}; in particular, diam𝐂𝐖(x0,r)22r\operatorname{\mathrm{diam}}\mathbf{C}_{\mathbf{W}}(x_{0},r)\leqslant 2\sqrt{2}r. We also notice that n(𝐂𝐖(x0,r))=𝜶(m)𝜶(nm)rn\mathscr{L}^{n}\left(\mathbf{C}_{\mathbf{W}}(x_{0},r)\right)=\boldsymbol{\alpha}(m)\boldsymbol{\alpha}(n-m)r^{n}.

7.3. —

Given 0<ε<1/30<\varepsilon<1/3, there exists 𝛅7.3(n,Λ,ε)>0\boldsymbol{\delta}_{7.3}(n,\Lambda,\varepsilon)>0 with the following property. If

  1. (1)

    0<r<𝜹7.3(n,Λ,ε)0<r<\boldsymbol{\delta}_{7.3}(n,\Lambda,\varepsilon);

  2. (2)

    u𝐂𝐖(x0,r)Uu\in\mathbf{C}_{\mathbf{W}}(x_{0},r)\subseteq U;

  3. (3)

    |g𝐯1,,𝐯nm,u(x0)|(13ε)r|g_{\mathbf{v}_{1},\ldots,\mathbf{v}_{n-m},u}(x_{0})|\leqslant(1-3\varepsilon)r;

  4. (4)

    CCεrC\subseteq C_{\varepsilon r} is closed;

then

n(𝐂𝐖(x0,r)g𝐯1,,𝐯nm,u1(C))11+ε𝜶(m)rmnm(C).\mathscr{L}^{n}\left(\mathbf{C}_{\mathbf{W}}(x_{0},r)\cap g_{\mathbf{v}_{1},\ldots,\mathbf{v}_{n-m},u}^{-1}(C)\right)\geqslant\frac{1}{1+\varepsilon}\boldsymbol{\alpha}(m)r^{m}\mathscr{L}^{n-m}(C).
7.4 Remark. —

With hopes that the following will help the reader form a geometrical imagery: Under the circumstances 7.3, 𝐂𝐖(x0,r)g𝐯1,,𝐯nm,u1(C)\mathbf{C}_{\mathbf{W}}(x_{0},r)\cap g_{\mathbf{v}_{1},\ldots,\mathbf{v}_{n-m},u}^{-1}(C) may be seen as a “nonlinear stripe”, “horizontal” with respect to 𝐖0(x0)\mathbf{W}_{0}(x_{0}), “at height” g𝐯1,,𝐯nm,u(x0)g_{\mathbf{v}_{1},\ldots,\mathbf{v}_{n-m},u}(x_{0}) with respect to x0x_{0}, and of “width” CC.

Proof of 7.3.

Given z𝐖0(x0)𝐁(0,r)z\in\mathbf{W}_{0}(x_{0})\cap\mathbf{B}(0,r), we define

Vz=n{x0+z+i=1nmyi𝐯i(x0):yCr}𝐂𝐖(x0,r)V_{z}=\mathbb{R}^{n}\cap\left\{x_{0}+z+\sum_{i=1}^{n-m}y_{i}\mathbf{v}_{i}(x_{0}):y\in C_{r}\right\}\subseteq\mathbf{C}_{\mathbf{W}}(x_{0},r)

and we consider the isometric parametrization γz:CrVz\gamma_{z}:C_{r}\to V_{z} defined by the formula

γz(y)=x0+z+i=1nmyi𝐯i(x0).\gamma_{z}(y)=x_{0}+z+\sum_{i=1}^{n-m}y_{i}\mathbf{v}_{i}(x_{0}).

We also abbreviate fz,u=g𝐯1,,𝐯nm,uγzf_{z,u}=g_{\mathbf{v}_{1},\ldots,\mathbf{v}_{n-m},u}\circ\gamma_{z}.

Claim #1. Lipfz,u(1+ε)1nm\operatorname{\mathrm{Lip}}f_{z,u}\leqslant(1+\varepsilon)^{\frac{1}{n-m}}.

Since γz\gamma_{z} is an isometry, it suffices to obtain an upper bound for Lipg𝐯1,,𝐯nm,u|𝐂𝐖(x0,r)\operatorname{\mathrm{Lip}}g_{\mathbf{v}_{1},\ldots,\mathbf{v}_{n-m},u}|_{\mathbf{C}_{\mathbf{W}}(x_{0},r)}. Let x,x𝐂𝐖(x0,r)x,x^{\prime}\in\mathbf{C}_{\mathbf{W}}(x_{0},r) and note that

|g𝐯1,,𝐯nm,u(x)g𝐯1,,𝐯nm,u(x)|=i=1nm|𝐯i(x),xu𝐯i(x),xu|2i=1nm(|𝐯i(x)𝐯i(x),xu|+|𝐯i(x),xx|)2i=1nm|𝐯i(x)𝐯i(x),xu|2+i=1nm|𝐯i(x),xx|2nmΛ|xx||xu|+|P𝐖0(x)(xx)|(1+nmΛ22r)|xx|.\begin{split}\big{|}g_{\mathbf{v}_{1},\ldots,\mathbf{v}_{n-m},u}(x)&-g_{\mathbf{v}_{1},\ldots,\mathbf{v}_{n-m},u}(x^{\prime})\big{|}=\sqrt{\sum_{i=1}^{n-m}\left|\langle\mathbf{v}_{i}(x),x-u\rangle-\langle\mathbf{v}_{i}(x^{\prime}),x^{\prime}-u\rangle\right|^{2}}\\ &\leqslant\sqrt{\sum_{i=1}^{n-m}\left(\left|\langle\mathbf{v}_{i}(x)-\mathbf{v}_{i}(x^{\prime}),x-u\rangle\right|+\left|\langle\mathbf{v}_{i}(x^{\prime}),x-x^{\prime}\rangle\right|\right)^{2}}\\ &\leqslant\sqrt{\sum_{i=1}^{n-m}\left|\langle\mathbf{v}_{i}(x)-\mathbf{v}_{i}(x^{\prime}),x-u\rangle\right|^{2}}+\sqrt{\sum_{i=1}^{n-m}\left|\langle\mathbf{v}_{i}(x^{\prime}),x-x^{\prime}\rangle\right|^{2}}\\ &\leqslant\sqrt{n-m}\Lambda|x-x^{\prime}||x-u|+\left|P_{\mathbf{W}_{0}(x^{\prime})^{\perp}}(x-x^{\prime})\right|\\ &\leqslant\left(1+\sqrt{n-m}\Lambda 2\sqrt{2}r\right)|x-x^{\prime}|.\end{split}

Recalling hypothesis (1), it is now apparent that 𝜹7.3\boldsymbol{\delta}_{\ref{53}} can be chosen small enough, according to nn, Λ\Lambda, and ε\varepsilon, so that Claim #1 holds.

Claim #2. For nm\mathscr{L}^{n-m}-almost every yCry\in C_{r}, one has Dfz,u(y)idnmε\|Df_{z,u}(y)-\operatorname{\mathrm{id}}_{\mathbb{R}^{n-m}}\|\leqslant\varepsilon.

Let yCry\in C_{r} be such that fz,uf_{z,u} is differentiable at yy. We shall estimate the coefficients of the matrix representing Dfz,u(y)Df_{z,u}(y) with respect to the canonical basis. Fix i,j=1,,nmi,j=1,\ldots,n-m and recall (7):

yifu,z(y),ej=yig𝐯1,,𝐯nm,u(γz(y)),ej=(yi)g𝐯j,u(γz(y))=g𝐯j,u(γz(y)),γz(y)yi=D𝐯j(γz(y))(𝐯i(x0)),γz(y)u+𝐯j(γz(y)),𝐯i(x0)=I+II.\begin{split}\frac{\partial}{\partial y_{i}}\langle f_{u,z}(y),e_{j}\rangle&=\frac{\partial}{\partial y_{i}}\langle g_{\mathbf{v}_{1},\ldots,\mathbf{v}_{n-m},u}(\gamma_{z}(y)),e_{j}\rangle\\ &=\left(\frac{\partial}{\partial y_{i}}\right)g_{\mathbf{v}_{j},u}(\gamma_{z}(y))\\ &=\left\langle\nabla g_{\mathbf{v}_{j},u}(\gamma_{z}(y)),\frac{\partial\gamma_{z}(y)}{\partial y_{i}}\right\rangle\\ &=\left\langle D\mathbf{v}_{j}(\gamma_{z}(y))(\mathbf{v}_{i}(x_{0})),\gamma_{z}(y)-u\right\rangle+\left\langle\mathbf{v}_{j}(\gamma_{z}(y)),\mathbf{v}_{i}(x_{0})\right\rangle\\ &=\mathrm{I}+\mathrm{II}.\end{split}

Next, note that

|IIδij|=|II𝐯j(x0),𝐯i(x0)|=|𝐯j(γz(y))𝐯j(x0),𝐯i(x0)|Λ|γz(y)x0|Λ22rε2(nm),\left|\mathrm{II}-\delta_{ij}\right|=\left|\mathrm{II}-\langle\mathbf{v}_{j}(x_{0}),\mathbf{v}_{i}(x_{0})\rangle\right|=\left|\langle\mathbf{v}_{j}(\gamma_{z}(y))-\mathbf{v}_{j}(x_{0}),\mathbf{v}_{i}(x_{0})\rangle\right|\\ \leqslant\Lambda\left|\gamma_{z}(y)-x_{0}\right|\leqslant\Lambda 2\sqrt{2}r\leqslant\frac{\varepsilon}{2(n-m)},

where the last inequality follows from hypothesis (1), upon choosing 𝜹7.3\boldsymbol{\delta}_{\ref{53}} small enough, according to nn, Λ\Lambda, and ε\varepsilon. Moreover,

|I|Λ|γz(y)u|Λ22rε2(nm).\left|\mathrm{I}\right|\leqslant\Lambda|\gamma_{z}(y)-u|\leqslant\Lambda 2\sqrt{2}r\leqslant\frac{\varepsilon}{2(n-m)}.

Therefore, if (aij)i,j=1,,nm(a_{ij})_{i,j=1,\ldots,n-m} is the matrix representing Dfz,u(y)Df_{z,u}(y) with respect to the canonical basis, we have shown that |aijδij|εnm|a_{ij}-\delta_{ij}|\leqslant\frac{\varepsilon}{n-m}, for all i,j=1,,nmi,j=1,\ldots,n-m. This completes the proof of Claim #2.

Claim #3. Cεrfz,u(Cr)C_{\varepsilon r}\subseteq f_{z,u}(C_{r}).

We shall show that |yfz,u(y)|(1ε)r|y-f_{z,u}(y)|\leqslant(1-\varepsilon)r, for every yBdryCry\in\operatorname{\mathrm{Bdry}}C_{r}, and the conclusion will become a consequence of the intermediate value theorem, in case m=n1m=n-1, or a standard application of homology theory, see, e.g., [5, 4.6.1], in case m<n1m<n-1. If m<n1m<n-1, it is clearly enough to establish this inequality only for nm1\mathscr{H}^{n-m-1}-almost every yBdryCry\in\operatorname{\mathrm{Bdry}}C_{r}. In that case, owing to the coarea theorem [9, 3.2.22], we may choose such a yy so that that fz,uf_{z,u} is differentiable 1\mathscr{H}^{1}-almost everywhere on the line segment nm{sy:0s1}\mathbb{R}^{n-m}\cap\{sy:0\leqslant s\leqslant 1\}. Whether m<n1m<n-1, or m=n1m=n-1, it then follows from Claim #2 that

|fz,u(y)fz,u(0)y|=|01Dfz,u(sy)(y)𝑑1(s)y|01|Dfz,u(sy)(y)y|𝑑1(s)ε|y|=εr.\left|f_{z,u}(y)-f_{z,u}(0)-y\right|=\left|\int_{0}^{1}Df_{z,u}(sy)(y)d\mathscr{L}^{1}(s)-y\right|\\ \leqslant\int_{0}^{1}\left|Df_{z,u}(sy)(y)-y\right|d\mathscr{L}^{1}(s)\leqslant\varepsilon|y|=\varepsilon r.

Accordingly,

|fz,u(y)y||fz,u(y)fz,u(0)y|+|fz,u(0)|εr+|fz,u(0)|\left|f_{z,u}(y)-y\right|\leqslant\left|f_{z,u}(y)-f_{z,u}(0)-y\right|+\left|f_{z,u}(0)\right|\leqslant\varepsilon r+\left|f_{z,u}(0)\right|

and the claim will be established upon showing that |fz,u(0)|(12ε)r\left|f_{z,u}(0)\right|\leqslant(1-2\varepsilon)r. Note that fz,u(0)=g𝐯1,,𝐯nm,u(x0+z)f_{z,u}(0)=g_{\mathbf{v}_{1},\ldots,\mathbf{v}_{n-m},u}(x_{0}+z); we shall use hypothesis (3) to bound its norm from above. Given j=1,,nmj=1,\ldots,n-m, recall that 𝐯j(x0),z=0\langle\mathbf{v}_{j}(x_{0}),z\rangle=0, thus,

|g𝐯j,u(x0+z)g𝐯j,u(x0)|=|𝐯j(x0+z),x0+zu𝐯j(x0),x0u|=|𝐯j(x0+z),x0+zu𝐯j(x0),x0+zu|Λ|z||x0+zu|Λr22rεrnm,\left|g_{\mathbf{v}_{j},u}(x_{0}+z)-g_{\mathbf{v}_{j},u}(x_{0})\right|=\left|\langle\mathbf{v}_{j}(x_{0}+z),x_{0}+z-u\rangle-\langle\mathbf{v}_{j}(x_{0}),x_{0}-u\rangle\right|\\ =\left|\langle\mathbf{v}_{j}(x_{0}+z),x_{0}+z-u\rangle-\langle\mathbf{v}_{j}(x_{0}),x_{0}+z-u\rangle\right|\leqslant\Lambda|z||x_{0}+z-u|\\ \leqslant\Lambda r2\sqrt{2}r\leqslant\frac{\varepsilon r}{\sqrt{n-m}},

where the last inequality holds, according to hypothesis (1), provided 𝜹7.3\boldsymbol{\delta}_{\ref{53}} is chosen sufficiently small. In turn,

|fz,u(0)||g𝐯1,,𝐯nm,u(x0+z)g𝐯1,,𝐯nm,u(x0)|+|g𝐯1,,𝐯nm,u(x0)|εr+(13ε)r=(12ε)r,\left|f_{z,u}(0)\right|\leqslant\left|g_{\mathbf{v}_{1},\ldots,\mathbf{v}_{n-m},u}(x_{0}+z)-g_{\mathbf{v}_{1},\ldots,\mathbf{v}_{n-m},u}(x_{0})\right|+\left|g_{\mathbf{v}_{1},\ldots,\mathbf{v}_{n-m},u}(x_{0})\right|\\ \leqslant\varepsilon r+(1-3\varepsilon)r=(1-2\varepsilon)r,

by virtue of hypothesis (3).

Claim #4. For every z𝐖0(x0)𝐁(0,r)z\in\mathbf{W}_{0}(x_{0})\cap\mathbf{B}(0,r) and every closed CCεrC\subseteq C_{\varepsilon r}, one has nm(C)(1+ε)nm(g𝐯1,,𝐯nm,u1(C)Vz)\mathscr{H}^{n-m}(C)\leqslant(1+\varepsilon)\mathscr{H}^{n-m}\left(g_{\mathbf{v}_{1},\ldots,\mathbf{v}_{n-m},u}^{-1}(C)\cap V_{z}\right).

First, notice that

g𝐯1,,𝐯nm,u1(C)Vz=γz(γz1(g𝐯1,,𝐯nm,u1(C)Vz))=γz(fz,u1(C))g_{\mathbf{v}_{1},\ldots,\mathbf{v}_{n-m},u}^{-1}(C)\cap V_{z}=\gamma_{z}\bigg{(}\gamma_{z}^{-1}\left(g_{\mathbf{v}_{1},\ldots,\mathbf{v}_{n-m},u}^{-1}(C)\cap V_{z}\right)\bigg{)}=\gamma_{z}\left(f_{z,u}^{-1}(C)\right)

and, therefore,

nm(g𝐯1,,𝐯nm,u1(C)Vz)=nm(fz,u1(C)),\mathscr{H}^{n-m}\left(g_{\mathbf{v}_{1},\ldots,\mathbf{v}_{n-m},u}^{-1}(C)\cap V_{z}\right)=\mathscr{H}^{n-m}\left(f_{z,u}^{-1}(C)\right),

since γz\gamma_{z} is an isometry. Now, since CCεrfz,u(Cr)C\subseteq C_{\varepsilon r}\subseteq f_{z,u}(C_{r}), according to Claim #3, we have

C=fz,u(fz,u1(C)).C=f_{z,u}\left(f_{z,u}^{-1}(C)\right).

It therefore follows from Claim #1 that

nm(C)(Lipfz,u)nmnm(fz,u1(C))(1+ε)nm(g𝐯1,,𝐯nm,u1(C)Vz).\begin{split}\mathscr{H}^{n-m}(C)&\leqslant\left(\operatorname{\mathrm{Lip}}f_{z,u}\right)^{n-m}\mathscr{H}^{n-m}\left(f_{z,u}^{-1}(C)\right)\\ &\leqslant(1+\varepsilon)\mathscr{H}^{n-m}\left(g_{\mathbf{v}_{1},\ldots,\mathbf{v}_{n-m},u}^{-1}(C)\cap V_{z}\right).\end{split}

We are now ready to finish the proof, by an application of Fubini’s theorem:

n(𝐂𝐖(x0,r)g𝐯1,,𝐯nm,u1(C))=𝐖0(x0)𝐁(0,r)𝑑m(z)nm(g𝐯1,,𝐯nm,u1(C)Vz)11+ε𝜶(m)rmnm(C).\begin{split}\mathscr{L}^{n}\bigg{(}\mathbf{C}_{\mathbf{W}}(x_{0},r)\cap&g_{\mathbf{v}_{1},\ldots,\mathbf{v}_{n-m},u}^{-1}(C)\bigg{)}\\ &=\int_{\mathbf{W}_{0}(x_{0})\cap\mathbf{B}(0,r)}d\mathscr{L}^{m}(z)\mathscr{H}^{n-m}\left(g_{\mathbf{v}_{1},\ldots,\mathbf{v}_{n-m},u}^{-1}(C)\cap V_{z}\right)\\ &\geqslant\frac{1}{1+\varepsilon}\boldsymbol{\alpha}(m)r^{m}\mathscr{H}^{n-m}(C).\end{split}

7.5Lower bound for 𝒴E𝐖\mathscr{Y}_{E}\mathbf{W}. —

Given 0<ε<1/30<\varepsilon<1/3, there exists 𝛅7.5(n,Λ,ε)>0\boldsymbol{\delta}_{7.5}(n,\Lambda,\varepsilon)>0 with the following property. If

  1. (1)

    0<r<𝜹7.5(n,Λ,ε)0<r<\boldsymbol{\delta}_{7.5}(n,\Lambda,\varepsilon);

  2. (2)

    𝐂𝐖(x0,r)U\mathbf{C}_{\mathbf{W}}(x_{0},r)\subseteq U;

  3. (3)

    AUA\subseteq U is closed;

  4. (4)

    n(A𝐂𝐖(x0,r))(1ε)n(𝐂𝐖(x0,r))\mathscr{L}^{n}\left(A\cap\mathbf{C}_{\mathbf{W}}(x_{0},r)\right)\geqslant(1-\varepsilon)\mathscr{L}^{n}\left(\mathbf{C}_{\mathbf{W}}(x_{0},r)\right);

then

𝐂𝐖(x0,r)𝒴A𝐂𝐖(x0,r)𝐖(u)𝑑n(u)(1𝐜7.5(n)ε)𝜶(m)rmn(𝐂𝐖(x0,r)),\int_{\mathbf{C}_{\mathbf{W}}(x_{0},r)}\mathscr{Y}_{A\cap\mathbf{C}_{\mathbf{W}}(x_{0},r)}\mathbf{W}(u)d\mathscr{L}^{n}(u)\geqslant(1-\mathbf{c}_{7.5}(n)\varepsilon)\boldsymbol{\alpha}(m)r^{m}\mathscr{L}^{n}\left(\mathbf{C}_{\mathbf{W}}(x_{0},r)\right),

where 𝐜7.5(n)\mathbf{c}_{7.5}(n) is a positive integer depending only upon nn.

Proof.

Similarly to the proof of 7.3, we will first establish a lower bound for 𝒴A𝐂𝐖(x0,r)𝐖\mathscr{Y}_{A\cap\mathbf{C}_{\mathbf{W}}(x_{0},r)}\mathbf{W} on “vertical slices” VzV_{z} of the given polyball, followed, next, by an application of Fubini. Given z𝐖0(x0)𝐁(0,r)z\in\mathbf{W}_{0}(x_{0})\cap\mathbf{B}(0,r), we let VzV_{z} and γz\gamma_{z} be as in 7.3, and we consider the set

Vˇz=n{x0+z+i=1nmyi𝐯i(x0):yC(13ε)r}\check{V}_{z}=\mathbb{R}^{n}\cap\left\{x_{0}+z+\sum_{i=1}^{n-m}y_{i}\mathbf{v}_{i}(x_{0}):y\in C_{(1-3\varepsilon)r}\right\}

(notice Vˇz\check{V}_{z} is slightly smaller than VzV_{z} used in the proof of 7.3) and the isometric parametrization γˇz:C(13ε)rVˇz\check{\gamma}_{z}:C_{(1-3\varepsilon)r}\to\check{V}_{z} defined by

γˇz(y)=x0+z+i=1nmyi𝐯i(x0).\check{\gamma}_{z}(y)=x_{0}+z+\sum_{i=1}^{n-m}y_{i}\mathbf{v}_{i}(x_{0}).

For part of the proof, we find it convenient to abbreviate E=A𝐂𝐖(x0,r)E=A\cap\mathbf{C}_{\mathbf{W}}(x_{0},r). We also put 𝒴ˇE𝐖=(𝒴E𝐖)γˇz\check{\mathscr{Y}}_{E}\mathbf{W}=\left(\mathscr{Y}_{E}\mathbf{W}\right)\circ\check{\gamma}_{z}.

By definition of 𝒴E𝐖\mathscr{Y}_{E}\mathbf{W}, for each γˇz(y)Vˇz\check{\gamma}_{z}(y)\in\check{V}_{z}, there exists a collection 𝒞y\mathscr{C}_{y} of closed balls in nm\mathbb{R}^{n-m} with the following properties: For every C𝒞yC\in\mathscr{C}_{y}, CC is a ball centered at 0, CCεrC\subseteq C_{\varepsilon r},

𝒴E𝐖(γˇz(y))+εCm(Eg𝐯1,,𝐯nm,γˇz(y)1{h})𝑑nm(h),\mathscr{Y}_{E}\mathbf{W}\left(\check{\gamma}_{z}(y)\right)+\varepsilon\geqslant\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int_{C}\mathscr{H}^{m}\left(E\cap g_{\mathbf{v}_{1},\ldots,\mathbf{v}_{n-m},\check{\gamma}_{z}(y)}^{-1}\{h\}\right)d\mathscr{L}^{n-m}(h),

and inf{diamC:C𝒞y}=0\inf\{\operatorname{\mathrm{diam}}C:C\in\mathscr{C}_{y}\}=0. Furthermore, 𝒴ˇE𝐖\check{\mathscr{Y}}_{E}\mathbf{W} being nm\mathscr{L}^{n-m}-summable, according to 6.3, there exists NC(13ε)rN\subseteq C_{(1-3\varepsilon)r} such that nm(N)=0\mathscr{L}^{n-m}(N)=0 and every yNy\not\in N is a Lebesgue point of 𝒴ˇE𝐖\check{\mathscr{Y}}_{E}\mathbf{W}. For such a yy, we may reduce 𝒞y\mathscr{C}_{y}, if necessary, keeping all the previously stated properties valid, while enforcing also that

y+C𝒴ˇE𝐖𝑑nm+ε(𝒴ˇE𝐖)(y),\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int_{y+C}\check{\mathscr{Y}}_{E}\mathbf{W}d\mathscr{L}^{n-m}+\varepsilon\geqslant\left(\check{\mathscr{Y}}_{E}\mathbf{W}\right)(y),

whenever C𝒞yC\in\mathscr{C}_{y}. We infer that, for each yC(13ε)rNy\in C_{(1-3\varepsilon)r}\setminus N and each C𝒞yC\in\mathscr{C}_{y},

y+C𝒴ˇE𝐖𝑑nm+2εnm(y+C)Cm(Eg𝐯1,,𝐯nm,γˇz(y)1{h})𝑑nm(h).\int_{y+C}\check{\mathscr{Y}}_{E}\mathbf{W}d\mathscr{L}^{n-m}+2\varepsilon\mathscr{L}^{n-m}(y+C)\geqslant\int_{C}\mathscr{H}^{m}\left(E\cap g_{\mathbf{v}_{1},\ldots,\mathbf{v}_{n-m},\check{\gamma}_{z}(y)}^{-1}\{h\}\right)d\mathscr{L}^{n-m}(h). (26)

It ensues from the Vitali covering theorem that there is a sequence (yk)k(y_{k})_{k} in C(13ε)rNC_{(1-3\varepsilon)r}\setminus N, and Ck𝒞ykC_{k}\in\mathscr{C}_{y_{k}}, such that the balls yk+Cky_{k}+C_{k}, k=1,2,,k=1,2,\ldots, are pairwise disjoint and nm(C(13ε)rk=1(yk+Ck))=0\mathscr{L}^{n-m}\left(C_{(1-3\varepsilon)r}\setminus\cup_{k=1}^{\infty}(y_{k}+C_{k})\right)=0. It therefore follows, from (26) and the fact that γz\gamma_{z} is an isometry, that

Vz𝒴E𝐖𝑑nm+2εnm(Vz)k=1Ckm(Eg𝐯1,,𝐯nm,uk1{y})𝑑nm(y),\int_{V_{z}}\mathscr{Y}_{E}\mathbf{W}d\mathscr{H}^{n-m}+2\varepsilon\mathscr{H}^{n-m}(V_{z})\geqslant\sum_{k=1}^{\infty}\int_{C_{k}}\mathscr{H}^{m}\left(E\cap g_{\mathbf{v}_{1},\ldots,\mathbf{v}_{n-m},u_{k}}^{-1}\{y\}\right)d\mathscr{L}^{n-m}(y), (27)

where we have abbreviated uk=γˇz(yk)u_{k}=\check{\gamma}_{z}(y_{k}). We also abbreviate Sk=g𝐯1,,𝐯nm,uk1(Ck)S_{k}=g_{\mathbf{v}_{1},\ldots,\mathbf{v}_{n-m},u_{k}}^{-1}(C_{k}) and we infer from the coarea formula that, for each k=1,2,k=1,2,\ldots,

Ckm(Eg𝐯1,,𝐯nm,uk1{y})𝑑nm(y)=ESkJg𝐯1,,𝐯nm,uk𝑑n(1ε)n(ESk),\int_{C_{k}}\mathscr{H}^{m}\left(E\cap g_{\mathbf{v}_{1},\ldots,\mathbf{v}_{n-m},u_{k}}^{-1}\{y\}\right)d\mathscr{L}^{n-m}(y)=\int_{E\cap S_{k}}Jg_{\mathbf{v}_{1},\ldots,\mathbf{v}_{n-m},u_{k}}d\mathscr{L}^{n}\\ \geqslant(1-\varepsilon)\mathscr{L}^{n}\left(E\cap S_{k}\right), (28)

where the last inequality follows from 3.12, applied with U=Int𝐂𝐖(x0,r)U=\operatorname{\mathrm{Int}}\mathbf{C}_{\mathbf{W}}(x_{0},r), provided that 𝜹7.5(n,Λ,ε)\boldsymbol{\delta}_{\ref{54}}(n,\Lambda,\varepsilon) is chosen smaller than (22)1𝜹3.12(n,Λ,ε)(2\sqrt{2})^{-1}\boldsymbol{\delta}_{\ref{jac.g}}(n,\Lambda,\varepsilon). Letting S=k=1SkS=\cup_{k=1}^{\infty}S_{k} and recalling that E=A𝐂𝐖(x0,r)E=A\cap\mathbf{C}_{\mathbf{W}}(x_{0},r), we infer, from (27) and (28), that

Vz𝒴E𝐖𝑑nm+2εnm(Vz)(1ε)n(ES)(1ε)(n(𝐂𝐖(x0,r)S)n(𝐂𝐖(x0,r)A)).\int_{V_{z}}\mathscr{Y}_{E}\mathbf{W}d\mathscr{H}^{n-m}+2\varepsilon\mathscr{H}^{n-m}(V_{z})\geqslant(1-\varepsilon)\mathscr{L}^{n}(E\cap S)\\ \geqslant(1-\varepsilon)\big{(}\mathscr{L}^{n}(\mathbf{C}_{\mathbf{W}}(x_{0},r)\cap S)-\mathscr{L}^{n}(\mathbf{C}_{\mathbf{W}}(x_{0},r)\setminus A)\big{)}. (29)

Applying 7.3 to each SkS_{k} does not immediately yield a lower bound for n(𝐂𝐖(x0,r)S)\mathscr{L}^{n}(\mathbf{C}_{\mathbf{W}}(x_{0},r)\cap S), because the SkS_{k} are not necessarily pairwise disjoint. This is why we now introduce slightly smaller versions of these:

Cˇk=(1ε)Ck and Sˇk=g𝐯1,,𝐯nm,uk1(Cˇk).\check{C}_{k}=(1-\varepsilon)C_{k}\quad\text{ and }\quad\check{S}_{k}=g_{\mathbf{v}_{1},\ldots,\mathbf{v}_{n-m},u_{k}}^{-1}\left(\check{C}_{k}\right).

Claim. The sets Sˇk𝐂𝐖(x0,r)\check{S}_{k}\cap\mathbf{C}_{\mathbf{W}}(x_{0},r), k=1,2,k=1,2,\ldots, are pairwise disjoint.

Assume, if possible, that there are jkj\neq k and xSˇjSˇk𝐂𝐖(x0,r)x\in\check{S}_{j}\cap\check{S}_{k}\cap\mathbf{C}_{\mathbf{W}}(x_{0},r). Letting ρj\rho_{j} and ρk\rho_{k} denote, respectively, the radius of CjC_{j}, and of CkC_{k}, we notice that ρj+ρk<|yjyk|\rho_{j}+\rho_{k}<|y_{j}-y_{k}|, because (yj+Cj)(yk+Ck)=(y_{j}+C_{j})\cap(y_{k}+C_{k})=\emptyset. Since γˇz\check{\gamma}_{z} is an isometry, we have |ujuk|=|γˇz(yj)γˇz(yk)|=|yjyk||u_{j}-u_{k}|=\left|\check{\gamma}_{z}(y_{j})-\check{\gamma}_{z}(y_{k})\right|=|y_{j}-y_{k}| and, therefore, also

|g𝐯1,,𝐯nm,uj(x)g𝐯1,,𝐯nm,uk(x)||g𝐯1,,𝐯nm,uj(x)|+|g𝐯1,,𝐯nm,uk(x)|(1ε)ρj+(1ε)ρk<(1ε)|ujuk|.\left|g_{\mathbf{v}_{1},\ldots,\mathbf{v}_{n-m},u_{j}}(x)-g_{\mathbf{v}_{1},\ldots,\mathbf{v}_{n-m},u_{k}}(x)\right|\leqslant\left|g_{\mathbf{v}_{1},\ldots,\mathbf{v}_{n-m},u_{j}}(x)\right|+\left|g_{\mathbf{v}_{1},\ldots,\mathbf{v}_{n-m},u_{k}}(x)\right|\\ \leqslant(1-\varepsilon)\rho_{j}+(1-\varepsilon)\rho_{k}<(1-\varepsilon)\left|u_{j}-u_{k}\right|. (30)

We now introduce the following vectors of nm\mathbb{R}^{n-m}:

hj=i=1nm𝐯i(x0),ujei and hk=i=1nm𝐯i(x0),ukeih_{j}=\sum_{i=1}^{n-m}\langle\mathbf{v}_{i}(x_{0}),u_{j}\rangle e_{i}\quad\text{ and }\quad h_{k}=\sum_{i=1}^{n-m}\langle\mathbf{v}_{i}(x_{0}),u_{k}\rangle e_{i}

and we notice that

|hjhk|=|P𝐖0(x0)(ujuk)|=|ujuk|,\left|h_{j}-h_{k}\right|=\left|P_{\mathbf{W}_{0}(x_{0})^{\perp}}(u_{j}-u_{k})\right|=|u_{j}-u_{k}|,

where the second equality holds because ujuk𝐖0(x0)u_{j}-u_{k}\in\mathbf{W}_{0}(x_{0})^{\perp}, as clearly follows from the definition of γˇz\check{\gamma}_{z}. Furthermore,

|(g𝐯1,,𝐯nm,uj(x)g𝐯1,,𝐯nm,uk(x))(hkhj)|=i=1nm|𝐯i(x)𝐯i(x0),ukuj|2nmΛ2r|ujuk|ε|ujuk|,\bigg{|}\big{(}g_{\mathbf{v}_{1},\ldots,\mathbf{v}_{n-m},u_{j}}(x)-g_{\mathbf{v}_{1},\ldots,\mathbf{v}_{n-m},u_{k}}(x)\big{)}-\big{(}h_{k}-h_{j}\big{)}\bigg{|}=\sqrt{\sum_{i=1}^{n-m}\left|\langle\mathbf{v}_{i}(x)-\mathbf{v}_{i}(x_{0}),u_{k}-u_{j}\rangle\right|^{2}}\\ \leqslant\sqrt{n-m}\Lambda\sqrt{2}r\left|u_{j}-u_{k}\right|\leqslant\varepsilon\left|u_{j}-u_{k}\right|,

since we may choose 𝜹7.5(n,Λ,ε)\boldsymbol{\delta}_{\ref{54}}(n,\Lambda,\varepsilon) to be so small that the last inequality holds, in view of hypothesis (1). Whence,

|g𝐯1,,𝐯nm,uj(x)g𝐯1,,𝐯nm,uk(x)||hjhk|ε|ujuk|=(1ε)|ujuk|,\left|g_{\mathbf{v}_{1},\ldots,\mathbf{v}_{n-m},u_{j}}(x)-g_{\mathbf{v}_{1},\ldots,\mathbf{v}_{n-m},u_{k}}(x)\right|\geqslant\left|h_{j}-h_{k}\right|-\varepsilon\left|u_{j}-u_{k}\right|=(1-\varepsilon)\left|u_{j}-u_{k}\right|,

contradicting (30). The Claim is established.

Thus,

n(𝐂𝐖(x0,r)S)=n(𝐂𝐖(x0,r)k=1Sk)n(𝐂𝐖(x0,r)k=1Sˇk)=k=1n(𝐂𝐖(x0,r)Sˇk)=k=1n(𝐂𝐖(x0,r)g𝐯1,,𝐯nm,uk1(Cˇk))11+ε𝜶(m)rmk=1nm(Cˇk),\begin{split}\mathscr{L}^{n}\left(\mathbf{C}_{\mathbf{W}}(x_{0},r)\cap S\right)&=\mathscr{L}^{n}\left(\mathbf{C}_{\mathbf{W}}(x_{0},r)\cap\cup_{k=1}^{\infty}S_{k}\right)\\ &\geqslant\mathscr{L}^{n}\left(\mathbf{C}_{\mathbf{W}}(x_{0},r)\cap\cup_{k=1}^{\infty}\check{S}_{k}\right)\\ &=\sum_{k=1}^{\infty}\mathscr{L}^{n}\left(\mathbf{C}_{\mathbf{W}}(x_{0},r)\cap\check{S}_{k}\right)\\ &=\sum_{k=1}^{\infty}\mathscr{L}^{n}\left(\mathbf{C}_{\mathbf{W}}(x_{0},r)\cap g_{\mathbf{v}_{1},\ldots,\mathbf{v}_{n-m},u_{k}}^{-1}\left(\check{C}_{k}\right)\right)\\ &\geqslant\frac{1}{1+\varepsilon}\boldsymbol{\alpha}(m)r^{m}\sum_{k=1}^{\infty}\mathscr{L}^{n-m}\left(\check{C}_{k}\right),\end{split} (31)

where the last inequality follows from 7.3. We notice that, indeed, 7.3 applies, since CˇkCkCεr\check{C}_{k}\subseteq C_{k}\subseteq C_{\varepsilon r} and |g𝐯1,,𝐯nm,uk(x0)|=|P𝐖0(x0)(ukx0)|=|yk|(13ε)r\left|g_{\mathbf{v}_{1},\ldots,\mathbf{v}_{n-m},u_{k}}(x_{0})\right|=\left|P_{\mathbf{W}_{0}(x_{0})^{\perp}}(u_{k}-x_{0})\right|=\left|y_{k}\right|\leqslant(1-3\varepsilon)r.

Now,

k=1nm(Cˇk)=(1ε)nmk=1nm(Ck)=(1ε)nmk=1nm(yk+Ck)(1ε)nmnm(C(13ε)r)(13ε)2(nm)𝜶(nm)rnm.\sum_{k=1}^{\infty}\mathscr{L}^{n-m}\left(\check{C}_{k}\right)=(1-\varepsilon)^{n-m}\sum_{k=1}^{\infty}\mathscr{L}^{n-m}\left(C_{k}\right)=(1-\varepsilon)^{n-m}\sum_{k=1}^{\infty}\mathscr{L}^{n-m}\left(y_{k}+C_{k}\right)\\ \geqslant(1-\varepsilon)^{n-m}\mathscr{L}^{n-m}\left(C_{(1-3\varepsilon)r}\right)\geqslant(1-3\varepsilon)^{2(n-m)}\boldsymbol{\alpha}(n-m)r^{n-m}. (32)

We infer, from (31) and (32), that

n(𝐂𝐖(x0,r)S)(13ε)2(nm)1+εn(𝐂𝐖(x0,r)).\mathscr{L}^{n}\left(\mathbf{C}_{\mathbf{W}}(x_{0},r)\cap S\right)\geqslant\frac{(1-3\varepsilon)^{2(n-m)}}{1+\varepsilon}\mathscr{L}^{n}\left(\mathbf{C}_{\mathbf{W}}(x_{0},r)\right).

It therefore ensues, from (29) and hypothesis (4), that

Vz𝒴E𝐖𝑑nm+2εnm(Vz)(1ε)((13ε)2(nm)1+εε)n(𝐂𝐖(x0,r)).\int_{V_{z}}\mathscr{Y}_{E}\mathbf{W}d\mathscr{H}^{n-m}+2\varepsilon\mathscr{H}^{n-m}(V_{z})\geqslant(1-\varepsilon)\left(\frac{(1-3\varepsilon)^{2(n-m)}}{1+\varepsilon}-\varepsilon\right)\mathscr{L}^{n}\left(\mathbf{C}_{\mathbf{W}}(x_{0},r)\right).

Integrating over zz, we infer from Fubini’s theorem that

𝐂𝐖(x0,r)𝒴A𝐂𝐖(x0,r)𝐖𝑑n=𝐖0(x0)𝐁(0,r)𝑑n(z)Vz𝒴E𝐖𝑑nm[(1ε)((13ε)2(nm)1+εε)2ε]𝜶(m)rmn(𝐂𝐖(x0,r)).\int_{\mathbf{C}_{\mathbf{W}}(x_{0},r)}\mathscr{Y}_{A\cap\mathbf{C}_{\mathbf{W}}(x_{0},r)}\mathbf{W}d\mathscr{L}^{n}=\int_{\mathbf{W}_{0}(x_{0})\cap\mathbf{B}(0,r)}d\mathscr{L}^{n}(z)\int_{V_{z}}\mathscr{Y}_{E}\mathbf{W}d\mathscr{H}^{n-m}\\ \geqslant\left[(1-\varepsilon)\left(\frac{(1-3\varepsilon)^{2(n-m)}}{1+\varepsilon}-\varepsilon\right)-2\varepsilon\right]\boldsymbol{\alpha}(m)r^{m}\mathscr{L}^{n}\left(\mathbf{C}_{\mathbf{W}}(x_{0},r)\right).

7.6 Proposition. —

Given 0<ε<1/30<\varepsilon<1/3, there exist 𝛅7.6(n,Λ,ε)>0\boldsymbol{\delta}_{7.6}(n,\Lambda,\varepsilon)>0 and 𝐜7.6(n)1\mathbf{c}_{7.6}(n)\geqslant 1 with the following property. If

  1. (1)

    0<r<𝜹7.6(n,Λ,ε)0<r<\boldsymbol{\delta}_{7.6}(n,\Lambda,\varepsilon);

  2. (2)

    𝐂𝐖(x0,r)U\mathbf{C}_{\mathbf{W}}(x_{0},r)\subseteq U;

  3. (3)

    AUA\subseteq U is closed;

  4. (4)

    n(A𝐂𝐖(x0,r))(1ε)n(𝐂𝐖(x0,r))\mathscr{L}^{n}\left(A\cap\mathbf{C}_{\mathbf{W}}(x_{0},r)\right)\geqslant(1-\varepsilon)\mathscr{L}^{n}\left(\mathbf{C}_{\mathbf{W}}(x_{0},r)\right);

then

A𝐂𝐖(x0,r)𝒴A𝐂𝐖(x0,r)𝐖(u)𝑑n(u)(1𝐜7.6(n)ε)𝜶(m)rmn(𝐂𝐖(x0,r)).\int_{A\cap\mathbf{C}_{\mathbf{W}}(x_{0},r)}\mathscr{Y}_{A\cap\mathbf{C}_{\mathbf{W}}(x_{0},r)}\mathbf{W}(u)d\mathscr{L}^{n}(u)\geqslant(1-\mathbf{c}_{7.6}(n)\varepsilon)\boldsymbol{\alpha}(m)r^{m}\mathscr{L}^{n}\left(\mathbf{C}_{\mathbf{W}}(x_{0},r)\right).
7.7 Remark. —

The difference with 7.5 is the domain of integration (being smaller) in the integral, on the left hand side in the conclusion.

Proof of 7.6.

The reader will happily check that

𝜹7.6(n,Λ,ε)=min{𝜹7.5(n,Λ,ε),(22)1𝜹6.3(n,Λ)}\boldsymbol{\delta}_{\ref{lower.bound}}(n,\Lambda,\varepsilon)=\min\left\{\boldsymbol{\delta}_{\ref{54}}(n,\Lambda,\varepsilon),\left(2\sqrt{2}\right)^{-1}\boldsymbol{\delta}_{\ref{upper.bound}}(n,\Lambda)\right\}

suits their needs. ∎

7.8 Proposition. —

There exists 𝛅7.8(n,Λ)>0\boldsymbol{\delta}_{7.8}(n,\Lambda)>0 with the following property. If diamE𝛅7.8(n,Λ)\operatorname{\mathrm{diam}}E\leqslant\boldsymbol{\delta}_{7.8}(n,\Lambda), then

𝒵E𝐖(u)>0,\mathscr{Z}_{E}\mathbf{W}(u)>0,

for n\mathscr{L}^{n}-almost every uEu\in E.

Proof.

We let

𝜹7.8(n,Λ)=min{𝜹7.6(n,Λ,14𝐜7.6(n)),𝜹5.16(n,Λ,1/2)}.\boldsymbol{\delta}_{\ref{Z.positive}}(n,\Lambda)=\min\left\{\boldsymbol{\delta}_{\ref{lower.bound}}\left(n,\Lambda,\frac{1}{4\mathbf{c}_{\ref{lower.bound}}(n)}\right),\boldsymbol{\delta}_{\ref{lb.2}}(n,\Lambda,1/2)\right\}.

According to 5.16, it suffices to show that 𝒴E𝐖(u)>0\mathscr{Y}_{E}\mathbf{W}(u)>0, for n\mathscr{L}^{n}-almost every uEu\in E. Define Z=E{u:𝒴E𝐖(u)=0}Z=E\cap\{u:\mathscr{Y}_{E}\mathbf{W}(u)=0\} and assume, if possible, that n(Z)>0\mathscr{L}^{n}(Z)>0. Since ZZ is n\mathscr{L}^{n}-measurable (recall 5.12), there exists a compact set AZA\subseteq Z such that n(A)>0\mathscr{L}^{n}(A)>0. Observe that the sets 𝐂𝐖(x,r)\mathbf{C}_{\mathbf{W}}(x,r), for xUx\in U and r>0r>0, form a density basis for n\mathscr{L}^{n}-measurable subsets of UU – because their eccentricity is bounded away from zero – thus, there exists x0Ax_{0}\in A and r0>0r_{0}>0 such that

n(A𝐂𝐖(x0,r))(114𝐜7.6(n))n(𝐂𝐖(x0,r)),\mathscr{L}^{n}\left(A\cap\mathbf{C}_{\mathbf{W}}(x_{0},r)\right)\geqslant\left(1-\frac{1}{4\mathbf{c}_{\ref{lower.bound}}(n)}\right)\mathscr{L}^{n}\left(\mathbf{C}_{\mathbf{W}}(x_{0},r)\right),

whenever 0<r<r00<r<r_{0}. There is no restriction to assume that r0r_{0} is small enough for 𝐂𝐖(x0,r0)U\mathbf{C}_{\mathbf{W}}(x_{0},r_{0})\subseteq U. Thus, if we let r=min{r0,𝜹7.6(n,Λ,1/(4𝐜7.6(n)))}r=\min\{r_{0},\boldsymbol{\delta}_{\ref{lower.bound}}(n,\Lambda,1/(4\mathbf{c}_{\ref{lower.bound}(n)}))\}, it follows from 7.6 that

A𝐂𝐖(x0,r)𝒴A𝐂𝐖(x0,r)𝐖(u)𝑑n(u)(114)𝜶(m)rmn(𝐂𝐖(x0,r))>0.\int_{A\cap\mathbf{C}_{\mathbf{W}}(x_{0},r)}\mathscr{Y}_{A\cap\mathbf{C}_{\mathbf{W}}(x_{0},r)}\mathbf{W}(u)d\mathscr{L}^{n}(u)\geqslant\left(1-\frac{1}{4}\right)\boldsymbol{\alpha}(m)r^{m}\mathscr{L}^{n}\left(\mathbf{C}_{\mathbf{W}}(x_{0},r)\right)>0. (33)

On the other hand, recalling 5.12 and the fact that A𝐂𝐖(x0,r)EA\cap\mathbf{C}_{\mathbf{W}}(x_{0},r)\subseteq E, we infer that 𝒴A𝐂𝐖(x0,r)𝐖(u)𝒴E𝐖(u)\mathscr{Y}_{A\cap\mathbf{C}_{\mathbf{W}}(x_{0},r)}\mathbf{W}(u)\leqslant\mathscr{Y}_{E}\mathbf{W}(u), for all unu\in\mathbb{R}^{n}. In particular, 𝒴A𝐂𝐖(x0,r)𝐖(u)=0\mathscr{Y}_{A\cap\mathbf{C}_{\mathbf{W}}(x_{0},r)}\mathbf{W}(u)=0, for all uA𝐂𝐖(x0,r)Zu\in A\cap\mathbf{C}_{\mathbf{W}}(x_{0},r)\subseteq Z, contradicting (33). ∎

8. Proof of the theorems

8.1 Theorem. —

Assume that SnS\subseteq\mathbb{R}^{n}, 𝐖0:S𝔾(n,m)\mathbf{W}_{0}:S\to\mathbb{G}(n,m) is Lipschitzian, and ASA\subseteq S is Borel measurable. The following are equivalent.

  1. (1)

    n(A)=0\mathscr{L}^{n}(A)=0.

  2. (2)

    For n\mathscr{L}^{n} almost every xAx\in A, m(A𝐖(x))=0\mathscr{H}^{m}(A\cap\mathbf{W}(x))=0.

  3. (3)

    For n\mathscr{L}^{n} almost every xSx\in S, m(A𝐖(x))=0\mathscr{H}^{m}(A\cap\mathbf{W}(x))=0.

Recall our convention that 𝐖(x)=x+𝐖0(x)\mathbf{W}(x)=x+\mathbf{W}_{0}(x).

Proof.

Since 𝔾(n,m)\mathbb{G}(n,m) is complete, we can extend 𝐖0\mathbf{W}_{0} to the closure of SS. Furthermore, if the theorem holds for ClosS\operatorname{\mathrm{Clos}}S, then it also holds for SS. Thus, there is no restriction to assume that SS is closed.

(1)(3)(1)\Rightarrow(3). It follows from 3.8 that each xSx\in S admits an open neighborhood UxU_{x} in n\mathbb{R}^{n} such that 𝐖(x)\mathbf{W}(x) can be associated with a Lipschitzian orthonormal frame satisfying all the conditions of 4.1, for some Λx>0\Lambda_{x}>0. Since SS is Lindelöf, there are countably many x1,x2,x_{1},x_{2},\ldots such that SjUxjS\subseteq\cup_{j}U_{x_{j}}. Letting Ej=SUxjE_{j}=S\cap U_{x_{j}}, we infer from 5.5 that ϕEj,𝐖\phi_{E_{j},\mathbf{W}} is absolutely continuous with respect to n\mathscr{L}^{n}. Thus, if n(A)=0\mathscr{L}^{n}(A)=0, then m(A𝐖(x))=0\mathscr{H}^{m}\left(A\cap\mathbf{W}(x)\right)=0, for n\mathscr{L}^{n}-almost every xEjx\in E_{j}, by definition of ϕEj,𝐖\phi_{E_{j},\mathbf{W}}. Since jj is arbitrary, the proof is complete.

(3)(2)(3)\Rightarrow(2) is trivial.

(2)(1)(2)\Rightarrow(1). Let AA satisfy condition (2). It is enough to show that n(A𝐁(0,r))=0\mathscr{L}^{n}(A\cap\mathbf{B}(0,r))=0, for each r>0r>0. Fix r>0r>0 and define Sr=S𝐁(0,r)S_{r}=S\cap\mathbf{B}(0,r). Consider the UxjU_{x_{j}} defined in the second paragraph of the present proof; since SrS_{r} is compact, finitely many of those, say Ux1,,UxNU_{x_{1}},\ldots,U_{x_{N}}, cover SrS_{r}. Let Λ=maxj=1,,NΛxj\Lambda=\max_{j=1,\ldots,N}\Lambda_{x_{j}}. Partition each UxjU_{x_{j}}, j=1,,Nj=1,\ldots,N, into Borel measurable sets Ej,kE_{j,k}, k=1,,Kjk=1,\ldots,K_{j}, such that diamEj,k𝜹7.8(n,Λ)\operatorname{\mathrm{diam}}E_{j,k}\leqslant\boldsymbol{\delta}_{\ref{Z.positive}}(n,\Lambda). It then follows from 7.8 that

(𝒵AEj,k𝐖)(u)>0,\left(\mathscr{Z}_{A\cap E_{j,k}}\mathbf{W}\right)(u)>0, (34)

for n\mathscr{L}^{n}-almost every uAEj,ku\in A\cap E_{j,k}. Now, fix jj and kk. Observe that m(AEj,k𝐖(x))=0\mathscr{H}^{m}\left(A\cap E_{j,k}\cap\mathbf{W}(x)\right)=0, for n\mathscr{L}^{n}-almost every xAEj,kx\in A\cap E_{j,k}. Thus, ϕAEj,k,𝐖(AEj,k)=0\phi_{A\cap E_{j,k},\mathbf{W}}(A\cap E_{j,k})=0. Moreover,

0=ϕAEj,k,𝐖(AEj,k)=AEj,k(𝒵AEj,k𝐖)(u)𝑑n(u).0=\phi_{A\cap E_{j,k},\mathbf{W}}\left(A\cap E_{j,k}\right)=\int_{A\cap E_{j,k}}\left(\mathscr{Z}_{A\cap E_{j,k}}\mathbf{W}\right)(u)d\mathscr{L}^{n}(u).

It follows from (34) that n(AEj,k)=0\mathscr{L}^{n}(A\cap E_{j,k})=0. Since jj and kk are arbitrary, n(A)=0\mathscr{L}^{n}(A)=0. ∎

8.2 Remark. —

Alternatively, one can prove the principal implication (2)(1)(2)\Rightarrow(1) in two other ways. One way – more involved – consists in applying our main result 8.4 below. A second – simpler – way, along the following lines, avoids reference to the estimates we obtained for the functions 𝒴E𝐖\mathscr{Y}_{E}\mathbf{W} and 𝒵E𝐖\mathscr{Z}_{E}\mathbf{W}. Consider xAx\in A, η>0\eta>0, and V𝔾(n,nm)V\in\mathbb{G}(n,n-m) such that d(V,𝐖0(x))<ηd\left(V^{\perp},\mathbf{W}_{0}(x)\right)<\eta, and put Vx=x+VV_{x}=x+V. Define

Φ:(UxVx)×mn:(ξ,t)ξ+i=1mti𝐰i(ξ).\Phi:(U_{x}\cap V_{x})\times\mathbb{R}^{m}\to\mathbb{R}^{n}:(\xi,t)\mapsto\xi+\sum_{i=1}^{m}t_{i}\mathbf{w}_{i}(\xi)\,.

Φ\Phi is locally Lipschitzian and one checks that, in fact, Φ\Phi is a lipeomorphism between BB and Φ(B)\Phi(B), where B=𝐁(x,ρ)B=\mathbf{B}(x,\rho), for some ρ>0\rho>0 depending upon η\eta and Λx\Lambda_{x}, because its differential is close to the identity. Referring to Fubini’s theorem, one then further checks that

n(A)=0 if and only if m(A𝐖(ξ))=0, for nm-almost every ξVx.\mathscr{L}^{n}(A^{\prime})=0\text{ if and only if }\mathscr{H}^{m}(A^{\prime}\cap\mathbf{W}(\xi))=0\text{, for $\mathscr{H}^{n-m}$-almost every }\xi\in V_{x}.

Finally, using Fubini again, with respect to the decomposition n=𝐖0(x)𝐖0(x)\mathbb{R}^{n}=\mathbf{W}_{0}(x)\oplus\mathbf{W}_{0}(x)^{\perp}, one shows that m(A𝐖(ζ))=0\mathscr{H}^{m}(A\cap\mathbf{W}(\zeta))=0, for nm\mathscr{H}^{n-m}-almost every ζx+𝐖0(x)\zeta\in x^{\prime}+\mathbf{W}_{0}(x)^{\perp}, where xx^{\prime} is as close as we wish to xx. Applying the previous construction with xx^{\prime} replacing xx, using V=𝐖0(x)V=\mathbf{W}_{0}(x)^{\perp}, we find that n(A𝐁(x,r))=0\mathscr{L}^{n}(A\cap\mathbf{B}(x,r))=0, for some r>0r>0 depending on Λx\Lambda_{x}.

Notwithstanding, it seems that the (simpler) change of variable described here is not enough to yield the (stronger) theorem below.

8.3Polyballs. —

Recalling 7.2, we notice that

𝐂𝐖(x0,r)=n{νx0(xx0)r},\mathbf{C}_{\mathbf{W}}(x_{0},r)=\mathbb{R}^{n}\cap\left\{\nu_{x_{0}}(x-x_{0})\leqslant r\right\},

where νx0\nu_{x_{0}} is a norm on n\mathbb{R}^{n} defined by the formula

νx0(x)=max{|P𝐖0(x0)(x)|,|P𝐖0(x0)(x)|},\nu_{x_{0}}(x)=\max\left\{\left|P_{\mathbf{W}_{0}(x_{0})}(x)\right|,\left|P_{\mathbf{W}_{0}(x_{0})^{\perp}}(x)\right|\right\},

for xnx\in\mathbb{R}^{n}. It is readily observed that Lipνx01\operatorname{\mathrm{Lip}}\nu_{x_{0}}\leqslant 1.

  1. (1)

    One has |νx0(x)|=1|\nabla\nu_{x_{0}}(x)|=1, for n\mathscr{L}^{n}-almost every xnx\in\mathbb{R}^{n}.

Abbreviate P=P𝐖0(x0)P=P_{\mathbf{W}_{0}(x_{0})} and Q=P𝐖0(x0)Q=P_{\mathbf{W}_{0}(x_{0})^{\perp}} and define S=n{x:|P(x)|=|Q(x)|}S=\mathbb{R}^{n}\cap\{x:|P(x)|=|Q(x)|\}, so that n(S)=0\mathscr{L}^{n}(S)=0. Let xnSx\in\mathbb{R}^{n}\setminus S and notice νx0\nu_{x_{0}} is differentiable at xx. We henceforth assume that |P(x)|<|Q(x)||P(x)|<|Q(x)|, whence, νx0(x)=|Q(x)|\nu_{x_{0}}(x)=|Q(x)| – the proof in the other case is similar. Define ε=|Q(x)||P(x)|>0\varepsilon=|Q(x)|-|P(x)|>0, e=Q(x)|Q(x)|e=\frac{Q(x)}{|Q(x)|}, and let 0<t<ε0<t<\varepsilon. Note that |Q(x+te)|=|Q(x)|(1+t|Q(x)|)>|Q(x)||Q(x+te)|=|Q(x)|\left(1+\frac{t}{|Q(x)|}\right)>|Q(x)| and |P(x+te)||P(x)|+t<|Q(x)|P(x+te)|\leqslant|P(x)|+t<|Q(x). Thus, νx0(x+te)=|Q(x+te)|\nu_{x_{0}}(x+te)=|Q(x+te)| and, in turn, νx0(x+te)νx0(x)=|Q(x+te)||Q(x)|=t\nu_{x_{0}}(x+te)-\nu_{x_{0}}(x)=|Q(x+te)|-|Q(x)|=t. It follows that νx0(x),e=1\langle\nabla\nu_{x_{0}}(x),e\rangle=1. Since ee is a unit vector and Lipνx01\operatorname{\mathrm{Lip}}\nu_{x_{0}}\leqslant 1, we conclude that |νx0(x)|=1|\nabla\nu_{x_{0}}(x)|=1.

  1. (2)

    Let UU, 𝐖\mathbf{W} and Λ>0\Lambda>0 be as in 4.1. Assume 𝐂𝐖(x0,r)U\mathbf{C}_{\mathbf{W}}(x_{0},r)\subseteq U, x𝐂𝐖(x0,r)x\in\mathbf{C}_{\mathbf{W}}(x_{0},r), and 0t10\leqslant t\leqslant 1 is so that νx0(xx0)=tr\nu_{x_{0}}(x-x_{0})=tr. It follows that

    𝐂𝐖(x0,r)𝐖(x)𝐁(x,r(1+t)+8mΛr2)𝐖(x).\mathbf{C}_{\mathbf{W}}(x_{0},r)\cap\mathbf{W}(x)\subseteq\mathbf{B}\left(x,r(1+t)+8m\Lambda r^{2}\right)\cap\mathbf{W}(x).

First notice that |z|2νx0(z)|z|\leqslant\sqrt{2}\nu_{x_{0}}(z), for every znz\in\mathbb{R}^{n}, so that |xx0|2νx0(xx0)=2tr|x-x_{0}|\leqslant\sqrt{2}\nu_{x_{0}}(x-x_{0})=\sqrt{2}tr and, for every x𝐂𝐖(x0,r)x^{\prime}\in\mathbf{C}_{\mathbf{W}}(x_{0},r), |xx|2(νx0(xx0)+νx0(x0x))2r(1+t)|x-x^{\prime}|\leqslant\sqrt{2}\left(\nu_{x_{0}}(x-x_{0})+\nu_{x_{0}}(x_{0}-x^{\prime})\right)\leqslant\sqrt{2}r(1+t). Next notice that, for each hnh\in\mathbb{R}^{n},

|(P𝐖0(x0)P𝐖0(x))(h)|=|i=1m𝐰i(x0),h𝐰i(x0)i=1m𝐰i(x),h𝐰i(x)|i=1m|𝐰i(x0),h||𝐰i(x0)𝐰i(x)|+i=1m|𝐰i(x0)𝐰i(x),h||𝐰i(x)|2mΛ|xx0||h|.\left|\left(P_{\mathbf{W}_{0}(x_{0})}-P_{\mathbf{W}_{0}(x)}\right)(h)\right|=\left|\sum_{i=1}^{m}\langle\mathbf{w}_{i}(x_{0}),h\rangle\mathbf{w}_{i}(x_{0})-\sum_{i=1}^{m}\langle\mathbf{w}_{i}(x),h\rangle\mathbf{w}_{i}(x)\right|\\ \leqslant\sum_{i=1}^{m}\left|\langle\mathbf{w}_{i}(x_{0}),h\rangle\right|\left|\mathbf{w}_{i}(x_{0})-\mathbf{w}_{i}(x)\right|+\sum_{i=1}^{m}\left|\langle\mathbf{w}_{i}(x_{0})-\mathbf{w}_{i}(x),h\rangle\right|\left|\mathbf{w}_{i}(x)\right|\\ \leqslant 2m\Lambda\left|x-x_{0}\right||h|.

We now assume that x𝐂𝐖(x,r)𝐖(x)x^{\prime}\in\mathbf{C}_{\mathbf{W}}(x,r)\cap\mathbf{W}(x), in particular, xx𝐖0(x)x-x^{\prime}\in\mathbf{W}_{0}(x), whence,

|xx|=|P𝐖0(x)(xx)||P𝐖0(x0)(xx)|+|(P𝐖0(x0)P𝐖0(x))(xx)|νx0(xx0)+νx0(xx0)+2mΛ|xx0||xx|r(1+t)+4mΛt(1+t)r2.\left|x-x^{\prime}\right|=\left|P_{\mathbf{W}_{0}(x)}(x-x^{\prime})\right|\leqslant\left|P_{\mathbf{W}_{0}(x_{0})}(x-x^{\prime})\right|+\left|\left(P_{\mathbf{W}_{0}(x_{0})}-P_{\mathbf{W}_{0}(x)}\right)(x-x^{\prime})\right|\\ \leqslant\nu_{x_{0}}(x-x_{0})+\nu_{x_{0}}(x^{\prime}-x_{0})+2m\Lambda|x-x_{0}||x-x^{\prime}|\leqslant r(1+t)+4m\Lambda t(1+t)r^{2}.
8.4 Theorem. —

Assume that AnA\subseteq\mathbb{R}^{n} is Borel measurable and that 𝐖0:A𝔾(n,m)\mathbf{W}_{0}:A\to\mathbb{G}(n,m) is Lipschitzian. It follows that

lim supr0+m(A𝐁(x,r)𝐖(x))𝜶(m)rm12n,\limsup_{r\to 0^{+}}\frac{\mathscr{H}^{m}\left(A\cap\mathbf{B}(x,r)\cap\mathbf{W}(x)\right)}{\boldsymbol{\alpha}(m)r^{m}}\geqslant\frac{1}{2^{n}},

for n\mathscr{L}^{n}-almost every xAx\in A.

Recall our convention that 𝐖(x)=x+𝐖0(x)\mathbf{W}(x)=x+\mathbf{W}_{0}(x).

Proof.

Extend 𝐖0\mathbf{W}_{0} to n\mathbb{R}^{n} in a Borel measurable way, for instance, to be an arbitrary constant outside of AA. It follows from 3.15 that, for each r>0r>0, the function n[0,]:xm(A𝐁(x,r)𝐖(x))𝜶(m)rm\mathbb{R}^{n}\to[0,\infty]:x\mapsto\frac{\mathscr{H}^{m}(A\cap\mathbf{B}(x,r)\cap\mathbf{W}(x))}{\boldsymbol{\alpha}(m)r^{m}} is Borel measurable. Thus, for each j=1,2,j=1,2,\ldots, the function gj:n[0,]g_{j}:\mathbb{R}^{n}\to[0,\infty] defined by

gj(x)=sup0<r<1jm(A𝐁(x,r)𝐖(x))𝜶(m)rm=sup0<r<1jrrationalm(A𝐁(x,r)𝐖(x))𝜶(m)rmg_{j}(x)=\sup_{0<r<\frac{1}{j}}\frac{\mathscr{H}^{m}\left(A\cap\mathbf{B}(x,r)\cap\mathbf{W}(x)\right)}{\boldsymbol{\alpha}(m)r^{m}}=\sup_{\begin{subarray}{c}0<r<\frac{1}{j}\\ r\,\text{rational}\end{subarray}}\frac{\mathscr{H}^{m}\left(A\cap\mathbf{B}(x,r)\cap\mathbf{W}(x)\right)}{\boldsymbol{\alpha}(m)r^{m}}

is Borel measurable as well, and so is g=limjgj=infjgjg=\lim_{j}g_{j}=\inf_{j}g_{j}.

Abbreviate 𝜼(n,m)=2(nm)\boldsymbol{\eta}(n,m)=2^{-(n-m)}. Arguing reductio ad absurdum, we henceforth assume that AA and 𝐖0\mathbf{W}_{0} fail the conclusion of the theorem. Thus, the set Z0=A{x:g(x)<𝜼(n,m)2m}Z_{0}=A\cap\left\{x:g(x)<\frac{\boldsymbol{\eta}(n,m)}{2^{m}}\right\} is Borel measurable and non Lebesgue null. Accordingly, there exists ε1>0\varepsilon_{1}>0 such that the set Z0=A{x:g(x)<(1ε1)𝜼(n,m)(2+ε1)m}Z_{0}^{\prime}=A\cap\left\{x:g(x)<(1-\varepsilon_{1})\frac{\boldsymbol{\eta}(n,m)}{(2+\varepsilon_{1})^{m}}\right\} is also Borel measurable and of positive Lebesgue measure. It therefore ensues from Egoroff’s theorem [9, 2.3.7] that there exists a closed set ZZ0AZ\subseteq Z^{\prime}_{0}\subseteq A such that n(Z)>0\mathscr{L}^{n}(Z)>0 and that there exists a positive integer j0j_{0} such that

m(Z𝐁(x,r)𝐖(x))𝜶(m)rmgj0(x)<(1ε1)𝜼(n,m)(2+ε1)m,\frac{\mathscr{H}^{m}\left(Z\cap\mathbf{B}(x,r)\cap\mathbf{W}(x)\right)}{\boldsymbol{\alpha}(m)r^{m}}\leqslant g_{j_{0}}(x)<(1-\varepsilon_{1})\frac{\boldsymbol{\eta}(n,m)}{(2+\varepsilon_{1})^{m}}, (35)

for each xZx\in Z and each 0<r<1j00<r<\frac{1}{j_{0}}. Choose 0<ε2<100<\varepsilon_{2}<10 such that

1ε11ε21ε12\frac{1-\varepsilon_{1}}{1-\varepsilon_{2}}\leqslant 1-\frac{\varepsilon_{1}}{2} (36)

and choose 0<ε3<1/30<\varepsilon_{3}<1/3 such that

1ε12<1𝐜7.6(n)ε3.1-\frac{\varepsilon_{1}}{2}<1-\mathbf{c}_{\ref{lower.bound}}(n)\varepsilon_{3}. (37)

As in the proof of 7.8, we recall that the family 𝐂𝐖(x,r)\mathbf{C}_{\mathbf{W}}(x,r), for xnx\in\mathbb{R}^{n} and r>0r>0, is a density basis of n\mathscr{L}^{n}-measurable sets. Since n(Z)>0\mathscr{L}^{n}(Z)>0, there exists x0Zx_{0}\in Z such that

limr0+n(Z𝐂𝐖(x0,r))n(𝐂𝐖(x0,r))=1.\lim_{r\to 0^{+}}\frac{\mathscr{L}^{n}\left(Z\cap\mathbf{C}_{\mathbf{W}}(x_{0},r)\right)}{\mathscr{L}^{n}\left(\mathbf{C}_{\mathbf{W}}(x_{0},r)\right)}=1.

In particular, there exists R>0R>0 such that

(1ε3)n(𝐂𝐖(x0,r))n(Z𝐂𝐖(x0,r)),(1-\varepsilon_{3})\mathscr{L}^{n}\left(\mathbf{C}_{\mathbf{W}}(x_{0},r)\right)\leqslant\mathscr{L}^{n}\left(Z\cap\mathbf{C}_{\mathbf{W}}(x_{0},r)\right), (38)

whenever 0<r<R0<r<R.

We let UU be an open neighborhood of x0x_{0} in n\mathbb{R}^{n} associated with AA and 𝐖0\mathbf{W}_{0} in 3.8, so that 𝐖0\mathbf{W}_{0} and 𝐖0\mathbf{W}_{0}^{\perp} are associated with orthonormal frames as in 4.1, for some Λ>0\Lambda>0. Define

r0=min{1,1j0(2+8mΛ),ε18mΛ,𝜹5.15(n,Λ,ε2)22,dist(x0,nU)22,𝜹7.6(n,Λ,ε3),R}.r_{0}=\min\bigg{\{}1,\frac{1}{j_{0}(2+8m\Lambda)},\frac{\varepsilon_{1}}{8m\Lambda},\frac{\boldsymbol{\delta}_{\ref{lb.1}}(n,\Lambda,\varepsilon_{2})}{2\sqrt{2}},\frac{\operatorname{\mathrm{dist}}(x_{0},\mathbb{R}^{n}\setminus U)}{2\sqrt{2}},\boldsymbol{\delta}_{\ref{lower.bound}}(n,\Lambda,\varepsilon_{3}),R\bigg{\}}.

Let 0<r<r00<r<r_{0} and observe that

(1𝐜7.6(n)ε3)n(𝐂𝐖(x0,r))Z𝐂𝐖(x0,r)𝒴Z𝐂𝐖(x0,r)𝐖(u)𝜶(m)rm𝑑n(u)(by 7.6 applied with =εε3 and =A∩Z⁢CW(x0,r))1(1ε2)𝜼(n,m)Z𝐂𝐖(x0,r)m(Z𝐂𝐖(x0,r)𝐖(x))𝜶(m)rm𝑑n(x)\begin{split}\left(1-\mathbf{c}_{\ref{lower.bound}}(n)\varepsilon_{3}\right)&\mathscr{L}^{n}\left(\mathbf{C}_{\mathbf{W}}(x_{0},r)\right)\leqslant\int_{Z\cap\mathbf{C}_{\mathbf{W}}(x_{0},r)}\frac{\mathscr{Y}_{Z\cap\mathbf{C}_{\mathbf{W}}(x_{0},r)}\mathbf{W}(u)}{\boldsymbol{\alpha}(m)r^{m}}d\mathscr{L}^{n}(u)\intertext{(by \ref{lower.bound} applied with $\varepsilon=\varepsilon_{3}$ and $A=Z\cap\mathbf{C}_{\mathbf{W}}(x_{0},r)$)}&\leqslant\frac{1}{(1-\varepsilon_{2})\boldsymbol{\eta}(n,m)}\int_{Z\cap\mathbf{C}_{\mathbf{W}}(x_{0},r)}\frac{\mathscr{H}^{m}\left(Z\cap\mathbf{C}_{\mathbf{W}}(x_{0},r)\cap\mathbf{W}(x)\right)}{\boldsymbol{\alpha}(m)r^{m}}d\mathscr{L}^{n}(x)\end{split} (39)

(by 5.15, applied with ε=ε2\varepsilon=\varepsilon_{2} and E=B=Z𝐂𝐖(x0,r)E=B=Z\cap\mathbf{C}_{\mathbf{W}}(x_{0},r)). We also note that

Z𝐂𝐖(x0,r)m(Z𝐂𝐖(x0,r)𝐖(x))𝜶(m)rm𝑑n(x)=Z{νx0r}m(Z𝐂𝐖(x0,r)𝐖(x))𝜶(m)rm|νx0(x)|𝑑n(x)(by 8.3(1))=0r𝑑1(ρ)Z{νx0=ρ}m(Z𝐂𝐖(x0,r)𝐖(x))𝜶(m)rm𝑑n1(x)(by [6, 3.4.3])=r01𝑑1(t)Z{νx0=tr}m(Z𝐂𝐖(x0,r)𝐖(x))𝜶(m)rm𝑑n1(x)r01𝑑1(t)Z{νx0=tr}m(Z𝐁(x,r(1+t)+8mΛr2)𝐖(x))𝜶(m)rm𝑑n1(x)(by 8.3(2))r01𝑑1(t)Z{νx0=tr}(1ε1)𝜼(n,m)(2+ε1)m(1+t+8mΛr)m𝑑n1(by (35))(1ε1)𝜼(n,m)0r𝑑1(ρ)Z{νx0=ρ}𝑑n1=(1ε1)𝜼(n,m)Z{νx0r}|νx0(x)|𝑑n(x)(by [6, 3.4.3])=(1ε1)𝜼(n,m)n(Z𝐂𝐖(x0,r))(1ε1)𝜼(n,m)n(𝐂𝐖(x0,r))\begin{split}&\int_{Z\cap\mathbf{C}_{\mathbf{W}}(x_{0},r)}\frac{\mathscr{H}^{m}\left(Z\cap\mathbf{C}_{\mathbf{W}}(x_{0},r)\cap\mathbf{W}(x)\right)}{\boldsymbol{\alpha}(m)r^{m}}d\mathscr{L}^{n}(x)\\ &=\int_{Z\cap\{\nu_{x_{0}}\leqslant r\}}\frac{\mathscr{H}^{m}\left(Z\cap\mathbf{C}_{\mathbf{W}}(x_{0},r)\cap\mathbf{W}(x)\right)}{\boldsymbol{\alpha}(m)r^{m}}\left|\nabla\nu_{x_{0}}(x)\right|d\mathscr{L}^{n}(x)\intertext{(by \ref{pb.complement}(1))}&=\int_{0}^{r}d\mathscr{L}^{1}(\rho)\int_{Z\cap\{\nu_{x_{0}}=\rho\}}\frac{\mathscr{H}^{m}\left(Z\cap\mathbf{C}_{\mathbf{W}}(x_{0},r)\cap\mathbf{W}(x)\right)}{\boldsymbol{\alpha}(m)r^{m}}d\mathscr{H}^{n-1}(x)\intertext{(by \cite[cite]{[\@@bibref{}{EVANS.GARIEPY}{}{}, 3.4.3]})}&=r\int_{0}^{1}d\mathscr{L}^{1}(t)\int_{Z\cap\{\nu_{x_{0}}=tr\}}\frac{\mathscr{H}^{m}\left(Z\cap\mathbf{C}_{\mathbf{W}}(x_{0},r)\cap\mathbf{W}(x)\right)}{\boldsymbol{\alpha}(m)r^{m}}d\mathscr{H}^{n-1}(x)\\ &\leqslant r\int_{0}^{1}d\mathscr{L}^{1}(t)\int_{Z\cap\{\nu_{x_{0}}=tr\}}\frac{\mathscr{H}^{m}\left(Z\cap\mathbf{B}\left(x,r(1+t)+8m\Lambda r^{2}\right)\cap\mathbf{W}(x)\right)}{\boldsymbol{\alpha}(m)r^{m}}d\mathscr{H}^{n-1}(x)\intertext{(by \ref{pb.complement}(2))}&\leqslant r\int_{0}^{1}d\mathscr{L}^{1}(t)\int_{Z\cap\{\nu_{x_{0}}=tr\}}(1-\varepsilon_{1})\frac{\boldsymbol{\eta}(n,m)}{(2+\varepsilon_{1})^{m}}(1+t+8m\Lambda r)^{m}d\mathscr{H}^{n-1}\intertext{(by \eqref{final.1})}&\leqslant(1-\varepsilon_{1})\boldsymbol{\eta}(n,m)\int_{0}^{r}d\mathscr{L}^{1}(\rho)\int_{Z\cap\{\nu_{x_{0}}=\rho\}}d\mathscr{H}^{n-1}\\ &=(1-\varepsilon_{1})\boldsymbol{\eta}(n,m)\int_{Z\cap\{\nu_{x_{0}}\leqslant r\}}\left|\nabla\nu_{x_{0}}(x)\right|d\mathscr{L}^{n}(x)\intertext{(by \cite[cite]{[\@@bibref{}{EVANS.GARIEPY}{}{}, 3.4.3]})}&=(1-\varepsilon_{1})\boldsymbol{\eta}(n,m)\mathscr{L}^{n}\left(Z\cap\mathbf{C}_{\mathbf{W}}(x_{0},r)\right)\leqslant(1-\varepsilon_{1})\boldsymbol{\eta}(n,m)\mathscr{L}^{n}\left(\mathbf{C}_{\mathbf{W}}(x_{0},r)\right)\end{split} (40)

(by 8.3(1)). Plugging (40) into (39), we obtain

(1𝐜7.6(n)ε3)n(𝐂𝐖(x0,r))(1ε11ε2)n(𝐂𝐖(x0,r))<(1𝐜7.6(n)ε3)n(𝐂𝐖(x0,r))\begin{split}\left(1-\mathbf{c}_{\ref{lower.bound}}(n)\varepsilon_{3}\right)\mathscr{L}^{n}\left(\mathbf{C}_{\mathbf{W}}(x_{0},r)\right)&\leqslant\left(\frac{1-\varepsilon_{1}}{1-\varepsilon_{2}}\right)\mathscr{L}^{n}\left(\mathbf{C}_{\mathbf{W}}(x_{0},r)\right)\\ &<\left(1-\mathbf{c}_{\ref{lower.bound}}(n)\varepsilon_{3}\right)\mathscr{L}^{n}\left(\mathbf{C}_{\mathbf{W}}(x_{0},r)\right)\end{split}

(by (36) and (37)), a contradiction. ∎

References

  • [1] J. Bourgain, A remark on the maximal function associated to an analytic vector field, Analysis at Urbana, Vol. I (Urbana, IL, 1986–1987), London Math. Soc. Lecture Note Ser., vol. 137, Cambridge Univ. Press, Cambridge, 1989, pp. 111–132. MR 1009171
  • [2] A. M. Bruckner, Differentiation of integrals, Amer. Math. Monthly 78 (1971), no. 9, part II, ii+51. MR 293044
  • [3] D. L. Cohn, Measure theory, second ed., Birkhäuser Advanced Texts: Basler Lehrbücher. [Birkhäuser Advanced Texts: Basel Textbooks], Birkhäuser/Springer, New York, 2013. MR 3098996
  • [4] M. de Guzmán, Real variable methods in Fourier analysis, North-Holland Mathematics Studies, vol. 46, North-Holland Publishing Co., Amsterdam-New York, 1981, Notas de Matemática [Mathematical Notes], 75. MR 596037
  • [5] Th. De Pauw, Concentrated, nearly monotonic, epiperimetric measures in Euclidean space, J. Differential Geom. 77 (2007), no. 1, 77–134. MR 2344355
  • [6] L. C. Evans and R. F. Gariepy, Measure theory and fine properties of functions, Studies in Advanced Mathematics, CRC Press, Boca Raton, FL, 1992. MR 1158660 (93f:28001)
  • [7] K. J. Falconer, Sets with prescribed projections and Nikodým sets, Proc. London Math. Soc. (3) 53 (1986), no. 1, 48–64. MR 842156
  • [8] H. Federer, Curvature measures, Trans. Amer. Math. Soc. 93 (1959), 418–491. MR 110078
  • [9] by same author, Geometric measure theory, Die Grundlehren der mathematischen Wissenschaften, Band 153, Springer-Verlag New York Inc., New York, 1969. MR 0257325 (41 #1976)
  • [10] A. B. Kharazishvili, Nonmeasurable sets and functions, North-Holland Mathematics Studies, vol. 195, Elsevier Science B.V., Amsterdam, 2004. MR 2067444 (2005d:28001)
  • [11] M. Lacey and X. Li, On a conjecture of E. M. Stein on the Hilbert transform on vector fields, Mem. Amer. Math. Soc. 205 (2010), no. 965, viii+72. MR 2654385
  • [12] O. Nikodým, Sur la mesure des ensembles plans dont tous les points sont rectilinéairement accessibles., Fundam. Math. 10 (1927), 116–168 (French).
  • [13] S. M. Srivastava, A course on Borel sets, Graduate Texts in Mathematics, vol. 180, Springer-Verlag, New York, 1998. MR 1619545 (99d:04002)