796
Views
7
CrossRef citations to date
0
Altmetric
Articles

Triangular ratio metric in the unit disk

ORCID Icon & ORCID Icon
Pages 1299-1325 | Received 01 Sep 2020, Accepted 28 Dec 2020, Published online: 14 Jan 2021

Abstract

The triangular ratio metric is studied in a domain GRn, n2. Several sharp bounds are proven for this metric, especially in the case where the domain is the unit disk of the complex plane. The results are applied to study the Hölder continuity of quasiconformal mappings.

AMS Subject Classifications:

1. Introduction

In geometric function theory, metrics are often used to define new types of geometries of subdomains of the Euclidean, Hilbert, Banach and other metric spaces [Citation1–4]. One can introduce a metric topology and build new types of geometries of a domain GRn, n2, based on metrics. Since the local behaviour of functions defined on G is an important area of study, it is natural to require that, given a point in G, a metric recognises points close to it from the boundary G.

Thus, certain constraints on metrics are necessary. A natural requirement is that the distance defined by a metric for given two points x,yG takes into account both how far the points are from each other and also their location with respect to the boundary. Indeed, we require that the closures of the balls defined by the metrics do not intersect the boundary G of the domain. We call these types of metrics intrinsic metrics. A generic example of an intrinsic metric is the hyperbolic metric [Citation5] of a planar domain or its generalisation, the quasihyperbolic metric [Citation6] defined in all proper subdomains of GRn, n2.

In the recent years, new kinds of intrinsic geometries have been introduced by numerous authors, see Ref. [Citation7, pp. 18–19]. Papadopoulos lists in Ref. [Citation8, pp. 42–48] 12 metrics recurrent in function theory. Because there are differences how these metrics catch certain intricate features of functions, using several metrics is often imperative. We might further specify the properties of the metrics by requiring that the intrinsic metric should be compatible with the function classes studied. For instance, some kind of quasi-invariance property is often valuable. Recall that the hyperbolic metric of a planar domain G is invariant under conformal automorphisms of G.

In 2002, Hästö [Citation9] introduced the triangular ratio metric, defined in a domain GRn as the function sG:G×G[0,1] sG(x,y)=|xy|infzG(|xz|+|zy|).

This metric was studied recently in Refs. [Citation10–12], and our goal here is to continue this investigation. We introduce new methods for estimating the triangular ratio metric in terms of several other metrics and establish several results with sharp constants.

In order to compute the value of the triangular ratio metric between points x and y in a domain G, we must find a point z on the boundary of G that gives the infimum for the sum |xz|+|zy|. This is a very simple task if the domain is, for instance, a half-plane or a polygon, but solving the triangular ratio distance in the unit disk is a complicated problem with a very long history (see Ref. [Citation11]). However, there are two special cases where this problem becomes trivial: if the points x and y in the unit disk are collinear with the origin or at the same distance from the origin, there are explicit formulas for the triangular ratio metric.

Since the points x and y can be always rotated around their midpoint to end up into one of these two special cases, we can estimate the value of the triangular ratio metric, regardless of how the original points are located in the unit disk. This rotation can be done either by using Euclidean or hyperbolic geometry, and the main result of this article is to prove that both these ways give lower and upper limits for the value of the triangular ratio metric. Note that while we study the midpoint rotation only in the two-dimensional disk, our results can be directly extended into the case Bn, n3, for the point z giving the infimum is always on the same two-dimensional disk as x, y and the origin.

The structure of this article is as follows. First, we show a few simple ways to find bounds for the triangular ratio metric in Section 3. We define the Euclidean midpoint rotation and prove the inequalities related to it in Section 4 and then do the same for the hyperbolic midpoint rotation in Section 5. Finally, in Section 6, we explain how finding better bounds for the triangular ratio metric can be useful for studying K-quasiconformal mappings in the unit disk.

2. Preliminaries

Let G be some non-empty, open, proper and connected subset of Rn. For all xG, dG(x) is the Euclidean distance d(x,G)=inf{|xz||zG}. Other than the triangular ratio metric defined earlier, we will need the following hyperbolic type metrics:

The jG-metric jG:G×G[0,1], jG(x,y)=|xy||xy|+2min{dG(x),dG(y)}, the point pair function pG:G×G[0,1], pG(x,y)=|xy||xy|2+4dG(x)dG(y) and the Barrlund metric bG,p:G×G[0,), bG,p(x,y)=supzG|xy|(|xz|p+|zy|p)1p. Note that the function pG is not a metric in all domains [Citation10, Remark 3.1, p. 689].

The hyperbolic metric is defined as chρHn(x,y)=1+|xy|22dHn(x)dHn(y),x,yHn,sh2ρBn(x,y)2=|xy|2(1|x|2)(1|y|2),x,yBn, in the upper half-plane Hn and in the Poincaré unit ball Bn, respectively [Citation7, (4.8), p. 52; (4.14), p. 55]. In the two-dimensional unit disk thρB2(x,y)2=th(12log(|1xy¯|+|xy||1xy¯||xy|))=|xy1xy¯|=|xy|A[x,y], where y¯ is the complex conjugate of y and A[x,y]=|xy|2+(1|x|2)(1|y|2) is the Ahlfors bracket [Citation7, (3.17) p. 39]. The hyperbolic segment between points x and y is denoted by J[x,y], while Euclidean lines, line segments, balls and spheres are written in forms L(x,y), [x,y], Bn(x,r) and Sn1(x,r), respectively, just like in Ref. [Citation7, pp. vii–xi]. Note that if the centre x or the radius r is not specified in the notations Bn(x,r) and Sn1(x,r), it means that x = 0 and r = 1. The hyperbolic ball is denoted by Bρn(q,R), as in the following lemma.

Lemma 2.1

[Citation7, (4.20) p. 56]

The equality Bρn(q,R)=Bn(j,h) holds, if j=q(1t2)1|q|2t2,h=(1|q|2)t1|q|2t2andt=th(R2).

For the results of Section 5, the formula for the hyperbolic midpoint is needed.

Theorem 2.2

[Citation13, Theorem 1.4, p. 3]

For all x,yB2, the hyperbolic midpoint q of J[x,y] with ρB2(x,q)=ρB2(q,y)=ρB2(x,y)2 is given by q=y(1|x|2)+x(1|y|2)1|x|2|y|2+A[x,y](1|x|2)(1|y|2).

Furthermore, the next results will be useful when studying the triangular ratio metric in the unit disk.

Theorem 2.3

[Citation7, p. 460]

For all x,yBn, thρBn(x,y)4jBn(x,y)sBn(x,y)pBn(x,y)thρBn(x,y)22thρBn(x,y)4.

Theorem 2.4

[Citation11, p. 138]

For all x,yBn, the radius drawn to the point z giving the infimum infzSn1(|xz|+|zy|) bisects the angle XZY.

Lemma 2.5

[Citation7, 11.2.1(1), p. 205]

For all x,yBn, sBn(x,y)|xy|2|x+y|, where the equality holds if the points x, y are collinear with the origin.

Theorem 2.6

[Citation14, Theorem 3.1, p. 276]

If x=h+kiB2 with h, k>0, then sB2(x,x¯)=|x| if |x12|>12,sB2(x,x¯)=k(1h)2+k2|x|otherwise.

Remark 2.7

If x,yBn such that |x|=|y| and there is only one point zS giving the infimum infzSn1(|xz|+|zy|), then it can be verified with Theorem 2.6 that z=(x+y)|x+y|.

3. Bounds for triangular ratio metric

In this section, we will introduce a few different upper and lower bounds for the triangular ratio metric in the unit disk B2, using the Barrlund metric and a special lower limit function. There are numerous similar results already in the literature, but we complement them and prove that our inequalities are sharp by showing that they have the best possible constant. First, we introduce the following inequality:

Lemma 3.1

For all yG, the inequality sG(x,y)|xy|dG(x)+|xy|2+dG(x)22dG(x)|xy|2dG(y)2 holds, if the domain G is starlike with respect to xG and dG(x)+dG(y)|xy|.

Proof.

Let G be starlike with respect to xG and consider an arbitrary point yG. Clearly, Bn(x,dG(x)),Bn(y,dG(y))G. It also follows from the starlikeness of G that the convex hull uBn(y,dG(y))[x,u] must belong to G. Fix u,vSn1(y,dG(y)), uv, on the same plane with the points x, y so that the lines L(x,u) and L(y,v) are tangents of Sn1(y,dG(y)), and fix z1Sn1(x,dG(x))[x,u].

By the starlikeness of G, sBn(y,dG(y))[x,s]G, so it follows that z1 fulfils |xz1|+|z1y|infzG(|xz|+|zy|)sG(x,y)|xy||xz1|+|z1y|. Here, |xz1|=dG(x) and, with the information that |uy|=|yv|=dG(y) and XUY=YVX=π2, we can conclude that |z1y|=|xy|2+dG(x)22dG(x)|xy|2dG(y)2. Thus, the lemma follows.

Remark 3.2

The same method as in the proof of Lemma 3.1 can be also applied into the case where G is convex. In that case, J=sBn(x,dG(x)),tBn(y,dG(y))[s,t]G for all x,yG, so sG(x,y)|xy||xz1|+|z1y|, where z1 is chosen from J so that |xz1|+|z1y| is at minimum. By finding the value of this sum, we end up with the result sG(x,y)pG(x,y), which holds by Ref. [Citation7, Lemma 11.6(1), p. 197].

Let us now focus on the Barrlund metric.

Lemma 3.3

[Citation15, Theorem 3.6, p. 7]

For all x,yGRn, sG(x,y)bG,p(x,y)211psG(x,y).

Theorem 3.4

[Citation15, Theorem 3.15, p. 11]

For all x,yB2, bB2,2(x,y)=|xy|2+|x|2+|y|22|x+y|.

Lemma 3.5

For all x,yB2, 12bB2,2(x,y)sB2(x,y)bB2,2(x,y). Furthermore, this inequality is sharp.

Proof.

The inequality follows from Lemma 3.3. Let x = 0 and y = k with 0<k<1. By Lemma 2.5 and Theorem 3.4, sB2(x,y)=k2kandbB2,2(x,y)=k2+k22k, so we will have the following limit values limk0+sB2(x,y)bB2,2(x,y)=limk0+(2+k22k2k)=12andlimk1sB2(x,y)bB2,2(x,y)=1. Thus, the sharpness follows.

Let us next study the connection between the Barrlund metric and two other hyperbolic type metrics that can be used to bound the value of the triangular ratio metric in the unit disk, see Theorem 2.3.

Theorem 3.6

For all x,yB2, the sharp inequality 12bB2,2(x,y)jB2(x,y)bB2,2(x,y) holds.

Proof.

The inequality follows from Lemma 3.5, Theorem 2.3 and Ref. [Citation12, Theorem 2.9(1), p. 1129]. By Theorem 3.4, jB2(x,y)bB2,2(x,y)=2+|x|2+|y|22|x+y||xy|+2|x||y|. For x = 0 and y = k with 0<k<1, limk1jB2(x,y)bB2,2(x,y)=limk1(2+k22k2)=12, and for x = −k and y = k with 0<k<1, limk1jB2(x,y)bB2,2(x,y)=limk1(1+k22)=1. Thus, the sharpness follows.

Theorem 3.7

For all x,yB2, the sharp inequality 12bB2,2(x,y)pB2(x,y)10+24bB2,2(x,y) holds.

Proof.

Consider now the quotient (1) pB2(x,y)bB2,2(x,y)=2+|x|2+|y|22|x+y||xy|2+4(1|x|)(1|y|).(1) By Lemma 3.5 and Ref. [Citation7, 11.16(1), p. 203], bB2,2(x,y)2sB2(x,y)pB2(x,y) holds for all x,yB2. This inequality is sharp, because, for x = 0 and y = k, limk0+pB2(x,y)bB2,2(x,y)=limk0+(k2+2k+22k)=12. Without loss of generality, fix x = h and y=jeμi with 0hj<1 and 0<μ<2π. The quotient (Equation1) is now pB2(x,y)bB2,2(x,y)=2+h2+j22h2+j2+2hjcos(μ)h2+j22hjcos(μ)+4(1h)(1j). This is decreasing with respect to cos(μ), so we can assume that μ=π and cos(μ)=1, when looking for the maximum of this quotient. It follows that pB2(x,y)bB2,2(x,y)=(1+h)2+(1j)2(h+j)2+4(1h)(1j)=(1+h)2+(1hq)2(2h+q)2+4(1h)(1hq), where q=jh0. The quotient above is clearly decreasing with respect to q. Thus, let us fix j = h. It follows that pB2(x,y)bB2,2(x,y)=2+2h28h28h+4=1+h24h24h+2f(h), where f:[0,1)R, f(h)=(1+h2)(4h24h+2). By differentiation, for 0h<1, f(h)=h(1+h24h24h+2)=(h2+h1)(2h22h+1)2=0h=512. Since f(0.1)>1 and f(0.9)<0, the quotient (Equation1) has a maximum f((51)2)=(10+2)4 and the other part of the theorem follows.

Finally, we will introduce one special function defined in the punctured unit disk.

Definition 3.8

For x,yB2{0}, define low(x,y)=|xy|min{|xy|,|xy|}, where x=x|x|2 and y=y|y|2.

Remark 3.9

The low-function is not a metric on the punctured unit disk: by choosing points x = 0.3, y = −0.1 and z = 0.1, we will have 0.117low(x,y)>low(x,z)+low(z,y)0.0817, so the triangle inequality does not hold.

Furthermore, because A[x,y]=|x||yx| for x,yBn{0}, it follows that (2) thρB2(x,y)2=|xy||x||yx|low(x,y),(2) see Ref. [Citation16, 7.44(20)]. Note also that, by Ref. [Citation16, 7.42(1)], the left-hand side of (Equation2) defines a metric.

This low-function is a suitable lower bound for the triangular ratio metric, as the next theorem states.

Lemma 3.10

For all x,yB2{0}, the inequality sB2(x,y)low(x,y) holds.

Proof.

Suppose that |xy||xy| and fix z1[x,y]S1. Clearly, d(y,S1)<d(y,S1)1|y|<|y|1=1|y|1|y|2+1|y|=1|y|(|y|1)2>0. It follows from this that sB2(x,y)|xy||xz1|+|z1y||xy||xz1|+|z1y|=|xy||xy|=low(x,y).

As a lower bound for sB2(x,y), the low-function is essentially sharp, when max{|x|,|y|}1. However, the low-function does not give any useful upper limits for the triangular ratio metric, unless we limit from below the absolute value of the points inspected. This can be seen in our next theorem.

Theorem 3.11

For all x,yB2{0}, the triangular ratio metric and its lower bound fulfil sup{sB2(x,y)low(x,y)|max{|x|,|y|}r}1+r2r, where the equality holds if max{|x|,|y|}=r.

Proof.

Consider the quotient (3) sB2(x,y)low(x,y)=min{|xy|,|xy|}infzS1(|xz|+|zy|).(3) Fix x,yB2 such that 0<|x||y| and choose zS1 so that it gives the infimum in the denominator of the quotient (Equation3). Let k0=ZOX and k1=ZOY, where the point o is the origin. Note that, by Theorem 2.4, XZO=OZY, so it follows that 0k1k0π2. We can write that infzS1(|xz|+|zy|)=|x|2+12|x|cos(k0)+|y|2+12|y|cos(k1). Furthermore, |xy|=|x|2+1|y|22|x||y|cos(k0+k1),|xy|=|y|2+1|x|22|y||x|cos(k0+k1). Now, we can find an upper bound for the quotient (Equation3): (4) sB2(x,y)low(x,y)|xy|infzS1(|xz|+|zy|)sup0k1k0π2|x|2+1|y|22(|x||y|)cos(k0+k1)|x|2+12|x|cos(k0)+|y|2+12|y|cos(k1)=(inf0k1k0π2|x|2+12|x|cos(k0)+|y|2+12|y|cos(k1)|x|2+1|y|22(|x||y|)cos(k0+k1))1(inf0k1k0π2|x|2+12|x|cos(k0)|x|2+1|y|22(|x||y|)cos(k0+k1)+inf0k1k0π2|y|2+12|y|cos(k1)|x|2+1|y|22(|x||y|)cos(k0+k1))1=(|x|2+12|x||x|2+1|y|22(|x||y|)+|y|2+12|y||x|2+1|y|22(|x||y|))1=(1|x|1|y||x|+1|y|1|y||x|)1=1|y||x|2|x||y|.(4) Let us yet find another upper bound for the quotient (Equation4). It can be shown by differentiation that the function f:(0,1)R, f(|x|)=1|y||x|2|x||y| is increasing. It follows from this that |x||y|f(|x|)f(|y|)1|y||x|2|x||y|1|y||y|2|y||y|=1+|y|2|y|. Thus, for all x,yB2 such that 0<|x||y|, the quotient (Equation3) fulfils the inequality (5) sB2(x,y)low(x,y)1|y||x|2|x||y|1+|y|2|y|.(5) Fix now x=12 and y=12+j with 0<j<12. The quotient (Equation3) is now sB2(x,y)low(x,y)=3+2j(2+4j)(1j)=1+|y|2|y|(1j)=11j1+|y|2|y|, and it has a limit value limj0+sB2(x,y)low(x,y)=1+|y|2|y|. Thus, the inequality (Equation5) is sharp and this result proves that supsB2(x,y)low(x,y)=1+max{|x|,|y|}2max{|x|,|y|}. Since the quotient (1+k)(2k) is decreasing for k(0,1), the theorem follows.

The low-function yields a lower limit also for other hyperbolic type metrics.

Lemma 3.12

For all x,yB2{0}, the following inequalities hold and are sharp:

  1. low(x,y)2jB2(x,y),

  2. low(x,y)pB2(x,y),

  3. low(x,y)bB2,2(x,y).

Furthermore, there is no c>0 such that low(x,y)cd(x,y) for all x,yB2{0}, where d{jB2,pB2,bB2,2}.

Proof.

The inequalities follow from Theorem 2.3, Lemmas 3.5 and 3.10, and Ref. [Citation12, Theorem 2.9(1), p. 1129]. Let 0<k<1. Since limk1low(k,ke2(1k)i)jB2(k,ke2(1k)i)=limk1(2k(ksin(1k)+1k)k4+12k2cos(2(1k)))=2,limk1low(k,k)pB2(k,k)=limk1(2k2k22k+1k2+1)=1,limk1low(k,k)bB2,2(k,k)=limk1(2kk2+1)=1, the inequalities are sharp. The latter part of the lemma follows from the fact that the limit values above are all 0 if k0 instead.

4. Euclidean midpoint rotation

In this section, we introduce the Euclidean midpoint rotation. Finding the value of the triangular ratio distance for two points in the unit disk is a trivial problem, if the points are collinear with the origin or at same distance from it, see Lemma 2.5 and Theorem 2.6. Since any two points can always be rotated around their midpoint into one of these two positions, this transformation gives us a simple way to estimate the value of the triangular ratio metric of the original points.

Definition 4.1

Euclidean midpoint rotation. Choose distinct points x,yB2. Let k=(x+y)2 and r=|xk|=|yk|. Let x0,y0S1(k,r), x0y0, so that |x0|=|y0| and the points x0,k,y0 are collinear. Fix then x1,y1S1(k,r) so that x1,k,y1 are collinear, |x1|=|k|+r and |y1|=|k|r. Note that x0,y0,y1B2 always but x1 is not necessarily in B2. See Figure .

Figure 1. Euclidean midpoint rotation.

Figure 1. Euclidean midpoint rotation.

For all x,yB2, xy, such that x1B2, the inequality sB2(x0,y0)sB2(x,y)sB2(x1,y1) holds, as we will prove in Theorems 4.11 and 4.12. If x1B2, sB2(x1,y1) is not defined but the first part of this inequality holds. In order to prove this result, let us next introduce a few results needed to find the value of sG-diameter of a closed disk in some domain G.

Proposition 4.2

For a fixed point xG and a fixed direction of xy, the value of sG(x,y) is increasing with respect to |xy|.

Proof.

Let x,yG and t[x,y]G. Choose zG so that sG(x,t)=|xt||xz|+|zt|. Because the function f:(0,)R,f(μ)=(u+μ)(v+μ) with constants 0<uv is increasing, sG(x,t)|xt|+|ty||xz|+|zt|+|ty|=|xy||xz|+|zt|+|ty|sG(x,y). Thus, the result follows.

Proposition 4.3

The function f:[0,π2]R, f(μ)=uvcos(μ)+u+vcos(μ), where u, v>0 are constants, is increasing on the interval μ[0,π2].

Proof.

Let s=cos(μ), so that the function f can be written as g:[0,1]R, g(s)=uvs+u+vs. By differentiation, g(s)=v2(1u+vs1uvs)0, and it follows that the function g is decreasing on the interval s[0,1]. Because s=cos(μ) is decreasing, too, with respect to μ, the function f is increasing.

Theorem 4.4

Fix j,r,k,zR such that jk<j+r<z. Choose x,yS1(j,r) so that ZKX=μ with 0μπ2 and k[x,y]. Then the quotient (6) |xy||xz|+|zy|(6) is decreasing with respect to μ.

Proof.

Suppose without loss of generality that j = 0 and r = 1. First, we will consider the special case where k = 0. From the condition k[x,y], it follows that x, y are the endpoints of a diameter of S1 and therefore |xy|=2 for all angles μ. Since |x|=|y|=1 and z = 1 + d, we obtain by the law of cosines |xz|=1+(1+d)22(1+d)cos(μ),|zy|=1+(1+d)2+2(1+d)cos(μ). The sum |xz|+|zy| can be described with the function f of Proposition 4.3 if the constants u, v are replaced with 1+(1+d)2>0 and 2(1+d)>0, respectively. By Proposition 4.3, this function f is increasing with respect to μ[0,π2]. Since the quotient (Equation6) can be clearly written as 2f(μ), it follows that it must be decreasing with respect to μ.

Suppose now that S1(j,r) is still the unit circle S1, but let 0<k<1. The equation of the line L(x,y) can be written as (7) t+xyt¯=x+y(7) with tC as variable. Here, x can be written as eθi with 0θ<π2. Furthermore, the line L(x,y) must contain k and, by substituting t = k in (Equation7), we will have y=xkkx1=eθikkeθi1. Consider now a function h:[0,2π)R, h(θ)=|eθi(eθik)(keθi1)||eθiz|+|z(eθik)(keθi1)|, which clearly depicts the values of the quotient (Equation6). For all θ[0,π2], by symmetry, (8) y=eφi=eθikkeθi1h(θ)=h(φ).(8) The function h fulfils h(0)=h(π)=1z, which is clearly its maximum value. If θ=0, then so is μ, so the maximum of the quotient (Equation6) is at μ=0. By Rolle's theorem, there is a critical point ~θ such that f(~θ)=0. By the property (Equation8), ~θ is the solution of eθi=eθikkeθi1. Thus, eθi+eθi2=kRe(eθ)=kμ=π2. Consequently, the quotient (Equation6) attains its minimum value at μ=π2. Because there are no other points where the derivative h is 0 at the open interval 0<θ<π2 than the one found above, the quotient is monotonic on the interval μ[0,π2]. To be more specific, the quotient must be decreasing because its maximum is at μ=0 and minimum at μ=π2.

Thus, we have proved that the quotient (Equation6) is decreasing with respect to μ, regardless of if k = j or k>j.

Theorem 4.5

Fix Sn1(j,r)Rn and zRn so that d=|zj|r>0. Then, supx,ySn1(j,r)|xy||xz|+|zy|=rr+d.

Proof.

Suppose without loss of generality that n = 2, j = 0, r = 1 and z=d+1(1,). By symmetry, we can assume that the points x,yS1 fulfil 0arg(x)π2 and arg(x)<arg(y)<2π. We will next prove the theorem by inspecting the quotient (Equation6) in a few different cases separately.

Consider first the case where arg(x)=0. Now, x = 1 and y=eφi for some 0<φ<2π. It follows that |xy||xz|+|zy|=|1eφi|d+|1+deφi|=(d|1eφi|+|1+deφi||1eφi|)1. Since both of the quotients d|1eφi| and |1+deφi||1eφi| obtain clearly their minimum with φ=π, the quotient (Equation6) is at maximum within limitation x = 1 when y = −1.

Suppose then that arg(x)=θ0 and arg(y)π. Now, we can rotate the points x, y by the angle θ clockwise about the origin. This transformation does not affect the distance |xy| but decreases distances |xz| and |zy|, so it increases the value of the quotient (Equation6). Since x maps into 1 in the rotation, this transformation leads to the first case studied above.

Finally, consider the case where arg(x)0 and π<arg(y)<2π. Now, (x,y)(1,1), so we can choose a point k(x,y)(1,1). If 1<k<0, we can always reflect the points x, y over the imaginary axis so that the quotient (Equation6) increases. Thus, we can suppose that 0k<1. By Theorem 4.4, the quotient is decreasing with respect to ZKX=μ[0,π2], so its maximum is at μ=0. It follows that x = 1 and y = −1.

Thus, the quotient (Equation6) obtains its highest value with x = 1 and y = −1. In the general case x,yS1(j,r), this means that x = j + r and y = jr. Since the value of the quotient (Equation6) is now r(r+d), the result follows.

Corollary 4.6

The sG-diameter of a closed ball J=B¯n(k,r) in a domain GRn is sG(J)=r(r+d), where d=d(J,G).

Proof.

Clearly, sG(J)=supx,yJsG(x,y)=supx,yJ(supzG|xy||xz|+|zy|)=supzG(supx,yJ|xy||xz|+|zy|)=supzG(supx,yJsRn{z}(x,y))=supzGsRn{z}(J). Trivially, sRn{z}(J) is at maximum when the distance d(z,J) is at minimum. Thus, (9) sG(J)=supx,yJ|xy||xz|+|zy|,(9) where zG such that d=d(z,J)=d(J,G). It follows from Proposition 4.2 that for all distinct x,yJ, we can choose s,tJ, st, such that [s,t]=L(x,y)J and sG(s,t)sG(x,y). Thus, the points x, y giving the supremum in (Equation9) must belong to Sn1(k,r). By Theorem 4.5, it follows from this that sG(J)=supx,ySn1(k,r)|xy||xz|+|zy|=rr+d.

Corollary 4.7

The sBn-diameter of a ball J=B¯n(k,r)Bn is sBn(J)=r(1|k|).

Proof.

Follows directly from Corollary 4.6.

Corollary 4.8

For all x,yBn such that |y||x|, the inequality sBn(x,y)|x| holds.

Proof.

Since yJ=Bn(|x|), sBn(x,y)sBn(J) and, by Corollary 4.7, sBn(J)=|x|.

Consider yet the following situation.

Lemma 4.9

For all points xB2{0} and yB2(|x|) non-collinear with the origin, sB2(x,y)<sB2(x,y), where y=xe2ψi and ψ=arcsin(|xy|2|x|).

Proof.

Since |y|=|xe2ψi|=|x| and |xy|=|x||1e2ψi|=2|x|sin(ψ)=|xy|, the point y is chosen from S1(|x|)S1(x,|xy|). By symmetry, we can assume that y is the intersection point closer to y. Fix z so that it gives the infimum infzS1(|xz|+|zy|). If μ=ZXY, then μ=ZXY=μ+YXY>μ. Clearly, by the law of cosines sB2(x,y)=|xy||xz|+|zy|=|xy||xz|+|xy|2+|xz|22|xy||xz|cos(μ)<|xy||xz|+|xy|2+|xz|22|xy||xz|cos(μ)=|xy||xz|+|zy|sB2(x,y), so the lemma follows.

Let us now focus on the results related to the Euclidean midpoint rotation.

Proposition 4.10

Consider two triangles YXZ and Y0X0Z0 with obtuse angles YXZ and Y0X0Z0. Let k and k0 be the midpoints of sides XY and X0Y0, respectively. Suppose that |xy|=|x0y0|, |kz||k0z0| and ZKXZ0K0X0. Then, |xz|+|zy||x0z0|+|z0y0|.

Proof.

Let r=|xk|=|x0k0|, m=|kz|, m0=|k0z0|, μ=ZKX and μ0=Z0K0X0, see Figure . By the law of cosines |xz|+|zy|=r2+m22rmcos(μ)+r2+m2+2rmcos(μ),|x0z0|+|z0y0|=r2+m022rm0cos(μ0)+r2+m02+2rm0cos(μ0). Furthermore, by Proposition 4.3, the function f:[0,π2]R, f(μ)=uvcos(μ)+u+vcos(μ), where u, v>0, is increasing with respect to μ[0,π2]. Note that here μ,μ0[0,π2] because the triangles already have obtuse angles YXZ and Y0X0Z0. Thus, it follows from μμ0 and mm0 that |xz|+|zy|=r2+m22rmcos(μ)+r2+m2+2rmcos(μ)r2+m22rmcos(μ0)+r2+m2+2rmcos(μ0)r2+m022rm0cos(μ0)+r2+m02+2rm0cos(μ0)=|x0z0|+|z0y0|.

Figure 2. The triangles YXZ and Y0X0Z0 of Proposition 4.10.

Figure 2. The triangles △YXZ and △Y0X0Z0 of Proposition 4.10.

Theorem 4.11

For all x,yB2, sB2(x,y)sB2(x0,y0)|xy||xy|2+(2|x+y|)2.

Proof.

Fix k=(x+y)2 and r=|xk|. Suppose that k0, for otherwise sB2(x,y)=sB2(x0,y0) holds trivially. Without loss of generality, let 0<k<1 and XKZ=ν[0,π2]. Now, YKZ=π+ν, x0=k+ri and y0=kri. There are two possible cases; either the infimum infz0S1(|x0z0|+|z0y0|) is given by one point z0S1 or there are two possible points z0S1.

Suppose first that the infimum infz0S1(|x0z0|+|z0y0|) is given by only one point. By Remark 2.7, this point must be z0=1. Fix u=r2+(1k)2 and v=2r(1k) and consider the function f of Proposition 4.3 for a variable ν. Now, we will have infzS1(|xz|+|zy|)|x1|+|1y|=f(ν)f(π2)=|x01|+|1y0|=infz0S1(|x0z|+|zy0|), from which the inequality sB2(x,y)sB2(x0,y0) follows.

Consider yet the case where there are two points giving the infimum infz0S1(|x0z0|+|z0y0|). By symmetry, we can fix z0 so that 0<arg(z0)π2. Now, the infimum infzS1(|xz|+|zy|) is given by some point z such that 0arg(z)arg(z0). If x, y are collinear with the origin, by Lemma 2.5 and Corollary 4.7, sB2(x,y)=|xy|2|x+y|=r1k=sB2(B¯n(k,r))sB2(x0,y0). If x, y, 0 are non-collinear instead, the triangles YXZ and Y0X0Z0 exists. The sides XY and X0Y0 are both the length of 2r and have a common midpoint k. It follows from Theorem 2.4 and the inequality 0<arg(z)arg(z0)π2 that angles YXZ and Y0X0Z0 are obtuse, |kz||kz0| and ZKXZ0KX0. By Proposition 4.10, |xz|+|zy||x0z0|+|z0y0|, so the inequality sB2(x,y)sB2(x0,y0) follows.

Thus, sB2(x,y)sB2(x0,y0) holds in every cases and, by Theorem 2.6, sB2(x0,y0)rr2+(1k)2=|xy||xy|2+(2|x+y|)2, which proves the latter part of the theorem.

Theorem 4.12

Let x,yB2 with k=(x+y)2 and r=|xk|. If r + k<1, sB2(x,y)sB2(x1,y1)=|xy|2|x+y|<1.

Proof.

If r + k<1, then x1,y1B2 and, by Lemma 2.5, sBn(x,y)|xy|2|x+y|=|x1y1|2|x1+y1|=sB2(x1,y1).

5. Hyperbolic midpoint rotation

In this section, we consider the hyperbolic midpoint rotation. The idea behind it is the same as the one of the Euclidean midpoint rotation, for our aim is still to rotate the points around their midpoint in order to estimate their triangular ratio distance. However, now the rotation is done by using the hyperbolic geometry of the unit circle instead of the simpler Euclidean method.

Definition 5.1

Hyperbolic midpoint rotation. Choose distinct points x,yB2. Let q be their hyperbolic midpoint and R=ρB2(x,q)=ρB2(y,q). Let x2,y2Sρ1(q,R) so that |x2|=|y2| but x2y2. Fix then x3,y3Sρ1(q,R) so that x3,y3 are collinear with the origin and |y1|<|q|<|x1|. See Figure .

Figure 3. Hyperbolic midpoint rotation

Figure 3. Hyperbolic midpoint rotation

The main result of this section is the inequality sB2(x2,y2)sB2(x,y)sB2(x3,y3). This inequality is well-defined for all distinct x,yB2 because the values of sB2(x2,y2) and sB2(x3,y3) are always defined. The first part of this inequality is proved in Theorem 5.11 and the latter part in Theorem 5.12, and the formula for the value of sB2(x2,y2) is in Theorem 5.3. Note that, according to numerical tests, the hyperbolic midpoint rotation gives better estimates for sB2(x,y) than the Euclidean midpoint rotation or the point pair function, see Conjecture 5.13.

Lemma 5.2

Choose x,yB2 so that their hyperbolic midpoint is 0<q<1. Let t=th(R2)=th(ρB2(x,y)4). Then, x2=q(1+t2)1+q2t2+t(1q2)1+q2t2i and y2=x2¯.

Proof.

By Lemma 2.1, Sρ1(q,R)=S1(j,h) with j=q(1t2)1q2t2 and h=(1q2)t1q2t2. To find x2 and y2, we need to find the intersection points of S1(j,h) and S1(c,d), where S1(c,d)S1 and c>1. Now, c2=(q+d)2=1+d2, from which it follows that d=1q22q and c=1+q22q. Clearly, x2=y2¯ since both j,cR. Let x2=u+ri and y2=uri. Now, h2=r2+(uj)2 and d2=r2+(cu)2. Thus, h2(uj)2=d2(cu)2h2u2+2juj2=d2c2+2cuu2u=h2j2d2+c22(cj). Since h2j2=(1q2)2t2(1q2t2)2q2(1t2)2(1q2t2)2=(t2q2)(1q2t2)(1q2t2)2=t2q21q2t2,d2+c2=(1q2)24q2+(1+q2)24q2=4q24q2=1,h2j2d2+c2=t2q21q2t2+1=(1+t2)(1q2)1q2t2,2(cj)=2(1+q22qq(1t2)1q2t2)=(1+q2)(1q2t2)2q2(1t2)q(1q2t2)=(1q2)(1+q2t2)q(1q2t2), we will have u=h2j2d2+c22(cj)=q(1+t2)(1q2)(1q2t2)(1q2)(1q2t2)(1+q2t2)=q(1+t2)1+q2t2. From the equality h2=r2+(uj)2, it follows that r=h2(uj)2=(1q2)2t2(1q2t2)2(q(1+t2)1+q2t2q(1t2)1q2t2)2=(1q2)2t2(1q2t2)2(q(1+t2)(1q2t2)q(1t2)(1+q2t2)1q4t4)2=(1q2)2t2(1q2t2)2(2qt2(1q2)1q4t4)2=(1q2)2t2(1+q2t2)2(1q4t4)24q2t4(1q2)2(1q4t4)2=t2(1q2)2(1q2t2)2(1q4t4)2=t2(1q2)2(1+q2t2)2=t(1q2)1+q2t2.

Theorem 5.3

For all x,yB2 with a hyperbolic midpoint qB2{0} and t=th(ρB2(x,y)4), sB2(x2,y2)=|q|2+t21+|q|2t2 if |q|<t2,sB2(x2,y2)=t(1+|q|)(1+t2)(1+|q|2t2)|q|2+t21+|q|2t2 otherwise.

Proof.

Suppose without loss of generality that 0<q<1, x2=u+ri and y2=x2¯. From Lemma 5.2, it follows that |x2|=u2+r2=q2(1+t2)2(1+q2t2)2+t2(1q2)2(1+q2t2)2=(q2+t2)(1+q2t2)(1+q2t2)2=q2+t21+q2t2, |x212|=|u+ri12|>12(u12)2+r2>14u2+r2>uq2(1+t2)2(1+q2t2)2+t2(1q2)2(1+q2t2)2>q(1+t2)1+q2t2q2(1+t2)2+t2(1q2)2=(t2+q2)(1+q2t2)>q(1+t2)(1+q2t2)t2+q2>q(1+t2)q(1q)<t2(1q)q<t2 and (1u)2+r2=(1q(1+t2)1+q2t2)2+t2(1q2)2(1+q2t2)2=(1q)2(1+t2)(1+q2t2)(1+q2t2)2=(1q)2(1+t2)1+q2t2r(1u)2+r2=t(1q2)1+q2t2(1q)(1+q2t2)1+t2=t(1+q)(1+t2)(1+q2t2). The result follows now from Theorem 2.6.

Theorem 5.4

[Citation17, Proposition 3.1, p. 447]

The hyperbolic midpoint of J[0,b] is [0,b]J[c,d] for all c,dS1 such that bL(c,d) and c, d are non-collinear with the origin.

Theorem 5.5

If hyperbolic segments J[ui,vi]B2, i=1,,n, are of the same hyperbolic length and have a common hyperbolic midpoint q, all their Euclidean counterparts [ui,vi] intersect at the same point.

Proof.

Choose distinct points u1,v1B2 that are non-collinear with the origin. Let q be their hyperbolic midpoint, R=ρB2(u1,v1) and k=L(0,q)L(u1,v1). Fix j, h as in Lemma 2.1. Now, u1,v1S1(j,h), J[u1,v1]S1(j,h) and u1,v1,j are non-collinear. It follows from Theorem 5.4 that the hyperbolic midpoint of J[j,k] is [j,k]J[u1,v1]. Since 0, j, q are collinear and kL(0,q), [j,k]J[u1,v1]=L(0,q)J[u1,v1]=q. Thus, q is the hyperbolic midpoint of J[j,k]. It follows that k only depends on q and j, so the intersection point L(0,q)L(ui,vi) must be the same for all indexes i, as can be seen in Figure . If ui,vi are collinear with the origin for some index i, then k[ui,vi] trivially. Thus, the theorem follows.

Figure 4. Hyperbolic circle Sρ1(q,R) with the points j, q, k of Theorem 5.5.

Figure 4. Hyperbolic circle Sρ1(q,R) with the points j, q, k of Theorem 5.5.

Corollary 5.6

For all x,yB2, there is a point k=L(x,y)L(x2,y2)L(x3,y3).

Proof.

Follows from Theorem 5.5.

Theorem 5.7

For all x,yB2 that are non-collinear with the origin and have a hyperbolic midpoint q, the distance |xy| is decreasing with respect to the smaller angle between L(x,y) and L(0,q).

Proof.

Consider a hyperbolic circle Sρ1(q,R), where R=ρB2(x,q), and let S1(j,h) be the corresponding Euclidean circle. By Lemma 2.1, we see that the points 0,j,q are collinear. Fix k as in Corollary 5.6 and let u=|jk|. Denote θ=(L(x,y),L(0,q))=(L(x,y),L(j,k))[0,π2]. Clearly, the distance $u$ does not depend on the angle θ. It follows that |xy|=2h2u2sin2(θ) is decreasing with respect to θ.

Corollary 5.8

For all x,yB2, |x2y2||xy||x3y3|.

Proof.

Follows from Theorem 5.7.

Corollary 5.9

For all x,yB2, |xy|2(1|q|2)t1|q|2t22th(ρB2(x,y)4), where q is the hyperbolic midpoint of J[x,y], and t=th(ρB2(x,y)4).

Proof.

By fixing h as in Lemma 2.1, we will have |x3y3|=2h=2(1|q|2)t1|q|2t22t=2th(ρB2(x,y)4), so the result follows from Corollary 5.8.

Remark 5.10

The inequality |xy|2th(ρB2(x,y)4) can be also found in Ref. [Citation7, (4.25), p. 57].

Theorem 5.11

For all x,yB2, sB2(x,y)sB2(x2,y2).

Proof.

Let q be the hyperbolic midpoint of J[x,y] and R=ρB2(x,q). If q = 0, sB2(x,y)=sB2(x2,y2) holds trivially. Thus, choose x,yB2 so that 0<q<1. Now, either the infimum infz2S1(|x2z2|+|z2y2|) is given by one point z2S1 or two points on S1.

If there is only one point giving the infimum infz2S1(|x2z2|+|z2y2|), it must be z2=1. by Remark 2.7 like in Lemma 2.1, and fix k as in Corollary 5.6. By symmetry, we can assume that 1KX=μ[0,π2]. Note that, if μ=π2, then x=x2 and y=y2. Now, it follows from Theorem 4.4 that sB2(x,y)|xy||x1|+|1y||x2y2||x21|+|1y2|=sB2(x2,y2).

Suppose now that there are two possible points z2S1 for infz2S1(|x2z2|+|z2y2|). By symmetry, let Im(x2)>0 and 0arg(x)arg(x2). Fix z2 so that Im(z2)>0 and OZ2X2=Y2Z2O, where o is the origin. By Theorem 2.4, this point z2 gives the infimum infz2S1(|x2z2|+|z2y2|). Denote yet ψ=Y2X2Z2, which is clearly an obtuse angle.

By Corollary 5.8, we can fix y[x,y] so that |xy|=|x2y2|. Let zS1 with Im(z)<Im(x) so that YXZ=ψ. Clearly, |xz||x2y2|. By Proposition 4.2, it follows that sB2(x,y)sB2(x,y)|xy||xz|+|zy|=|xy||xz|+|xy|2+|xz|22|xy||xz|cos(ψ)|x2y2||x2z2|+|x2y2|2+|x2z2|22|x2y2||x2z2|cos(ψ)=|x2y2||x2z2|+|z2y2|=sB2(x2,y2).

Theorem 5.12

For all x,yB2, sB2(x,y)sB2(x3,y3)=(1+|q|)t1+|q|t2, where q is the hyperbolic midpoint of J[x,y], and t=th(ρB2(x,y)4).

Proof.

Let q be the hyperbolic midpoint of J[x,y]. Fix then R=ρB2(x,q) and j, h, t as in Lemma 2.1. Now, B¯2(j,h)=B¯ρ2(q,R) and t=th(ρB2(x,y)4). By Corollary 4.7, sB2(x,y)sB2(B¯ρ2(q,R))=sB2(B¯2(j,h))=h1|j|=(1|q|2)t1|q|2t2|q|(1t2)=(1|q|2)t1|q|+|q|t2|q|2t2=(1|q|)(1+|q|)t(1|q|)(1+|q|t2)=(1+|q|)t1+|q|t2. Since |j|=|x3+y3|2 and h=|x3y3|2, by Lemma 2.5, sB2(x3,y3)=|x3y3|2|x3+y3|=h1|j|, so the theorem follows.

According to numerous computer tests, the following result holds.

Conjecture 5.13

For all x,yB2,

  1. sB2(x2,y2)sB2(x0,y0),

  2. sB2(x3,y3)sB2(x1,y1),

  3. sB2(x3,y3)pB2(x,y),

where the points xi,yi, i=0,,3, are as in Definitions 4.1 and 5.1.

Thus, by this conjecture, the hyperbolic midpoint rotation gives sharper estimations for sB2(x,y) than the Euclidean midpoint rotation or the point pair function.

6. Hölder continuity

In this section, we show how finding better upper bounds for the triangular ratio metric in the unit disk is useful when studying quasiconformal mappings. The behaviour of the distance between two points x,yBn under a K-quasiconformal homeomorphism f:BnBn=f(Bn) has been studied earlier in numerous works, for instance, see Ref. [Citation7, Theorem 16.14, p. 304]. Our next theorem illustrates how finding a good upper limit for the value of the triangular ratio metric can give new information regarding this question.

Theorem 6.1

If f:B2B2=f(B2) is a K-quasiconformal map, the inequality |f(x)f(y)|231K(sB2(x,y)1+sB2(x,y)2)1K holds for all x,yB2.

Proof.

Define a homeomorphism φK:[0,1][0,1] as in Ref. [Citation7, (9.13), p. 167] for K>0. By Refs. [Citation7, Theorem 9.32(1), p. 167] and [Citation7, (9.6), p. 158], (10) φK(r)411Kr1K=411(2K)(r2)1K,(10) where 0r1 and K1. Let f be as above, x,yB2 and t=th(ρB2(x,y)4). By Theorem 2.3, Schwarz lemma (see Ref. [Citation7, Theorem 16.2, p. 300]) and the inequality (Equation10), sB2(f(x),f(y))thρB2(f(x),f(y))2φK(thρB2(x,y)2)=φK(2t1+t2)411(2K)(t1+t2)1K411(2K)(sB2(x,y)1+sB2(x,y)2)1K. By Ref. [Citation7, Lemma 11.12, p. 201; Proposition 11.15, p. 202], it follows from the inequality above that |f(x)f(y)|2sB2(f(x),f(y))231K(sB2(x,y)1+sB2(x,y)2)1K, which proves the theorem.

Thus, as we see from Theorem 6.1, finding a suitable upper bound for the value of sB2(x,y) can help us estimating the distance of the points x, y under the K-quasiconformal mapping f.

Corollary 6.2

If f is as in Theorem 6.1, the inequality |f(x)f(y)|232K(|xy|2+4(1|x|)(1|y|)|xy||xy|2+2(1|x|)(1|y|))1K holds for all x,yB2.

Proof.

It follows from Theorems 6.1 and 2.3 that |f(x)f(y)|231K(sB2(x,y)1+sB2(x,y)2)1K231K(pB2(x,y)1+pB2(x,y)2)1K=231K(|xy|2+4(1|x|)(1|y|)|xy|2|xy|2+4(1|x|)(1|y|))1K=232K(|xy|2+4(1|x|)(1|y|)|xy||xy|2+2(1|x|)(1|y|))1K.

Corollary 6.3

If f is as in Theorem 6.1, the inequality |f(x)f(y)|232K((2|x+y|)|xy|22|x+y|+|x|2+|y|2)1K holds for all x,yB2.

Proof.

Follows from Theorem 6.1 and Lemma 2.5 and the fact that |x+y|2+|xy|2=2|x|2+2|y|2.

Corollary 6.4

If f is as in Theorem 6.1, all x,yB2 fulfil |f(x)f(y)|231K((1+|q|)(1+|q|t2)t(1+|q|t2)2+(1+|q|)2t2)1K, where q is the hyperbolic midpoint of J[x,y], and t=th(ρB2(x,y)4).

Proof.

Follows from Theorems 6.1 and 5.12.

Remark 6.5

Neither of Corollaries 6.3 and 6.2 is better than the other for all points x,yB2. For x = 0.3 and y = 0.3i, the limit in Corollary 6.3 is sharper than the one in Corollary 6.2 and for x = 0.9 and y = 0.9i, the opposite holds. However, according to numerical tests related to Conjecture 5.13, the result in Corollary 6.4 is always better than the ones in Corollaries 6.3 and 6.2.

By restricting how the point pair x, y is chosen from B2, we can find yet better estimates.

Corollary 6.6

If f is as in Theorem 6.1, the inequality |f(x)f(y)|232K(|xy|1r)1K holds for all x,yB2 such that |x+y|2r.

Proof.

Now, 2|x+y|22|x+y|+|x|2+|y|22|x+y|22|x+y|+|x+y|22=11|x+y|211r, so the result follows from Corollary 6.3.

Corollary 6.7

For all x,yB2 such that |x+y|1, |f(x)f(y)|231K|xy|1K, where f is as in Theorem 6.1.

Proof.

Follows from Corollary 6.6.

Remark 6.8

The proof of Theorem 6.1 is based on the Schwarz lemma of quasiregular mappings [Citation7, Thm 16.2, p. 300] and therefore the results of this section hold also for quasiregular mappings with minor modifications.

Acknowledgements

The authors are indebted to Professor Masayo Fujimura and Professor Marcelina Mocanu for their kind help in connection with the proof of Theorem 4.4.

Disclosure statement

No potential conflict of interest was reported by the author(s).

Correction Statement

This article has been republished with minor changes. These changes do not impact the academic content of the article.

Additional information

Funding

The research of the first author was supported by Finnish Concordia Fund.

References

  • Gehring FW, Hag K. The ubiquitous quasidisk. Providence, RI: American Mathematical Society; 2012 (Mathematical surveys and monographs, Vol. 184) with contributions by Ole Jacob Broch.
  • Gehring FW, Martin GJ, Palka BP. An introduction to the theory of higher dimensional quasiconformal mappings. Providence, RI: American Mathematical Society; 2017 (Mathematical surveys and monographs, Vol. 216).
  • Heinonen J, Koskela P, Shanmugalingam N, et al. Sobolev spaces on metric measure spaces. Cambridge: Cambridge University Press; 2015 (New mathematical monographs, Vol. 27), an approach based on upper gradients.
  • Väisälä J. The free quasiworld. Freely quasiconformal and related maps in Banach spaces. In: Quasiconformal geometry and dynamics (Lublin, 1996), 55–118, Banach Center Publ., 48, Polish Acad. Sci. Inst. Math., Warsaw, 1999.
  • Beardon AF, Minda D. The hyperbolic metric and geometric function theory. In: Ponnusamy S, Sugawa T, Vuorinen M, editors. Proceedings International Workshop on Quasiconformal Mappings and Their Applications (IWQCMA05), IIT-Madras, Chennai, India; 2007. p. 9–56.
  • Gehring FW, Palka BP. Quasiconformally homogeneous domains. J Analyse Math. 1976;30:172–199.
  • Hariri P, Klén R, Vuorinen M. Conformally invariant metrics and quasiconformal mappings. Cham, Switzerland: Springer; 2020.
  • Papadopoulos A. Metric spaces convexity and non-positive curvature. Second ed. Zürich: European Mathematical Society (EMS); 2014 (IRMA lectures in mathematics and theoretical physics, Vol. 6).
  • Hästö P. A new weighted metric the relative metric I. J Math Anal Appl. 2002;274:38–58.
  • Chen J, Hariri P, Klén R, et al. Lipschitz conditions triangular ratio metric and quasiconformal maps. Ann Acad Sci Fenn Math. 2015;40:683–709.
  • Fujimura M, Hariri P, Mocanu M, et al. The Ptolemy–Alhazen problem and spherical mirror reflection. Comput Methods Funct Theory. 2019;19:135–155.
  • Hariri P, Vuorinen M, Zhang X. Inequalities and bilipschitz conditions for triangular ratio metric. Rocky Mountain J Math. 2017;47(4):1121–1148.
  • Wang G, Vuorinen M, Zhang X. On cyclic quadrilaterals in Euclidean and hyperbolic geometries. Publ. Math. Debrecen (to appear). Preprint arXiv:1908.10389.
  • Hariri P, Klén R, Vuorinen M, et al. Some remarks on the Cassinian metric. Publ Math Debrecen. 2017;90(3–4):269–285.
  • Fujimura M, Mocanu M, Vuorinen M. Barrlund's distance function and quasiconformal maps. Complex Var Elliptic Equ. 2020;1–31. https://doi.org/https://doi.org/10.1080/17476933.2020.1751137
  • Anderson G, Vamanamurthy M, Vuorinen M. Conformal invariants inequalities and quasiconformal maps. New York: Wiley-Interscience; 1997.
  • Vuorinen M, Wang G. Hyperbolic lambert quadrilaterals and quasiconformal mappings. Ann Acad Sci Fenn Math. 2013;38:433–453.