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

AN ANALOGUE OF KOEBE’S THEOREM AND THE OPENNESS OF A LIMIT MAP IN ONE CLASS

EVGENY SEVOST’YANOV, VALERY TARGONSKII
Abstract

We study mappings that satisfy the inverse modulus inequality of Poletsky type in a fixed domain. It is shown that, under some additional restrictions, the image of a ball under such mappings contains a fixed ball uniformly over the class. This statement can be interpreted as the well-known analogue of Koebe’s theorem for analytic functions. As an application of the obtained result, we show that, if a sequence of mappings belonging to the specified class converges locally uniformly, then the limit mapping is open.


2010 Mathematics Subject Classification: Primary 30C65; Secondary 31A15, 31B25

Key words: mappings with a finite and bounded distortion, moduli, capacity

1 Introduction

Let us recall the formulation of the classical Koebe theorem, see, for example, [CG, Theorem 1.3].

Theorem A. Let f:𝔻f:{\mathbb{D}}\rightarrow{\mathbb{C}} be an univalent analytic function such that f(0)=0f(0)=0 and f(0)=1.f^{\,\prime}(0)=1. Then the image of ff covers the open disk centered at 0 of radius one-quarter, that is, f(𝔻)B(0,1/4).f({\mathbb{D}})\supset B(0,1/4).

The main fact contained in the paper is the statement that something similar has been done for a much more general class of spatial mappings. Below dm(x)dm(x) denotes the element of the Lebesgue measure in n.{\mathbb{R}}^{n}. Everywhere further the boundary A\partial A of the set AA and the closure A¯\overline{A} should be understood in the sense of the extended Euclidean space n¯.\overline{{\mathbb{R}}^{n}}. Recall that, a Borel function ρ:n[0,]\rho:{\mathbb{R}}^{n}\,\rightarrow[0,\infty] is called admissible for the family Γ\Gamma of paths γ\gamma in n,{\mathbb{R}}^{n}, if the relation

γρ(x)|dx|1\int\limits_{\gamma}\rho(x)\,|dx|\geqslant 1 (1.1)

holds for all (locally rectifiable) paths γΓ.\gamma\in\Gamma. In this case, we write: ρadmΓ.\rho\in{\rm adm}\,\Gamma. The modulus of Γ\Gamma is defined by the equality

M(Γ)=infρadmΓnρn(x)𝑑m(x).M(\Gamma)=\inf\limits_{\rho\in\,{\rm adm}\,\Gamma}\int\limits_{{\mathbb{R}}^{n}}\rho^{n}(x)\,dm(x)\,. (1.2)

Let y0n,y_{0}\in{\mathbb{R}}^{n}, 0<r1<r2<0<r_{1}<r_{2}<\infty and

A=A(y0,r1,r2)={yn:r1<|yy0|<r2}.A=A(y_{0},r_{1},r_{2})=\left\{y\,\in\,{\mathbb{R}}^{n}:r_{1}<|y-y_{0}|<r_{2}\right\}\,. (1.3)

Given x0n,x_{0}\in{\mathbb{R}}^{n}, we put

B(x0,r)={xn:|xx0|<r},𝔹n=B(0,1),B(x_{0},r)=\{x\in{\mathbb{R}}^{n}:|x-x_{0}|<r\}\,,\quad{\mathbb{B}}^{n}=B(0,1)\,,
S(x0,r)={xn:|xx0|=r}.S(x_{0},r)=\{x\,\in\,{\mathbb{R}}^{n}:|x-x_{0}|=r\}\,.

A mapping f:Dnf:D\rightarrow{\mathbb{R}}^{n} is called discrete if the pre-image {f1(y)}\{f^{-1}\left(y\right)\} of any point yny\,\in\,{\mathbb{R}}^{n} consists of isolated points, and open if the image of any open set UDU\subset D is an open set in n.{\mathbb{R}}^{n}.

Given sets E,E, Fn¯F\subset\overline{{\mathbb{R}}^{n}} and a domain DnD\subset{\mathbb{R}}^{n} we denote by Γ(E,F,D)\Gamma(E,F,D) the family of all paths γ:[a,b]n¯\gamma:[a,b]\rightarrow\overline{{\mathbb{R}}^{n}} such that γ(a)E,γ(b)F\gamma(a)\in E,\gamma(b)\in\,F and γ(t)D\gamma(t)\in D for t(a,b).t\in(a,b). Given a mapping f:Dn,f:D\rightarrow{\mathbb{R}}^{n}, a point y0f(D)¯{},y_{0}\in\overline{f(D)}\setminus\{\infty\}, and 0<r1<r2<r0=supyf(D)|yy0|,0<r_{1}<r_{2}<r_{0}=\sup\limits_{y\in f(D)}|y-y_{0}|, we denote by Γf(y0,r1,r2)\Gamma_{f}(y_{0},r_{1},r_{2}) a family of all paths γ\gamma in DD such that f(γ)Γ(S(y0,r1),S(y0,r2),A(y0,r1,r2)).f(\gamma)\in\Gamma(S(y_{0},r_{1}),S(y_{0},r_{2}),A(y_{0},r_{1},r_{2})). Let Q:n[0,]Q:{\mathbb{R}}^{n}\rightarrow[0,\infty] be a Lebesgue measurable function. We say that ff satisfies the inverse Poletsky inequality at a point y0f(D)¯{}y_{0}\in\overline{f(D)}\setminus\{\infty\} if the relation

M(Γf(y0,r1,r2))A(y0,r1,r2)f(D)Q(y)ηn(|yy0|)𝑑m(y)M(\Gamma_{f}(y_{0},r_{1},r_{2}))\leqslant\int\limits_{A(y_{0},r_{1},r_{2})\cap f(D)}Q(y)\cdot\eta^{n}(|y-y_{0}|)\,dm(y) (1.4)

holds for any Lebesgue measurable function η:(r1,r2)[0,]\eta:(r_{1},r_{2})\rightarrow[0,\infty] such that

r1r2η(r)𝑑r1.\int\limits_{r_{1}}^{r_{2}}\eta(r)\,dr\geqslant 1\,. (1.5)

The definition of the relation (1.4) at the point y0=y_{0}=\infty may be given by the using of the inversion ψ(y)=y|y|2\psi(y)=\frac{y}{|y|^{2}} at the origin.

Note that conformal mappings preserve the modulus of families of paths, so that we may write

M(Γ)=M(f(Γ)).M(\Gamma)=M(f(\Gamma))\,.

It is not difficult to see from this that conformal mappings from Koebe theorem satisfy the relation (1.4) with Q1Q\equiv 1 for any function η\eta in (1.5).

Remark 1.1. It is known that the quasiregular mappings satisfy the inequality

M(Γ)N(f,D)KO(f)M(f(Γ)),M(\Gamma)\leqslant N(f,D)K_{O}(f)M(f(\Gamma))\,,

where 1KO(f)<1\leqslant K_{O}(f)<\infty is some number, and N(f,D)N(f,D) denotes the multiplicity function,

N(y,f,E)=card{xE:f(x)=y},N(y,f,E)\,=\,{\rm card}\,\left\{x\in E:f(x)=y\right\}\,,
N(f,E)=supynN(y,f,E),N(f,E)\,=\,\sup\limits_{y\in{\mathbb{R}}^{n}}\,N(y,f,E)\,, (1.6)

see [MRV1, Theorem 3.2]. There are also mappings in which the distortion of the modulus of families of paths is much more complex. Say, for homeomorphisms fWloc1,nf\in W^{1,n}_{\rm loc} such that f1Wloc1,nf^{\,-1}\in W^{1,n}_{\rm loc} we have the inequality

M(Γ)f(D)KI(y,f1)ρn(y)𝑑m(x)M(\Gamma)\leqslant\int\limits_{f(D)}K_{I}(y,f^{\,-1})\cdot\rho_{*}^{n}(y)\,dm(x) (1.7)

for any ρadmf(Γ)\rho_{*}\in{\rm adm}\,f(\Gamma) (see below), where

KI(y,f1)=xf1(y)KO(x,f),K_{I}(y,f^{\,-1})\quad=\sum\limits_{x\in f^{\,-1}(y)}K_{O}(x,f)\,, (1.8)
KO(x,f)={f(x)n|J(x,f)|,J(x,f)0,1,f(x)=0,,otherwise,K_{O}(x,f)\quad=\quad\left\{\begin{array}[]{rr}\frac{\|f^{\,\prime}(x)\|^{n}}{|J(x,f)|},&J(x,f)\neq 0,\\ 1,&f^{\,\prime}(x)=0,\\ \infty,&\text{otherwise}\end{array}\right.\,\,,

see [MRSY, Theorems 8.1, 8.6].

All of the above allows us to assert that relation (1.4) is satisfied by a fairly large number of mappings. In general, for practically all currently known classes, including conformal and quasiconformal mappings, quasiregular mappings, mappings with finite distortion, etc. such inequalities are satisfied.

Set

qy0(r)=1ωn1rn1S(y0,r)Q(y)𝑑n1(y),q_{y_{0}}(r)=\frac{1}{\omega_{n-1}r^{n-1}}\int\limits_{S(y_{0},r)}Q(y)\,d\mathcal{H}^{n-1}(y)\,, (1.9)

and ωn1\omega_{n-1} denotes the area of the unit sphere 𝕊n1{\mathbb{S}}^{n-1} in n.{\mathbb{R}}^{n}.

We say that a function φ:D{\varphi}:D\rightarrow{\mathbb{R}} has a finite mean oscillation at a point x0D,x_{0}\in D, write φFMO(x0),\varphi\in FMO(x_{0}), if

lim supε01ΩnεnB(x0,ε)|φ(x)φ¯ε|𝑑m(x)<,\limsup\limits_{\varepsilon\rightarrow 0}\frac{1}{\Omega_{n}\varepsilon^{n}}\int\limits_{B(x_{0},\,\varepsilon)}|{\varphi}(x)-\overline{{\varphi}}_{\varepsilon}|\ dm(x)<\infty\,,

where φ¯ε=1ΩnεnB(x0,ε)φ(x)𝑑m(x)\overline{{\varphi}}_{\varepsilon}=\frac{1}{\Omega_{n}\varepsilon^{n}}\int\limits_{B(x_{0},\,\varepsilon)}{\varphi}(x)\,dm(x) and Ωn\Omega_{n} is the volume of the unit ball 𝔹n{\mathbb{B}}^{n} in n.{\mathbb{R}}^{n}. We also say that a function φ:D{\varphi}:D\rightarrow{\mathbb{R}} has a finite mean oscillation at AD¯,A\subset\overline{D}, write φFMO(A),{\varphi}\in FMO(A), if φ{\varphi} has a finite mean oscillation at any point x0A.x_{0}\in A. Let hh be a chordal metric in n¯,\overline{{\mathbb{R}}^{n}},

h(x,)=11+|x|2,h(x,\infty)=\frac{1}{\sqrt{1+{|x|}^{2}}}\,,
h(x,y)=|xy|1+|x|21+|y|2xy.h(x,y)=\frac{|x-y|}{\sqrt{1+{|x|}^{2}}\sqrt{1+{|y|}^{2}}}\qquad x\neq\infty\neq y\,. (1.10)

and let h(E):=supx,yEh(x,y)h(E):=\sup\limits_{x,y\in E}\,h(x,y) be a chordal diameter of a set En¯E\subset\overline{{\mathbb{R}}^{n}} (see, e.g., [Va, Definition 12.1]).

Given a continuum ED,E\subset D, δ>0\delta>0 and a Lebesgue measurable function Q:n[0,]Q:{\mathbb{R}}^{n}\rightarrow[0,\infty] we denote by 𝔉E,δ(D)\mathfrak{F}_{E,\delta}(D) the family of all mapping f:Dn,f:D\rightarrow{\mathbb{R}}^{n}, n2,n\geqslant 2, satisfying relations (1.4)–(1.5) at any point y0n¯y_{0}\in\overline{{\mathbb{R}}^{n}} such that h(f(E))δ.h(f(E))\geqslant\delta. The following statement holds.

Theorem 1.1. Let DD be a domain in n,{\mathbb{R}}^{n}, n2,n\geqslant 2, and let B(x0,ε1)DB(x_{0},\varepsilon_{1})\subset D for some ε1>0.\varepsilon_{1}>0.

Assume that, QL1(n)Q\in L^{1}({\mathbb{R}}^{n}) and, in addition, one of the following conditions hold:

1) QFMO(n¯);Q\in FMO(\overline{{\mathbb{R}}^{n}});

2) for any y0n¯y_{0}\in\overline{{\mathbb{R}}^{n}} there is δ(y0)>0\delta(y_{0})>0 such that

0δ(y0)dttqy01n1(t)=.\int\limits_{0}^{\delta(y_{0})}\frac{dt}{tq_{y_{0}}^{\frac{1}{n-1}}(t)}=\infty\,. (1.11)

Then there is r0>0,r_{0}>0, which does not depend on f,f, such that

f(B(x0,ε1))B(f(x0),r0)f𝔉E,δ(D).f(B(x_{0},\varepsilon_{1}))\supset B(f(x_{0}),r_{0})\qquad\forall\,\,f\in\mathfrak{F}_{E,\delta}(D)\,.

Remark 1.2. The condition QFMO()Q\in FMO(\infty) of the condition (1.11) for y0=y_{0}=\infty must be understood as follows: these conditions hold for y0=y_{0}=\infty if and only if the function Q~:=Q(y|y|2)\widetilde{Q}:=Q\left(\frac{y}{|y|^{2}}\right) satisfies similar conditions at the origin.

Note that the above analogue of Koebe’s theorem has an important application in the field of convergence of mappings. Recall that, a mapping f:Dnf:D\rightarrow{\mathbb{R}}^{n} is called a KK-quasiregular mapping, if the following conditions hold:

1) fWloc1,n(D),f\in W_{loc}^{1,n}(D),

2) the Jacobian J(x,f)J(x,f) of ff at xDx\in D preserves the sign almost everywhere in D,D,

3) f(x)nK|J(x,f)|\|f^{\,\prime}(x)\|^{n}\leqslant K\cdot|J(x,f)| for almost any xDx\in D and some constant K<,K<\infty, where f(x)=maxhn\{0}|f(x)h||h|,J(x,f)=detf(x),\|f^{\,\prime}(x)\|\,=\,\max\limits_{h\in{\mathbb{R}}^{n}\backslash\{0\}}\frac{|f^{\,\prime}(x)h|}{|h|}\,,\quad J(x,f)=\det f^{\,\prime}(x), see e.g. [Re, Section 4, Ch. I], cf. [Ri, Definition 2.1, Ch. I]. As is known, the class of mappings with bounded distortion is closed under locally uniform convergence. In particular, the following statement is true (see, for example, [Re, Theorem 9.2.II]).

Theorem B. Let fj:Dn,f_{j}:D\rightarrow{\mathbb{R}}^{n}, n2,n\geqslant 2, j=1,2,,j=1,2,\ldots, be a sequence of KK-quasiregular mappings converging to some mapping f:Dnf:D\rightarrow{\mathbb{R}}^{n} as jj\rightarrow\infty locally uniformly in D.D. Then either ff is KK-quasiregular, of ff is a constant. In particular, in the first case ff is discrete and open (see [Re, Theorems 6.3.II and 6.4.II]).

As for the classes we are studying in (1.4)–(1.5), the following analogue of Theorem B is valid for them.

Theorem 1.2.   Let DD be a domain in n,{\mathbb{R}}^{n}, n2.n\geqslant 2. Let fj:Dn,f_{j}:D\rightarrow{\mathbb{R}}^{n}, n2,n\geqslant 2, j=1,2,,j=1,2,\ldots, be a sequence of open discrete mappings satisfying the conditions (1.4)–(1.5) at any point y0n¯y_{0}\in\overline{{\mathbb{R}}^{n}} and converging to some mapping f:Dnf:D\rightarrow{\mathbb{R}}^{n} as jj\rightarrow\infty locally uniformly in D.D. Assume that the conditions on the function QQ from Theorem 1 hold. Then either ff is a constant, or ff is light and open.

Remark 1.3. The lightness of the mapping ff in Theorem 1 was established earlier, see [Sev1], cf. [Cr]. The goal of the paper is to obtain the openness of this mapping, which will follow from Theorem 1. Note that mappings that satisfy conditions (1.4)–(1.5) may not be open. For example, let x=(x1,,xn).x=(x_{1},\ldots,x_{n}). We define ff as the identical mapping in the closed domain {xn0}\{x_{n}\geqslant 0\} and set f(x)=(x1,,xn)f(x)=(x_{1},\ldots,-x_{n}) for xn<0.x_{n}<0. Observe that, the mapping ff satisfies conditions (1.4)–(1.5) for Q(y)2.Q(y)\equiv 2. Indeed, ff preserves the lengths of paths, is differentiable almost everywhere and has Luzin’s NN and N1N^{\,-1}-properties. Therefore, ff is a mapping with a finite length distortion (for the definition see [MRSY, section 8]). Now, ff satisfies (1.7) with Q:=KI(y,f1)=xf1(y)KO(x,f)1+1=2Q:=K_{I}(y,f^{\,-1})\quad=\sum\limits_{x\in f^{\,-1}(y)}K_{O}(x,f)\leqslant 1+1=2 by [MRSY, Theorems 8.1, 8.6]. Therefore ff satisfies conditions (1.4)–(1.5) for Q(y)2,Q(y)\equiv 2, as well.

As for the discreteness of the limit mapping ff in Theorem 1, whether this mapping will be such is currently unknown.

2 Preliminaries

The following statement was proved in [Na, Lemma 2.1].

Proposition 2.1.   The number δn(r)=infM(Γ(F,F,n¯)),\delta_{n}(r)=\inf M(\Gamma(F,F_{*},\overline{{\mathbb{R}}^{n}})), where the infimum is taken over all continua FF and FF_{*} in n¯\overline{{\mathbb{R}}^{n}} with h(F)rh(F)\geqslant r and h(F)r,h(F_{*})\geqslant r, is positive for each r>0r>0 and zero for r=0.r=0.

The following statement also may be found in [Na, Theorem 4.1]).

Proposition 2.2.   Let 𝔉\mathfrak{F} be a collection of connected sets in a domain DD and let infh(F)>0,\inf h(F)>0, F𝔉.F\in\mathfrak{F}. Then infF𝔉M(Γ(F,A,D))>0\inf\limits_{F\in\mathfrak{F}}M(\Gamma(F,A,D))>0 either for each or for no continuum AA in D.D.

The following statement may be found in [Vu, Lemma 4.3].

Proposition 2.3.   Let DD be an open half space or an open ball in n{\mathbb{R}}^{n} and let EE and FF be subsets of D.D. Then

M(Γ(E,F,D))12M(Γ(E,F,n¯)).M(\Gamma(E,F,D))\geqslant\frac{1}{2}\cdot M(\Gamma(E,F,\overline{{\mathbb{R}}^{n}}))\,.

For a domain Dn,D\subset{\mathbb{R}}^{n}, n2,n\geqslant 2, and a Lebesgue measurable function Q:n[0,],Q:{\mathbb{R}}^{n}\rightarrow[0,\infty], Q(y)0Q(y)\equiv 0 for ynf(D),y\in{\mathbb{R}}^{n}\setminus f(D), we denote by 𝔉Q(D)\mathfrak{F}_{Q}(D) the family of all open discrete mappings f:Dnf:D\rightarrow{\mathbb{R}}^{n} such that relations (1.4)–(1.5) hold for each point y0f(D).y_{0}\in f(D). The following result holds (see [SSD, Theorem 1.1]).

Proposition 2.4. Let n2,n\geqslant 2, and let QL1(n).Q\in L^{1}({\mathbb{R}}^{n}). Then for any x0Dx_{0}\in D and any r0>0r_{0}>0 such that 0<r0<dist(x0,D)0<r_{0}<{\rm dist}(x_{0},\partial D) the inequality

|f(x)f(x0)|Cn(Q1)1/nlog1/n(1+r02|xx0|)|f(x)-f(x_{0})|\leqslant\frac{C_{n}\cdot(\|Q\|_{1})^{1/n}}{\log^{1/n}\left(1+\frac{r_{0}}{2|x-x_{0}|}\right)} (2.1)

holds for any x,yB(x0,r0)x,y\in B(x_{0},r_{0}) and f𝔉Q(D),f\in\mathfrak{F}_{Q}(D), where Q1\|Q\|_{1} denotes the L1L^{1}-norm of QQ in n,{\mathbb{R}}^{n}, and Cn>0C_{n}>0 is some constant depending only on n.n. In particular, 𝔉Q(D)\mathfrak{F}_{Q}(D) is equicontinuous in D.D.

Let Dn,D\subset{\mathbb{R}}^{n}, f:Dnf:D\rightarrow{\mathbb{R}}^{n} be a discrete open mapping, β:[a,b)n\beta:[a,\,b)\rightarrow{\mathbb{R}}^{n} be a path, and xf1(β(a)).x\in\,f^{\,-1}(\beta(a)). A path α:[a,c)D\alpha:[a,\,c)\rightarrow D is called a maximal ff-lifting of β\beta starting at x,x, if (1)α(a)=x;(1)\quad\alpha(a)=x\,; (2)fα=β|[a,c);(2)\quad f\circ\alpha=\beta|_{[a,\,c)}; (3)(3) for c<cb,c<c^{\prime}\leqslant b, there is no a path α:[a,c)D\alpha^{\prime}:[a,\,c^{\prime})\rightarrow D such that α=α|[a,c)\alpha=\alpha^{\prime}|_{[a,\,c)} and fα=β|[a,c).f\circ\alpha^{\,\prime}=\beta|_{[a,\,c^{\prime})}. If β:[a,b)n¯\beta:[a,b)\rightarrow\overline{{\mathbb{R}}^{n}} is a path and if Cn¯,C\subset\overline{{\mathbb{R}}^{n}}, we say that βC\beta\rightarrow C as tb,t\rightarrow b, if the spherical distance h(β(t),C)0h(\beta(t),C)\rightarrow 0 as tbt\rightarrow b (see [MRV2, section 3.11]), where h(β(t),C)=infxCh(β(t),x).h(\beta(t),C)=\inf\limits_{x\in C}h(\beta(t),x). The following assertion holds (see [MRV2, Lemma 3.12]).

Proposition 2.5. Let f:Dn,f:D\rightarrow{\mathbb{R}}^{n}, n2,n\geqslant 2, be an open discrete mapping, let x0D,x_{0}\in D, and let β:[a,b)n\beta:[a,\,b)\rightarrow{\mathbb{R}}^{n} be a path such that β(a)=f(x0)\beta(a)=f(x_{0}) and such that either limtbβ(t)\lim\limits_{t\rightarrow b}\beta(t) exists, or β(t)f(D)\beta(t)\rightarrow\partial f(D) as tb.t\rightarrow b. Then β\beta has a maximal ff-lifting α:[a,c)D\alpha:[a,\,c)\rightarrow D starting at x0.x_{0}. If α(t)x1D\alpha(t)\rightarrow x_{1}\in D as tc,t\rightarrow c, then c=bc=b and f(x1)=limtbβ(t).f(x_{1})=\lim\limits_{t\rightarrow b}\beta(t). Otherwise α(t)D\alpha(t)\rightarrow\partial D as tc.t\rightarrow c.

The following statement may be found in [Sev2, Lemma 1.3].

Proposition 2.6.   Let Q:n[0,],Q:{\mathbb{R}}^{n}\rightarrow[0,\infty], n2,n\geqslant 2, be a Lebesgue measurable function and let x0n.x_{0}\in{\mathbb{R}}^{n}. Assume that either of the following conditions holds

(a) QFMO(x0),Q\in FMO(x_{0}),

(b) qx0(r)=O([log1r]n1)q_{x_{0}}(r)\,=\,O\left(\left[\log{\frac{1}{r}}\right]^{n-1}\right) as r0,r\rightarrow 0,

(c) for some small δ0=δ0(x0)>0\delta_{0}=\delta_{0}(x_{0})>0 we have the relations

δδ0dttqx01n1(t)<,0<δ<δ0,\int\limits_{\delta}^{\delta_{0}}\frac{dt}{tq_{x_{0}}^{\frac{1}{n-1}}(t)}<\infty,\qquad 0<\delta<\delta_{0}, (2.2)

and

0δ0dttqx01n1(t)=.\int\limits_{0}^{\delta_{0}}\frac{dt}{tq_{x_{0}}^{\frac{1}{n-1}}(t)}=\infty\,. (2.3)

Then there exist a number ε0(0,1)\varepsilon_{0}\in(0,1) and a function ψ(t)0\psi(t)\geqslant 0 such that the relation

ε<|xb|<ε0Q(x)ψn(|xb|)𝑑m(x)=o(In(ε,ε0)),\int\limits_{\varepsilon<|x-b|<\varepsilon_{0}}Q(x)\cdot\psi^{n}(|x-b|)\ dm(x)=o(I^{n}(\varepsilon,\varepsilon_{0}))\,, (2.4)

holds as ε0,\varepsilon\rightarrow 0, where ψ:(0,ε0)[0,)\psi:(0,\varepsilon_{0})\rightarrow[0,\infty) is some function such that, for some 0<ε1<ε0,0<\varepsilon_{1}<\varepsilon_{0},

0<I(ε,ε0)=εε0ψ(t)𝑑t<ε(0,ε1).0<I(\varepsilon,\varepsilon_{0})=\int\limits_{\varepsilon}^{\varepsilon_{0}}\psi(t)\,dt<\infty\qquad\forall\quad\varepsilon\in(0,\varepsilon_{1})\,. (2.5)

3 Main Lemmas

Lemma 3.1.   Let DD be a domain in n,{\mathbb{R}}^{n}, let x0D,x_{0}\in D, let AA be a (non-degenerate) continuum in D,D, and let ε1>0\varepsilon_{1}>0 be such that B(x0,ε1)D.B(x_{0},\varepsilon_{1})\subset D. Let r>0r>0 and let Cj,C_{j}, j=1,2,,j=1,2,\ldots, be a sequence of continua in B(x0,ε1)B(x_{0},\varepsilon_{1}) such that h(Cj)r,h(C_{j})\geqslant r, h(Cj)=supx,yCjh(x,y).h(C_{j})=\sup\limits_{x,y\in C_{j}}h(x,y). Then there is R0>0R_{0}>0 such that

M(Γ(Cj,A,D))R0j.M(\Gamma(C_{j},A,D))\geqslant R_{0}\qquad\forall\,\,j\in{\mathbb{N}}\,.

Proof.   Let A1A_{1} be an arbitrary (non-degenerate) continuum in B(x0,ε1).B(x_{0},\varepsilon_{1}). By Proposition 2, there is R>0R_{*}>0 such that M(Γ(Cj,A1,n¯))RM(\Gamma(C_{j},A_{1},\overline{{\mathbb{R}}^{n}}))\geqslant R_{*} for any j.j\in{\mathbb{N}}. Now, by Proposition 2 M(Γ(Cj,A1,B(x0,ε1)))12M(Γ(Cj,A1,n¯))R/2.M(\Gamma(C_{j},A_{1},B(x_{0},\varepsilon_{1})))\geqslant\frac{1}{2}\cdot M(\Gamma(C_{j},A_{1},\overline{{\mathbb{R}}^{n}}))\geqslant R_{*}/2. Now M(Γ(Cj,A1,D))R/2M(\Gamma(C_{j},A_{1},D))\geqslant R_{*}/2 for any j.j\in{\mathbb{N}}. Finally, M(Γ(Cj,A,D))R/2M(\Gamma(C_{j},A,D))\geqslant R_{*}/2 for any jj\in{\mathbb{N}} by Proposition 2. For the completeness of the proof, we may put R0:=R.R_{0}:=R_{*}. \Box

Lemma 3.2. Let DD be a domain in n,{\mathbb{R}}^{n}, n2,n\geqslant 2, let AA be a set in D,D, and let B(x0,ε1)AB(x_{0},\varepsilon_{1})\subset A for some ε1>0.\varepsilon_{1}>0.

Assume that, QL1(n)Q\in L^{1}({\mathbb{R}}^{n}) and, in addition, for any y0n¯y_{0}\in\overline{{\mathbb{R}}^{n}} there is ε0=ε0(y0)>0\varepsilon_{0}=\varepsilon_{0}(y_{0})>0 and a Lebesgue measurable function ψ:(0,ε0)[0,]\psi:(0,\varepsilon_{0})\rightarrow[0,\infty] such that

I(ε,ε0):=εε0ψ(t)𝑑t<ε(0,ε0),I(ε,ε0)приε0,I(\varepsilon,\varepsilon_{0}):=\int\limits_{\varepsilon}^{\varepsilon_{0}}\psi(t)\,dt<\infty\quad\forall\,\,\varepsilon\in(0,\varepsilon_{0})\,,\quad I(\varepsilon,\varepsilon_{0})\rightarrow\infty\quad\text{при}\quad\varepsilon\rightarrow 0\,, (3.1)

and, in addition,

A(y0,ε,ε0)Q(y)ψn(|yy0|)𝑑m(x)=o(In(ε,ε0)),\int\limits_{A(y_{0},\varepsilon,\varepsilon_{0})}Q(y)\cdot\psi^{\,n}(|y-y_{0}|)\,dm(x)=o(I^{n}(\varepsilon,\varepsilon_{0}))\,, (3.2)

as ε0,\varepsilon\rightarrow 0, where A(y0,ε,ε0)A(y_{0},\varepsilon,\varepsilon_{0}) is defined in (1.3). Then there is r0>0,r_{0}>0, which does not depend on f,f, such that

f(B(x0,ε1))B(f(x0),r0)f𝔉E,δ(D).f(B(x_{0},\varepsilon_{1}))\supset B(f(x_{0}),r_{0})\qquad\forall\,\,f\in\mathfrak{F}_{E,\delta}(D)\,. (3.3)

Remark 3.1. If y0=,y_{0}=\infty, the relation (3.2) must be understood by the using the inversion ψ(y)=y|y|2\psi(y)=\frac{y}{|y|^{2}} at the origin. In other words, instead of

A(y0,ε,ε0)Q(y)ψn(|yy0|)𝑑m(y)=o(In(ε,ε0))\int\limits_{A(y_{0},\varepsilon,\varepsilon_{0})}Q(y)\cdot\psi^{\,n}(|y-y_{0}|)\,dm(y)=o(I^{n}(\varepsilon,\varepsilon_{0}))

we need to consider the condition

A(0,ε,ε0)Q(y|y|2)ψn(|y|)𝑑m(y)=o(In(ε,ε0)).\int\limits_{A(0,\varepsilon,\varepsilon_{0})}Q\left(\frac{y}{|y|^{2}}\right)\cdot\psi^{\,n}(|y|)\,dm(y)=o(I^{n}(\varepsilon,\varepsilon_{0}))\,.

Proof of Lemma 3. Let us prove the lemma by contradiction. Assume that its conclusion is wrong, i.e., the relation (3.3) does not hold for any r0>0.r_{0}>0. Then for any mm\in{\mathbb{N}} there is ymny_{m}\in{\mathbb{R}}^{n} and fm𝔉E,δ(D)f_{m}\in\mathfrak{F}_{E,\delta}(D) such that |fm(x0)ym|<1/m|f_{m}(x_{0})-y_{m}|<1/m and ymfm(B(x0,ε1)).y_{m}\not\in f_{m}(B(x_{0},\varepsilon_{1})). Due to the compactness of n¯\overline{{\mathbb{R}}^{n}} we may consider that ymy0y_{m}\rightarrow y_{0} as m,m\rightarrow\infty, where y0n¯.y_{0}\in\overline{{\mathbb{R}}^{n}}. Then also fm(x0)y0f_{m}(x_{0})\rightarrow y_{0} as m.m\rightarrow\infty. We may consider that y0.y_{0}\neq\infty.

Since by the assumption h(fm(E))δh(f_{m}(E))\geqslant\delta for any mm\in{\mathbb{N}} and d(fm(E))h(fm(E)),d(f_{m}(E))\geqslant h(f_{m}(E)), where d(fm(E))d(f_{m}(E)) denotes the Euclidean diameter of fm(E),f_{m}(E), there is ε2>0\varepsilon_{2}>0 such that

fm(E)B(y0,ε2),m=1,2,.f_{m}(E)\setminus B(y_{0},\varepsilon_{2})\neq\varnothing,\qquad m=1,2,\ldots\,. (3.4)

By (3.4), there is wm=fm(zm)n¯B(y0,ε2),w_{m}=f_{m}(z_{m})\in\overline{{\mathbb{R}}^{n}}\setminus B(y_{0},\varepsilon_{2}), where zmE.z_{m}\in E. Since EE is a continuum, n¯\overline{{\mathbb{R}}^{n}} is a compactum and the set n¯B(y0,ε2)\overline{{\mathbb{R}}^{n}}\setminus B(y_{0},\varepsilon_{2}) is closed, we may consider that zmz0Ez_{m}\rightarrow z_{0}\in E as mm\rightarrow\infty and wmw0n¯B(y0,ε2).w_{m}\rightarrow w_{0}\in\overline{{\mathbb{R}}^{n}}\setminus B(y_{0},\varepsilon_{2}). Obviously, w0y0.w_{0}\neq y_{0}.

By Proposition 2 the family fmf_{m} is equicontinuous. Now, for any ε>0\varepsilon>0 there is δ=δ(z0)>0\delta=\delta(z_{0})>0 such that h(fm(z0),fm(z))<εh(f_{m}(z_{0}),f_{m}(z))<\varepsilon whenever |zz0|δ.|z-z_{0}|\leqslant\delta. Then, by the triangle inequality

h(fm(z),w0)h(fm(z),fm(z0))+h(fm(z0),fm(zm))+h(fm(zm),w0)<3εh(f_{m}(z),w_{0})\leqslant h(f_{m}(z),f_{m}(z_{0}))+h(f_{m}(z_{0}),f_{m}(z_{m}))+h(f_{m}(z_{m}),w_{0})<3\varepsilon (3.5)

for |zz0|<δ,|z-z_{0}|<\delta, some M1M_{1}\in{\mathbb{N}} and all mM1.m\geqslant M_{1}. We may consider that latter holds for any m=1,2,.m=1,2,\ldots. Since w0n¯B(y0,ε2),w_{0}\in\overline{{\mathbb{R}}^{n}}\setminus B(y_{0},\varepsilon_{2}), we may choose ε>0\varepsilon>0 such that Bh(w0,3ε)¯B(y0,ε2)¯=,\overline{B_{h}(w_{0},3\varepsilon)}\cap\overline{B(y_{0},\varepsilon_{2})}=\varnothing, where Bh(w0,ε)={wn¯:h(w,w0)<ε}.B_{h}(w_{0},\varepsilon)=\{w\in\overline{{\mathbb{R}}^{n}}:h(w,w_{0})<\varepsilon\}. Then (3.5) implies that

fm(E1)B(y0,ε2)¯=,m=1,2,,f_{m}(E_{1})\cap\overline{B(y_{0},\varepsilon_{2})}=\varnothing,\qquad m=1,2,\ldots\,, (3.6)

where E1:=B(z0,δ)¯.E_{1}:=\overline{B(z_{0},\delta)}.

Join the points ymy_{m} and fm(x0)f_{m}(x_{0}) by a segment βm:[0,1]B(fm(x0),1/m)¯\beta_{m}:[0,1]\rightarrow\overline{B(f_{m}(x_{0}),1/m)} such that βm(0)=fm(x0)\beta_{m}(0)=f_{m}(x_{0}) and βm(0)=ym.\beta_{m}(0)=y_{m}. Let αm,\alpha_{m}, αm:[0,cm)B(x0,ε1),\alpha_{m}:[0,c_{m})\rightarrow B(x_{0},\varepsilon_{1}), be a maximal fmf_{m}-lifting of βm\beta_{m} in B(x0,ε1)B(x_{0},\varepsilon_{1}) starting at x0.x_{0}. The lifting αm\alpha_{m} exists by Proposition 2. By the same Proposition either αm(t)x1B(x0,ε1)\alpha_{m}(t)\rightarrow x_{1}\in B(x_{0},\varepsilon_{1}) as tcm0t\rightarrow c_{m}-0 (in this case, cm=1c_{m}=1 and fm(x1)=ymf_{m}(x_{1})=y_{m}), or αm(t)S(x0,ε1)\alpha_{m}(t)\rightarrow S(x_{0},\varepsilon_{1}) as tcm.t\rightarrow c_{m}. Observe that, the first situation is excluded. Indeed, if fm(x1)=ym,f_{m}(x_{1})=y_{m}, then ymfm(B(x0,ε1)),y_{m}\in f_{m}(B(x_{0},\varepsilon_{1})), that contradicts the choice of ym.y_{m}. Thus, αm(t)S(x0,ε1)\alpha_{m}(t)\rightarrow S(x_{0},\varepsilon_{1}) as tcm.t\rightarrow c_{m}. Observe that, |αm|¯\overline{|\alpha_{m}|} is a continuum in B(x0,ε1)¯\overline{B(x_{0},\varepsilon_{1})} and h(|αm|¯)h(0,S(x0,ε1)).h(\overline{|\alpha_{m}|})\geqslant h(0,S(x_{0},\varepsilon_{1})). Let us to apply Lemma 3 for A:=E1:=B(z0,δ),A:=E_{1}:=B(z_{0},\delta), Cm:=|αm|C_{m}:=|\alpha_{m}| and r=h(0,S(x0,ε1)).r=h(0,S(x_{0},\varepsilon_{1})). By this lemma we may find R0>0R_{0}>0 such that

M(Γ(|αm|¯,E1,D))R0,m=1,2,.M(\Gamma(\overline{|\alpha_{m}|},E_{1},D))\geqslant R_{0}\,,\qquad m=1,2,\ldots\,. (3.7)

Let us show that the relation (3.7) contradicts the definition of the mapping fmf_{m} in (1.4)–(1.5). Indeed, since fm(x0)y0f_{m}(x_{0})\rightarrow y_{0} as m,m\rightarrow\infty, for any kk\in{\mathbb{N}} there is a number mkm_{k}\in{\mathbb{N}} such that

B(fmk(x0),1/k)B(y0,2k).B(f_{m_{k}}(x_{0}),1/k)\subset B(y_{0},2^{\,-k})\,. (3.8)

Since |βm|B(fm(x0),1/m),|\beta_{m}|\in B(f_{m}(x_{0}),1/m), by (3.8) we obtain that

|βmk|B(y0,2k),k=1,2,.|\beta_{m_{k}}|\subset B(y_{0},2^{\,-k})\,,\qquad k=1,2,\ldots\,. (3.9)

Let k0k_{0}\in{\mathbb{N}} be such that 2k<ε2,2^{\,-k}<\varepsilon_{2}, where ε2\varepsilon_{2} is a number from (3.6), and let Γk:=Γ(|αmk|,E1,D).\Gamma_{k}:=\Gamma(|\alpha_{m_{k}}|,E_{1},D). In this case, we observe that

fmk(Γk)>Γ(S(y0,ε2),S(y0,2k),A(y0,2k,ε2)),f_{m_{k}}(\Gamma_{k})>\Gamma(S(y_{0},\varepsilon_{2}),S(y_{0},2^{\,-k}),A(y_{0},2^{\,-k},\varepsilon_{2}))\,, (3.10)

see Figure 1 for the scheme of the proof.

Refer to caption

Figure 1: To the proof of Lemma 3

Indeed, let γ~fmk(Γk).\widetilde{\gamma}\in f_{m_{k}}(\Gamma_{k}). Then γ~(t)=fmk(γ(t)),\widetilde{\gamma}(t)=f_{m_{k}}(\gamma(t)), where γΓk,\gamma\in\Gamma_{k}, γ:[0,1]D,\gamma:[0,1]\rightarrow D, γ(0)|αmk|,\gamma(0)\in|\alpha_{m_{k}}|, γ(1)E1.\gamma(1)\in E_{1}. By the relation (3.6), we obtain that fmk(γ(0))nB(y0,ε2).f_{m_{k}}(\gamma(0))\in{\mathbb{R}}^{n}\setminus B(y_{0},\varepsilon_{2}). In addition, by (3.9) we have that fmk(γ(1))B(y0,2k)B(y0,ε2)f_{m_{k}}(\gamma(1))\in B(y_{0},2^{\,-k})\subset B(y_{0},\varepsilon_{2}) for kk0.k\geqslant k_{0}. Thus, |fmk(γ(t))|B(y0,ε2)|fmk(γ(t))|(nB(y0,ε2)).|f_{m_{k}}(\gamma(t))|\cap B(y_{0},\varepsilon_{2})\neq\varnothing\neq|f_{m_{k}}(\gamma(t))|\cap({\mathbb{R}}^{n}\setminus B(y_{0},\varepsilon_{2})). Now, by [Ku, Theorem 1.I.5.46] we obtain that, there is 0<t1<10<t_{1}<1 such that fmk(γ(t1))S(y0,ε2).f_{m_{k}}(\gamma(t_{1}))\in S(y_{0},\varepsilon_{2}). Set γ1:=γ|[t1,1].\gamma_{1}:=\gamma|_{[t_{1},1]}. We may consider that fmk(γ(t))B(y0,ε2)f_{m_{k}}(\gamma(t))\in B(y_{0},\varepsilon_{2}) for any tt1.t\geqslant t_{1}. Arguing similarly, we obtain t2[t1,1]t_{2}\in[t_{1},1] such that fmk(γ(t2))S(y0,2k).f_{m_{k}}(\gamma(t_{2}))\in S(y_{0},2^{\,-k}). Put γ2:=γ|[t1,t2].\gamma_{2}:=\gamma|_{[t_{1},t_{2}]}. We may consider that fmk(γ(t))B(y0,2k)f_{m_{k}}(\gamma(t))\not\in B(y_{0},2^{\,-k}) for any t[t1,t2].t\in[t_{1},t_{2}]. Now, a path fmk(γ2)f_{m_{k}}(\gamma_{2}) is a subpath of f(γ)=γ~,f(\gamma)=\widetilde{\gamma}, which belongs to Γ(S(y0,2k),S(y0,ε2),A(y0,2k,ε2)).\Gamma(S(y_{0},2^{\,-k}),S(y_{0},\varepsilon_{2}),A(y_{0},2^{\,-k},\varepsilon_{2})). The relation (3.10) is established.

It follows from (3.10) that

Γk>Γfmk(S(y0,2k),S(y0,ε2),A(y0,2k,ε2)).\Gamma_{k}>\Gamma_{f_{m_{k}}}(S(y_{0},2^{\,-k}),S(y_{0},\varepsilon_{2}),A(y_{0},2^{\,-k},\varepsilon_{2}))\,. (3.11)

Since I(ε,ε0)I(\varepsilon,\varepsilon_{0})\rightarrow\infty as ε0,\varepsilon\rightarrow 0, we may consider that I(2k,ε2)>0I(2^{\,-k},\varepsilon_{2})>0 for sufficiently large k.k\in{\mathbb{N}}. Set

ηk(t)={ψ(t)/I(2k,ε2),t(2k,ε2),0,t(2k,ε2),\eta_{k}(t)=\left\{\begin{array}[]{rr}\psi(t)/I(2^{\,-k},\varepsilon_{2}),&t\in(2^{\,-k},\varepsilon_{2})\,,\\ 0,&t\not\in(2^{\,-k},\varepsilon_{2})\,,\end{array}\right.

where I(2k,ε2)=2kε2ψ(t)𝑑t.I(2^{\,-k},\varepsilon_{2})=\int\limits_{2^{\,-k}}^{\varepsilon_{2}}\,\psi(t)\,dt. Observe that 2kε2ηk(t)𝑑t=1.\int\limits_{2^{\,-k}}^{\varepsilon_{2}}\eta_{k}(t)\,dt=1. Now, by the relations (3.2) and (3.11), and due to the definition of fmkf_{m_{k}} in (1.4)–(1.5), we obtain that

M(Γk)=M(Γ(|αmk|,E1,D))M(Γfmk(S(y0,2k),S(y0,ε2),A(y0,2k,ε2)))M(\Gamma_{k})=M(\Gamma(|\alpha_{m_{k}}|,E_{1},D))\leqslant M(\Gamma_{f_{m_{k}}}(S(y_{0},2^{\,-k}),S(y_{0},\varepsilon_{2}),A(y_{0},2^{\,-k},\varepsilon_{2})))\leqslant
1In(2k,ε2)A(y0,2k,ε2)Q(y)ψn(|yy0|)𝑑m(y)0ask.\leqslant\frac{1}{I^{n}(2^{\,-k},\varepsilon_{2})}\int\limits_{A(y_{0},2^{\,-k},\varepsilon_{2})}Q(y)\cdot\psi^{\,n}(|y-y_{0}|)\,dm(y)\rightarrow 0\quad\text{as}\quad k\rightarrow\infty\,. (3.12)

The relation (3.12) contradicts with (3.7). The contradiction obtained above proves the lemma. \Box

Proof of Theorem 1 immediately follows by Lemma 3 and Proposition 2\Box

Remark 3.2. If, under the conditions of Theorem 1, the mapped domain f(D)=Df(D)=D^{\,\prime} is fixed and bounded, then the condition QL1(n)Q\in L^{1}({\mathbb{R}^{n}}) may be slightly weakened.

Given domain D,Dn,D,D^{\,\prime}\subset{\mathbb{R}}^{n}, n2,n\geqslant 2, a continuum ED,E\subset D, δ>0\delta>0 and a Lebesgue measurable function Q:n[0,]Q:{\mathbb{R}}^{n}\rightarrow[0,\infty] we denote by 𝔉E,δ(D,D)\mathfrak{F}_{E,\delta}(D,D^{\,\prime}) the family of all mapping f:Dn,f:D\rightarrow{\mathbb{R}}^{n}, n2,n\geqslant 2, satisfying relations (1.4)–(1.5) at any point y0n¯y_{0}\in\overline{{\mathbb{R}}^{n}} such that h(f(E))δ.h(f(E))\geqslant\delta. The following statement holds.

Assume that, DD^{\,\prime} is bounded and for each point y0Dy_{0}\in D^{\,\prime} and for every 0<r1<r2<r0:=supyD|yy0|0<r_{1}<r_{2}<r_{0}:=\sup\limits_{y\in D^{\,\prime}}|y-y_{0}| there is a set E[r1,r2]E\subset[r_{1},r_{2}] of a positive linear Lebesgue measure such that the function QQ is integrable with respect to n1\mathcal{H}^{n-1} over the spheres S(y0,r)S(y_{0},r) for every rE.r\in E. In addition, assume that one of the following conditions hold:

1) QFMO(n¯);Q\in FMO(\overline{{\mathbb{R}}^{n}});

2) for any y0n¯y_{0}\in\overline{{\mathbb{R}}^{n}} there is δ(y0)>0\delta(y_{0})>0 such that (1.11) holds. Then there is r0>0,r_{0}>0, which does not depend on f,f, such that

f(B(x0,ε1))B(f(x0),r0)f𝔉E,δ(D,D).f(B(x_{0},\varepsilon_{1}))\supset B(f(x_{0}),r_{0})\qquad\forall\,\,f\in\mathfrak{F}_{E,\delta}(D,D^{\,\prime})\,.

The proof of this statement is exactly the same as the proof of Theorem 1. The above condition on the function Q,Q, replacing the condition QL1(n),Q\in L^{1}({\mathbb{R}^{n}}), ensures equicontinuity of the family of mappings 𝔉E,δ(D,D)\mathfrak{F}_{E,\delta}(D,D^{\,\prime}) (see [SevSkv, Theorem 1.1]). In all other respects, the proof scheme is the same.

Proof of Theorem 1. Assume that ff is not a constant. Now, the lightness of ff follows by [Sev1, Theorem]. It remains to show that ff is open. Let AA be an open set and let x0A.x_{0}\in A. We need to show that, there is ε>0\varepsilon^{*}>0 such that B(f(x0),ε)f(A).B(f(x_{0}),\varepsilon^{*})\subset f(A). Since AA is open, there is ε1>0\varepsilon_{1}>0 such that B(x0,ε1)¯A.\overline{B(x_{0},\varepsilon_{1})}\subset A.

Since ff is not constant, there are a,bDa,b\in D such that f(a)f(b).f(a)\neq f(b). Let us join the points aa and bb by a path γ\gamma in D.D. We set E:=|γ|.E:=|\gamma|. Now, h(fm(a),fm(b))12h(f(a),f(b)):=δh(f_{m}(a),f_{m}(b))\geqslant\frac{1}{2}\cdot h(f(a),f(b)):=\delta for sufficiently large m.m\in{\mathbb{N}}.

By Theorem 1 there is r0>0,r_{0}>0, which does not depend on m,m, such that B(fm(x0),r0)fm(B(x0,ε1)),B(f_{m}(x_{0}),r_{0})\subset f_{m}(B(x_{0},\varepsilon_{1})), m=1,2,.m=1,2,\ldots.

Set ε:=r0/2.\varepsilon^{*}:=r_{0}/2. Let yB(f(x0),r0/2).y\in B(f(x_{0}),r_{0}/2). Since by the assumption fm(x)f(x)f_{m}(x)\rightarrow f(x) locally uniformly in D,D, by the triangle inequality we obtain that

|fm(x0)y||fm(x0)f(x0)|+|f(x0)y|<r0|f_{m}(x_{0})-y|\leqslant|f_{m}(x_{0})-f(x_{0})|+|f(x_{0})-y|<r_{0}

for sufficiently large m.m\in{\mathbb{N}}. Thus, yB(fm(x0),r0)fm(B(x0,ε1)).y\in B(f_{m}(x_{0}),r_{0})\subset f_{m}(B(x_{0},\varepsilon_{1})). Consequently, y=fm(xm)y=f_{m}(x_{m}) for some xmB(x0,ε1).x_{m}\in B(x_{0},\varepsilon_{1}). Due to the compactness of B(x0,ε1)¯,\overline{B(x_{0},\varepsilon_{1})}, we may consider that xmz0B(x0,ε1)¯x_{m}\rightarrow z_{0}\in\overline{B(x_{0},\varepsilon_{1})} as m.m\rightarrow\infty. By the continuity of ff in A,A, since B(x0,ε1)¯A,\overline{B(x_{0},\varepsilon_{1})}\subset A, we obtain that f(xm)f(z0)f(x_{m})\rightarrow f(z_{0}) as m.m\rightarrow\infty. So, we have that f(xm)f(z0)f(x_{m})\rightarrow f(z_{0}) as mm\rightarrow\infty and simultaneously y=fm(xm)y=f_{m}(x_{m}) for sufficiently large m.m\in{\mathbb{N}}. Thus

|yf(z0)|=|fm(xm)f(z0)||y-f(z_{0})|=|f_{m}(x_{m})-f(z_{0})|\leqslant
|fm(xm)f(xm)|+|f(xm)f(z0)|0,\leqslant|f_{m}(x_{m})-f(x_{m})|+|f(x_{m})-f(z_{0})|\rightarrow 0\,,

m.m\rightarrow\infty. Thus, y=f(z0)f(B(x0,ε1)¯)f(A).y=f(z_{0})\in f(\overline{B(x_{0},\varepsilon_{1})})\subset f(A). So, yf(A),y\in f(A), i.e., B(f(x0),r0/2)f(A),B(f(x_{0}),r_{0}/2)\subset f(A), as required. \Box

Theorem 3.1.   Let DD be a domain in n,{\mathbb{R}}^{n}, n2.n\geqslant 2. Let fj:Dn,f_{j}:D\rightarrow{\mathbb{R}}^{n}, n2,n\geqslant 2, j=1,2,,j=1,2,\ldots, be a sequence of open discrete mappings satisfying the conditions (1.4)–(1.5) at any point y0n¯y_{0}\in\overline{{\mathbb{R}}^{n}} and converging to some mapping f:Dn¯f:D\rightarrow\overline{{\mathbb{R}}^{n}} as jj\rightarrow\infty locally uniformly in DD with respect to the chordal metric h.h. Assume that the conditions on the function QQ from Theorem 1 hold. Then either ff is a constant in n¯\overline{{\mathbb{R}}^{n}}, or ff is light and open mapping f:Dn.f:D\rightarrow{\mathbb{R}}^{n}.

Proof.   Assume that ff is not a constant. Then there are a,bDa,b\in D such that f(a)f(b).f(a)\neq f(b). Let us join the points aa and bb by a path γ\gamma in D.D. We set E:=|γ|.E:=|\gamma|. Now, h(fm(a),fm(b))12h(f(a),f(b)):=δh(f_{m}(a),f_{m}(b))\geqslant\frac{1}{2}\cdot h(f(a),f(b)):=\delta for sufficiently large m.m\in{\mathbb{N}}.

Let x0Dx_{0}\in D and let y0=f(x0).y_{0}=f(x_{0}). By Theorem 1 there is r0>0,r_{0}>0, which does not depend on m,m, such that B(fm(x0),r0)fm(B(x0,ε1)),B(f_{m}(x_{0}),r_{0})\subset f_{m}(B(x_{0},\varepsilon_{1})), m=1,2,.m=1,2,\ldots. Then also Bh(fm(x0),r)fm(B(x0,ε1)),B_{h}(f_{m}(x_{0}),r_{*})\subset f_{m}(B(x_{0},\varepsilon_{1})), m=1,2,,m=1,2,\ldots, for some r>0.r_{*}>0. Let yBh(y0,r/2)=Bh(f(x0),r/2).y\in B_{h}(y_{0},r_{*}/2)=B_{h}(f(x_{0}),r_{*}/2). By the converges of fmf_{m} to ff and by the triangle inequality, we obtain that

h(y,fm(x0))h(y,f(x0))+h(f(x0),fm(x0))<r/2+r/2=rh(y,f_{m}(x_{0}))\leqslant h(y,f(x_{0}))+h(f(x_{0}),f_{m}(x_{0}))<r_{*}/2+r_{*}/2=r_{*}

for sufficiently large m.m\in{\mathbb{N}}. Thus,

Bh(f(x0),r/2)Bh(fm(x0),r)fm(B(x0,ε1))n.B_{h}(f(x_{0}),r_{*}/2)\subset B_{h}(f_{m}(x_{0}),r_{*})\subset f_{m}(B(x_{0},\varepsilon_{1}))\subset{\mathbb{R}}^{n}\,.

In particular, y0=f(x0)n,y_{0}=f(x_{0})\in{\mathbb{R}}^{n}, as required. The lightness and the openness of ff follows by Theorem 1\Box

Open problem. Is it possible to assert that, under the conditions of Theorems 1 and 3, the mapping ff is open and discrete?

References

  • [CG] Carleson, L., T.W. Gamelin: Complex dynamics, Universitext: Tracts in Mathematics. - Springer-Verlag, New York etc., 1993.
  • [Cr] Cristea, M.: Open discrete mappings having local ACLnACL^{n} inverses. - Complex Variables and Elliptic Equations 55: 1–3, 2010, 61–90.
  • [Ku] Kuratowski, K.: Topology, v. 2. – Academic Press, New York–London, 1968.
  • [MRV1] Martio, O., S. Rickman, and J. Väisälä: Definitions for quasiregular mappings. - Ann. Acad. Sci. Fenn. Ser. A1 448, 1969, 1–40.
  • [MRV2] Martio, O., S. Rickman, and J. Väisälä: Topological and metric properties of quasiregular mappings. - Ann. Acad. Sci. Fenn. Ser. A1. 488, 1971, 1–31.
  • [MRSY] Martio, O., V. Ryazanov, U. Srebro, and E. Yakubov: Moduli in modern mapping theory. - Springer Science + Business Media, LLC, New York, 2009.
  • [Na] Näkki, R.: Extension of Loewner’s capacity theorem. - Trans. Amer. Math. Soc. 180, 1973, 229–236.
  • [Re] Reshetnyak, Yu.G.: Space Mappings with Bounded Distortion. Transl. of Math. Monographs 73, AMS, 1989.
  • [Ri] Rickman, S.: Quasiregular mappings. Springer-Verlag, Berlin, 1993.
  • [Sev1] Sevost’yanov, E.A.: On the zero-dimensionality of the limit of the sequence of generalized quasiconformal mappings. - Math. Notes 102:4, 2017, 547–555.
  • [Sev2] Sevost’yanov, E.A: Mappings with Direct and Inverse Poletsky Inequalities. Developments in Mathematics (DEVM, volume 78). - Springer Nature Switzerland AG, Cham, 2023.
  • [SevSkv] Sevost’yanov, E.A., S.O. Skvortsov: Logarithmic Hölder continuous mappings and Beltrami equation. - Analysis and Mathematical Physics 11:3, 2021, Article number 138.
  • [SSD] Sevost’yanov, E.A., S.O. Skvortsov, O.P. Dovhopiatyi: On nonhomeomorphic mappings with the inverse Poletsky inequality. - Journal of Mathematical Sciences 252:4, 2021, 541–557.
  • [Vu] Vuorinen, M.: On the existence of angular limits of nn-dimensional quasiconformal mappings. - Ark. Math. 18, 1980, 157–180.
  • [Va] Väisälä J.: Lectures on nn-dimensional quasiconformal mappings. - Lecture Notes in Math. 229, Springer-Verlag, Berlin etc., 1971.
  • [1]

CONTACT INFORMATION

Evgeny Sevost’yanov
1. Zhytomyr Ivan Franko State University,
40 Bol’shaya Berdichevskaya Str., 10 008 Zhytomyr, UKRAINE
2. Institute of Applied Mathematics and Mechanics
of NAS of Ukraine,
19 Henerala Batyuka Str., 84 116 Slavyansk, UKRAINE
[email protected]

Valery Targonskii
Zhytomyr Ivan Franko State University,
40 Bol’shaya Berdichevskaya Str., 10 008 Zhytomyr, UKRAINE
[email protected]