Sbornik: Mathematics
RUS  ENG    JOURNALS   PEOPLE   ORGANISATIONS   CONFERENCES   SEMINARS   VIDEO LIBRARY   PACKAGE AMSBIB  
General information
Latest issue
Forthcoming papers
Archive
Impact factor
Guidelines for authors
License agreement
Submit a manuscript

Search papers
Search references

RSS
Latest issue
Current issues
Archive issues
What is RSS



Mat. Sb.:
Year:
Volume:
Issue:
Page:
Find






Personal entry:
Login:
Password:
Save password
Enter
Forgotten password?
Register


Sbornik: Mathematics, 2022, Volume 213, Issue 9, Pages 1157–1186
DOI: https://doi.org/10.4213/sm9702e
(Mi sm9702)
 

This article is cited in 1 scientific paper (total in 1 paper)

Coincidence of set functions in quasiconformal analysis

S. K. Vodopyanov

Sobolev Institute of Mathematics, Siberian Branch of the Russian Academy of Sciences, Novosibirsk, Russia
References:
Abstract: It is known that mappings occurring in quasiconformal analysis can be defined in several equivalent ways: 1) as homeomorphisms inducing bounded composition operators between Sobolev spaces; 2) as Sobolev-class homeomorphisms with bounded distortion whose operator distortion function is integrable; 3) as homeomorphism changing the capacity of the image of a condenser in a controllable way in terms of the weighted capacity of the condenser in the source space; 4) as homeomorphism changing the modulus of the image of a family of curves in a controllable way in terms of the weighted modulus of the family of curves in the source space. A certain set function, defined on open subsets, can be associated with each of these definitions. The main result consists in the fact that all these set functions coincide.
Bibliography: 48 titles.
Keywords: quasiconformal analysis, Sobolev space, composition operator, condenser capacity, outer operator distortion function, set function.
Funding agency Grant number
Ministry of Science and Higher Education of the Russian Federation FWNF-2022-0006
This research was carried out in the framework of the state assignment of the Ministry of Education and Science of the Russian Federation for the Institute of Mathematics of the Siberian Branch of the Russian Academy of Sciences (project no. FWNF-2022-0006).
Received: 28.11.2021 and 27.01.2022
Russian version:
Matematicheskii Sbornik, 2022, Volume 213, Number 9, Pages 3–33
DOI: https://doi.org/10.4213/sm9702
Bibliographic databases:
Document Type: Article
MSC: Primary 30C65; Secondary 28A10
Language: English
Original paper language: Russian

§ 1. Introduction

It is known that quasiconformal mappings have several equivalent descriptions (for instance, see [1] and [2]): a metric one, an analytic one, a geometric description in the language of moduli (see [2]), a geometric description in the language of capacities (see [3]), a functional one (see [4]). Associated with each of these descriptions is a set function defined on open subsets of the domain of definition of the mapping. The main result of our paper is that all these set functions coincide. This result is apparently new even for classical quasiconformal mappings (of course, apart from conformal ones).

More precisely, given a homeomorphism $\varphi\colon D\to D'$ of domains $D,D'\subset \mathbb R^n$, $n\!\geqslant\!2$, assume that it induces (by the change of variables formula ${\varphi^*(f)\!=\!f\!\circ\!\varphi}$) a bounded composition operator1

$$ \begin{equation*} \varphi^*\colon {L}^1_p(D';\omega) \cap \operatorname{Lip}_{\mathrm{loc}}(D')\to L^1_q(D) \end{equation*} \notag $$
with parameters $n-1< q\leqslant p<\infty$ for $n\geqslant3$, or $1\leqslant q\leqslant p<\infty$ for $n=2$ and with weight function $\omega\in L_{1,\mathrm{loc}}(D')$.

Then the following hold.

I. For $q<p$ the set functions2

are equal;

$$ \begin{equation} \|\varphi^*_W\|^\sigma=\bigl\|K^{1,\omega}_{q,p}(\,\cdot\,)\mid L_\sigma(\varphi^{-1}(W))\bigr\|^\sigma \end{equation} \tag{1.1} $$
for each open set $W\subset D'$.

II. For $q=p$ the two quantities,

are equal:

$$ \begin{equation} \|\varphi^*\|=\bigl\|K^{1,\omega}_{p,p}(\,\cdot\,)\mid L_\infty(D)\bigr\|. \end{equation} \tag{1.2} $$

For classical quasiconformal mappings (when $q=p=n$ and $\omega\equiv1$) the last equality says that the norm of the composition operator

$$ \begin{equation*} \varphi^*\colon {L}^1_n(D')\to L^1_n(D) \end{equation*} \notag $$
coincides with
$$ \begin{equation*} \bigl\|K^{1,\omega}_{p,p}(\,\cdot\,)\mid L_\infty(D)\bigr\| =\operatorname*{ess\,sup}_{x\in D}\frac{|D\varphi(x)|}{|{\det D\varphi (x)}|^{1/n}}. \end{equation*} \notag $$

Problems in quasiconformal analysis stated more than 50 years ago has led to the concept of mappings characterized by a controllable change of the capacity of the image of a condenser in terms of the weighted capacity of the condenser in the source space: for instance, see [5]. At the same time authors from Finland, Ukraine, Israel and other countries investigated maps (see [6]) whose definition, in place of capacity, involved another geometric characteristic, the modulus of a family of curves. Note that for arbitrary weights there are no two-sided estimates connecting weighted capacity with weighted modulus (with the exception of weights in the Muckenhoupt class $\mathcal A_n$; see [7]–[9]). So both approaches were investigated independently, particularly since the formal impression was that by using the modulus-based approach (see [6]) one can treat a wider class of maps than in the framework of the capacity-based approach.

The central result of our paper (cf. [10]) is that, in fact, whether we use the capacity-based characteristic or the modulus-based characteristic, we obtain the same classes of maps. A striking feature of this result is that we do not obtain it by comparing a weighted capacity and a weighted modulus, but we use a functional analytic description of mappings characterized by a controllable change of the capacity in the target space in terms of the capacity in the source space, which was developed in 2020 (see [11]–[14]).

In § 3 we define two other set functions (one is based on bounds for capacities and the other on bounds for moduli) and show that, depending on their parameters, they either coincide with those in (1.1) or with those in (1.2).

This paper can be viewed s a natural part of the cycle [11]–[14], and also of the cycle [15]–[19], where the history of the problem was exposed and a comprehensive bibliography was presented. Before those papers [4], [20] and [21] were published, where methods from the theory of Sobolev function spaces (see [22] and [23]) and geometric function theory (see [1]–[3] and [24]–[27]) were synthesized. Some results obtained in that cycle of papers found applications to nonlinear elasticity theory: see [28].

§ 2. Classes of $\mathcal Q_{q,p}$-homeomorphisms

Here and below $D$ and $D'$ are domains (connected open sets) in $\mathbb{R}^n$.

2.1. The definitions of Sobolev spaces and of condenser capacity

Recall that a function $u\colon D\to\mathbb R$ belongs to the Sobolev class $L^1_{p}(D)$ if $u$ is locally integrable in $D$ (that is, $u\in L_1(U)$ for each compactly embedded subdomain $U\Subset D$), has generalized derivatives ${\partial u}/{dx_j}\in L_{1,\mathrm{loc}}(D) $ for each $j=1,\dots,n$ and has a finite seminorm

$$ \begin{equation*} \|u\mid L^1_{p}(D)\|=\biggl(\int_{D}|\nabla u(y)|^p\,dy\biggr)^{1/p}, \qquad 1\leqslant p\leqslant \infty. \end{equation*} \notag $$
A map $\varphi=(\varphi_1,\dots, \varphi_n) $ belongs to the Sobolev class $W^1_{p,\mathrm{loc}}(D; \mathbb{R}^n)$ if ${\varphi_j(x) \in L_{p,\mathrm{loc}}(D)}$ and the generalized derivatives ${\partial\varphi_j}/{dx_i}$ belong to $L_{p,\mathrm{loc}}(D) $ for all ${j,i=1,\dots,n}$.

A map $\varphi\colon D\to \mathbb R^n$ in the Sobolev class $W^1_{1,\mathrm{loc}}(D;\mathbb{R}^n)$ is called a map with bounded distortion if3

$$ \begin{equation*} D\varphi(x)=0 \quad\text{a.e. on } Z=\{x\in D\colon \det D\varphi (x)=0\}. \end{equation*} \notag $$
Here and throughout, $D\varphi (x)=({\partial\varphi_j}/{\partial x_i}(x))$ is the Jacobian matrix of $\varphi$ at $x\in D$, $|D\varphi (x)|$ is the Euclidean operator norm of this matrix and $\det D\varphi (x)$ is its determinant (Jacobian).

A locally integrable function $\omega\colon D'\!\to\!\mathbb R$ is called a weight function if ${0\!<\!\omega(y)\!<\!\infty}$ for almost all $y\in D'$. Recall that $u\colon D'\to\mathbb R$ belongs to the weighted Sobolev class $L^1_{p}(D';\omega)$, $p\in[1,\infty)$, if $u$ is locally integrable in $D'$ and has generalized derivatives ${\partial u}/{\partial y_j}$ in $D'$ that belong to $L_{p}(D';\omega)$ for all $j=1,\dots,n$. The seminorm of $u\in L^1_{p}(D';\omega)$ is the quantity

$$ \begin{equation*} \|u\mid L^1_{p}(D';\omega)\|=\biggl(\int_{D'}|\nabla u|^p(y)\omega(y)\,dy\biggr)^{1/p}. \end{equation*} \notag $$
For $\omega\equiv 1$ we simply write $L^1_{p}(D')$ instead of $L^1_{p}(D';1)$.

In what follows we let $\operatorname{Lip}_{\mathrm{loc}}(D')$ denote the space of locally Lipschitz functions on the domain $D'$. Clearly, $\operatorname{Lip}_{\mathrm{loc}}(D')=W^1_{\infty,\mathrm{loc}}(D')\cap C(D')$.

Recall that a homeomorphism $\varphi \colon D \to D'$ between domains $D$ and $D'$ in $\mathbb{R}^n$ induces a bounded composition operator

$$ \begin{equation*} \varphi^* \colon {L}^1_p(D';\omega) \cap \operatorname{Lip}_{\mathrm{loc}}(D')\to L^1_q(D), \qquad 1\leqslant q \leqslant p < \infty, \end{equation*} \notag $$
which acts by the formula $D\ni x\mapsto(\varphi^*u)(x)=u(\varphi(x))$, provided that for some constant $K_{q,p}<\infty$ we have
$$ \begin{equation*} \begin{gathered} \, \|\varphi^*u\mid L^1_q(D)\|\leqslant K_{q,p}\|u\mid L^1_p(D';\omega)\| \\ \text{for each } u\in L^1_p(D')\cap \operatorname{Lip}_{\mathrm{loc}}(D'). \end{gathered} \end{equation*} \notag $$

Definition 1. A condenser in a domain $D\subset \mathbb{R}^n$ is a pair $E=(F_1,F_0)$ of connected compact sets (continua) in $D$: $F_1,F_0\subset D$. If $F$ lies in a connected compactly embedded subset $U\Subset D$, then we denote the condenser $E=(F,\partial U)$ by $E=(F,U)$.

A condenser $E=(F,\partial U)$ is called a ring condenser if $U\Subset D'$ is a connected open set and $F\subset U$ is a continuum such that $\mathbb R^n\setminus F$ is a connected open set and the complement $\overline{\mathbb R^n}\setminus (U\setminus F)$ has two connected components, $F$ and $ \overline{\mathbb R^n}\setminus U$ (here $\overline{\mathbb R^n}=\mathbb R^n\cup\{\infty\}$ is a one-point compactification of $\mathbb R^n$).

A ring condenser $E=(F,\partial U)$ in $\mathbb R^n$ is said to be spherical (cubic) if $U$ is a ball4 $B(x,R)=\{y\in\mathbb R^n\colon |y-x|_2< R\}$ (a cube $Q(x,R)=\{y\in\mathbb R^n\colon |y-x|_\infty< R\}$) and $F\subset U$ is the closure $F=\{y\in\mathbb R^n\colon |y-x|_2\leqslant r\}$ of a ball $B(x,r)$ (the closure $F=\{y\in\mathbb R^n\colon |y-x|_\infty\leqslant r\}$ of a cube $Q(x,r)$), where $r\in (0,R)$.

A continuous function $u\colon D\to\mathbb R$ in the class $W^1_{1,\mathrm{loc}}(D)$ is said to be admissible for a condenser $E=(F_1,F_0)\subset D$ if $u\equiv 1$ on $F_1$ and $u \equiv 0$ on $F_0$. We denote the set of functions admissible for $E=(F_1,F_0)$ by $\mathcal A(E)$.

We define the capacity of the condenser $E=(F_1,F_0)$ in the space $L^1_q(D)$, ${q\in[1,\infty)}$, by

$$ \begin{equation*} \operatorname{cap}(E; L^1_q(D))=\inf_{u\in \mathcal A(E)}\|u\mid L^1_{q}(D)\|^q; \end{equation*} \notag $$
here we take the lower bound over all functions admissible for $E = (F_1,F_0) \subset D$ that belong to $ L^1_{q}(D)$.

The weighted capacity of a condenser $E=(F_1,F_0)\subset D'$ in $L^1_p(D';\omega)$ is defined similarly, by

$$ \begin{equation*} \operatorname{cap}(E; L^1_p(D';\omega))=\inf_{u\in \mathcal A(E)\cap\operatorname{Lip}_{\mathrm{loc}}(D')}\|u\mid L^1_{p}(D';\omega)\|^p, \end{equation*} \notag $$
where the lower bound is taken over all functions admissible for $E=(F_1,F_0)$ that belong to $ \operatorname{Lip}_{\mathrm{loc}}(D')\cap L^1_{p}(D';\omega)$. For details on the properties of weighted capacities (for a certain special class of weight functions), see. [23], Ch. 2.

The following principle is a direct consequence of the definition of capacity,

Subordination principle. Let $E'=(F'_1,F'_0)$ and $E=(F_1,F_0)$ be two condensers in a domain $D'$ such that the plates of the first condenser lie in the plates of the second: $F'_1\subset F_1 $ and $F'_0\subset F_0$. Then

$$ \begin{equation*} \operatorname{cap}(E'; L^1_p(D';\omega))\leqslant\operatorname{cap}(E; L^1_p(D';\omega)). \end{equation*} \notag $$

A similar property also holds for the $q$-capacity of a condenser in a domain $D$.

To prove the subordination principle it is sufficient to observe that

$$ \begin{equation*} \mathcal A(E')\cap \operatorname{Lip}_{\mathrm{loc}}(D')\cap L^1_{p}(D';\omega)\subset \mathcal A(E)\cap \operatorname{Lip}_{\mathrm{loc}}(D')\cap L^1_{p}(D';\omega). \end{equation*} \notag $$

The definition of the class of $\mathcal Q_{q,p}$-homeomorphisms presented in § 2.4 below is based on an estimate for the capacity of the image in $D'$ of a cubic condenser in terms of the weighted capacity of the original condenser.

2.2. A quasiadditive set function and its properties

Let ${\mathcal O}(D)$ be a system of open subsets of $D$ with the following properties:

The choice of the ball or a cube in this definition depends on the choice of the system of elementary sets used for the differentiation of set functions (see Proposition 1 below).

Definition 2. The map $\Phi\colon {\mathcal O}(D)\to[0,\infty]$ is called a quasiadditive set function if

1) for each point $x\in D$ there exists $\delta$, $0<\delta<\mathrm{dist}(x, \partial D)$, such that

$$ \begin{equation*} 0\leqslant \Phi(B(x,\delta))<\infty \end{equation*} \notag $$
(if $D=\mathbb R^n$, then $0\leqslant\Phi(D(x,{\delta}))<\infty$ must hold for all $\delta\in(0, \delta(x))$, where $\delta(x)$ is a positive number which can depend on $x$); we can replace balls by cubes in this condition;

2) for each finite disjoint system $\{U_i\in{\mathcal O}(D)$, $i=1,\dots,l\}$ of open sets such that $\bigcup_{i=1}^lU_i\subset U$, where $U\in{\mathcal O}(D)$, we have

$$ \begin{equation*} \sum_{i=1}^{l}\Phi(U_i)\leqslant \Phi(U). \end{equation*} \notag $$

If for each finite system $\{U_i\in{\mathcal O}(D)\}$ of mutually disjoint open sets we have

$$ \begin{equation} \sum_{i=1}^{n}\Phi(U_i)=\Phi\biggl(\bigcup_{i=1}^{n} U_i \biggr), \end{equation} \tag{2.1} $$
then the quasiadditive set function with the corresponding property is said to be finitely additive, and if (2.1) holds for each countable system $\{U_i\in{\mathcal O}(D)\}$ of mutually disjoint open sets, then it is said to be countably additive.

A function $\Phi$ is monotone if $\Phi(U_1)\leqslant \Phi(U_2)$ for $ U_1\subset U_2 \subset D$, $U_1,U_2\in{\mathcal O}(D)$. It is obvious that all quasiadditive set functions are monotone.

A quasiadditive set function $\Phi\colon {\mathcal O}(D)\to[0,\infty]$ is called a bounded quasiadditive set function if $\sup_{U\in {\mathcal O}(D)}\Phi(U)<\infty$.

A quasiadditive set function $\Phi$ as defined above is known to be differentiable in the following sense.

Proposition 1 (see [30]–[32]). Let $\Phi$ be a quasiadditive set function defined on some system ${\mathcal O}(D')$ of open subsets of $D'$. Then

1) for almost all $y\in D'$ the finite derivative 5

$$ \begin{equation*} \Phi'(y)= \lim_{\delta\to 0,\, y\in B_\delta}\frac{\Phi(B_\delta)}{|B_{\delta}|} \end{equation*} \notag $$
exists;

2) for each open set $U\in \mathcal O(D')$

$$ \begin{equation*} \int_{U}\Phi'(y)\,dy \leqslant \Phi(U). \end{equation*} \notag $$

2.3. The formula for a change of variable

We present, in the form we require, the change of variable formula from [13] and [14], which is a modification of the formula for a change of variable in a Lebesgue integral from [33] and [34].

Let $\varphi\colon \Omega\to \mathbb R^n$ be a map and let $E \subseteq D$. Then the function $\mathcal N (y, \varphi, E)\colon \mathbb R^n \to\mathbb {N} \cup \{0,\infty \} $ defined by

$$ \begin{equation*} \mathbb {R}^n\ni y\mapsto \mathcal N (y, \varphi, E) = \begin{cases} 0 \quad&\text{if $\varphi^{-1}(y) \cap E$ is empty,}\\ \infty \quad&\text{if $\varphi^{-1}(y) \cap E$ is infinite,}\\ \#(\varphi^{-1}(y) \cap E) \quad&\text{otherwise}, \end{cases} \end{equation*} \notag $$
is called the Banach indicatrix of $\varphi$. Here the symbol $\#(\varphi^{-1}(y) \cap E)$ stands for the number of points in the preimage $\varphi^{-1}(y) \cap E$ of $y$.

Proposition 2. Let $\varphi\colon \Omega\to \mathbb R^n$ be a map in the Sobolev class $W^1_{1,\mathrm{loc}}(\Omega)$ (or in the class $\operatorname{ACL}(\Omega)$). Then

1) there exists a Borel set $\Sigma\subset \Omega$ of zero measure such that $\varphi\colon \Omega\setminus\Sigma\to\mathbb R^n$ has the Luzin $\mathcal N$-property;

2) the functions

$$ \begin{equation*} \Omega\setminus\Sigma\ni x\mapsto (u\circ \varphi)(x) |{\det D\varphi (x)}|\quad\textit{and} \quad \mathbb R^n \ni y\mapsto u(y)\mathcal N(y,\varphi, \Omega\setminus \Sigma) \end{equation*} \notag $$
are measurable if so is $u\colon \mathbb R^n \to\mathbb R$;

3) if $A\subset \Omega\setminus \Sigma$ is a measurable set, then the following area formula holds:

$$ \begin{equation*} \int_{A} |{\det D\varphi (x)}|\, dx=\int_{\mathbb{R}^{n}} \mathcal N (y, \varphi, A) \, dy; \end{equation*} \notag $$

4) if a measurable function $h$ is nonnegative, then the integrands in (2.2) are measurable and the following formula for a change of variable in the Lebesgue integral holds:

$$ \begin{equation} \int_{\Omega\setminus \Sigma} u(\varphi(x))|{\det D\varphi (x)}|\,dx= \int_{\mathbb{R}^{n}}\sum_{x \in \varphi^{-1}(y)\setminus \Sigma} u(x)\, dy; \end{equation} \tag{2.2} $$

5) if one of the functions

$$ \begin{equation*} \Omega\setminus\Sigma \ni x\mapsto (u\circ \varphi)(x) | \det D\varphi (x) |\quad\textit{and} \quad \mathbb R^n \ni y\mapsto u(y)\mathcal N(y,\varphi, \Omega\setminus \Sigma) \end{equation*} \notag $$
is integrable, then the other is too and
$$ \begin{equation} \int_{\Omega\setminus\Sigma} u(\varphi(x))|{\det D\varphi (x)}|\,dx =\int_{\mathbb R^n}u(y)\mathcal N(y,\varphi, \Omega\setminus \Sigma)\,dy. \end{equation} \tag{2.3} $$

2.4. The definition of the class of $\mathcal Q_{q,p}(D',\omega;D)$-homeomorphisms, and their properties

As the system of open sets $\mathcal O_c(D')$ on which we define a quasiadditive set function $\Psi$ we take the minimal system of open subsets of $D'$ (see Definition 2) that contains

In the next definition and in Theorem 1, as a bounded quasiadditive set function we take a map $\Phi\colon \mathcal O_c(D')\to[0,\infty)$.

Definition 3 (see [11] and [14]). 1) Let $D$ and $D'$ be two domains in $\mathbb R^n$, ${n\geqslant2}$. We say that a homeomorphism $f\colon D'\to D$ belongs to the class $Q\mathcal{RQ}_{q,p}(D',\omega;D)$, where ${1< q\leqslant p<\infty}$ for $n\geqslant3$, or $1\leqslant q\leqslant p<\infty$ for $n=2$ and $\omega\in L_{1,\mathrm{loc}}(D')$ is a weight function, if there exists

or such that for each cubic condenser $E=(\overline{Q(x,r)}, Q(x,R))$, $0<r<R$, lying in $D'$ whose image $f(E)=(f(\overline{Q(x,r)}), f(Q(x,R))$ lies in $D$ we have
$$ \begin{equation} \begin{cases} \operatorname{cap}^{1/p}(f(E); L^1_p(D)) \leqslant K_p\operatorname{cap}^{1/p}(E; L^1_p(D';\omega)) &\text{for }q=p, \\ \operatorname{cap}^{1/q}(f(E); L^1_q(D)) \\ \ \ \leqslant \Psi_{q,p}(Q(x,R)\setminus \overline{Q(x,r)})^{1/\sigma} \operatorname{cap}^{1/p}(E; L^1_p(D';\omega)) &\text{for }q<p, \end{cases} \end{equation} \tag{2.4} $$
where $1/\sigma=1/q-1/p$.

2) Let $ D$ and $D'$ be domains in $\mathbb R^n$, $n\geqslant2$. We say that a homeomorphism is in the class $\mathcal{Q}_{q,p}(D',\omega;D)$, where $1< q\leqslant p<\infty$ for $n\geqslant3$, or $1\leqslant q\leqslant p<\infty$ for $n=2$ and $\omega\in L_{1,\mathrm{loc}}(D')$ is a weight function, if there exists

($\widetilde{\rm{a}}$) a constant $\widetilde K_p>0$ in the case $q=p$,

or

($\widetilde{\rm{b}}$) a bounded quasiadditive function $ \widetilde\Psi_{q,p}$ defined on the system $\mathcal O(D')$ of all open subsets of $D'$ in the case $q<p$

such that for each condenser $E=(F_1,F_0)$ lying in $D'$ and its image $f(E)=(f(F_1), f(F_0))$ lying in $D$ we have

$$ \begin{equation} \begin{aligned} \, \notag &\operatorname{cap}^{1/q}(f(E); L^1_q(D)) \\ &\qquad\leqslant \begin{cases} \widetilde K_p \operatorname{cap}^{1/p}(E; L^1_p(D';\omega)) &\text{for }1<q=p<\infty, \\ \widetilde\Psi_{q,p}(D'\setminus(F_0\cup F_1))^{1/\sigma} \operatorname{cap}^{1/p}(E; L^1_p(D';\omega)) &\text{for }1<q<p<\infty, \end{cases} \end{aligned} \end{equation} \tag{2.5} $$
where ${1}/{\sigma}={1}/{q}-{1}/{p}$.

3) If (2.5) holds for all ring condensers $E=(F,\partial U)$ in $D'$, then we say that the mapping $f$ belongs to the class $\mathcal{RQ}_{q,p}(D',\omega;D)$.

The following inclusions are obvious:

$$ \begin{equation*} Q\mathcal{RQ}_{q,p}(D',\omega;D)\subset \mathcal{RQ}_{q,p}(D',\omega;D)\subset \mathcal{Q}_{q,p}(D',\omega;D). \end{equation*} \notag $$

Definition 4. Let $\varphi\colon D\to D'$ be a homeomorphism between domains $D, D'\subset \mathbb R^n$ that induces a bounded composition operator

$$ \begin{equation*} \varphi^*\colon {L}^1_p(D';\omega) \cap \operatorname{Lip}_{\mathrm{loc}}(D')\to L^1_q(D), \qquad 1\leqslant q \leqslant p<\infty, \end{equation*} \notag $$
where $\omega\colon D'\to (0,\infty)$ is a locally integrable weight function.

Fix an arbitrary open set $W\subset D'$ and consider the restriction of $\varphi^*$ to the subspace6

$$ \begin{equation} \mathcal R(W)={L}^1_p(W;\omega) \cap \mathring{\operatorname{Lip}}_{\mathrm{loc}}(W)\subset{L}^1_p(D';\omega) \cap \operatorname{Lip}_{\mathrm{loc}}(D'). \end{equation} \tag{2.6} $$
(Here $\mathring{\operatorname{Lip}}_{\mathrm{loc}}(W)\subset \operatorname{Lip}_{\mathrm{loc}}(D')$ is a subspace of the space of locally Lipschitz functions on $D'$ that vanish identically outside $W$.) It is obvious that the norm of the restriction $\varphi_W\colon {L}^1_p(W;\omega) \cap \mathring{\operatorname{Lip}}_{\mathrm{loc}}(W)\to L^1_q(D)$ can depend on $W$:
$$ \begin{equation} \|\varphi^*_W\|=\sup_{\substack{u\in \mathcal R(W)\\ u\ne0}} \frac{\|\varphi_W^*u\mid L^1_q(D)\|}{\|u\mid L^1_p(W;\omega)\|}, \qquad1\leqslant q \leqslant p<\infty \end{equation} \tag{2.7} $$
(here we assume that the denominator in (2.7) is distinct from zero). For $1\leqslant q<p<\infty$ we define a set function by assigning to $W\subset D'$ the quantity
$$ \begin{equation} \Phi(W)=\|\varphi^*_W\|^\sigma, \quad\text{where } \frac1\sigma=\frac1q-\frac1p. \end{equation} \tag{2.8} $$

In the following theorem we describe analytically those mappings whose inverse maps belong to $Q\mathcal{RQ}_{q,p}(D',\omega;D)$.

Theorem 1 (see [11]–[14]). A homeomorphism $f\colon D' \to D$ belongs to the class $Q\mathcal{RQ}_{q,p}(D',\omega;D)$, where $1<q\leqslant p<\infty$ for $n\geqslant 3$, or $1\leqslant q\leqslant p<\infty$ for $n=2$, if and only if the inverse homeomorphism $\varphi=f^{-1}\colon D\to D'$ has one of the following properties 1)–4).

1) The composition operator $\varphi^*\colon {L}^1_p(D';\omega) \cap \operatorname{Lip}_{\mathrm{loc}}(D')\to L^1_q(D)$, $1\!<\!q\!\leqslant\!p\!<\!\infty$, is bounded.

2) For any condenser $E=(F_1,F_0)$ in $D'$ with inverse image $\varphi^{-1}(E)=(\varphi^{-1}(F_1), \varphi^{-1}(F_0))$ in $D$,

$$ \begin{equation} \operatorname{cap}^{1/q}(\varphi^{-1}(E); L^1_q(D)) \leqslant \begin{cases} \widetilde K_p \operatorname{cap}^{1/p}(E; L^1_p(D';\omega)), \\ \widetilde\Psi_{q,p}(D'\setminus(F_0\cup F_1))^{1/\sigma} \operatorname{cap}^{1/p}(E; L^1_p(D';\omega)), \end{cases} \end{equation} \tag{2.9} $$
where the ranges of the parameters $q$ and $p$ are as indicated in Definition 3, ${{1}/{\sigma}={1}/{q}- {1}/{p}}$, and $\widetilde\Psi_{q,p}$ is a bounded quasiadditive set function defined on open subsets of $D'$.

3) For each ring condenser $E=(F,U)$ in $D'$ with inverse image $\varphi^{-1}(E)=(\varphi^{-1}(F),\varphi^{-1}(U))$ in $D$

$$ \begin{equation} \begin{aligned} \, \notag &\operatorname{cap}^{1/q}(\varphi^{-1}(E); L^1_q(D)) \\ &\qquad\leqslant \begin{cases} K_p \operatorname{cap}^{1/p}(E; L^1_p(D';\omega)), &1<q=p<\infty, \\ \Psi(U\setminus F)^{1/\sigma} \operatorname{cap}^{1/p}(E; L^1_p(D';\omega)), &1<q<p<\infty, \end{cases} \end{aligned} \end{equation} \tag{2.10} $$
where $K_p $ is a constant and $\Psi$ is a bounded quasiadditive set function defined on some system7 $\mathcal O(D')$ of open subsets of $D'$.

4) For the homeomorphism $\varphi \colon D \to D'$ the following hold:

$\qquad$(a) it belongs to the Sobolev class $W^1_{q, \operatorname{loc}}(D)$;

$\qquad$(b) it has a finite distortion: $D\varphi(x)=0$ almost everywhere on the set $Z={\{x\in D \mid \det D\varphi(x)=0\}}$;

$\qquad$(c) the operator distortion function

$$ \begin{equation} D\ni x \mapsto K^{1,\omega}_{q,p}(x,\varphi) = \begin{cases} \dfrac{|D\varphi(x)|}{|{\det D\varphi (x)}|^{1/p}\omega^{1/p}(\varphi(x))} &\textit{for }\det D\varphi (x)\neq 0, \\ 0 &\textit{for }\det D\varphi (x) = 0, \end{cases} \end{equation} \tag{2.11} $$
belongs to $L_{\sigma}(D)$, where ${1}/{\sigma}={1}/{q}-{1}/{p}$ for $1\leqslant q<p<\infty$ and $\sigma=\infty$ for ${q=p}$.

Moreover, $\varphi\in W^1_{q, \operatorname{loc}}(D)$ and

$$ \begin{equation} \begin{aligned} \, &\begin{cases} 2^{-n/p}\biggl(\dfrac{3n}2\biggr)^{-1}\|K^{1,\omega}_{p,p}(\,\cdot\,)\mid L_\sigma(\varphi^{-1}(W))\|, \\ 2^{-n/q}\biggl(\dfrac{3n}2\biggr)^{-1}\|K^{1,\omega}_{q,p}(\,\cdot\,)\mid L_\sigma(\varphi^{-1}(W))\| \end{cases}\notag \\ &\quad\leqslant\|\varphi_W^*\|\leqslant\|K^{1,\omega}_{q,p}(\,\cdot\,)\mid L_\sigma(\varphi^{-1}(W))\| \notag \\ &\quad\leqslant\begin{cases} 7^{n/p} nK_p, &1< q= p<\infty, \\ 7^{n/q} n\Psi_{q,p}(W)^{1/\sigma}, &1< q< p<\infty, \end{cases} \quad\textit{for } \varphi\in Q\mathcal{RQ}_{q,p}(D',\omega;D), \notag \\ &\quad\leqslant\begin{cases} {3n} 2^{(n-p)/p}\widetilde K_p, &q=p, \\ {3n} 2^{(n-q)/q}\widetilde\Psi(W)^{1/\sigma}, &q<p, \end{cases} \quad\textit{for } \varphi\in \mathcal{Q}_{q,p}(D',\omega;D)\qquad \end{aligned} \end{equation} \tag{2.12} $$
for each open set $W\in \mathcal O(D')$. (The quantity $\|\varphi_W^*\|$ was defined above in (2.7).)

5) Properties 1)4) also hold for $n=2$ and $1=q \leqslant p<\infty$.

The proof of Theorem 1 was presented in [11], Theorem 1, [12], [13] and [14], Theorem 1. Here are some relevant comments to these references.

That conditions 1) and 4) in Theorem 1 are necessary was shown in [13], Theorem 18, under the assumption that $f\in Q\mathcal{RQ}_{q,p}(D',\omega;D)$, where $1<q\leqslant p<\infty$ for $n\geqslant 3$ and $1\leqslant q\leqslant p<\infty$ for $n=2$. The implication 1) $\Rightarrow$ 2) is proved just as the implication 1) $\Rightarrow$ 2) in Theorem 1 from [14].

To prove that condition 1) is sufficient, it was shown in [14], Theorem 1 (in [13], Theorem 18) that condition 2) also holds for $1<q\leqslant p<\infty$ and $n\geqslant 3$ (for $1\leqslant q\leqslant p<\infty$ and $n=2$, respectively). If a homeomorphism $\varphi\colon D\to D'$ satisfies condition 2), then it also clearly satisfies condition 1).

If condition 3) holds, then the inverse homeomorphism $f=\varphi^{-1}\colon D'\to D$ belongs to $Q\mathcal{RQ}_{q,p}(D',\omega;D)$.

It was shown in [14], Theorem 1 (in [13], Theorem 18, respectively) that when condition 4) holds, so does condition 1), and therefore $f=\varphi^{-1}\in Q\mathcal{RQ}_{q,p}(D',\omega;D)$ by the above.

The second line of inequality (2.12) was proved8 in [13], while the first and third lines were proved in [14].

Corollary 1. Let $f\colon D'\to D$ be a homeomorphism in the class $Q\mathcal{RQ}_{q,p}(D',\omega;D)$, where $1<q\leqslant p<\infty$ for $n\geqslant 3$, or $1\leqslant q\leqslant p<\infty$ for $n=2$. Then ${f\in\mathcal{Q}_{q,p}(D',\omega;D)}$, that is, (2.9) holds for each condenser $E=(F_1,F_0)$ in $D'$ with image $f(E)=(f(F_1),f(F_0))$ in $D$. Furthermore, the least constant $\widetilde K_p$ for $q=p$ (the quantity

$$ \begin{equation*} \sup_E\frac{\operatorname{cap}^{1/q}(f(E); L^1_p(D))}{\operatorname{cap}^{1/p}(E; L^1_p(D';\omega))}, \end{equation*} \notag $$
where for $q<p$ the upper bound in taken over all condensers $E=(F_1,F_0)$ in $D'$ such that the denominator is distinct from zero), just as in Theorem 1, is estimated in terms of the least constant $K_p$ (the quantity $ \Psi_{q,p}(D')^{1/\sigma}$ in (2.4)) times a multiplicative constant depending only on $n$ and $p$ (on $n$, $p$ and $q$, respectively).

The differentiability properties of maps in the classes $\mathcal Q_{q,p}(D',\omega;D)$ were established in [12] and [13], Theorem 2.

The homeomorphisms $\varphi\colon D \to D'$ in Theorem 1 have the following properties:

We can extract two classes of $Q_{q,p}$-homeomorphisms from Theorem 1.

Example 1 (see [11] and [14]). If the homeomorphism $\varphi\colon D \to D'$ induces a bounded composition operator $\varphi^*\colon {L}^1_p(D';\omega) \cap \operatorname{Lip}_{\mathrm{loc}}(D')\to L^1_q(D)$, where $1<q\leqslant p<\infty$ for $n\geqslant 3$, or $1\leqslant q\leqslant p<\infty$ for $n=2$, then the inverse homeomorphism $f=\varphi^{-1}\colon D'\to D$ belongs to the class $Q\mathcal{RQ}_{q,p}(D',\omega;D)$.

Example 2 (see [11] and [14]). Consider a homeomorphism $\varphi \colon D \to D'$ with the following properties:

Then the inverse homeomorphism $f=\varphi^{-1}\colon D'\to D$ belongs to the class $Q\mathcal{RQ}_{q,p}(D',\omega;D)$.

In addition to Examples 1 and 2, the reader can find in [13] further examples of classes of maps in the family $\mathcal Q_{q,p}(D',\omega;D)$.

Example 3 (see [13], Example 3). Let $\varphi\colon D\to D'$ be a homeomorphism in the Sobolev class $W^1_{p,\mathrm{loc}}(D)$, where $1<p<\infty$ for $n\geqslant3$, or $1\leqslant p<\infty$ for $n=2$, which has a finite distortion. Then the inverse homeomorphism $f=\varphi^{-1}\colon D'\to D$ belongs to the class $\mathcal Q_{p,p}(D',\omega;D)$ with constant $K_p=1$ and weight function

$$ \begin{equation} D'\ni y\mapsto \omega(y)= \begin{cases} \dfrac{|D\varphi(\varphi^{-1}(y))|^p}{|{\det D\varphi (\varphi^{-1}(y))}|} &\text{if } y\in D'\setminus (Z'\cup\Sigma'), \\ 1&\text{otherwise}. \end{cases} \end{equation} \tag{2.13} $$

Remark 1. It was shown in [13], Theorem 5, that the weight function (2.13) is locally integrable.

Example 4 (see [13], Example 4). Let $n-1< s<\infty$, and let $f \colon D' \to D$ be a homeomorphism between open domains $D', D\subset \mathbb{R}^n$, $n\geqslant 2$, such that

Then the inverse homeomorphism $\varphi=f^{-1}\colon D\!\to\! D'$ has the following properties:

while the direct homeomorphism $f\colon D'\to D$

Example 5 (see [13], Example 5). Let $n-1< s<\infty$, and let $f \colon D' \to D$ be a homeomorphism between open domains $D', D\subset \mathbb{R}^n$, $n\geqslant 2$, such that

Then the inverse homeomorphism $\varphi=f^{-1}\colon D\!\to\! D'$ has the following properties:

while the direct homeomorphism $f\colon D'\to D$

Example 6 (see [16], Definition 11 and Theorem 34). A homeomorphism $f\colon D'\to D$ is called a homeomorphism with bounded inner $\theta$-weighted $(s,r)$-distortion (belongs to the class $\mathcal{ID}(D;s,r;\theta,1)$), where $n-1< s\leqslant r<\infty$, if

Set $\mathcal K^{\theta,1}_{s,r}(f;D')=\|\mathcal K_{s,r}^{\theta,1}(\,\cdot\,,f)\mid L_{\varrho}(D')\|$.

If $n-1< s\leqslant r<\infty$ and the function $\omega(x)=\theta^{-(n-1)/(s-(n-1))}(x)$ is locally integrable, then $f\colon D'\to D$ belongs to the family

$$ \begin{equation*} \mathcal Q_{q,p}(D',\omega;D), \end{equation*} \notag $$
where $q=r/(r-(n-1))$ and $p=s/(s-(n-1))$, $1<q\leqslant p<\infty$. Furthermore, the coefficients on the right-hand sides of relations in (2.4) are $K_p={\|\mathcal K_{r,r}^{\theta,1}(\,\cdot\,,f)\,{\mid}\, L_{\infty}(\Omega)\|}$ for $q=p$ and
$$ \begin{equation*} \Psi_{q,p}(Q(x,R)\setminus \overline{Q(x,r)})^{1/\sigma}= \bigl\|\mathcal K_{s,r}^{\theta,1}(\,\cdot\,,f)\mid L_{\varrho}(Q(x,R)\setminus \overline{Q(x,r)})\bigr\| \quad\text{for } q<p, \end{equation*} \notag $$
where $1/\sigma=1/q-1/p={1}/{\varrho}$.

Example 7 (see [17], Definition 3 and Theorem 19). A homeomorphism $f\colon D'\to D$ belongs to the class $\mathcal{OD}(D';s,r;\theta,1)$, where $n-1< s\leqslant r<\infty$ (is called a map with bounded outer $\theta$-weighted $(s,r)$-distortion) if

Then, provided that $n-1< s\leqslant r<\infty$ and the function $\omega(x)=\theta^{-(n-1)/(s-(n-1))}(x)$ is locally integrable, the homeomorphism $f\colon D'\!\to\! D$ belongs to

$$ \begin{equation*} \mathcal Q_{q,p}(D',\omega;D), \end{equation*} \notag $$
where $q=r/(r-(n-1))$ and $p=s/(s-(n-1))$, $1<q\leqslant p<\infty$. Moreover, the coefficients on the right-hand sides of (2.4) are equal to $K_p=\|K_{r,r}^{\theta,1}(\,\cdot\,,f) | L_{\infty}(D')\|^{n-1}$ for $q= p$ and
$$ \begin{equation*} \Psi_{q,p}(Q(x,R)\setminus \overline{Q(x,r)})^{1/\sigma}= \bigl\|K_{s,r}^{\theta,1}(\,\cdot\,,f)\mid L_{\rho}(Q(x,R)\setminus \overline{Q(x,r)})\bigr\|^{n-1} \quad\text{for } q<p, \end{equation*} \notag $$
where $1/\sigma=1/q-1/p=(n-1)/{\varrho}$.

It was proved in [17], Theorem 8, that for $n-1< s\leqslant r<\infty$ we have

$$ \begin{equation*} \mathcal{OD}(\Omega;s,r;\theta,1)\subset\mathcal{ID}(\Omega;s,r;\theta,1). \end{equation*} \notag $$
Moreover, for each homeomorphism $f\colon D'\to D$ in the class $\mathcal{OD}(D';s,r;\theta,1)$, where $n-1< s\leqslant r<\infty$, we have
$$ \begin{equation*} \bigl\|\mathcal K_{s,r}^{\theta,1}(\,\cdot\,,f)\mid L_{\sigma}(D')\bigr\| \leqslant\bigl\|K_{s,r}^{\theta,1}(\,\cdot\,,f)\mid L_\rho(D')\bigr\|^{n-1}, \end{equation*} \notag $$
where $\rho$ and $\sigma$ were defined in Examples 6 and 7.

2.5. The moduli of families of curves and homeomorphisms in the class $ \mathcal Q_{q,p}(D',\omega)$

Let $D'$ is a domain in $ \mathbb{R}^{n}$, $n \geqslant 2$, and $\omega\colon D'\to (0, \infty) $ be a weight function in $L_{1,\mathrm{loc}}$. Also let $\Gamma $ be an arbitrary family of locally rectifiable, continuous curves (briefly, paths) $\gamma\colon [a,b]\to D'$.

Recall that, given a family of curves $\Gamma$ в $D'$ and a real number $p\geqslant 1$, the (weighted) $p$-modulus of the family $\Gamma$ is defined by

$$ \begin{equation*} \operatorname{mod}_{p}(\Gamma)=\inf_\rho \int_{D'} \rho^{p}\, dx \qquad \biggl( \operatorname{mod}_{p}(\Gamma)=\inf_\rho \int_{D'} \rho^{p}\omega(x)\, dx\biggr), \end{equation*} \notag $$
where the infimum is taken over all nonnegative Borel functions $ \rho\colon D' \to [0, \infty]$ such that
$$ \begin{equation} \int_{\gamma} \rho\, ds \geqslant 1 \end{equation} \tag{2.18} $$
for all paths $ \gamma \in \Gamma$. Recall that, for a rectifiable curve $\gamma\colon [a,b]\to D'$, the integral in (2.18) is defined as the quantity
$$ \begin{equation*} \int_{0}^{l(\gamma)} \rho(\widetilde{\gamma}(t))\, dt, \end{equation*} \notag $$
where $l(\gamma)$ is the length of the curve $\gamma\colon [a,b]\to D'$ and $\widetilde{\gamma}\colon [0,l(\gamma)]\to D'$ is the natural parametrization of this curve, that is, the unique continuous map such that $\gamma=\widetilde{\gamma}\,{\circ}\, S_{\gamma}$, where $S_{\gamma}\colon [a,b]\to[0,l(\gamma)]$ is the length function, whose value at $t\in [a,b]$ is defined by $S_{\gamma}(t)=l(\gamma\vert_{[a,t]})$. If $\gamma$ is a locally rectifiable curve, then we set
$$ \begin{equation*} \int_{\gamma} \rho\,ds = \sup \int_{\gamma'} \rho\,ds; \end{equation*} \notag $$
here the supremum is taken over all rectifiable subcurves $ \gamma'\colon [a', b'] \to D' $ of $\gamma$, where $[a', b']\subset(a,b)$ and $ \gamma'= \gamma_{[a', b']}$.

Functions $\rho$ satisfying (2.18) are called admissible functions or metrics for the family $\Gamma$.

Theorem 2 (inequalities between moduli for maps in the class $\mathcal{OD}(D,D';q,p;1,\omega)$; see [10]). Given a homeomorphism $\varphi\colon D\to D'$ between two domains $D,D'\subset\mathbb R^n$, $n\geqslant2$, and a locally integrable weight function $\omega\colon D'\to (0,\infty)$, assume that $\varphi\colon {D \to D'}$ belongs to the family

$$ \begin{equation*} \mathcal{OD}(D,D'; q,p; 1,\omega), \quad \textit{where}\ \begin{cases} n-1< q \leqslant p<\infty &\textit{for } n>2, \\ 1\leqslant q \leqslant p<\infty &\textit{for }n=2. \end{cases} \end{equation*} \notag $$
Then for the same values of the parameters $q\leqslant p$ the following inequality holds for each family $\Gamma$ of paths in $D$:
$$ \begin{equation} (\operatorname{mod}_q(\Gamma))^{1/q}\leqslant K_{q,p}(\operatorname{mod}^\omega_p(\varphi \Gamma))^{1/p}, \end{equation} \tag{2.19} $$
where $K_{q,p}=\|K^{1,\omega}_{q,p}(\,\cdot\,,\varphi)\mid L_\sigma(D)\|$, $K^{1,\omega}_{q,p}(\,\cdot\,,\varphi)$ is the distortion function (2.11), and ${1}/{\sigma}={1}/{q}-{1}/{p}$ for $q<p<\infty$, while $\sigma=\infty$ for $q=p$.

The claim also holds for $1=q \leqslant p<\infty$ and $n=2$.

As a corollary to Theorem 2 (see [14], Lemma 2.3, and [13], Theorem 18), we obtain the following.

Theorem 3 (modulus-based description of maps in the classes $Q\mathcal{RQ}_{q,p}(D',\omega)$; see [10]). Fix a homeomorphism $f\colon D'\to D$ between domains $D',D\subset \mathbb R^n$, $n\geqslant2$, and a locally integrable weight function $\omega\colon D'\to (0,\infty)$.

1) Assume that $f\colon D'\to D$ belongs to the family

$$ \begin{equation*} Q\mathcal{RQ}_{q,p}(D',\omega), \qquad n-1< q \leqslant p<\infty. \end{equation*} \notag $$
Then for each family $\Gamma$ of paths in $D'$
$$ \begin{equation} (\operatorname{mod}_q(f\Gamma))^{1/q}\leqslant \begin{cases} K_{p,p} (\operatorname{mod}^\omega_p(\Gamma))^{1/p} &\textit{for } n-1<q=p<\infty, \\ K_{q,p}(\operatorname{mod}^\omega_p(\Gamma))^{1/p} &\textit{for } n-1<q<p<\infty, \end{cases} \end{equation} \tag{2.20} $$
where $K_{q,p}$ in the constant from (2.19), and also
$$ \begin{equation} (\operatorname{mod}_q(f\Gamma))^{1/q}\leqslant \begin{cases} K_{p,p}(E) (\operatorname{mod}^\omega_p(\Gamma))^{1/p} &\textit{for } n-1<q=p<\infty, \\ K_{q,p}(E)(\operatorname{mod}^\omega_p(\Gamma))^{1/p} &\textit{for } n-1<q<p<\infty, \end{cases} \end{equation} \tag{2.21} $$
for the family $\Gamma$ of all curves10 in the condenser $E=(F,U)$, where
$$ \begin{equation} K_{q,p}(E)=\bigl\|K^{1,\omega}_{q,p}(\,\cdot\,,\varphi)\mid L_\sigma(f(U\setminus F))\bigr\|, \end{equation} \tag{2.22} $$
$K^{1,\omega}_{q,p}(\,\cdot\,,\varphi)$ is the distortion function (2.11), and ${1}/{\sigma}={1}/{q}-{1}/{p}$ for $n-1< q<p< \infty$, or $\sigma=\infty$ for $q=p$.

2) Assume that $f\colon D'\to D$ satisfies the relations

$$ \begin{equation} \begin{aligned} \, &(\operatorname{mod}_q(f\Gamma))^{1/q}\notag \\ &\quad\leqslant \begin{cases} K_{p,p} (\operatorname{mod}^\omega_p(\Gamma))^{1/p} &\textit{for } n-1<q=p<\infty, \\ \Psi_{q,p}(Q(x,R)\setminus\overline{Q(x,r)})^{1/\sigma}(\operatorname{mod}^\omega_p(\Gamma))^{1/p} &\textit{for } n-1<q<p<\infty, \end{cases} \end{aligned} \end{equation} \tag{2.23} $$
involving the constant $K_{p,p}$ for $1< q=p<\infty$ and the bounded quasiadditive function $\Psi_{q,p}$ for $1< q<p<\infty$, for all cubic condensers $(\overline{Q(x,r)}, Q(x,R))$ in $D'$, $r\in (0,R)$, whose plates are concentric cubes, and for the family $\Gamma$ of all curves $\gamma\colon [a,b]\to D'$ in the condenser $(\overline{Q(x,r)},Q(x,R))$ such that $\gamma(a)\in\overline{Q(x,r)}$ and $\gamma(b)\in\partial{Q(x,R)}$. Then

$\qquad$ (a) the homeomorphism $f$ belongs to the class $ \mathcal Q_{q,p}(D',\omega) $, $n-1< q \leqslant p<\infty$;

$\qquad$ (b) relations (2.20) hold for $f$.

The least constants $K_{q,p}$ in (2.20), the quantities11

$$ \begin{equation} \sup_\Gamma\frac{(\operatorname{mod}_q(f\Gamma))^{1/q}}{(\operatorname{mod}^\omega_p(\Gamma))^{1/p}}\quad\textit{and} \quad \sup_E\frac{(\operatorname{mod}_q(f\Gamma))^{1/q}}{(\operatorname{mod}^\omega_p(\Gamma))^{1/p}}, \end{equation} \tag{2.24} $$
corresponding to all possible choices of the family of curves in (2.20) and all possible condensers in (2.21), have an upper estimate in terms of the quantity $\|K^{1,\omega}_{q,p}(\,\cdot\,,\varphi)\,|\, L_\sigma(D)\|$ in (2.11), with coefficients depending only on $q$, $p$ and $n$ (the denominators of the expressions in (2.24) are distinct from zero).

The above claims also hold for $n=2$ and $1\leqslant q\leqslant p<\infty$.

Remark 2. In the proof of Theorem 3 presented in [10] it was shown that the homeomorphism $f\colon D'\to D$ satisfies the estimates

$$ \begin{equation} \begin{aligned} \, \notag &\bigl(\operatorname{cap}(f(\overline{Q(x,r)}),f(Q(x,R));L^1_q(D))\bigr)^{1/q} =(\operatorname{mod}_q(f\Gamma))^{1/q} \\ \notag &\qquad\leqslant\begin{cases} K_{p,p} (\operatorname{mod}^\omega_p(\Gamma))^{1/p}, & n-1<q=p<\infty, \\ \Psi_{q,p}(Q(x,R)\setminus\overline{Q(x,r)})^{1/\sigma}(\operatorname{mod}^\omega_p(\Gamma))^{1/p}, &n-1<q<p<\infty, \end{cases} \\ &\qquad\leqslant\begin{cases} K_{p,p} \bigl({\operatorname{cap}}(\overline{Q(x,r)},Q(x,R); L^1_p(D';\omega))^{1/p}, \\ \Psi_{q,p}(Q(x,R)\setminus\overline{Q(x,r)})^{1/\sigma} \bigl({\operatorname{cap}}\big(\overline{Q(x,r)},Q(x,R); L^1_p(D';\omega)\bigr)^{1/p} \end{cases} \end{aligned} \end{equation} \tag{2.25} $$
provided that (2.23) holds for $f$.

From Theorems 3 and 1 we can make the following striking conclusion.

Corollary 2. Homeomorphisms $f\colon D'\to D$ in the family $\mathcal Q_{q,p}(D',\omega)$, where ${n-1}< q \leqslant p<\infty$ or $1=q \leqslant p<\infty$ for $n=2$, have two equivalent descriptions, (2.4) in the language of capacities and (2.20) in the language of moduli.

Remark 3. For $q=p=n$ ($n - 1<q=p<n$) we can deduce from (2.25) that the class $\mathcal Q_{n,n}(D',\omega;D)$ ($\mathcal Q_{p,p}(D',\omega;D)$, respectively) contains (see [14], § 4.4) the class of so-called $Q$-homeomorphisms ($(p, Q)$-homeomorphisms)12, which are defined in terms of a controllable change of the modulus of a family of curves; see [6] (and [39], respectively).

From Theorem 3 above we can conclude that the class $\mathcal Q_{n,n}(D',\omega)$ coincides in fact with the family of $Q$-homeomorphisms from [6], § 4.1. Let $D'$ and $ D$ be two domains in $ \mathbb{R}^{n}$, $n \geqslant 2$, and let $Q\colon D' \to [1, \infty) $ be a function in $L_{1,\mathrm{loc}}$. Recall that a homeomorphism $ f\colon D' \to D$ is a $Q$-homeomorphism (see [6], § 4.1) if

$$ \begin{equation} \operatorname{mod}_n(f \Gamma) \leqslant \int_{D'} Q(x) \cdot \rho^{n}(x)\,dx \end{equation} \tag{2.26} $$
for each family $\Gamma $ of paths in $D'$ and any function $\rho$ admissible for $\Gamma$. By Theorem 3 the homeomorphisms satisfying (2.26) coincide with the homeomorphisms $f\colon {D'\to D}$ in the class $\mathcal Q_{n,n}(D',\omega)$ for $\omega=Q$.

Some properties of homeomorphisms in $\mathcal Q_{q,p}(D',\omega)$ were studied in [40] (for $n-1<q<p=n$, $\Psi_{q,n}(U)$ in place of $\Psi_{q,n}(U\setminus F)$ and $\omega\equiv1$), [6], [41]–[45] (for ${q=p=n}$ and $\omega=Q$), [46], [47] (for $1<q=p<n$ and $\omega=Q$) and other papers. In all these works, apart from [40], the distortion of the geometry of condensers was described using the language of the moduli of families of curves. In some cases such a characterisation is more restrictive in its essential capabilities than a description using the language of capacities.

§ 3. New set functions

It this section we define two new set functions; the first is defined in terms of ratios of capacities and the second in terms of ratios of moduli.

3.1. A set function based on capacities

Let $\varphi\colon D\to D'$ be a homomorphism such that the composition operator $\varphi^*\colon {L}^1_p(D';\omega) \cap \operatorname{Lip}_{\mathrm{loc}}(D')\to L^1_q(D)$, ${1< q < p<\infty}$, is bounded. For the pair of numbers $q$, $p$, $1<q<p<\infty$, consider the set function $N_c(W)$ defined on connected open sets $W\subset D'$ by

$$ \begin{equation} W\mapsto N_c(W)=\sup_{E=(F,U)}\frac{\operatorname{cap}^{\sigma/q}(\varphi^{-1}(E); L^1_q(D))}{\operatorname{cap}^{\sigma/p}(E; L^1_p(D';\omega))}, \end{equation} \tag{3.1} $$
where the supremum is taken over all condensers $E=(F,U)$ such that $U\subset W$ and the denominator in (3.1) is distinct from zero. By (2.10) the denominator in (3.1) is always nonzero for $n-1<q$ if $n\geqslant3$ and for $1\leqslant q$ if $n\geqslant2$, provided that the plate $F$ of the condenser $E=(F,U)$ contains a continuum: for instance, see inequality (3.2) in [14]. In this case, for the condenser $E=(F, U)$ in $D$ we have
$$ \begin{equation*} \operatorname{cap}_q^{n-1}(E; L^1_q(U)) \geqslant c\,\frac{(\operatorname{diam} F)^{q}}{|U|^{q-(n-1)}}>0, \end{equation*} \notag $$
where $c$ is a constant depending only on $n$ and $q$.

Note the following properties of the function $N_c(W)$ for connected open sets $W\subset D'$, which are easy to verify:

Consider the variation $(N_c, U)$ of the above set function $D'\supset W\mapsto N_c(W)$. As is known (for instance, see [30]), the variation $V(N_c, U)$ of $N_c$ with respect to an open set $U\subset D'$ is defined by

$$ \begin{equation} V(N_c, U)=\sup \sum_{k=1}^{m} N_c(W_{k}), \end{equation} \tag{3.2} $$
where the supremum is taken over all possible finite systems of disjoint connected open sets $W_{1}, \dots, W_{m}$ lying in $U$. The variation $V(N_c, U)$ is a quasiadditive function of the open set $U \subset D'$. It is obvious that

Note that $V(N_c, U)$, $U\subset D'$, is a bounded quasiadditive function of the open set $U \subset D'$:

In fact, as $\Phi$ is finitely additive, we obtain

$$ \begin{equation} V(N_c, U)=\sup \sum_{k=1}^{m} N_c(W_{k})\leqslant \sup \sum_{k=1}^{m} \Phi(W_{k}) = \Phi\biggl(\bigcup_{k=1}^{m}W_{k}\biggr)\leqslant\Phi(U). \end{equation} \tag{3.3} $$
(Here $W_{1},\dots,W_{m}$ are the disjoint connected open sets from (3.2).)

Proposition 3. For each ring condenser $E=(F,U)$ in $D'$ relation (2.10) holds for $1<q<p<\infty$ and the quasiadditive function $V(N_c, U\setminus F)$ in place of $\Phi(U\setminus F)$.

Proof. Set $W=U\setminus F$. For the condenser $E=(F,U)$ in $W\subset D'$ we have
$$ \begin{equation} \begin{aligned} \, \notag \frac{\operatorname{cap}^{1/q}(\varphi^{-1}(E); L^1_q(D))}{\operatorname{cap}^{1/p}(E; L^1_p(D';\omega))} &=\sup_{u\in \mathcal A(E;L^1_p(D';\omega))} \frac{\displaystyle\biggl(\int_{\varphi^{-1}(U\setminus F)} |\nabla v_e (x)|^q\,dx\biggr)^{1/q}} {\displaystyle\biggl(\int_{U\setminus F} |\nabla u(y)|^p\omega(y)\,dy\biggr)^{1/p}} \\ &=\sup_{u\in \mathcal A(E;L^1_p(D';\omega))} \frac{\displaystyle\biggl(\int_{\varphi^{-1}(U\setminus F)} |\nabla \widetilde v_e(x)|^q\,dx\biggr)^{1/q}} {\displaystyle\biggl(\int_{U\setminus F} |\nabla\widetilde u(y)|^p\omega(y)\,dy\biggr)^{1/p}}, \end{aligned} \end{equation} \tag{3.4} $$
where $\mathcal A(E;\operatorname{Lip}_{\mathrm{loc}}(D'))$ is the class of functions admissible for $E=(F,U)$ in $D'$, $v_e$ is the extremal function for the capacity $\operatorname{cap}(\varphi^{-1}(E); L^1_q(D))$, and $\widetilde u$ ($\widetilde v_e$) is defined in terms of the function $u$ (of $v_e$, respectively), by the formula
$$ \begin{equation} \widetilde u=\min(2u,1 )-\max(2u,1)+1. \end{equation} \tag{3.5} $$
In verifying the last equality in (3.4) note that the equalities $|\nabla \widetilde v_e(x)|=2|\nabla v_e(x)|$ and $|\nabla \widetilde u(x)|=2|\nabla u(x)|$ hold for almost all $x\in U$.

Note that13

$$ \begin{equation} \int_{U\setminus F} |\nabla\widetilde u(y)|^p\omega(y)\,dy \geqslant\operatorname{cap}\biggl(\biggl(u^{-1}\biggl(\frac12\biggr),U\setminus F\biggr); L^1_p(D';\omega)\biggr) \end{equation} \tag{3.6} $$
and
$$ \begin{equation} \int_{\varphi^{-1}(U\setminus F)} |\nabla \widetilde v_e(x)|^q\,dx =\operatorname{cap}\biggl(\biggl(\varphi^{-1}\biggl(u^{-1}\biggl(\frac12\biggr)\biggr), \varphi^{-1}(U\setminus F)\biggr); L^1_q(D)\biggr). \end{equation} \tag{3.7} $$
Inequality (3.6) holds because $\widetilde u$ is admissible for the capacity $\operatorname{cap}((u^{-1}(1/2), {U\,{\setminus}\, F}); L^1_q(D))$. Inequality (3.7) holds because otherwise, starting from the extremal function for the capacity $\operatorname{cap}((\varphi^{-1}(u^{-1}(1/2)),\varphi^{-1}(U\setminus F)); L^1_q(D))$ we can construct (by inverting the operation in (3.5)) an admissible function for $\operatorname{cap}(\varphi^{-1}(E); L^1_q(D))$ and thus arrive at a contradiction with the fact that $v_e$ is an extremal function. Next, substituting the smaller value from (3.6) into the denominator in (3.4) and the greater value from (3.7) into the numerator, using inequalities 3) and 4) we obtain an upper estimate for the left-hand side on (3.4):
$$ \begin{equation} \begin{aligned} \, \notag &\frac{\operatorname{cap}^{\sigma/q}(\varphi^{-1}(E); L^1_q(D))}{\operatorname{cap}^{\sigma/p}(E; L^1_p(D';\omega))} \\ \notag &\qquad\leqslant\sup_{u\in \mathcal A(E;L^1_p(D';\omega))} \frac{\operatorname{cap}^{\sigma/q}\bigl((\varphi^{-1}(u^{-1}(1/2)),\varphi^{-1}(U\setminus F)); L^1_q(D)\bigr)}{\operatorname{cap}^{\sigma/p}\bigl((u^{-1}(1/2),U\setminus F); L^1_p(D';\omega)\bigr)} \\ &\qquad\leqslant N_c(U\setminus F)\leqslant V(N_c, U\setminus F)\leqslant\|\varphi_{U\setminus F}^*\|^\sigma\leqslant \bigl\|K^{1,\omega}_{q,p}(\,\cdot\,)\mid L_\sigma(\varphi^{-1}(U\setminus F))\bigr\|^\sigma. \end{aligned} \end{equation} \tag{3.8} $$
From (3.8) we obtain
$$ \begin{equation*} N_c(W)=\sup_{E=(F,U)}\frac{\operatorname{cap}^{\sigma/q}(\varphi^{-1}(E); L^1_q(D))}{\operatorname{cap}^{\sigma/p}(E; L^1_p(D';\omega))}= N_c(U\setminus F)\leqslant V(N_c, U\setminus F)\leqslant\|\varphi_{U\setminus F}^*\|^\sigma, \end{equation*} \notag $$
where the upper bound is taken over all condensers $E=(F,U)$ such that $U\subset W$.

Proposition 3 is proved.

Remark 4. Since (2.10) holds for $1<q<p<\infty$ and the bounded quasiadditive function $V(N_c, U\setminus F)$ in place of $\Phi(U\setminus F)$, we obtain the following inequalities (see the second line in (2.12) and inequality 4) above):

$$ \begin{equation} \|\varphi_{W}^*\|\leqslant\bigl\|K^{1,\omega}_{q,p}(\,\cdot\,)\mid L_\sigma(\varphi^{-1}(W))\bigr\| \leqslant n7^{n/q} V(N_c, W)^{1/\sigma}\leqslant n7^{n/q}\|\varphi_{W}^*\| \end{equation} \tag{3.9} $$
for each open set $W\subset D'$.

3.2. A set function based on moduli

In this subsection our considerations rely on the following result, which is a direct consequence of Theorem 3.

Proposition 4 (modulus-based description of maps in the classes $\mathcal{RQ}_{q,p}(D',\omega)$; see [10]). Fix a homeomorphism $f\colon D'\to D$ between domains $D',D\subset \mathbb R^n$, $n\geqslant2$, and a locally integrable weight function $\omega\colon D'\to (0,\infty)$.

1) Assume that $f\colon D'\to D$ belongs to the class

$$ \begin{equation*} \mathcal{RQ}_{q,p}(D',\omega), \qquad n-1< q \leqslant p<\infty. \end{equation*} \notag $$
Then for each family $\Gamma$ of curves 14 in an arbitrary condenser $E=(F,U)\subset D'$,
$$ \begin{equation} (\operatorname{mod}_q(f\Gamma))^{1/q}\leqslant \begin{cases} K_{p,p}(E) (\operatorname{mod}^\omega_p(\Gamma))^{1/p} &\textit{for } n-1<q=p<\infty, \\ K_{q,p}(E)(\operatorname{mod}^\omega_p(\Gamma))^{1/p}&\textit{for } n-1<q<p<\infty, \end{cases} \end{equation} \tag{3.10} $$
where
$$ \begin{equation} K_{q,p}(E)=\bigl\|K^{1,\omega}_{q,p}(\,\cdot\,,\varphi)\mid L_\sigma(f(U\setminus F))\bigr\|, \end{equation} \tag{3.11} $$
$K^{1,\omega}_{q,p}(\,\cdot\,,\varphi)$ is the distortion function (2.11), and ${1}/{\sigma}={1}/{q}-{1}/{p}$ for $n-1< q<p< \infty$, or $\sigma=\infty$ for $q=p$.

2) Assume that $f\colon D'\to D$ satisfies the relations

$$ \begin{equation} \begin{aligned} \, &(\operatorname{mod}_q(f\Gamma))^{1/q}\notag \\ &\qquad\leqslant \begin{cases} K_{p,p} (\operatorname{mod}^\omega_p(\Gamma))^{1/p} &\textit{for } n-1<q=p<\infty, \\ \Psi_{q,p}(Q(x,R)\setminus\overline{Q(x,r)})^{1/\sigma}(\operatorname{mod}^\omega_p(\Gamma))^{1/p} &\textit{for } n-1<q<p<\infty, \end{cases} \end{aligned} \end{equation} \tag{3.12} $$
involving the constant $K_{p,p}$ for $n-1< q=p<\infty$ and the bounded quasiadditive function $\Psi_{q,p}$ for $n-1< q<p<\infty$, for all condensers $E=(F,U)$ and the family $\Gamma$ of all curves $\gamma\colon [a,b]\to D'$ in $E=(F,U)$ such that $\gamma(a)\in F$ and $\gamma(b)\in\partial U$. Then

$\qquad$ (a) $f\colon D'\to D$ belongs to the class $ \mathcal{RQ}_{q,p}(D',\omega) $ for $n-1< q \leqslant p<\infty$;

$\qquad$ (b) relations (2.20) hold for $f\colon D'\to D$.

These results also hold for $n=2$ and $1\leqslant q\leqslant p<\infty$.

Let $\varphi\colon D\to D'$ be a homeomorphism such that the composition operator $\varphi^*$: ${L}^1_p(D';\omega) \cap \operatorname{Lip}_{\mathrm{loc}}(D')\to L^1_q(D)$, $n-1< q < p<\infty$, is bounded. Then $f=\varphi^{-1}$: $D'\to D$ belongs to $\mathcal{RQ}_{q,p}(D',\omega) $ for $n-1< q \leqslant p<\infty$.

For a pair of numbers $q$, $p$, $n-1<q<p<\infty$, consider the set function $N_m(W)$ defined on connected open sets $W\subset D'$ by

$$ \begin{equation} W\mapsto N_m(W)=\sup_{E=(F,U)}\sup_{\Gamma}\frac{(\operatorname{mod}_q(f\Gamma))^{\sigma/q}} {(\operatorname{mod}^\omega_p(\Gamma))^{\sigma/p}}, \end{equation} \tag{3.13} $$
where the inner supremum is taken over the families $\Gamma$ of curves in the condenser $E=(F,U)$, and the outer one is taken over the condensers $E=(F,U)$ such that $U\subset W$ (in view of (3.10) the denominator in (3.13) is distinct from zero, provided that the plate $F$ of $E=(F,U)$ contains a continuum).

Note the following properties of $N_m(W)$ for connected open sets $W\subset D'$:

Consider the variation $V(N_m, U)$ of the above set function $D'\supset W\mapsto N_m(W)$. As is known (for instance, see [30]), the variation $V(N_m, U)$ of a set function $N_m$ with respect to an open set $U\subset D'$ is defined by

$$ \begin{equation} V(N_m, U)=\sup \sum_{k=1}^{m} N_m(W_{k}), \end{equation} \tag{3.14} $$
where the upper bound is taken over various finite systems of disjoint connected open sets $W_{1}, \dots, W_{m}$ lying in $U$. The variation $V(N_m, U)$ is a quasiadditive function of the open set $U \subset D'$. It is obvious that

Note that $V(N_m, U)$, $U\subset D'$, is a bounded quasiadditive function of the open set $U \subset D'$:

Really, from the fact that $\Phi$ is finitely additive we deduce

$$ \begin{equation} V(N_m, U)=\sup \sum_{k=1}^{m} N_m(W_{k})\leqslant \sup \sum_{k=1}^{m} K_{q,p}(W)^\sigma = K_{q,p}\biggl(\bigcup_{k=1}^{m}W_{k}\biggr)\leqslant K_{q,p}(U). \end{equation} \tag{3.15} $$
Here $W_{1},\dots,W_{m}$ are the disjoint connected open sets from (3.14).

§ 4. Coincidence of four set functions

In this section we establish some new properties of homeomorphisms $f\colon D'\to D$ between domains $D'$ and $D$ in $\mathbb R^n$, $n\geqslant2$, that belong to the class $Q\mathcal{RQ}_{q,p}(D',\omega;D)$.

Here we prove the central result of this paper.

Theorem 4 (coincidence of the set functions associated with a homeomorphism $Q\mathcal{RQ}_{q,p}(D',\omega)$). Let $\varphi\colon D\to D'$ be a homeomorphism between domains ${D,D'\!\subset\! \mathbb R^n}$, $n\geqslant2$, such that the composition operator

$$ \begin{equation*} \varphi^*\colon {L}^1_p(D';\omega) \cap \operatorname{Lip}_{\mathrm{loc}}(D')\to L^1_q(D) \end{equation*} \notag $$
with parameters $n-1< q\leqslant p<\infty$ for $n\geqslant3$, or $1\leqslant q\leqslant p<\infty$ for $n=2$ and weight function $\omega\in L_{1,\mathrm{loc}}(D')$ is bounded.

Then the following hold.

I. For $q<p$ the following set functions coincide:

$\qquad$ 1) $\mathcal O(D')\ni W\mapsto \Phi(W)=\|\varphi^*_W\|^\sigma$ (see (2.8));

$\qquad$ 2) $\mathcal O(D')\ni W\mapsto \|K^{1,\omega}_{q,p}(\,\cdot\,)\mid L_\sigma(\varphi^{-1}(W))\|^\sigma$ (see (2.11));

$\qquad$ 3) $\mathcal O(D')\ni W\mapsto V(N_c, W)$ (see (3.2));

$\qquad$ 4) $\mathcal O(D')\ni W\mapsto V(N_m, W)$ (see (3.14));

so that

$$ \begin{equation} \|\varphi^*_W\|^\sigma=\bigl\|K^{1,\omega}_{q,p}(\,\cdot\,)\mid L_\sigma(\varphi^{-1}(W))\bigr\|^\sigma=V(N_c, W)=V(N_m, W) \end{equation} \tag{4.1} $$
for each open set $W\in \mathcal O(D')$.

II. For $q=p$ the following quantities coincide:

$\qquad$ 5) the norm of the composition operator $\varphi^*\colon {L}^1_p(D';\omega) \cap \operatorname{Lip}_{\mathrm{loc}}(D')\to L^1_p(D)$ in Theorem 1;

$\qquad$ 6) $\|K^{1,\omega}_{p,p}(\,\cdot\,)\mid L_\infty(D)\|$;

$\qquad$ 7) the least constant $K_p $ in (2.10);

$\qquad$ 8) the least constant $K_{p,p}$ in (3.12);

so that

$$ \begin{equation} \|\varphi^*\|=\|K^{1,\omega}_{p,p}(\,\cdot\,)\mid L_\infty(D)\|=K_p=K_{p,p}. \end{equation} \tag{4.2} $$

Fix a homeomorphism $\varphi\colon D\to D'$ meeting the assumption of Theorem 4. Then by Theorem 1 the inverse homeomorphism $f=\varphi^{-1}$ belongs to the class $\mathcal{RQ}_{q,p}(D',\omega;D)$ with parameters $n-1< q\leqslant p<\infty$ for $n\geqslant3$, or $1\leqslant q\leqslant p<\infty$ for $n=2$ and with weight function $\omega\in L_{1,\mathrm{loc}}(D')$. The following result on the properties of the inverse homeomorphism $f=\varphi^{-1}\colon D'\to D$ was proved in [13].

Theorem 5 (see [13], Theorem 23). Let $n-1<q<\infty$ for $n\geqslant3$, or $1\leqslant q<\infty$ for $n=2$. Then each homeomorphism $f\colon D'\to D$ in the class $\mathcal{RQ}_{q,p}(D',\omega;D)$, $q\leqslant p<\infty$:

$\qquad$ 1) belongs to the Sobolev class $W^1_{1,\mathrm{loc}}(D')$;

$\qquad$ 2) has a finite distortion;

$\qquad$ 3) is differentiable almost everywhere in $D'$.

The proof of Theorem 4 is based on the following lemma.

Lemma 1. Let $f\colon D'\to D$ be a homeomorphism in the class $\mathcal{RQ}_{q,p}(D',\omega;D)$ with parameters satisfying $n-1< q\leqslant p<\infty$ for $n\geqslant3$, or $1\leqslant q\leqslant p<\infty$ for $n=2$ and with weight function $\omega\in L_{1,\mathrm{loc}}(D')$. Then the following inequality holds for almost all points $a\in D'$ such that $\det Df(a)\ne0$ and $\omega(a)\ne0$:

$$ \begin{equation} \frac{|{\operatorname{adj} Df(a)}|}{|{\det Df(a)}|^{(q-1)/q}\omega(a)^{1/p}} \leqslant \begin{cases} \Psi'(a)^{1/\sigma}&\textit{for } q<p, \\ K_p &\textit{for }q=p, \end{cases} \end{equation} \tag{4.3} $$
where $\Psi'(a)$ is the derivative of the quasiadditive set function $\Psi$ in (2.10).

Let $H_{I}(a, f)$ denote the quantity on the left-hand side of (4.3). This quantity is well defined at points $a\in D'$ such that $\det Df(a)\ne0$ and $\omega(a)\ne0$. At points $a\in D'$, where $\det Df(a)=0$, we have $\operatorname{adj} Df(a)= 0$; then we set $H_{I}(a, f)=0$.

Proof of Lemma 1. Consider the case $q<p$ (for $q=p$ the arguments are simpler). The mapping is differentiable almost everywhere in $ D'$. Hence the quantity $H_{I}(x, f)$ is well defined almost everywhere in $D'$. We prove that $H_{I}(x, f)^\sigma\leqslant \Psi'(x)$ almost everywhere in $D'$.

By Theorem 5, $f$ is differentiable at almost all points $x\in D'$. It is sufficient to verify (4.3) for almost all points $x\in D'$ at which $\det Df(x) \neq 0$ (if $\det Df(x)=0$, then this inequality is obvious) and $\omega(x)\ne0$ (almost all points have this property).

Let $a \in D'$ be a point at which

By the above, conditions 1)–3) hold simultaneously at almost all points ${a\!\in\! D'\setminus Z'}$, where $Z'=\{y\in D'\colon \det Df(a)=0\}$.

Now we use the following algebraic lemma.

Lemma 2. Let $L\colon \mathbb R^n\to\mathbb R^n$ be a nondegenerate linear operator. Then there exist orthonormal bases $\{u_1,\dots,u_n\}$ and $\{v_1,\dots,v_n\}$ in $\mathbb R^n$ and nonnegative numbers $\{\lambda_1,\dots,\lambda_n\}$ such that $L(u_i)=\lambda_iv_i$, $i=1,\dots,n$. The vectors $\{u_1,\dots,u_n\}$ are eigenvectors of the operator $L^*L$, and $\lambda_1^2,\dots,\lambda_n^2$ are its eigenvalues. Moreover, $|{\det L}|=\lambda_1\dotsb\lambda_n$.

In Lemma 2 we set $L=Df(a)$ and arrange the semiaxes of the ellipsoid obtained as the image of the unit ball under the linear map $Df(a)$ so that $\lambda_{1} \geqslant \lambda_{2} \geqslant \dots \geqslant \lambda_{n}>0$. To prove (4.3) it is sufficient to show that

$$ \begin{equation} \frac{|{\det Df(a)}|}{\lambda_{n}^{q}}=\frac{|{\operatorname{adj} Df(a)}|^q}{|{\det Df(a)}|^{q-1}} \leqslant \Psi_{p}'(a)^{q/\sigma}\omega(a)^{q/p}, \end{equation} \tag{4.5} $$
where $|{\det Df(a)}|=\lambda_{1} \dotsb \lambda_{n}$, $\lambda_{n}^{-1}={|{\operatorname{adj} Df(a)}|}/{|{\det Df(a)}|}$ and
$$ \begin{equation*} \operatorname{adj} Df(a)\cdot Df(a)=\det Df(a)\cdot E, \end{equation*} \notag $$
where $E$ is the identity matrix.

Let $T_a\mathbb R^n$ ($T_{f(a)}\mathbb R^n$) be the tangent space to $\mathbb R^n$ at the point $a$ (at $f(a)$) with basis vectors $\{u_i\}$ ($\{v_i\}$, respectively).

We fix an arbitrary $t\!\in\!(0,\lambda_{n})$ and select $r\!>\!0$ so that the condenser ${E_r\!=\!(F_r, U_r)}$, where

$$ \begin{equation} F_r=\biggl\{y=\sum_{i=1}^ny_iu_i\colon y_{n}=0,\ |y_{i}| \leqslant r,\ i=1, \dots, n-1\biggr\} \end{equation} \tag{4.6} $$
and
$$ \begin{equation} U_r=\biggl\{y=\sum_{i=1}^ny_iu_i \colon|y_{n}|<r t ,\ |y_{i}|<r(1+ t),\ i=1, \dots, n\biggl\}, \end{equation} \tag{4.7} $$
lies in $D'$. By the condition $f \in \mathcal{RQ}_{q,p}(D',\omega;D)$ we have
$$ \begin{equation} \begin{aligned} \, \notag &\operatorname{cap}(\varphi^{-1}(E_r); L^1_q(D)) \\ &\qquad\leqslant \begin{cases} K^p_p \operatorname{cap}(E_r; L^1_p(D';\omega)) &\text{for }1<q=p<\infty, \\ \Psi(U_r\setminus F_r)^{q/\sigma} \operatorname{cap}^{q/p}(E_r; L^1_p(D';\omega)) &\text{for }1<q<p<\infty, \end{cases} \end{aligned} \end{equation} \tag{4.8} $$
for each ring condenser $E_r=(F_r,U_r)$ in $D'$ with inverse image $\varphi^{-1}(E_r)\,{=}\,(\varphi^{-1}(F_r), \varphi^{-1}(U_r))$ in $D$, where $K_p$ is a constant and $\Psi$ is a bounded quasiadditive set function from (2.10), which is defined on a system of open subsets of $D'$ that contains the complements $U_r \setminus F_r$ for all sufficiently small $r$.

We estimate the capacities in (4.8). For the one on the right-hand side we use Lemma 19 in [13], which claims that

$$ \begin{equation} \begin{aligned} \, \notag \operatorname{cap}(E_r; L^1_p(D';\omega)) &\leqslant\frac{\omega(U_r)}{\operatorname{dist}(F_r, \partial U_r)^{p}} \\ & =\frac{\displaystyle\int_{U_r}\omega(y)\,dy}{(r t)^{p}} =\frac{\omega(a)\mathcal{H}^{n}(U_{r})+o(\mathcal{H}^{n}(U_{r}))}{(rt)^{p}}. \end{aligned} \end{equation} \tag{4.9} $$
On the other hand, by Proposition 5 in [40],
$$ \begin{equation} \operatorname{cap}(f(E_r); L^1_q(D)) \geqslant \frac{\bigl(\inf \mathcal{H}^{n-1}(S)\bigr)^{q}}{\mathcal{H}^{n}(f(U_r))^{q-1}}, \end{equation} \tag{4.10} $$
where $\mathcal{H}^{n-1}(S)$ is the $(n-1)$-dimensional Hausdorff measure of the $C^{\infty}$-manifold $S$, the boundary of an open set $A$ containing $f(F_r)$ and contained in $f(U_r)$ together with its closure, and the infimum is taken over all such sets $S$. (The proof of (4.10) is based on the inequality $\operatorname{cap}(f(E_r); L^1_1(D)) \geqslant \inf_S \mathcal{H}^{n-1}(S)$, established in [48].)

We continue the estimate from below in (4.10) by using the differentiability properties of the map $D'\ni y\mapsto f(y)$ at $x=a$. Let $r>0$ be sufficiently small so that in the expansion

$$ \begin{equation*} f(y)-f(a)=Df(a) + o(1)r \end{equation*} \notag $$
we have $|o(1)|<t$ for $y \in U_r$. Then the measure of $f(U_r)$ exhibits the asymptotic behaviour
$$ \begin{equation*} \mathcal{H}^{n}(f(U_r))=(|{\det Df(a)}|+o(1))\cdot\mathcal{H}^{n}(U_{r})\quad\text{as}\quad r\to0. \end{equation*} \notag $$
The projection of $f(F_r)$ onto the subspace $x_{n}=0$ contains the $(n-1)$-dimensional ellipsoid
$$ \begin{equation*} V_r=(Df(a)-o(1)\operatorname{Id})F_r, \end{equation*} \notag $$
where $\operatorname{Id}$ is the linear operator given by the formula $\operatorname{Id}(u_i)=v_i$, $i=1,\dots,n$, on the basis vectors.

Hence, in view of (4.6) and (4.7) we have

$$ \begin{equation} \begin{aligned} \, \notag \mathcal{H}^{n-1} (S) &\geqslant 2 \mathcal{H}^{n-1}(V_{r})=2\mathcal{H}^{n-1}(F_{r}) \prod_{i=1}^{n-1}(\lambda_{i}-o(1)) \\ &=2|{\operatorname{adj} Df(a)}|(1-o(1))\cdot\mathcal{H}^{n-1}(F_{r}), \end{aligned} \end{equation} \tag{4.11} $$
so that for the left-hand side of (4.10) we have the lower bound
$$ \begin{equation*} \operatorname{cap}(f(E_r); L^1_q(D)) \geqslant\frac{(2|{\operatorname{adj} Df(a)}|(1-o(1))\cdot\mathcal{H}^{n-1}(F_{r}) )^{q}}{((|{\det Df(a)}|+o(1))\cdot\mathcal{H}^{n}(U_{r}))^{q-1}} \quad\text{as } r\to0. \end{equation*} \notag $$
Substituting the estimates for capacities we have obtained into inequality (4.8) above we see that
$$ \begin{equation*} \frac{(2 |{\operatorname{adj} Df(a)}|\cdot \mathcal{H}^{n-1}(F_{r}))^{q}\cdot (1-o(1))} {((|{\det Df(a)}|+o(1))\cdot\mathcal{H}^{n}(U_{r})))^{q-1}}\leqslant \Psi(U_r)^{q/\sigma} \frac{((\omega(a)+o(1))\cdot\mathcal{H}^{n}(U_{r}))^{q/p} }{(r t )^{q}} \end{equation*} \notag $$
as $r\to0$. Multiplying both sides by $(r t )^{q}$ and dividing by $\mathcal{H}^{n}(U_{r}) =2rt\mathcal{H}^{n-1}(F_{r})(1+t)^{n-1}$ yields
$$ \begin{equation*} \begin{aligned} \, &\frac{(|{\operatorname{adj} Df(a)}|\cdot 2rt \mathcal{H}^{n-1}(F_{r})(1+t)^{n-1} )^{q}\cdot(1-o(1))}{((|{\det Df(a)}|+o(1)) \cdot\mathcal{H}^{n}(U_{r}))^{q}} \, \frac{(|{\det Df(a)}|+o(1))\cdot\mathcal{H}^{n}(U_{r})}{(1+t)^{q(n-1)} \cdot\mathcal{H}^{n}(U_{r})} \\ &\qquad\to\frac{|{\operatorname{adj} Df(a)}|^{q}}{(1+t)^{q(n-1)}|{\det Df(a)}|^{q-1}} \leqslant \Psi'(a)^{q/\sigma} \omega(a)^{q/p} \end{aligned} \end{equation*} \notag $$
as $r \to0$. Since $t > 0$ can be arbitrary, we conclude that
$$ \begin{equation*} \frac{|{\operatorname{adj} Df(a)}|^q}{|{\det Df(a)}|^{q-1}} \leqslant \Psi'(a)^{q/\sigma}\omega(a)^{q/p}, \end{equation*} \notag $$
which yields the required estimate (4.3).

For a proof for $q=p$ one must replace $q$ by $p$ and $\Psi_{p}'(a)^{q/\sigma}$ by $K_p^p$ in the above arguments.

Lemma 1 is proved.

Proof of Theorem 4. Combining (4.3) with the first line of (2.12) we obtain
$$ \begin{equation} \begin{aligned} \, \notag \|\varphi_W^*\|^\sigma &\leqslant\bigl\|K^{1,\omega}_{q,p}(\,\cdot\,)\mid L_\sigma(\varphi^{-1}(W))\bigr\|^\sigma =\int_{W\setminus Z'}\biggl(\frac{|{\operatorname{adj} Df(y)}|}{|{\det Df(y)}|^{(q-1)/q}\omega(y)^{1/p}}\biggr)^\sigma\,dy \\ &\leqslant\int_{W\setminus Z'}\Psi'(y)\,dy\leqslant\int_{W}\Psi'(y)\,dy\leqslant \Psi(W) \end{aligned} \end{equation} \tag{4.12} $$
for $q<p$ and $\|\varphi^*\|\leqslant \|K^{1,\omega}_{p,p}(\,\cdot\,)\mid L_\infty(D)\| \leqslant K_p$ for $q=p$.

Since we can take $\Phi(W)=\|\varphi^*_W\|^\sigma$ as the quasiadditive set function, from (4.12) we infer

$$ \begin{equation*} \|\varphi_W^*\|^\sigma\leqslant\bigl\|K^{1,\omega}_{q,p}(\,\cdot\,)\mid L_\sigma(\varphi^{-1}(W))\bigr\|^\sigma\leqslant \|\varphi_W^*\|^\sigma. \end{equation*} \notag $$
In a similar way, we can take $V(N_c, W)$ as the quasiadditive set function in (2.10). Then taking (3.3) into account, from (4.12) we obtain
$$ \begin{equation*} \|\varphi_W^*\|^\sigma \leqslant\bigl\|K^{1,\omega}_{q,p}(\,\cdot\,)\mid L_\sigma(\varphi^{-1}(W))\bigr\|^\sigma\leqslant V(N_c, W)\leqslant \|\varphi_W^*\|^\sigma \end{equation*} \notag $$
for $q<p$ and $\|\varphi^*\|\leqslant \|K^{1,\omega}_{p,p}(\,\cdot\,)\mid L_\infty(D)\| \leqslant K_p\leqslant\|\varphi^*\|$ for $q=p$.

This completes the proof of equalities (4.1) between the first three quasiadditive functions (of equalities (4.2) between the first three constants, respectively).

To finish the proof it remains to show that the quasiadditive function $V(N_m,W)$ coincides with any of the first three quasiadditive functions in (4.1) (the constant $K_{p,p}$ is equal to the first three quantities in (4.2)).

Note that the proof of Lemma 1 is based on inequalities (4.9) and (4.10) for capacities and on (4.8). In place of (4.8) we can now write an updated inequality (3.10) for the condenser $E_r=(F_r,U_r)$:

$$ \begin{equation} (\operatorname{mod}_q(f\Gamma))^{1/q}\leqslant \begin{cases} K_{p,p}(U_r) (\operatorname{mod}^\omega_p(\Gamma))^{1/p} &\text{for } n-1<q=p<\infty, \\ V(N_m, U_r) K_{q,p}(E_r)(\operatorname{mod}^\omega_p(\Gamma))^{1/p} &\text{for } n-1<q<p<\infty, \end{cases} \end{equation} \tag{4.13} $$
for the family $\Gamma$ of all curves15 in $E_r=(F_r,U_r)$. In this inequality
$$ \begin{equation} K_{p,p}(U_r)=\bigl\|K^{1,\omega}_{q,p}(\,\cdot\,,\varphi)\mid L_\infty(f(U_r))\bigr\| \end{equation} \tag{4.14} $$
and $K^{1,\omega}_{p,p}(\,\cdot\,,\varphi)$ is the distortion function (2.11) (for $n=2$ inequalities (4.13) also hold for $1\leqslant q\leqslant p< \infty$).

By means of the arguments used in [10] to verify (2.25) we can also establish the following relations for the family $\Gamma$ of all curves in $E_r=(F_r,U_r)\subset D'$:

$$ \begin{equation} \begin{aligned} \, \notag &\bigl(\operatorname{cap}(f(F_r),f(U_r);L^1_q(D))\bigr)^{1/q} =(\operatorname{mod}_q(f\Gamma))^{1/q} \\ \notag &\qquad\leqslant\begin{cases} K_{p,p}(D') (\operatorname{mod}^\omega_p(\Gamma))^{1/p} &\text{for } n-1<q=p<\infty, \\ V(N_m, U_r)(\operatorname{mod}^\omega_p(\Gamma))^{1/p} &\text{for } n-1<q<p<\infty, \end{cases} \\ &\qquad\leqslant\begin{cases} K_{p,p}(D') ({\operatorname{cap}}(F_r,U_r); L^1_p(D';\omega))^{1/p}, \\ V(N_m, U_r)({\operatorname{cap}}(F_r,U_r); L^1_p(D';\omega))^{1/p} \end{cases} \end{aligned} \end{equation} \tag{4.15} $$
provided that (2.21) holds for $f$. Hence, applying the upper estimate (4.10) to the right-hand side of (4.15) and the lower estimate (4.9) to the left-hand side we arrive at the same relations as those obtained for (4.8) and underlying the proof of (4.3). Since in (4.15) we take $V(N_m, \cdot)$ as the quasiadditive function, we have
$$ \begin{equation} \frac{|{\operatorname{adj} Df(a)}|}{|{\det Df(a)}|^{(q-1)/q}\omega(a)^{1/p}} \leqslant \begin{cases} V'(N_m,\,\cdot\,)(a)&\text{for } q<p, \\ K_{p,p}(D')&\text{for } q=p, \end{cases} \end{equation} \tag{4.16} $$
where $V'(N_m,\,\cdot\,)(a)$ is the derivative of the quasiadditive set function $V'(N_m,\,\cdot\,)$ at $a$. Hence, just as in (4.12), we obtain
$$ \begin{equation*} \|\varphi_W^*\|^\sigma\leqslant\int_{W\setminus Z'}V'(N_m,\,\cdot\,)(y)\,dy \leqslant\int_{W}V'(N_m,\,\cdot\,)(y)\,dy\leqslant V(N_m,W). \end{equation*} \notag $$

Taking account of inequality 4) at the end of § 3.2 and the equality of the first two quantities in (4.1), we arrive at the required result:

$$ \begin{equation*} \|\varphi_W^*\|^\sigma =V(N_m,W). \end{equation*} \notag $$

Theorem 4 is proved.

From (4.1) and (4.12) we can derive the following unexpected result.

Proposition 5. Under the assumptions of Theorem 4 the set functions in (4.1) can be represented analytically as follows:

$$ \begin{equation} \begin{aligned} \, \|\varphi^*_W\|^\sigma &=\|K^{1,\omega}_{q,p}(\,\cdot\,)\mid L_\sigma(\varphi^{-1}(W))\|^\sigma= V(N_c, W)=V(N_m, W) \nonumber \\ &=\int_{W\setminus Z'}\biggl(\frac{|{\operatorname{adj} Df(y)}|}{|{\det Df(y)}|^{(q-1)/q}\omega(y)^{1/p}}\biggr)^\sigma\,dy, \end{aligned} \end{equation} \tag{4.17} $$
where $W$ is an open subset of $ D'$, $f = \varphi^{-1}\colon D'\to D$ (the analytic properties of $f$ were stated in Theorem 5), and $Z' = \{y \in D'\colon \!\det Df(y) = 0\}$ is the zero set of the Jacobian of $f$.

In addition, the representation (4.17) ensures an extension of set functions to the Borel $\sigma$-algebra in $D'$ such that the extended functions are absolutely continuous.

Remark 5. The interesting part of results in § 4 is related to the case when one considers the quantities (3.1) and (3.2) in the context of a homeomorphism $\varphi\colon {D\to D'}$ inducing a bounded composition operator

$$ \begin{equation*} \varphi^*\colon {L}^1_p(D';\omega) \cap \operatorname{Lip}_{\mathrm{loc}}(D')\to L^1_q(D), \end{equation*} \notag $$
where $n-1< q < p<\infty$ for $n\geqslant3$, or $1\leqslant q <p<\infty$ for $n=2$.

Note that the quantities (3.1) and (3.2) can be defined for any homeomorphism $\varphi\colon D\to D'$, without assuming that $\varphi^*$ is bounded. Of course, such a definition is consistent only when the denominator in (3.1) is distinct from zero and the quantities (3.1) and (3.2) are bounded.

For our arguments in this paper to be valid it is sufficient to assume that condensers of two types have positive capacities in $L^1_p(D';\omega)$, namely, the spherical condensers $E=(B(y_0,r), B(y_0,R))\subset D' $, $0<r<R<\infty$, and the condensers $E_r=(F_r, U_r)\subset D'$ (see (4.6) and (4.7)), so that $\operatorname{cap}(E; L^1_p(D';\omega))>0$ and $\operatorname{cap}(E_r; L^1_p(D';\omega))>0$. This always holds if $\omega\equiv 1$. For an arbitrary weight we indicate sufficient conditions ensuring that the capacities of these condensers are positive for $p>1$. We can obtain such conditions from an estimate for an arbitrary function $u$ that is admissible for the condenser $E=(B(y_0,r), B(y_0,R)) \subset D'$ in $L^1_p(D';\omega)$:

$$ \begin{equation} \begin{aligned} \, \notag 0 &<\operatorname{cap}(E; L^1_1(D')) \\ &\leqslant \int_{B(y_0,R)}|\nabla u(y)|\,dy \leqslant \int_{B(y_0,R)}|\nabla u(y)|\omega^{1/p}(y)\omega^{-1/p}(y)\,dy \notag \\ &\leqslant\biggl( \int_{B(y_0,R)}|\nabla u(y)|^p\omega(y)\,dy\biggr)^{1/p}\biggl( \int_{B(y_0,R)}\omega^{-1/(p-1)}(y)\,dy\biggr)^{(p-1)/p}. \end{aligned} \end{equation} \tag{4.18} $$
The left-hand side of (4.18) is a consequence of estimate (4.10) for $q=1$. Hence, if the integral
$$ \begin{equation} \int_{B(y_0,R)}\omega^{-1/(p-1)}(y)\,dy<\infty \end{equation} \tag{4.19} $$
is finite, then $\operatorname{cap}(E; L^1_p(D';\omega))>0$ for each ball $B(y_0,R)\subset D'$. The last condition always holds for $p>1$, provided that $\omega$ belongs to the Muckenhoupt class $A_p$. Note that condition (4.19) also ensures that the capacity $\operatorname{cap}(E_r; L^1_p(D';\omega))$ is positive.

Based on the above, we make the following conclusion.

Let the weight function $\omega$ satisfy (4.19) and $\varphi\colon D\to D'$ be a homeomorphism. If the quantity (3.1), where the condenser capacity in the denominator is distinct from zero, defines a bounded quasiadditive function (3.2), then condition 2) in Theorem 1 holds for $\varphi\colon D\to D'$, so the conclusions of Theorem 1 hold for this homeomorphism.

We can also make similar conclusion for the quantities (3.13) and (3.14). In fact, if $\rho$ is an admissible metric for the family of curves $\Gamma$ in the condenser ${E=(B(y_0,r), B(y_0,R))\subset D'}$ that have their endpoints at the boundary spheres $S(y_0,r)$ and $S(y_0,R)$, then

$$ \begin{equation} \begin{aligned} \, \notag 0&<\operatorname{cap}(E; L^1_1(D'))=\operatorname{mod}_1(\Gamma) \\ \notag &\leqslant\int_{B(y_0,R)}\rho(y)\,dy\leqslant \int_{B(y_0,R)}\rho(y)\omega^{1/p}(y)\omega^{-1/p}(y)\,dy \\ &\leqslant\biggl( \int_{B(y_0,R)}\rho(y)^p\omega(y)\,dy\biggr)^{1/p}\biggl( \int_{B(y_0,R)}\omega^{-1/(p-1)}(y)\,dy\biggr)^{(p-1)/p}. \end{aligned} \end{equation} \tag{4.20} $$
Hence, if (4.19) holds, then the modulus $\operatorname{mod}^\omega_p(\Gamma)$ is positive. The equality of a capacity and a modulus in the first line of (4.20) was proved in [7].

Acknowledgement

I am grateful to the referee for their useful comments.


Bibliography

1. F. W. Gehring and J. Väisälä, “The coefficients of quasiconformality of domains in space”, Acta Math., 114 (1965), 1–70  crossref  mathscinet  zmath
2. J. Väisälä, Lectures on $n$-dimensional quasiconformal mappings, Lecture Notes in Math., 229, Springer-Verlag, Berlin–New York, 1971, xiv+144 pp.  crossref  mathscinet  zmath
3. G. D. Mostow, “Quasi-conformal mappings in $n$-space and the rigidity of hyperbolic space forms”, Inst. Hautes Études Sci. Publ. Math., 34 (1968), 53–104  crossref  mathscinet  zmath
4. S. K. Vodop'yanov and V. M. Gol'dshtein, “Lattice isomorphisms of the spaces $W_n^1$ and quasiconformal mappings”, Sibirsk. Mat. Zh., 16:2 (1975), 224–246  mathnet  mathscinet  zmath; English transl. in Siberian Math. J., 16:2 (1975), 174–189  crossref
5. V. M. Miklyukov, Conformal mapping of an irregular surface and its applications, Volgograd State University publishing house, Volgograd, 2005, 273 pp. (Russian)
6. O. Martio, V. Ryazanov, U. Srebro and E. Yakubov, Moduli in modern mapping theory, Springer Monogr. Math., Springer, New York, 2009, xii+367 pp.  crossref  mathscinet  zmath
7. J. Hesse, “A $p$-extremal length and $p$-capacity equality”, Ark. Mat., 13:1–2 (1975), 131–144  crossref  mathscinet  zmath
8. V. A. Shlyk, “The equality between $p$-capacity and $p$-modulus”, Sibirsk. Mat. Zh., 34:6 (1993), 216–221  mathnet  mathscinet  zmath; English transl. in Siberian Math. J., 34:6 (1993), 1196–1200  crossref
9. H. Aikawa and M. Ohtsuka, “Extremal length of vector measures”, Ann. Acad. Sci. Fenn. Math., 24:1 (1999), 61–88  mathscinet  zmath
10. S. K. Vodopyanov, “On the equivalence of two approaches to problems of quasiconformal analysis”, Sibirsk. Mat. Zh., 62:6 (2021), 1252–1270  mathnet  crossref  mathscinet  zmath; English transl. in Siberian Math. J., 62:6 (2021), 1010–1025  crossref
11. S. K. Vodopyanov, “Composition operators on weighted Sobolev spaces and the theory of $\mathscr{Q}_p$-homeomorphisms”, Dokl. Ross. Akad. Nauk. Mat. Inform. Protsessy Upr., 494 (2020), 21–25  mathnet  crossref  zmath; English transl. in Dokl. Math., 102:2 (2020), 371–375  crossref  mathscinet
12. S. K. Vodopyanov, “On the analytic and geometric properties of mappings in the theory of $\mathscr Q_{q,p}$-homeomorphisms”, Mat. Zametki, 108:6 (2020), 925–929  mathnet  crossref  mathscinet  zmath; English transl. in Math. Notes, 108:6 (2020), 889–894  crossref
13. S. K. Vodopyanov, “The regularity of inverses to Sobolev mappings and the theory of $\mathscr Q_{q,p}$-homeomorphisms”, Sibirsk. Mat. Zh., 61:6 (2020), 1257–1299  mathnet  crossref  mathscinet  zmath; English transl. in Siberian Math. J., 61:6 (2020), 1002–1038  crossref
14. S. K. Vodopyanov and A. O. Tomilov, “Functional and analytic properties of a class of mappings in quasi-conformal analysis”, Izv. Ross. Akad. Nauk Ser. Mat., 85:5 (2021), 58–109  mathnet  crossref  mathscinet  zmath; English transl. in Izv. Math., 85:5 (2021), 883–931  crossref  adsnasa
15. S. K. Vodopyanov, “Foundations of quasiconformal analysis of a two-index scale of spatial mappings”, Dokl. Ross. Akad. Nauk, 484:2 (2019), 142–146  zmath; English transl. in Dokl. Math., 99:1 (2019), 23–27  crossref  mathscinet
16. S. K. Vodopyanov, “Basics of the quasiconformal analysis of a two-index scale of spatial mappings”, Sibirsk. Mat. Zh., 59:5 (2018), 1020–1056  mathnet  crossref  mathscinet  zmath; English transl. in Siberian Math. J., 59:5 (2018), 805–834  crossref
17. S. K. Vodopyanov, “Differentiability of mappings of the Sobolev space $W^1_{n-1}$ with conditions on the distortion function”, Sibirsk. Mat. Zh., 59:6 (2018), 1240–1267  mathnet  crossref  mathscinet  zmath; English transl. in Siberian Math. J., 59:6 (2018), 983–1005  crossref
18. S. K. Vodop'yanov, “Regularity of mappings inverse to Sobolev mappings”, Mat. Sb., 203:10 (2012), 3–32  mathnet  crossref  mathscinet  zmath; English transl. in Sb. Math., 203:10 (2012), 1383–1410  crossref
19. S. K. Vodop'yanov and A. D. Ukhlov, “Superposition operators in Sobolev spaces”, Izv. Vyssh. Uchebn. Zaved. Mat., 2002, no. 10, 11–33  mathnet  mathscinet  zmath; English transl. in Russian Math. (Iz. VUZ), 46:10 (2002), 9–31
20. S. K. Vodopyanov, “Admissible changes of variables for Sobolev functions on (sub-)Riemannian manifolds”, Mat. Sb., 210:1 (2019), 63–112  mathnet  crossref  mathscinet  zmath; English transl. in Sb. Math., 210:1 (2019), 59–104  crossref  adsnasa
21. S. K. Vodopyanov, “Isomorphisms of Sobolev spaces on Riemannian manifolds and quasiconformal mappings”, Sibirsk. Mat. Zh., 60:5 (2019), 996–1034  mathnet  crossref  mathscinet  zmath; English transl. in Siberian Math. J., 60:5 (2019), 774–804  crossref
22. S. L. Sobolev, “Some transformation groups of $n$-dimensional space”, Dokl. Akad. Nauk SSSR, 32:6 (1941), 380–382 (Russian); French transl., S. Sobolev, “Sur quelques groupes de transformations de l'espace $n$-dimensionnel”, C. R. Acad. Sci. URSS, 32 (1941), 380–382  mathscinet  zmath
23. V. G. Maz'ya, Set classes and embedding theorems for function spaces. Some questions in the theory of elliptic equations, Kandidat Dissertation, Moscow State University, Moscow, 1962 (Russian)
24. Yu. G. Reshetnyak, Space mappings with bounded distortion, Nauka, Novosibirsk, 1982, 286 pp.  mathscinet  zmath; English transl., Transl. Math. Monogr., 73, Amer. Math. Soc., Providence, RI, 1989, xvi+362 pp.  crossref  mathscinet  zmath
25. H. M. Reimann, “Über harmonische Kapazität und quasikonforme Abbildungen im Raum”, Comment. Math. Helv., 44 (1969), 284–307  mathscinet  zmath
26. F. W. Gehring, “Lipschitz mappings and $p$-capacity of rings in $n$-space”, Advances in the theory of Riemann surfaces (Stony Brook, NY 1969), Ann. of Math. Studies, 66, Princeton Univ. Press, Princeton, NJ, 1971, 175–193  crossref  mathscinet  zmath
27. J. Lelong-Ferrand, “Étude d'une classe d'applications liées à des homomorphismes d'algébres de fonctions et généralisant les quasi-conformes”, Duke Math. J., 40 (1973), 163–186  crossref  mathscinet  zmath
28. A. Molchanova and S. Vodopyanov, “Injectivity almost everywhere and mappings with finite distortion in nonlinear elasticity”, Calc. Var. Partial Differential Equations, 59:1 (2020), 17, 25 pp.  crossref  mathscinet  zmath
29. J. Heinonen, T. Kilpeläinen and O. Martio, Nonlinear potential theory of degenerate elliptic equations, Oxford Math. Monogr., The Clarendon Press, Oxford Univ. Press, New York, 1993, vi+363 pp.  mathscinet  zmath
30. T. Rado and P. V. Reichelderfer, Continuous transformations in analysis. With an introduction to algebraic topology, Grundlehren Math. Wiss., LXXV, Springer-Verlag, Berlin–Göttingen–Heidelberg, 1955, vii+442 pp.  crossref  mathscinet  zmath
31. M. de Guzmán, Differentiation of integrals in $\mathbb{R}^n$, Lecture Notes in Math., 481, Springer-Verlag, Berlin–New York, 1975, xii+266 pp.  crossref  mathscinet  zmath
32. S. K. Vodop'yanov and A. D. Ukhlov, “Set functions and their applications in the theory of Lebesgue and Sobolev spaces. I”, Mat. Tr., 6:2 (2003), 14–65  mathnet  mathscinet  mathscinet  zmath; English transl. in Siberian Adv. Math., 14:4 (2004), 78–125
33. H. Federer, Geometric measure theory, Grundlehren Math. Wiss., 153, Springer-Verlag, New York, 1969, xiv+676 pp.  mathscinet  zmath
34. P. Hajłasz, “Change of variables formula under minimal assumptions”, Colloq. Math., 64:1 (1993), 93–101  crossref  mathscinet  zmath
35. S. K. Vodop'yanov, Taylor's formula and function spaces, Novosibirsk State University, Novosibirsk, 1988, 96 pp. (Russian)  mathscinet  zmath
36. A. D. Ukhlov, “On mappings generating the embeddings of Sobolev spaces”, Sibirsk. Mat. Zh., 34:1 (1993), 185–192  mathnet  mathscinet  zmath; English transl. in Siberian Math. J., 34:1 (1993), 165–171  crossref
37. S. K. Vodop'yanov and A. D. Ukhlov, “Sobolev spaces and $(P,Q)$-quasiconformal mappings of Carnot groups”, Sibirsk. Mat. Zh., 39:4 (1998), 776–795  mathnet  mathscinet  zmath; English transl. in Siberian Math. J., 39:4 (1998), 665–682  crossref
38. S. K. Vodop'yanov, “Monotone functions and quasiconformal mappings on Carnot groups”, Sibirsk. Mat. Zh., 37:6 (1996), 1269–1295  mathnet  mathscinet  zmath; English transl. in Siberian Math. J., 37:6 (1996), 1113–1136  crossref
39. R. R. Salimov and E. A. Sevost'yanov, “$ACL$ and differentiability of open discrete ring $(p, Q)$-mappings”, Mat. Stud., 35:1 (2011), 28–36  mathscinet  zmath
40. V. I. Kruglikov, “Capacity of condensers and spatial mappings quasiconformal in the mean”, Mat. Sb., 130(172):2(6) (1986), 185–206  mathnet  mathscinet  zmath; English transl. in Sb. Math., 58:1 (1987), 185–205  crossref
41. V. I. Ryazanov and E. A. Sevost'yanov, “Equicontinuity of mean quasiconformal mappings”, Sibirsk. Mat. Zh., 52:3 (2011), 665–679  mathnet  mathscinet  zmath; English transl. in Siberian Math. J., 52:3 (2011), 524–536  crossref
42. R. R. Salimov, “ACL and differentiability of a generalization of quasi-conformal maps”, Izv. Ross. Akad. Nauk Ser. Mat., 72:5 (2008), 141–148  mathnet  crossref  mathscinet  zmath; English transl. in Izv. Math., 72:5 (2008), 977–984  crossref  adsnasa
43. R. Salimov, “$ACL$ and differentiability of $Q$-homeomorphisms”, Ann. Acad. Sci. Fenn. Math., 33:1 (2008), 295–301  mathscinet  zmath
44. R. R. Salimov and E. A. Sevost'yanov, “The theory of shell-based $Q$-mappings in geometric function theory”, Mat. Sb., 201:6 (2010), 131–158  mathnet  crossref  mathscinet; English transl. in Sb. Math., 201:6 (2010), 909–934  crossref  zmath  adsnasa
45. E. A. Sevost'yanov and S. A. Skvortsov, On behavior of homeomorphisms with inverse modulus conditions, arXiv: 1801.01808v9
46. R. R. Salimov and E. A. Sevost'yanov, “On local properties of spatial generalized quasi-isometries”, Mat. Zametki, 101:4 (2017), 594–610  mathnet  crossref  mathscinet  zmath; English transl. in Math. Notes, 101:4 (2017), 704–717  crossref
47. R. Salimov, “On $Q$-homeomorphisms with respect to $p$-modulus”, Ann. Univ. Buchar. Math. Ser., 2(LX):2 (2011), 207–213  mathscinet  zmath
48. V. G. Maz'ja, Sobolev spaces, Leningrad University Publishing House, Leningrad, 1985, 416 pp.  mathscinet  zmath; English transl., Springer Ser. Soviet Math., Springer-Verlag, Berlin, 1985, xix+486 pp.  crossref  mathscinet  zmath

Citation: S. K. Vodopyanov, “Coincidence of set functions in quasiconformal analysis”, Mat. Sb., 213:9 (2022), 3–33; Sb. Math., 213:9 (2022), 1157–1186
Citation in format AMSBIB
\Bibitem{Vod22}
\by S.~K.~Vodopyanov
\paper Coincidence of set functions in quasiconformal analysis
\jour Mat. Sb.
\yr 2022
\vol 213
\issue 9
\pages 3--33
\mathnet{http://mi.mathnet.ru/sm9702}
\crossref{https://doi.org/10.4213/sm9702}
\mathscinet{http://mathscinet.ams.org/mathscinet-getitem?mr=4563374}
\adsnasa{https://adsabs.harvard.edu/cgi-bin/bib_query?2022SbMat.213.1157V}
\transl
\jour Sb. Math.
\yr 2022
\vol 213
\issue 9
\pages 1157--1186
\crossref{https://doi.org/10.4213/sm9702e}
\isi{https://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcApp=Publons&SrcAuth=Publons_CEL&DestLinkType=FullRecord&DestApp=WOS_CPL&KeyUT=000992271700001}
\scopus{https://www.scopus.com/record/display.url?origin=inward&eid=2-s2.0-85165898294}
Linking options:
  • https://www.mathnet.ru/eng/sm9702
  • https://doi.org/10.4213/sm9702e
  • https://www.mathnet.ru/eng/sm/v213/i9/p3
  • This publication is cited in the following 1 articles:
    Citing articles in Google Scholar: Russian citations, English citations
    Related articles in Google Scholar: Russian articles, English articles
    Математический сборник Sbornik: Mathematics
    Statistics & downloads:
    Abstract page:374
    Russian version PDF:37
    English version PDF:58
    Russian version HTML:203
    English version HTML:66
    References:48
    First page:20
     
      Contact us:
     Terms of Use  Registration to the website  Logotypes © Steklov Mathematical Institute RAS, 2024