Home Freely quasiconformal and locally weakly quasisymmetric mappings in metric spaces
Article Open Access

Freely quasiconformal and locally weakly quasisymmetric mappings in metric spaces

  • Hong-Jun Liu EMAIL logo , Sha-Sha Yan , Ling Xia and Zhi-Fa Yu
Published/Copyright: April 16, 2025

Abstract

In this article, we investigate the relationship between freely quasiconformal mappings and locally weakly quasisymmetric mappings in quasiconvex and complete metric spaces. We first prove the equivalence of freely quasiconformal mappings, ring properties, and locally weakly quasisymmetric mappings. Finally, we prove that the composition of two locally weakly quasisymmetric mappings in metric spaces is locally weakly quasisymmetric mapping.

MSC 2010: 30C65; 30F45; 30L10

1 Introduction and main results

The quasihyperbolic metric was introduced by Gerhing and Palka in the 1970s [1,2] in the setting of Euclidean spaces R n . Since its first appearance, the quasihyperbolic metric has become an important tool in geometric function theory, and the theory related to quasihyperbolic metrics has been generalized to metric spaces and Banach spaces [3]. During the past few decades, modern geometric function theory of quasisymmetric and quasiconformal mappings has been studied from several points of view. Quasisymmetric mappings on the real line were first introduced by Beurling and Ahlfors [4]. They found a way to extend each quasisymmetric self-mapping of the real line to a quasiconformal self-mapping of the upper half-planes. This concept was later promoted by Tukia and Väisälä [5], who introduced and studied quasisymmetric mappings between arbitrary metric spaces. In 1990, based on the idea of quasisymmetry, Väisälä developed a “dimension-free” theory of quasiconformal mappings in infinite-dimensional Banach spaces [3,69]. In 1998, Heinonen and Koskela [10] showed that these concepts, quasiconformality and quasisymmetry, are quantitatively equivalent in a large class of metric spaces, which includes Euclidean spaces. Since these two concepts are equivalent, mathematicians show much interest in the research of quasisymmetric mappings between suitable metric spaces.

The main goal of this article is to discuss the relation between freely quasiconformal mappings, ring properties, and locally weakly quasisymmetric mappings in metric spaces, and prove that the composition of two locally weakly quasisymmetric mappings in metric spaces is locally weakly quasisymmetric mapping. For a metric space X , we use notation x y to indicate the distance between x and y . Following analogous notations and terminologies of [3,1022]. We start with the definition of quasisymmetric mappings.

Definition 1.1

A homeomorphism f : X Y is said to be

  1. η -quasisymmetric if there exists a homeomorphism η : [ 0 , ) [ 0 , ) , η ( 0 ) = 0 , such that

    x a t x b implies f ( x ) f ( a ) η ( t ) f ( x ) f ( b ) ,

    for each t > 0 and for each triple x , a , and b of points in X ;

  2. weakly ( h , H ) -quasisymmetric if there exist constants h > 0 and H 1 so that

    (1) x a h x b implies f ( x ) f ( a ) H f ( x ) f ( b ) ,

    for each triple x , a , and b of points in X .

Remark 1.2

  1. A homeomorphism f : X Y is weakly H -quasisymmetric mapping if it satisfies (1) with h = 1 . The weakly ( h , H ) -quasisymmetric mapping is a generalization of the weakly H -quasisymmetric mapping since the weakly H -quasisymmetric mapping coincides with the weakly ( 1 , H ) -quasisymmetric mapping.

  2. The η -quasisymmetric mapping implies the weakly ( h , H ) -quasisymmetric mapping with H = η ( h ) . Obviously, η ( h ) h .

Definition 1.3

Let X and Y be two metric spaces and G X , G Y be two domains (open and connected). Let h > 0 and H 1 be two constants. A homeomorphism f : G G is said to be q -locally weakly ( h , H ) -quasisymmetric for some 0 < q < 1 , if the map f B x , q : B x , q f ( B x , q ) is weakly ( h , H ) -quasisymmetric for any point x G , where B x , q = B ( x , q δ G ( x ) ) .

We remark that, in general, the locally weakly quasisymmetric mapping means the locally weakly H -quasisymmetric mapping [12]. Obviously, the locally weakly ( h , H ) -quasisymmetric mapping is a generalization of the locally weakly H -quasisymmetric mapping. In this article, we always simplify locally weakly ( h , H ) -quasisymmetric mapping by locally weakly quasisymmetric mapping.

Recall that a metric space is proper if all its closed balls are compact. The quasihyperbolic metric of a domain in R n was introduced by Gehring and Palka [2]. For the concepts of quasihyperbolic metrics k G and k G , please see Section 2.

Definition 1.4

Let G X and G Y be two domains, and let φ : [ 0 , ) [ 0 , ) be a homeomorphism with φ ( t ) t . We say that a homeomorphism f : G G is

  1. φ -semisolid if

    k G ( f ( x ) , f ( y ) ) φ ( k G ( x , y ) ) ,

    for all x , y G ;

  2. φ -solid if f and f 1 are φ -semisolid, i.e.,

    φ 1 ( k G ( x , y ) ) k G ( f ( x ) , f ( y ) ) φ ( k G ( x , y ) ) ,

    for all x , y G ;

  3. fully φ -semisolid (resp., fully φ -solid) if f is φ -semisolid (resp., φ -solid) in every proper subdomain of G , i.e., the homeomorphism f D : D f ( D ) is φ -semisolid (resp., φ -solid) for each proper subdomain D G . Fully φ -solid mappings are also called freely φ -quasiconformal mappings.

In order to state our results, we introduce the following definitions.

Definition 1.5

Let G X be a domain in a metric space X . A point x G is called cut point if G \ { x } is not connected. A domain G is called non-cut-point domain if it has no cut points.

Definition 1.6

Let X be a metric space and B = B ( x , r ) be a metric open ball in X . Let 0 < λ 1 be a constant. We say that B is a John ball if for each y B , there exists a rectifiable curve γ joining x and y in B whose arc length parameterization γ s : [ 0 , l ( γ ) ] B satisfies: γ s ( 0 ) = y , γ s ( l ( γ ) ) = x and dist ( γ s ( u ) , X \ B ) λ u for all u [ 0 , l ( γ ) ] . Here, l ( γ ) is the length of γ , and curve γ s is the arc length parameterization of γ .

Definition 1.7

Let X be a metric space and G X be a domain. We called G a λ -John-ball domain if for every metric open ball, B ( x , r ) G is a John ball.

Väisälä [3] obtained several characterizations of locally quasisymmetric in Banach spaces; Huang et al. [14] investigated the relationship between semisolid and locally weakly quasisymmetric in quasiconvex and complete metric spaces. Under suitable geometric conditions (Section 2), in this article, we shall prove a more general result (Theorem 1.8) for metric spaces.

Theorem 1.8

Let X and Y be two c-quasiconvex, complete metric spaces, and let G X and G Y be two λ -John-ball and non-cut-point domains. Suppose that f : G G is a homeomorphism, for some 0 < q < 1 . Then, the following are quantitatively equivalent:

  1. f and f 1 are q-locally weakly ( h , H ) -quasisymmetric mappings;

  2. f and f 1 have the ( M , α , β ) -ring properties;

  3. f is freely φ -quasiconformal mapping, where the constants q, h, H, M, α , β , and φ depend on each other and constants c and λ .

As an application of Theorem 1.8, we show that the composite mapping of two locally weakly quasisymmetric mappings in a large class of metric spaces is a locally weakly quasisymmetric mapping.

Theorem 1.9

Let X, Y, and Z be three c-quasiconvex, complete metric spaces. Suppose that G 1 X , G 2 Y , and G 3 Z are three λ -John-ball and non-cut-point domains. If f : G 1 G 2 and f 1 : G 2 G 1 are q 1 -locally weakly ( h 1 , H 1 ) -quasisymmetric mappings, g : G 2 G 3 and g 1 : G 3 G 2 are q 2 -locally weakly ( h 2 , H 2 ) -quasisymmetric mappings, then the compositions g f and ( g f ) 1 are q-locally weakly ( h , H ) -quasisymmetric mappings, where q , h , and H depend on q 1 , h 1 , H 1 , q 2 , h 2 , H 2 , c , and λ .

This article is organized as follows. In Section 2, we will introduce some necessary notations and concepts, recall some known results, and prove a series of basic and useful results. The goal of Section 3 is to show the equivalence between (i), (ii), and (iii) in Theorem 1.8, and in Section 4, we shall prove that the composition of two locally weakly quasisymmetric mappings in metric spaces is locally weakly quasisymmetric mapping.

2 Preliminaries

Let X be a metric space. We always denote the metric open and close balls with center x X and radius r > 0 by

B ( x , r ) = { y X : y x < r } and B ¯ ( x , r ) = { y X : y x r } ,

and denote the metric sphere by

S ( x , r ) = { y X : y x = r } .

If B = B ( x , r ) (or B ¯ = B ¯ ( x , r ) , S = S ( x , r ) ), then λ B = B ( x , λ r ) (or λ B ¯ = B ¯ ( x , λ r ) , λ S = S ( x , λ r ) ) for any λ > 0 . For a set A in X , we always use A ¯ (resp., A ) to denote the closure (resp., the boundary) of A . The diameter of a set E X is the quantity

diam ( E ) = sup { x y : x , y E } ,

and the distance between E , F X is

dist ( E , F ) = inf { x y : x E , y F } .

2.1 Quasihyperbolic metrics

In this subsection, we give some basic properties of quasihyperbolic metric (see [3,6,16,2326], and references therein for more details).

Let γ X be a rectifiable curve of length l ( γ ) with endpoints a and b . If l ( γ ) < , then the curve γ is said to be a rectifiable curve. Suppose that γ : [ a , b ] X is a rectifiable curve. For each t [ a , b ] , we denote s γ ( t ) = l ( γ [ a , t ] ) , where the function s γ : [ a , b ] [ 0 , l ( γ ) ] is called the length function of γ . For any rectifiable curve γ : [ a , b ] X , there exists a unique curve γ s : [ 0 , l ( γ ) ] X such that γ = γ s s γ . The curve γ s is called the arc length parameterization of γ . This parameterization is characterized by the relation l ( γ s [ t 1 , t 2 ] ) = t 2 t 1 for all 0 t 1 < t 2 l ( γ ) .

Definition 2.1

Let X be a metric space and G X be a non-empty open set, and let γ be a rectifiable curve in a domain G X . The quasihyperbolic length of γ in G is

l q h ( γ ) = γ d s δ G ( x ) ,

where δ G ( x ) denotes the distance from x to X \ G , i.e., δ G ( x ) = dist ( x , X \ G ) .

The quasihyperbolic distance between x and y in G is defined by

k G ( x , y ) = inf γ l q h ( γ ) ,

where γ runs over all rectifiable curves in G joining x and y . If there is no rectifiable curve in G joining x and y , then we define k G ( x , y ) = + .

A non-empty open set G of a metric space X is said to be rectifiably connected if for any two points x , y G , there exists a rectifiable curve in G joining x and y . If G X is a rectifiably connected open set, it is clear that k G ( x , y ) < for any two points x , y G . Thus, it is easy to verify that k G is a metric in G , and we call it the quasihyperbolic metric of G , then ( G , k G ) is a quasihyperbolic metric space.

For c 1 , a metric space X is c -quasiconvex if each pair of points x , y X can be joined by a curve γ with length l ( γ ) c x y .

Lemma 2.2

(See, [25], lemma 2.5) Let X be a c-quasiconvex metric space, and let G X be a domain. For any x , y B ( z , q δ G ( z ) ) with 0 < q 1 2 c + 1 , there exists a rectifiable curve γ B ( z , ( 2 c + 1 ) q δ G ( z ) ) joining x and y such that l ( γ ) c x y .

Lemma 2.3

(See, [16], Theorem 2.7) Let X be a c-quasiconvex metric space, and let G X be a domain. Then,

  1. for each x , y G ,

    x y ( e k G ( x , y ) 1 ) δ G ( x ) ;

  2. if z G , 0 < t < 1 , and x , y B ¯ z , t δ G ( z ) 4 c , then

    1 1 + 2 t x y δ G ( z ) k G ( x , y ) c 1 t x y δ G ( z ) .

Lemma 2.4

(See, [15], Theorem 2.6) Let X be a c-quasiconvex metric space and G X be a domain. Suppose that x , y G and either x y δ G ( x ) 8 c or k G ( x , y ) 1 4 . Then,

1 2 x y δ G ( x ) k G ( x , y ) 2 c x y δ G ( x ) .

Lemma 2.5

(See, [15], Theorem 2.7) Let X be a c-quasiconvex metric space and G X be a domain. Then,

  1. l k G ( γ ) = l q h ( γ ) , where l k G ( γ ) is the length of γ in the metric space ( G , k G ) ;

  2. the metric space ( G , k G ( ) ) is a 2-quasiconvex metric space.

2.2 Ring properties, relative homeomorphisms, and uniformly continuous

Definition 2.6

Let x G , M > 0 , and 1 < α β be two positive real numbers. We say that a homeomorphism f : G G has ( M , α , β ) -ring property if

sup 0 < r < r x , β diam ( f ( B ¯ ) ) dist ( f ( B ¯ ) , G \ f ( α B ) ) M ,

where

r x , β = δ G ( x ) β , B = B ( x , r ) , and α B = B ( x , α r ) .

Definition 2.7

Let 0 < t 0 1 , and let θ : [ 0 , t 0 ) [ 0 , ) be an embedding with θ ( 0 ) = 0 . Suppose that G X and G Y be two domains. We say that a homeomorphism f : G G is

  1. ( θ , t 0 ) -relative if there is a constant t 0 ( 0,1 ] and a homeomorphism θ : [ 0 , t 0 ) [ 0 , ) such that

    f ( x ) f ( y ) δ G ( f ( x ) ) θ x y δ G ( x ) ,

    whenever x , y G and x y < t 0 δ G ( x ) ; in particular, if t 0 = 1 , then we say that f is θ -relative;

  2. fully ( θ , t 0 ) -relative (resp., fully θ -relative) if f is ( θ , t 0 ) -relative (resp., θ -relative) in every subdomain of G .

Definition 2.8

Let X and Y be metric spaces. A map f : X Y is a uniformly continuous if there is t 0 ( 0 , ] and a homeomorphism φ : [ 0 , t 0 ) [ 0 , ) such that

f ( x ) f ( y ) φ ( x y ) ,

for all x , y X with x y < t 0 . The function φ is a modulus of continuity of f , and we say that f is ( φ , t 0 ) -uniformly continuous. If t 0 = , we briefly say that f is φ -uniformly continuous.

Theorem 2.9

Let X and Y be two c-quasiconvex and complete metric spaces, and let G X and G Y be two domains. Suppose that f : G G is a homeomorphism. Then, the following conditions are quantitatively equivalent:

  1. f is ( φ , t 0 ) -uniformly continuous in the quasihyperbolic metric;

  2. f is φ -semisolid mapping,

where φ and φ depend on each other and constant t 0 .

Proof

Obviously, we only need to show the implication from (i) to (ii), since the implication from (ii) to (i) is a direct consequence of the definitions of uniformly continuous and semisolid mapping. To prove this implication from (i) to (ii), we assume that f is ( φ , t 0 ) -uniformly continuous in the quasihyperbolic metric, i.e.,

(2) k G ( f ( x ) , f ( y ) ) φ ( k G ( x , y ) ) ,

for all x , y G with k G ( x , y ) t 0 .

Since X is a c -quasiconvex complete metric space, by Observation 2.6 of [16], we have G rectifiably connected. By Lemma 2.5, we know that ( G , k G ) is 2-quasiconvex. Therefore, for any given x , y G , there is a path γ G joining x and y with

l k G ( γ ) 2 k G ( x , y ) .

Let m 0 be the unique integer satisfying

(3) m t 0 < l k G ( γ ) ( m + 1 ) t 0 .

Let γ s k : [ 0 , l k G ( γ ) ] G be the arc length parameterization of γ with metric k G . Denote

(4) τ j = j m + 1 l k G ( γ ) and x j = γ s k ( τ j ) ,

for all 1 j m + 1 . Thus, from (3) and (4), we know that

k G ( x j 1 , x j ) l k G ( γ s k [ τ j 1 , τ j ] ) = τ j τ j 1 = l k G ( γ ) m + 1 t 0 .

Since f is ( φ , t 0 ) -uniformly continuous in the quasihyperbolic metric, by (2), it follows that

k G ( f ( x j 1 ) , f ( x j ) ) φ ( k G ( x j 1 , x j ) ) φ ( t 0 ) ,

for all 1 j m + 1 . Hence, we deduce that

(5) k G ( f ( x ) , f ( y ) ) j = 1 m + 1 k G ( f ( x j 1 ) , f ( x j ) ) ( m + 1 ) φ ( t 0 ) .

For all x , y G , remembering that m t 0 < l k G ( γ ) 2 k G ( x , y ) , (5) implies that

(6) k G ( f ( x ) , f ( y ) ) 2 φ ( t 0 ) t 0 k G ( x , y ) + φ ( t 0 ) .

If k G ( x , y ) > t 0 , according to (6), we obtain that

k G ( f ( x ) , f ( y ) ) 2 φ ( t 0 ) t 0 k G ( x , y ) + φ ( t 0 ) .

If t 0 2 < k G ( x , y ) t 0 , by the definition of uniformly continuous in the quasihyperbolic metric, we deduce that

k G ( f ( x ) , f ( y ) ) φ ( k G ( x , y ) ) φ ( t 0 ) 4 φ ( t 0 ) t 0 k G ( x , y ) t 0 2 + φ ( t 0 ) .

If 0 < k ( x , y ) t 0 2 , applying the definition of uniformly continuous in the quasihyperbolic metric again, we know that

k G ( f ( x ) , f ( y ) ) φ ( k G ( x , y ) ) φ ( k G ( x , y ) ) + 2 k G ( x , y ) t 0 φ ( t 0 ) φ t 0 2 .

Therefore, we obtain that

k G ( f ( x ) , f ( y ) ) φ ( k G ( x , y ) ) ,

for all x , y G . Here,

φ ( t ) = 2 φ ( t 0 ) t 0 t + φ ( t 0 ) , for t 0 t ; 4 φ ( t 0 ) t 0 t t 0 2 + φ ( t 0 ) , for t 0 2 t t 0 ; φ ( t ) + 2 t t 0 φ ( t 0 ) φ t 0 2 , for 0 < t t 0 2 .

Hence, Theorem 2.9 is proved.□

3 Proof of Theorem 1.8

In what follows, we always assume that X and Y are two c -quasiconvex and complete metric spaces, and that G X and G Y are two domains. Furthermore, we suppose that f and f 1 are q -locally weakly ( h , H ) -quasisymmetric mappings for some 0 < q < 1 .

Now, we are ready to prove Theorem 1.8. We verify the implications indicated by the following routes:

( i ) ( ii ) ( iii ) ( i ) .

By Lemmas 5.3 and 5.4 of [14], the implication ( ii ) ( iii ) is true in all metric spaces. It suffices to prove that ( i ) ( ii ) and ( iii ) ( i ) .

The proof is based on a refinement of the method due to Väisälä [3] and Huang et al. [14]. The proof is given in the following two subsections.

3.1 Proof of the implication from (i) to (ii) in Theorem 1.8

Assume f and f 1 are q -locally weakly ( h , H ) -quasisymmetric mappings for some 0 < q < 1 . Let

r x , β = δ G ( x ) β , B = B ( x , r ) and α B = B ( x , α r ) .

We shall show that f and f 1 have the ( M , α , β ) -ring properties, i.e.,

sup 0 < r < r x , β diam ( f ( B ¯ ) ) dist ( f ( B ¯ ) , G \ f ( α B ) ) M ,

where

M = 2 H 2 ( H + 1 ) , α = max 3 h 2 , 3 and β = max 3 q h 2 , 3 q .

In what follows, we only need to verify the f has the ( M , α , β ) -ring property. The proof of the ( M , α , β ) -ring property of f 1 follows in a similar manner.

In order to prove that f has the ( M , α , β ) -ring property, we divide the discussions into two cases:

Case 3.1.1. 0 < h < 1 .

For x G , 0 < q < 1 , let B = B ( x , r ) and α = 3 ( h 2 ) , where 0 < r q 3 h 2 δ G ( x ) . Then,

B ¯ ( x , r ) B ¯ ( x , α r ) B ( x , q δ G ( x ) ) .

For all a , b B ¯ ( x , r ) , we assume that

y 1 h B ¯ ( x , r ) and z G \ 3 h 2 B ( x , r ) .

It is clear that

max { a x , b x } h x y .

Since f is q -locally weakly ( h , H ) -quasisymmetric mapping for some 0 < q < 1 , it follows from the definition that

f ( a ) f ( x ) H f ( x ) f ( y ) and f ( b ) f ( x ) H f ( x ) f ( y ) .

It follows from the aforementioned fact that

(7) f ( a ) f ( b ) f ( a ) f ( x ) + f ( b ) f ( x ) 2 H f ( x ) f ( y ) .

Using z G \ ( ( 3 ( h 2 ) ) B ( x , r ) ) and y ( ( 1 h ) B ¯ ( x , r ) ) , we obtain

z y z x x y 3 r h 2 r h 2 r h 2 > 1 h x y .

From the definition of q -locally weakly ( h , H ) -quasisymmetric mapping, we have

(8) f ( x ) f ( y ) H f ( z ) f ( y ) .

Together with inequalities (7) and (8), it follows that

(9) f ( a ) f ( b ) 2 H 2 f ( z ) f ( y ) .

Let u be any point in B ¯ ( x , r ) , we obtain that

(10) u y diam 1 h B ¯ ( x , r ) = 2 r h

and

(11) u z z x x u 3 r h 2 r 2 r h 2 .

Using inequalities (10) and (11), we now obtain

u y h u z .

Therefore, by the definition of q -locally weakly ( h , H ) -quasisymmetric mapping again, we have

f ( u ) f ( y ) H f ( u ) f ( z ) .

Thus,

(12) f ( z ) f ( y ) f ( z ) f ( u ) + f ( u ) f ( y ) ( H + 1 ) f ( u ) f ( z ) .

Combining inequalities (9) and (12), we have thus proved that

f ( a ) f ( b ) 2 H 2 ( H + 1 ) f ( u ) f ( z ) .

Since u and z are arbitrary, it follows from Remark 2.7 that

sup 0 < r < r x , β diam ( f ( B ¯ ) ) dist ( f ( B ¯ ) , G \ f ( ( 3 ( h 2 ) ) B ) ) 2 H 2 ( H + 1 ) ,

where β = 3 ( q h 2 ) .

Case 3.1.2. h 1 .

In this case, let α = 3 and 0 < r ( q 3 ) δ G ( x ) , for all x G , then

B ¯ ( x , r ) B ¯ ( x , α r ) B ( x , q δ G ( x ) ) .

Suppose that a and b are any two points in B ¯ ( x , r ) . Let

y ( B ¯ ( x , r ) ) and z G \ ( 3 B ( x , r ) ) ,

then we have

max { a x , b x } x y h x y .

From the assumption of z G \ ( 3 B ( x , r ) ) , it is clear that

x y z y h z y .

For any u B ¯ ( x , r ) , we obtain

(13) y u diam ( B ( x , r ) ) = 2 r

and

(14) z u z x x u 3 r r = 2 r .

By combining (13) and (14), we have

y u z u h z u .

Since f is q -locally weakly ( h , H ) -quasisymmetric mapping for some 0 < q < 1 , using a similar argument as in case 3.1.1, we obtain that

sup 0 < r < r x , β diam ( f ( B ¯ ) ) dist ( f ( B ¯ ) , G \ f ( 3 B ) ) 2 H 2 ( H + 1 ) ,

where β = 3 q .

Thus, as discussion earlier, it follows that f has the ( M , α , β ) -ring property. Here,

M = 2 H 2 ( H + 1 ) , α = max 3 h 2 , 3 and β = max 3 q h 2 , 3 q .

Hence, the proof of implication from (i) to (ii) is complete.

3.2 Proof of the implication from (iii) to (i) in Theorem 1.8

In this subsection, in order to prove the implication from (iii) to (i) in Theorem 1.8, we need the following lemma.

Lemma 3.1

Under the assumptions of Theorem 1.8, suppose that the homeomorphism f : G G is freely φ -quasiconformal mapping. Then, f and f 1 are fully θ -relative with θ depending only on λ , φ , and c .

Proof

Let D G be any subdomain of G and D = f ( D ) . Assume f : G G is freely φ -quasiconformal mapping, i.e., for all x , y D ,

φ 1 ( k D ( x , y ) ) k D ( f ( x ) , f ( y ) ) φ ( k D ( x , y ) ) .

Here, φ : [ 0 , ) [ 0 , ) is a homeomorphism with φ ( t ) t .

In what follows, we only need to verify that the f is fully θ -relative. The proof of fully θ -relative of f 1 follows in a similar manner.

To prove that f is fully θ -relative with θ depending only on λ , φ , and c , we suppose that x , y D with x y = t δ D ( x ) , 0 t < 1 . We have the following claim.

Claim 3.1. For any 0 t < 1 , if x , y D with x y = t δ D ( x ) , then

k D ( x , y ) θ 0 ( t ) ,

where θ 0 : [ 0,1 ) [ 0 , ) is a homeomorphism.

By Lemma 2.4, if 0 < t 1 8 c , then

k D ( x , y ) 2 c x y δ D ( x ) = 2 c t .

So we can assume that 1 8 c < t < 1 in the following. Since X is a c -quasiconvex metric space, by ([16], Observation 2.6), we know that D G is rectifiably connected. Since G X is a λ -John-ball domain, we have every metric open ball B x = B ( x , δ D ( x ) ) in D G is a John ball. Let 0 < λ 1 , from the definition of John ball, it is clear that

dist ( γ s ( l ( γ ) ) , X \ B x ) λ l ( γ ) .

Because B x = { y X : y x = δ D ( x ) } and B x D , it follows that

δ D ( x ) = dist ( x , X \ B x ) dist ( x , X \ D ) = δ D ( x ) .

Therefore, by δ B x ( x ) = dist ( x , X \ B x ) , we obtain that δ B x ( x ) = δ D ( x ) . Furthermore,

(15) l ( γ ) dist ( γ s ( l ( γ ) ) , X \ B x ) λ = dist ( x , X \ B x ) λ = δ D ( x ) λ .

Denote

u 1 sup u [ 0 , l ( γ ) ] : γ s [ 0 , u ] B ¯ y , 1 t 2 δ D ( x ) .

Obviously, if u 1 = l ( γ ) , then we have γ B ¯ y , 1 t 2 δ D ( x ) . We observe, for all u [ 0 , l ( γ ) ] ,

(16) δ D ( γ s ( u ) ) = dist ( γ s ( u ) , X \ D ) dist ( γ s ( u ) , X \ B x ) 1 t 2 δ D ( x ) .

By the definition of quasihyperbolic metric k D ( x , y ) and (15) and (16), it follows that

(17) k D ( x , y ) 0 l ( γ ) d u δ D ( γ s ( u ) ) l ( γ ) ( ( 1 t ) 2 ) δ D ( x ) 2 λ ( 1 t ) .

Now, we assume that u 1 < l ( γ ) . Denote

u 2 inf { u [ 0 , l ( γ ) ] : γ s [ u , l ( γ ) ] B ¯ ( x , t δ D ( x ) ) } .

Then, for any u [ u 2 , l ( γ ) ] , we have

(18) δ D ( γ s ( u ) ) δ D ( x ) x γ s ( u ) ( 1 t ) δ D ( x ) .

If u 2 u 1 , by combining (15), (16), and (18), we have

(19) k D ( x , y ) 0 l ( γ ) d u δ D ( γ s ( u ) ) 0 u 1 d u δ D ( γ s ( u ) ) + u 1 l ( γ ) d u δ D ( γ s ( u ) ) u 1 ( ( 1 t ) 2 ) δ D ( x ) + l ( γ ) u 1 ( 1 t ) δ D ( x ) 2 l ( γ ) ( 1 t ) δ D ( x ) 2 λ ( 1 t ) .

If u 2 > u 1 , then for any u [ u 1 , u 2 ] , by the definitions of u 1 and John ball, we have

(20) δ D ( γ s ( u ) ) = dist ( γ s ( u ) , X \ D ) dist ( γ s ( u ) , X \ B x ) λ u λ u 1 λ γ s ( u 1 ) y λ 1 t 2 δ D ( x ) .

Therefore,

(21) k D ( x , y ) 0 l ( γ ) d u δ D ( γ s ( u ) ) = 0 u 1 d u δ D ( γ s ( u ) ) + u 1 u 2 d u δ D ( γ s ( u ) ) + u 2 l ( γ ) d u δ D ( γ s ( u ) ) u 1 1 t 2 δ D ( x ) + u 2 u 1 λ 1 t 2 δ D ( x ) + l ( γ ) u 2 ( 1 t ) δ D ( x ) = ( 2 λ 2 ) u 1 + ( 2 λ ) u 2 + λ l ( γ ) λ ( 1 t ) δ D ( x ) ( since u 1 < u 2 l ( γ ) ) ( 2 λ ) u 2 + λ l ( γ ) λ ( 1 t ) δ D ( x ) ( since 0 < λ 1 ) 2 + λ λ 2 ( 1 t ) .

Therefore, remembering that 1 8 c < t < 1 , (17), (19), and (21) imply that

k D ( x , y ) 2 + λ λ 2 ( 1 t ) .

We consider function θ 0 ( t ) in the following:

θ 0 ( t ) = 64 ( 2 + λ ) c 2 t λ 2 ( 8 c 1 ) , for 0 < t 1 8 c , 2 + λ λ 2 ( 1 t ) , for 1 8 c < t < 1 .

Then, it follows that

k D ( x , y ) θ 0 ( t ) ,

for any x , y D and 0 t < 1 . Hence, the statements in Claim 3.1 are proved.

We are now turning to the proof of Lemma 3.1.

The proof of Lemma 3.1

For x , y D with x y = t δ D ( x ) . By applying the inequality (i) of Lemma 2.3, we have

(22) f ( x ) f ( y ) δ D ( f ( x ) ) e k D ( f ( x ) , f ( y ) ) 1 = ψ ( k D ( f ( x ) , f ( y ) ) ) .

Here, ψ : [ 0 , ) [ 0 , ) is a homeomorphism by ψ ( t ) = e t 1 .

Since f is freely φ -quasiconformal mapping, from the conclusion of Claim 3.1 and inequality (22), it follows that

f ( x ) f ( y ) δ D ( f ( x ) ) ψ ( k D ( f ( x ) , f ( y ) ) ) ψ ( φ ( k D ( x , y ) ) ) ψ φ θ 0 x y δ D ( x ) ,

which implies that the map f is θ -relative, where θ ( t ) = ψ ( φ ( θ 0 ( t ) ) ) is a homeomorphism from [ 0,1 ) to [ 0 , ) and θ depends only on λ , φ , and c . This completes the proof of Lemma 3.1.□

The proof of Theorem 1.8

Suppose that the homeomorphism f : G G is freely φ -quasiconformal mapping; by Lemma 3.1, f and f 1 are fully θ -relative, with θ depending only on λ , φ , and c .

In what follows, we only need to verify the f is q -locally weakly ( h , H ) -quasisymmetric mapping. The proof of q -locally weakly ( h , H ) -quasisymmetric mapping of f 1 follows in a similar manner.

Let z G , 0 < q < 1 . For x , a , b B ( z , q δ G ( z ) ) with a x = h b x and in order to prove that f is q -locally weakly ( h , H ) -quasisymmetric mapping, we need only to show that

(23) f ( a ) f ( x ) H f ( b ) f ( x ) ,

where H 1 .

We divide the discussions into three cases:

Case 3.2.1. 0 < h 2 c 2 c + 1 .

Set D = G \ { b } , then f ( D ) = G \ { f ( b ) } , and let D = f ( D ) . Since G is a non-cut-point domain, D is a subdomain. So the assumption on the fully θ -relative of f implies f D is θ -relative. Note that θ : [ 0 , 1 ) [ 0 , ) is a homeomorphism. Without loss of generality, we assume that

θ 2 c 2 c + 1 1 .

Assume that x , a , b B ( z , q δ G ( z ) ) with a x = h b x . For any 0 < q < 1 3 , it is clear that

δ G ( x ) δ G ( z ) x z δ G ( z ) q δ G ( z ) > 2 q δ G ( z ) x b .

Therefore, we have

a x = h b x h δ D ( x ) .

From the definition of θ -relative, it follows that

(24) f ( a ) f ( x ) θ ( h ) δ D ( f ( x ) ) θ ( h ) f ( b ) f ( x ) H 1 f ( b ) f ( x ) ,

where

H 1 = θ 2 c 2 c + 1 .

Case 3.2.2. 2 c 2 c + 1 < h 1 .

Let 0 < q 1 4 c + 3 . For each point x , a , b B ( z , q δ G ( z ) ) with a x = h b x , since X is c -quasiconvex metric space, by Lemma 2.2, there is a rectifiable curve γ : [ u , v ] G joining x and a with γ B ( z , ( 2 c + 1 ) q δ G ( z ) ) such that l ( γ ) c x a .

Define inductively the successive points x = z 0 , z 1 , , z n 1 , z n = a of γ as follows. Let t 0 = u ,

t j = sup t t [ u , v ] : γ ( t ) z j 1 2 c 2 c + 1 j x b , 1 j n ,

and z j = γ ( t j ) , 0 j n .

Furthermore, since 2 c 2 c + 1 < h 1 , for 1 j n 1 , we have n 2 ,

z j 1 z j = 2 c 2 c + 1 j x b and z n 1 z n 2 c 2 c + 1 n x b .

Therefore, for 1 j n 1 , it is clear that

(25) l ( γ [ z j 1 , z j ] ) z j 1 z j = 2 c 2 c + 1 j x b .

By summing (25) over 1 j n 1 , we have

j = 1 n 1 2 c 2 c + 1 j x b l ( γ ) c x a = c h x b c x b .

Hence, we obtain that n k with

(26) k ln 2 ln ( 2 c + 1 ) ln ( 2 c ) + 1 .

For each 1 j n 1 , since x , b B ( z , q δ G ( z ) ) , we see that

(27) z j z j 1 = 2 c 2 c + 1 j x b x b 2 q δ G ( z )

and

(28) i = 0 j 1 z i z i + 1 = i = 0 j 1 2 c 2 c + 1 i + 1 x b 2 c x b 4 c q δ G ( z ) .

Combining (27) and (28), we have

(29) δ G ( z j ) z j z j 1 δ G ( z 0 ) i = 0 j 1 z i z i + 1 z j z j 1 δ G ( z ) z 0 z i = 0 j 1 z i z i + 1 z j z j 1 δ G ( z ) q δ G ( z ) 4 c q δ G ( z ) 2 q δ G ( z ) = ( 1 ( 4 c + 3 ) q ) δ G ( z ) 0 .

Set D j = G \ { z j } . Since G is a non-point-cut domain, D j is a subdomain. So f ( D j ) = G \ { f ( z j ) } is also a subdomain. Let D j = f ( D j ) for 1 j n 1 ; according to (29), we obtain

z j z j + 1 2 c 2 c + 1 z j 1 z j = 2 c 2 c + 1 δ D j 1 ( z j ) < δ D j 1 ( z j ) .

Since f D j is θ -relative and z 1 z 0 = 2 c 2 c + 1 x b , by Case 3.2.1, it follows that

(30) f ( z 1 ) f ( z 0 ) θ 2 c 2 c + 1 f ( b ) f ( x ) = H 1 f ( b ) f ( x ) ; f ( z 2 ) f ( z 1 ) H 1 δ D 0 ( f ( z 1 ) ) H 1 f ( z 1 ) f ( z 0 ) H 1 2 f ( b ) f ( x ) ; f ( z n ) f ( z n 1 ) H 1 δ D n 2 ( f ( z n 1 ) ) H 1 n f ( b ) f ( x ) .

Summation of aforementioned formulas gives

f ( x ) f ( a ) ( H 1 + H 1 2 + + H 1 n ) f ( b ) f ( x ) n H 1 n f ( b ) f ( x ) .

Combining this estimate with (26), it follows immediately that

(31) f ( x ) f ( a ) H 2 f ( b ) f ( x ) ,

where

H 2 = k θ 2 c 2 c + 1 k and k = ln 2 ln ( 2 c + 1 ) ln ( 2 c ) + 1 .

Case 3.2.3. h > 1 .

For any x , a , b B z , 1 ( 2 c + 1 ) ( 4 c + 3 ) δ G ( z ) . Since X is a c -quasiconvex metric space, by Lemma 2.2, there is a rectifiable curve γ 1 : [ u , v ] G joining x and a with γ 1 B z , 1 4 c + 3 δ G ( z ) such that l ( γ 1 ) c x a .

Define inductively the successive points x = p 0 , p 1 , , p m 1 , p m = a of γ 1 as follows. Let t 0 = u ,

t i = sup t { t [ u , v ] : γ 1 ( t ) p i 1 x b , 1 i m } ,

and p i = γ 1 ( t i ) , 0 i m .

Moreover, since h > 1 , for 1 i m 1 , we have m 2 ,

p i 1 p i = x b and p m 1 p m x b .

Note that for 1 i m 1 ,

(32) l ( γ 1 [ p i 1 , p i ] ) p i 1 p i = x b .

By summing (32) over 1 i m 1 , we have

(33) ( m 1 ) x b i = 1 m 1 l ( γ 1 [ p i 1 , p i ] ) l ( γ 1 ) c x a = c h x b .

According to (32) and (33), it follows that

(34) m 1 + c h .

Note that x = p 0 , p 1 , , p m = a and b are contained in B z , 1 4 c + 3 δ G ( z ) . Using a similar argument as in Case 3.2.2, we obtain

(35) f ( a ) f ( x ) i = 1 m f ( p i ) f ( p i 1 ) m H 2 f ( x ) f ( b ) .

Combining (34) and (35), we have the desired estimate

(36) f ( a ) f ( x ) H 3 f ( x ) f ( b ) ,

where

H 3 = ( 1 + c h ) H 2 .

Therefore, Case 3.2.3 is completed.

Thus, in terms of (24), (31), and (36), for any x , a , b B z , 1 ( 2 c + 1 ) ( 4 c + 3 ) δ G ( z ) with x a = h b x , we obtain that

f ( a ) f ( x ) H f ( x ) f ( b ) .

Here,

H = max { H 1 , H 2 , H 3 } = ( 1 + c h ) k θ 2 c 2 c + 1 k ,

where

k = ln 2 ln ( 2 c + 1 ) ln ( 2 c ) + 1 .

Therefore, f is q -locally weakly ( h , H ) -quasisymmetric mapping with h and H depending only on θ and c . Here,

q = 1 ( 2 c + 1 ) ( 4 c + 3 ) .

Hence, the proof of implication from (iii) to (i) is complete.

4 Proof of Theorem 1.9

In this section, we assume that X , Y , and Z are three c -quasiconvex, complete metric spaces and that G 1 X , G 2 Y , and G 3 Z are three λ -John-ball and non-cut-point domains. Furthermore, we suppose that f : G 1 G 2 and f 1 : G 2 G 1 are q 1 -locally weakly ( h 1 , H 1 ) -quasisymmetric mappings; g : G 2 G 3 and g 1 : G 3 G 2 are q 2 -locally weakly ( h 2 , H 2 ) -quasisymmetric mappings.

Lemma 4.1

Under the assumptions of Theorem 1.9, if f : G 1 G 2 is a freely φ 1 -quasiconformal mapping and g : G 2 G 3 is a freely φ 2 -quasiconformal mapping, then, the composition g f : G 1 G 3 is a freely φ 2 φ 1 -quasiconformal mapping.

Proof

Let D G 1 be any subdomain of G 1 and D = f ( D ) G 2 , D = g f ( D ) G 3 . In what follows, we only need to verify that g f : G 1 G 3 is fully semisolid mapping, i.e.,

k D ( g f ( x ) , g f ( y ) ) φ 2 φ 1 ( k D ( x , y ) ) .

The proof of ( g f ) 1 is fully semisolid mapping that follows in a similar manner.

For any domain D G 1 , by the assumption, we see that f : D D is φ 1 -semisolid mapping and g : D D is φ 2 -semisoild mapping, which shows that

k D ( f ( x ) , f ( y ) ) φ 1 ( k D ( x , y ) )

and

k D ( g f ( x ) , g f ( y ) ) φ 2 ( k D ( f ( x ) , f ( y ) ) ) ,

for all x , y D . Therefore, we obtain that

k D ( g f ( x ) , g f ( y ) ) φ 2 φ 1 ( k D ( x , y ) ) .

Hence, the proof of Lemma 4.1 is complete.□

The proof of Theorem 1.9

Under the assumptions of Theorem 1.9, we see that f and g are freely φ i -quasiconformal mappings with some homeomorphisms φ i , i = 1, 2 . By Lemma 4.1, the composition g f : G 1 G 3 is freely φ 2 φ 1 -quasiconformal mapping. It follows from Theorem 1.8 that there exist constants q , h , and H such that g f and ( g f ) 1 are q -locally weakly ( h , H ) -quasisymmetric mappings. Hence, Theorem 1.9 is proved.

Acknowledgements

The authors thank the referee for their careful reading and valuable comments that led to the improvement of this article.

  1. Funding information: This work was supported by the National Natural Science Foundation of China (Grant No. 12461012) and the Guizhou Province Science and Technology Foundation (Grant No. QianKeHeJiChu[2020]1Y003).

  2. Author contributions: All authors have accepted responsibility for the entire content of this manuscript and consented to its submission to the journal, reviewed all the results, and approved the final version of the manuscript. All authors contributed to the study conception and design.

  3. Conflict of interest: The authors state no conflict of interest.

  4. Data availability statement: No datasets were generated or analyzed during the current study.

References

[1] F. W. Gehring and B. G. Osgood, Uniform domains and the quasi-hyperbolic metric, J. Anal. Math. 36 (1979), 50–74. 10.1007/BF02798768Search in Google Scholar

[2] F. W. Gehring and B. P. Palka, Quasiconformally homogeneous domains, J. Anal. Math. 30 (1976), 172–199. 10.1007/BF02786713Search in Google Scholar

[3] J. Väisälä, The free quasiworld: freely quasiconformal and related maps in Banach spaces, Banach Center Publ. 48 (1999), 55–118. 10.4064/-48-1-55-118Search in Google Scholar

[4] A. Beurling and L. Ahlfors, The boundary correspondence under quasiconformal mappings, Acta Math. 96 (1956), 125–142. 10.1007/BF02392360Search in Google Scholar

[5] P. Tukia and J. Väisälä, Quasisymmetric embeddings of metric spaces, Ann. Fenn. Math. 5 (1980), no. 1, 97–114. 10.5186/aasfm.1980.0531Search in Google Scholar

[6] J. Väisälä, Free quasiconformality in Banach spaces, I, Ann. Fenn. Math. 15 (1990), no. 2, 355–379. 10.5186/aasfm.1990.1527Search in Google Scholar

[7] J. Väisälä, Free quasiconformality in Banach spaces, II, Ann. Fenn. Math. 16 (1991), no. 2, 255–310. 10.5186/aasfm.1991.1629Search in Google Scholar

[8] J. Väisälä, Free quasiconformality in Banach spaces, III, Ann. Fenn. Math. 17 (1992), no. 2, 393–408. 10.5186/aasfm.1992.1732Search in Google Scholar

[9] J. Väisälä, Free quasiconformality in Banach spaces, IV, in: Analysis and Topology, World Scientific Publishing, River Edge, N.J., 1998, pp. 697–717. 10.1142/9789812817297_0040Search in Google Scholar

[10] J. Heinonen and P. Koskela, Quasiconformal maps in metric spaces with controlled geometry, Acta Math. 181 (1998), no. 1, 1–61. 10.1007/BF02392747Search in Google Scholar

[11] J. Heinonen, Lectures on Analysis on Metric Spaces, Springer-Verlag, New York, 2001. 10.1007/978-1-4613-0131-8Search in Google Scholar

[12] J. Heinonen and P. Koskela, Definitions of quasiconformality, Invent. Math. 120 (1995), no. 1, 61–79. 10.1007/BF01241122Search in Google Scholar

[13] J. Heinonen, P. Koskela, N. Shanmugalingam, and J. T. Tyson, Sobolev classes of Banach space-valued functions and quasiconformal mappings, J. Anal. Math. 85 (2001), 87–139. 10.1007/BF02788076Search in Google Scholar

[14] M. Huang, A. Rasila, X. Wang, and Q. Zhou, Semisolidity and locally weak quasisymmetry of homeomorphisms in metric spaces, Studia Math. 242 (2018), 267–301. 10.4064/sm8629-6-2017Search in Google Scholar

[15] X. Huang, H. Liu, and J. Liu, Local properties of quasihyperbolic mappings in metric spaces, Ann. Fenn. Math. 41 (2016), 23–40. 10.5186/aasfm.2016.4106Search in Google Scholar

[16] X. Huang and J. Liu, Quasihyperbolic metric and Quasisymmetric mappings in metric spaces, Trans. Amer. Math. Soc. 367 (2015), no. 9, 6225–6246. 10.1090/S0002-9947-2015-06240-0Search in Google Scholar

[17] P. Koskela and K. Wildrick, Analytic properties of quasiconformal mappings between metric spaces, in: X. Dai, X. Rong (eds.), Metric and Differential Geometry, Progress in Mathematics, vol. 297, Birkhäuser, Basel, 2012, pp. 163–174. 10.1007/978-3-0348-0257-4_6Search in Google Scholar

[18] J. Tyson, Quasiconformality and quasisymmetry in metric measure spaces, Ann. Fenn. Math. 23 (1998), 525–548. Search in Google Scholar

[19] J. Tyson, Metric and geometric quasiconformality in Ahlfors regular Loewner spaces, Conform. Geom. Dyn. 5 (2001), 21–73. 10.1090/S1088-4173-01-00064-9Search in Google Scholar

[20] J. Väisälä, Quasisymmetric embeddings in Euclidean spaces, Trans. Amer. Math. Soc. 264 (1981), no. 1, 191–204. 10.1090/S0002-9947-1981-0597876-7Search in Google Scholar

[21] X. Wang and Q. Zhou, Quasimöbius maps, weakly Quasimöbius maps and uniform perfectiness in quasi-metric spaces, Ann. Fenn. Math. 42 (2017), 257–284. 10.5186/aasfm.2017.4216Search in Google Scholar

[22] Q. Zhou, Y. Li, and A. Rasila, A note on Väisäläas problem concerning free quasiconformal mappings, Monatshefte für Mathematik 196 (2021), 607–616. 10.1007/s00605-021-01554-4Search in Google Scholar

[23] F. W. Gehring, K. Hag, and O. Martio, Quasihyperbolic geodesics in John domains, Math. Scand. 65 (1989), 75–92. 10.7146/math.scand.a-12267Search in Google Scholar

[24] P. Koskela and J. Lehrbäck, Quasihyperbolic growth conditions and compact embeddings of Sobolev spaces, Michigan Math. J. 55 (2007), 183–193. 10.1307/mmj/1177681992Search in Google Scholar

[25] H. Liu and X. Huang, The properties of quasisymmetric mappings in metric spaces, J. Math. Anal. Appl. 435 (2016), 1591–1606. 10.1016/j.jmaa.2015.11.035Search in Google Scholar

[26] J. Väisälä, Quasiconformal maps of cylindrical domains, Acta Math. 162 (1989), no. 3–4, 201–225. 10.1007/BF02392837Search in Google Scholar

Received: 2024-08-28
Revised: 2024-12-21
Accepted: 2025-01-21
Published Online: 2025-04-16

© 2025 the author(s), published by De Gruyter

This work is licensed under the Creative Commons Attribution 4.0 International License.

Articles in the same Issue

  1. Special Issue on Contemporary Developments in Graphs Defined on Algebraic Structures
  2. Forbidden subgraphs of TI-power graphs of finite groups
  3. Finite group with some c#-normal and S-quasinormally embedded subgroups
  4. Classifying cubic symmetric graphs of order 88p and 88p 2
  5. Simplicial complexes defined on groups
  6. Two-sided zero-divisor graphs of orientation-preserving and order-decreasing transformation semigroups
  7. Further results on permanents of Laplacian matrices of trees
  8. Special Issue on Convex Analysis and Applications - Part II
  9. A generalized fixed-point theorem for set-valued mappings in b-metric spaces
  10. Research Articles
  11. Dynamics of particulate emissions in the presence of autonomous vehicles
  12. The regularity of solutions to the Lp Gauss image problem
  13. Exploring homotopy with hyperspherical tracking to find complex roots with application to electrical circuits
  14. The ill-posedness of the (non-)periodic traveling wave solution for the deformed continuous Heisenberg spin equation
  15. Some results on value distribution concerning Hayman's alternative
  16. 𝕮-inverse of graphs and mixed graphs
  17. A note on the global existence and boundedness of an N-dimensional parabolic-elliptic predator-prey system with indirect pursuit-evasion interaction
  18. On a question of permutation groups acting on the power set
  19. Chebyshev polynomials of the first kind and the univariate Lommel function: Integral representations
  20. Blow-up of solutions for Euler-Bernoulli equation with nonlinear time delay
  21. Spectrum boundary domination of semiregularities in Banach algebras
  22. Statistical inference and data analysis of the record-based transmuted Burr X model
  23. A modified predictor–corrector scheme with graded mesh for numerical solutions of nonlinear Ψ-caputo fractional-order systems
  24. Dynamical properties of two-diffusion SIR epidemic model with Markovian switching
  25. Classes of modules closed under projective covers
  26. On the dimension of the algebraic sum of subspaces
  27. Periodic or homoclinic orbit bifurcated from a heteroclinic loop for high-dimensional systems
  28. On tangent bundles of Walker four-manifolds
  29. Regularity of weak solutions to the 3D stationary tropical climate model
  30. A new result for entire functions and their shifts with two shared values
  31. Freely quasiconformal and locally weakly quasisymmetric mappings in metric spaces
  32. On the spectral radius and energy of the degree distance matrix of a connected graph
  33. Solving the quartic by conics
  34. A topology related to implication and upsets on a bounded BCK-algebra
  35. On a subclass of multivalent functions defined by generalized multiplier transformation
  36. Local minimizers for the NLS equation with localized nonlinearity on noncompact metric graphs
  37. Approximate multi-Cauchy mappings on certain groupoids
  38. Multiple solutions for a class of fourth-order elliptic equations with critical growth
  39. A note on weighted measure-theoretic pressure
  40. Majorization-type inequalities for (m, M, ψ)-convex functions with applications
  41. Recurrence for probabilistic extension of Dowling polynomials
  42. Unraveling chaos: A topological analysis of simplicial homology groups and their foldings
  43. Global existence and blow-up of solutions to pseudo-parabolic equation for Baouendi-Grushin operator
  44. A characterization of the translational hull of a weakly type B semigroup with E-properties
  45. Some new bounds on resolvent energy of a graph
  46. Carmichael numbers composed of Piatetski-Shapiro primes in Beatty sequences
  47. The number of rational points of some classes of algebraic varieties over finite fields
  48. Singular direction of meromorphic functions with finite logarithmic order
  49. Pullback attractors for a class of second-order delay evolution equations with dispersive and dissipative terms on unbounded domain
  50. Eigenfunctions on an infinite Schrödinger network
  51. Boundedness of fractional sublinear operators on weighted grand Herz-Morrey spaces with variable exponents
  52. On SI2-convergence in T0-spaces
  53. Bubbles clustered inside for almost-critical problems
  54. Classification and irreducibility of a class of integer polynomials
  55. Existence and multiplicity of positive solutions for multiparameter periodic systems
  56. Averaging method in optimal control problems for integro-differential equations
  57. On superstability of derivations in Banach algebras
  58. Investigating the modified UO-iteration process in Banach spaces by a digraph
  59. The evaluation of a definite integral by the method of brackets illustrating its flexibility
  60. Existence of positive periodic solutions for evolution equations with delay in ordered Banach spaces
  61. Tilings, sub-tilings, and spectral sets on p-adic space
  62. The higher mapping cone axiom
  63. Continuity and essential norm of operators defined by infinite tridiagonal matrices in weighted Orlicz and l spaces
  64. A family of commuting contraction semigroups on l 1 ( N ) and l ( N )
  65. Pullback attractor of the 2D non-autonomous magneto-micropolar fluid equations
  66. Maximal function and generalized fractional integral operators on the weighted Orlicz-Lorentz-Morrey spaces
  67. On a nonlinear boundary value problems with impulse action
  68. Normalized ground-states for the Sobolev critical Kirchhoff equation with at least mass critical growth
  69. Decompositions of the extended Selberg class functions
  70. Subharmonic functions and associated measures in ℝn
  71. Some new Fejér type inequalities for (h, g; α - m)-convex functions
  72. The robust isolated calmness of spectral norm regularized convex matrix optimization problems
  73. Multiple positive solutions to a p-Kirchhoff equation with logarithmic terms and concave terms
  74. Joint approximation of analytic functions by the shifts of Hurwitz zeta-functions in short intervals
  75. Green's graphs of a semigroup
  76. Some new Hermite-Hadamard type inequalities for product of strongly h-convex functions on ellipsoids and balls
  77. Infinitely many solutions for a class of Kirchhoff-type equations
  78. On an uncertainty principle for small index subgroups of finite fields
Downloaded on 5.9.2025 from https://www.degruyterbrill.com/document/doi/10.1515/math-2025-0131/html
Scroll to top button