Next Article in Journal
Modeling the Impact of the Imperfect Vaccination of the COVID-19 with Optimal Containment Strategy
Next Article in Special Issue
2-Complex Symmetric Composition Operators on H2
Previous Article in Journal
Forward Order Law for the Reflexive Inner Inverse of Multiple Matrix Products
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

On the Semi-Group Property of the Perpendicular Bisector in a Normed Space

by
Gheorghiță Zbăganu
Institute for Mathematical Statistics and Applied Mathematics, 050711 Bucharest, Romania
Axioms 2022, 11(3), 125; https://doi.org/10.3390/axioms11030125
Submission received: 2 November 2021 / Revised: 2 March 2022 / Accepted: 2 March 2022 / Published: 10 March 2022
(This article belongs to the Special Issue Symmetry of Nonlinear Operators)

Abstract

:
Let (X,d) be a metric linear space and aX. The point a divides the space into three sets: Ha = {xX: d(0,x) < d(x,a)}, Ma = {xX: d(0,x) = d(x,a)} and La = {xX: d(0,x) > d(x,a)}. If the distance is generated by a norm, Ha is called the Leibnizian halfspace of a, Ma is the perpendicular bisector of the segment 0,a and La is the remaining set La = X\(Ha Ma). It is known that the perpendicular bisector of the segment [0,a] is an affine subspace of X for all aX if, and only if, X is an inner product space, that is, if and only if the norm is generated by an inner product. In this case, it is also true that if x,yLaMa, then x + yLa Ma. Otherwise written, the set La Ma is a semi-group with respect to addition. We investigate the problem: for what kind of norms in X the pair (LaMa,+) is a semi-group for all aX? In that case, we say that “ ( X , . ) has the semi-group property” or that “the norm . has the semi-group property”. This is a threedimensional property, meaning that if all the three-dimensional subspaces of X have it, then X also has it. We prove that for two-dimensional spaces, (La,+) is a semi-group for any norm, that ( X , . ) has the semi-group property if, and only if, the norm is strictly convex, and, in higher dimensions, the property fails to be true even if the norm is strictly convex. Moreover, studying the Lp norms in higher dimensions, we prove that the semi-group property holds if, and only if, p = 2. This fact leads us to the conjecture that in dimensions greater than three, the semi-group property holds if, and only if, X is an inner-product space.

1. Introduction

Let (X, ‖ ‖) be a real normed space.
In [1], the authors were interested in the following property of X
For every x,y,z ∈ X, the inequalities
(AFor every x,y,z ∈ X, the inequalities
x║ ≥ ║y + z║, ║y║ ≥ ║y + z║, ║z║ ≥ ║x + y║, imply the equalities
x║ = ║y + z║, ║y║ = ║y + z║, ║z║ = ║x + y
The typical examples of spaces with this property are the inner product spaces. Indeed, if there is an inner product <⋅,⋅> such that ‖x2 = <x,x>, then the inequalities in (A) become ‖x2 ≥ ‖y2 +z2 + 2<x,y>, ‖y2 ≥ ‖x2 +z2 + 2<x,z>, ‖z2 ≥ ‖y2 +x2 + 2<z,y>; if we add them, we obtain ‖x + y + z‖ ≤ 0 ⇔ x + y + z = 0. In this case, we obtain even more than the simple equalities ‖x‖ = ‖y + z‖, ‖y‖ = ‖z + x‖, ‖z‖ = ‖x + y‖, namely, x = −(y + z), y = −(x + z), z = −(y + x).
If we denote a = x + y + z, we can write property (A) as
(AFor every x,y,a ∈ X, the inequalities
x║ ≥ ║a − x║, ║y║ ≥ ║a − y║, ║a − x − y║ ≥ ║x + yimply the equalities
x║ = ║a − x║, ║y║ = ║a − y║, ║a − x − y║ = ║x + y
Written in this form, the property (A) receives a geometric flavor.
Notation. 
Let aX. Denote by La, Ma, Ha the sets
La = {xX|║x − a║ < ║x║}
Ma = {xX|║x − a║ = ║x║}
Ha = {xX|║x − a║ > ║x║}
Note that Ma is the perpendicular bisector of the segment [0,a], La is the set of points closer to a and Ha is the set of points from X closer to 0. Some authors [2] call the set Ha the Leibnizian halfspace of a. Horvath ([2,3]) writes H0,a instead of Ha and Ha,0 instead of Ma
In the sequel, we always assume that a ≠ 0, since a = 0 is nonsensical.
Using these notations, the property (A) becomes
(A) 
For any 0 ≠ a ∈ X, if x ∈ La ∪ Ma, y ∈ La ∪ Ma, x + y ∈ Ha ∪ Ma then x,y,x + y ∈ Ma
A weaker form of this property is
(B) 
For any 0 ≠ a ∈ X, if x ∈ La ∪ Ma, y ∈ La ∪ Ma, x + y ∈ Ha ∪ Ma then x + y ∈ Ma
However, is it possible that x,y,x + yMa? In an inner-product space, this is not possible. In this case, xLaMa, yLaMax + yLa.
In other words, in an inner-product space, (La Ma,+) is a semigroup. Indeed, this is obvious: squaring the inequalities ║x║ ≥ ║a − x║, ║y║ ≥ ║a − y║and adding them, we obtain ║x + y2 ≥ 2║a2 − 2<a,x + y> + ║x + y2 ≥ ║x + y − a2, meaning that x + y is in La.

2. The Main Result

A norm space is strictly convex ([1,2,3,4]) if and only if the unity ball B1 = {xX: ║x║ ≤ 1} is a strict convex set or, equivalently, the equality
x + y║ = ║x║ + ║y║ can hold if and only if x = 0 or y = 0 or y = λx for some λ > 0. An equivalent definition is that the norm is a strict convex function, i.e., for xy, the equality
║(1 − λ)x + λy║ = (1 − λ)║x║ + λ║y║ can hold if, and only if, λ ∈ {0,1}. Here λ ∈ [0,1].
Let us denote by ha : X → ℜ the mapping ha(x) = ║x − a║ − ║x║. Then
Ha = {ha > 0}
Ma = {ha = 0}
La = {ha < 0}.
As the function ha is continuous, Ma is a closed set and Ha, La are open.
Definition 1.
Say that (X, ║.║) has the semigroup property (or that the norm has the semigroup property) if it satisfies the condition
(M) (La Ma,+) is a semi-group for every aX
or, equivalently,
(M) (La Ma) + (La Ma) ⊆ (La Ma).
Explicitly, the semi group property means that
x║ ≥ ║x − a║,║y║ ≥ ║y − a║ ⇒ ║x + y║ ≥ ║x +y − a║ or any x,y,aX
We shall also consider the following similar property
(M°) (La,+) is a semigroup
Proposition 1.
(i)
The property (M) implies the property (B).
(ii)
A space X has the properties (A), (B), (M) or (M°) if, and only if, all three-dimensional subspaces of X have it.
Proof. 
(i)
Obvious. If La Ma is a semigroup, then xLaMa, yLa Max + yLaMa. Thus, if xLa Ma, yLa Ma, x + yHa Ma, then x + y ∈ (Ha Ma) ∩ (Ha Ma) = Ma.
(ii)
For instance, if X has the property (A) and Y = span({x,y,z}), then it obviously has the property (A) as well. Conversely, if we know that Y has the property (A), we know that X also has it. □
Remark 1.
Thus, the semigroup property is geometric; it deals with the shape of the unity ball in a three-dimensional subspace of X. In the case of inner product spaces, this is an ellipsoid. In fact, the inner product spaces satisfy a stronger assumption, namely
(C) ‖x‖ ≥ ‖xa‖, ‖y‖ ≥ ‖ya‖, ‖x + y‖ = ‖x + ya‖ ⇒ a = 0
Of course, (C) implies (A).
Any norm space has the following properties:
Proposition 2.
Let X be any norm space and a ∈ X, a ≠ 0.
(i)
Ifx‖ ≥ ‖x − aand 0 ≤ t ≤ 1 thenx‖ ≥ ‖xta‖.
(ii)
If xLa and t ≥ 1 then txLa.
If xMa and 0 ≤ t ≤ 1, then txMaHa.
If xMa, t ≥ 1 then txMaHa.
(iii)
If xLa, there exists 0 ≤ t ≤ 1 such that txMa
(iv)
If, moreover, the function X is strictly convex, then xMa, txMa if, and only if, t = 1. In other words, if xMa and t > 1, then txMa.
(v)
If X is strictly convex, the closures are Cl(Ha) = Ha ∪ Ma and Cl(La) = La ∪ Ma.
Proof. 
(i)
Let f: ℜ → ℜ be defined by f(t) = ║x║ − ║x − ta║. The function f is concave, f(0) = 0 and f(1) ≥ 0. Thus t ∈ [0,1] ⇒ f(t) ≥ 0 or, explicitly, ║x║ ≥ ║x − ta
(ii)
Obvious if we write the inequality ║x║ ≥ ║xta ║ as ║ 1 t x ║ ≥ ║ 1 t xa ║. If t ∈ (0,1], then 1 t ≥ 1. If xMa, then f(0) = f(1) = 0 ⇒ f(s) ≤ 0 ∀ s ∈ (−∞,0) ∪ (1,∞) ⇔ ║x║ ≤ ║xta ║ ∀ t ∈ (0,1). Thus ║tx║ ≤ ║txa║ ∀ t ∈ (0,1).
(iii)
The function f from (i) has the property that f(∞) = −∞. As f(1) ≥ 0, there must be t ≥ 1 such that f(t) = 0 ⇔ ║x║ = ║x − ta ║ ⇔ 1 t xMa
(iv)
If the norm is strictly convex, then the function f from 1 is strictly concave. Let s = 1 t . As xMaf(0) = f(1) = 0, the strict concavity implies s ∈ (0,1) ⇒ f(s) > 0 ⇔║x║ > ║xsa║ or ║tx║ > ║txa║.
(v)
The inclusion Cl(Ha) ⊆ HaMa holds in any norm space. The problem is to check that Ma ⊆ Cl(Ha), and this is not true in general. However, if the norm ║.║ is strictly convex, and xMa, then the points tnx belong to Ha if tn > 1. Now, it is obvious that if tn ↓ 1, then tnxx hence x ∈ Cl(Ha). The equality Cl(La) = LaMa has the same proof. □
The following result also holds in any norm space and simplifies the issue:
Proposition 3.
If (X, ║⋅║) has the property (M) and T:X → X is a bijective linear operator then (X, ║⋅║T) has the property (M), too, where the norm is defined by
║x║T = ║Tx║
Moreover, if (X, ║⋅║) is strictly convex, then (X, ║⋅║T) is strictly convex, too.
Proof. 
Let Ha(T) = {xX: ║xT > ║xaT }, Ma(T) = {xX: ║xT = ║xaT} and La(T) = {xX: ║xT = ║xaT}.
We claim that Ha(T) = T1 (HTa), Ma(T) = T1 (MTa) and La(T) = T1 (LTa).
Indeed, Ha(T) = {xX: ║Tx║ < ║TxTa║} ⇔ T(Ha(T)) = {Tx: ║Tx║ < ║TxTa║} = {yX: ║y║ < ║y − a║} = Ha and the same holds for the other two sets. The property (M) says that HaMa is a semigroup for any aX. Then HTaMTa is also a semigroup. it is obvious that if HX is a semigroup and U:XX is a linear operator, then U(H) is a semigroup, too. Therefore Ha(T)Ma(T) = T1(HTa MTa) is a semigroup.
The last assertion is obvious. □
Example 1.
If2 endowed with some norm that has the property (M), the same holds for 2 endowed with the norm ║(x1, x2)║* = ║(ax1 + bx2, cx1 + dx2)║ provided that  det ( a b c d ) ≠ 0. Now we can state our main result.
Theorem 1.
Let X = 2 endowed with some norm ║⋅║. Let aX. Then
(i)
La + LaLa
(ii)
If X is strictly convex, then (LaMa) + (LaMa) ⊆ LaMa
(iii)
Conversely, if the norm has the semi-group property, then it is strictly convex.
Thus, in the two-dimensional case, the semi-group property of a norm is equivalent with its strict convexity.
Proof. 
For the sake of a better understanding, the proof is divided into several steps. □
Step 1. There is no restriction on considering a = (0,1).
Indeed, if a = (a1, a2), we can choose a linear operator T such that Ta = (0,1). Next, we replace the norm ║.║ with the norm ║⋅║T and use Proposition 3.
Step 2. The two-dimensional norms have the following useful property (although it is well known, we prove it here since we did not find appropriate references to it in our research and it can be used to construct many norms on the plane):
Lemma 1.
For any norm on a two dimensional space there exists a convex function f: ℜ → (0,∞), such that
lim t f ( t ) t = lim t f ( t ) t = m > 0   and
( x , y ) = { | x | f ( y x ) i f x 0 m | y | i f x = 0 .   Notice   that   m = ( 0 , 1 )
Moreover, if X is strictly convex, then the function f is also strictly convex.
Conversely, for any convex function f satisfying (4), the equality (5) defines a norm.
All the norms in a two-dimensional space have this form. If f is strictly convex, then the norm given by (5) is also strictly convex.
For the proof of the Lemma, see Appendix A.
Combine Step 1 and Step 2. Thus a = (0,1) and f is a function, as in the above Lemma. Next, we drop the index a and write
L = {(x,y): x ≠ 0, f ( y x ) > f ( y 1 x ) or x = 0, m|y| > m|y − 1| ⇔ y ∈(½, ∞)},
M = {(x,y): x ≠ 0, f ( y x ) = f ( y 1 x ) or x = 0, m|y| = m|y − 1| ⇔ y = ½}
H = {(x,y): x ≠ 0, f ( y x ) < f ( y 1 x ) or x = 0, m|y| < m|y − 1| ⇔ y ∈(−∞, ½)}
Step 3. Describe the sets M, H, L.
Due to the property that lim t f ( t ) t = lim t f ( t ) t = m > 0 , the convex function f behaves as follows: there exists u1u2 ∈ ℜ, such that
f is decreasing on (−∞, u1)
f is constant on (u1, u2)
f is increasing on (u2, ∞)
Note that if f is strictly convex, then u1 = u2. Let v = f(u1) = f(u2).
Lemma 2.
Let f: ℜ → ℜ be a continuous function which satisfies the assumptions (6), (7), (8). Suppose further that  lim t f ( t ) t = m 2 > 0 , lim t f ( t ) t = m 1 > 0
Let h(x,y) = f ( y x ) f ( y 1 x ) , h: ℜ* × ℜ → ℜ. The sets M = {h = 0}, L = {h > 0} and H = {h < 0} may then be characterized as follows:
There are exist two functions φ1, φ2: ℜ* → ℜ, φ1 ≤ φ2 such that
L = {(x,y): φ1(x)∨φ2(x) < y}
M = {(x,y): φ1(x)∧φ2(x) ≤ y ≤ φ1(x)∨φ2(x)}
H= {(x,y): φ1(x)∧φ1(x) > y}
Moreover,
φj(−x) = 1 − φj(x) ∀ x ∈ ℜ*, j = 1,2
φ j ( 0 + 0 ) = m 1 m 1 + m 2 , φ j ( 0 0 ) = m 2 m 1 + m 2
φ 1 ( x ) = φ 2 ( x )     x     ( 1 u 2 u 1 , 1 u 2 u 1 ) \ { 0 }
If f is strictly convex, then φ1 = φ2
if ( x , y )     M   a n d ( tx , ty )     M   for   some   t   >   1 ,   then   either   x > 1 u 2 u 1   or   x < 1 u 2 u 1 .
For the proof, see Appendix A.
In our case, m1 = m2 allows us to extend the mapping of φj to 0, defining φj(0) = ½.
Recall our claim: we want to prove that
y > φ1(x)∨φ2(x), y′ > φ1(x′)∨φ2(x′) ⇒ y + y′ > φ1(x + x′)∨φ2(x + x′)
This implication would surely hold if we could prove that φj are sub-additive.
Indeed, in this case, y > φ1(x), y′ > φ1(x′) ⇒ y + y′ > φ1(x) + φ1(x′) ≥ φ1(x + x′) and, similarly, y + y′ > φ2(x + x′).
Step 4. The functions φj are sub-additive
We shall use the following criterion for sub-additivity:
Lemma 3.
(i)
Let g:ℜ → ℜ be such that gs) ≤ λg(s) ∀ λ ≥ 1 ∀ s ∈ ℜ. Next, g(s + t) ≤ g(s) + g(t) ∀ s,t such that st ≥ 0.
If, moreover, λ > 1 ⇒ gs) < λg(s) ∀ s ∈ ℜ, then g(s + t) < g(s) + g(t) ∀ s,t such that st ≥ 0.
(ii)
Let g be a function, as before. Suppose that g satisfies the following symmetry property:
(S)
there is a ∈ ℜ, b > 0, such that g(as) + g(s) = b.
Consequently, g is sub-additive: g(s + t) ≤ g(s) + g(t).
If, moreover, λ > 1 ⇒ gs) < λg(s) ∀ s ∈ ℜ, then g(s + t) < g(s) + g(t) ∀ s,t
The proof in Appendix A.
Proposition 4.
For 1 < p < ∞, the functions φj defined by Lemma 2 are sub-additive.
Moreover, if X is strictly convex, they coincide and are strictly sub-additive.
Proof. 
Because of the fact that m1 = m2, these two functions are defined on the whole real line (φj(0) = ½) and they coincide on the interval [ 1 u 2 u 1 , 1 u 2 u 1 ]. The equality (12) demonstrates that the symmetry property (S) is fulfilled with a = 0, b = 1. Thus, we only need to prove that
φjx) ≤ λφj(x) for any x.
Check that for φ1. Indeed,
-
if x < 1 u 2 u 1 , then φ1(x) = 1 + xu2, φ1(tx) = 1 + txu2 < t + txu2 = tφ1(x);
-
if x > 1 u 2 u 1 , φ1(x) = xu2 ⇒ φ1(tx) = tφ1(x).
-
if x ∈ [ 1 u 2 u 1 , 1 u 2 u 1 ], then y = φ1(x) ⇔ (x,y) ∈ M. Then (tx,ty) ∈ HM due to Proposition 2(ii). However, (tx,ty) does not belong to M because of (16); hence, in this case, (tx,ty) ∈ H. According to (9) this means that ty > (φ1∨φ2)(tx) ⇒ tφ1(x) > φ1(tx).
In the same way, one proves that φ2(tx) ≤ tφ2(x) ∀ t ≥ 1.
If X is strictly convex, then if (x,y) ∈ M and t > 1, then (tx, ty) ∈ M ⇔ φ(tx) < ty = tφ(x); hence, Lemma 3 points demonstrates that φ is strictly sub-additive.
This ends the proof. □
The proof of Theorem 1 immediately follows due to Lemma 4.
Step 5. End of proof of Theorem 1
(i)
According to (9), (x,y) ∈ Ly > φ(x) with φ =φ1∨φ2. The functions φj are sub-additive; hence, φ is also sub-additive. Therefore y > φ(x), y′ > φ(x′) ⇒ y + y′ > φ(x) + φ(x′) ≥ φ(x + x′) ⇒ (x + x′, y + y′) ∈ L.
(ii)
If the space X is strictly convex, Cl(L) = LM. However, obviously
L + LL ⇒ Cl(L) + Cl(L) ⊆ Cl(L).
(iii)
Suppose that X is not strictly convex. The unity sphere B = {zX: ║z║ = 1} contains a segment of line I = {(1 − λ)z1 + λz2: 0 ≤ λ ≤ 1} for some z1z2B. We suppose these two points to be extreme. Let T:ℜ2 → ℜ2 a linear operator such that Tz1 = ( 1 1 ) , Tz2 = ( 1 1 ) . Suppose that (X, ║⋅║) has the semi-group property. Consequently, (X, ║⋅║T) also has it. The new unity sphere BT = {zX: ║zT = 1} contains the segment {(1,t)|−1 ≤ t ≤ 1}. Let f(t) = ║ ( 1 t ) T. This function is convex and f is decreasing on (−∞,−1), constant on (−1,1), and increasing on (1,∞). Thus, u1 and u2 from Lemma 2 are u1 = −1, u2 = 1. According to (10), z = (x,y) belongs to M if, and only if, (φ1∧φ2)(x) ≤ y ≤ (φ1∨φ2)(x). Moreover, we see that x ∈[½, ∞) ⇒ φ1(x) = x, φ2(x) = 1 − x and x ∈ (−∞, ½] ⇒ φ1(x) = 1 + x, φ2(x) = −x. It follows that there are many pairs of z,z′, such that z,z′M but z + z′ = 0 ∈ L (for instance, z = (1,0) and z′ = (−1,0) are both in M). This contradicts the semi-group property: M + M should be included in LM.
To conclude: if a = T1e2, then one can find zMa such that −zMa, as well.
Furthermore, we can prove the following.
Theorem 2.
The space X = ℜ2 endowed with a strict convex norm satisfies the assumption (C)
Proof. 
We have to prove that the relations
z║≥║za║, ║z′║≥║z′ − a║, ║z + z′║=║z +z′ − a
for some z, z′X can hold if, and only if, a = 0. Suppose, ad absurdum, that this is not true. Let a = (a1, a2) ≠ 0. There is a one-to-one and onto linear operator T: XX, such that Ta = e2 = (0, 1). Let w = T1z, w′ = T−1z′. Consequently, (19) becomes
wT ≥ ║we2T,║wT ≥ ║w′ − e2T,║w + w′║T = ║w + w′ − e2T
where ║⋅║ is the norm defined at Proposition 3.
If X is strictly convex, the norm ║⋅║T is also strictly convex. Let w = (x,y), w′ = (x′,y′) and φ the function defined by Lemma 2. Therefore, (20) becomes
y ≥ φ(x), y′ ≥ φ(x′), y + y′ = φ(x + x′)
However, this is impossible, since y + y′ ≥ φ(x) + φ(x′) > φ(x + x′), as φ is strictly sub-additive. □

3. Conjectures, Open Problems, and Counterexamples

At a first glance, a possible conjecture would be that a strict convex normed space should have the mediator property. However, this not true: all the spaces Lp({1,2,3}) with 1 < p < ∞ are strictly convex and fail to have the property (M).
Proposition 5.
Let X = ℜ3 endowed with lp norm, x = ( j = 1 3 | x j | p ) 1 p .
(i)
If p ∈ (1,2) ∪ (2,∞), then X does not have the mediator property.
(ii)
However, all its proper subspaces have it (due to Theorem 1).
Proof. 
(i)
Case 1: p ∈ (1,2). We choose x = (3,t,−t), y = (3, −t, t), a = (4,4,4). Then
  • x + y = (6, 0, 0), x − a = (−1, t − 4, −t − 4), y − a = (−1, −t − 4, t − 4), x + y − a = (2, −4, −4); hence, if t > 4, ║xp = ║yp = 3 p + 2 t p , ║x − ap = ║y − ap = 1 + (t − 4)p + (t + 4)p, ║x + yp = 6p and ║x + y − ap = 2p + 2⋅4p. We claim that for any 1 ≤ p < 2, we can choose t > 4, such that ║x║ > ║x − a║and ║y║ > ║y − a║ but ║x + y ║ < ║x + y − a║. Indeed, this is an equivalent to 3p 1 > (t − 4)p + (t + 4)p 2tp but 6p < 2p + 2⋅4p. The function g(t) =(t − 4)p + (t + 4)p 2tp has the property that g(∞) = 0 for any 1 ≤ p < 2; hence, for any fixed p ∈ [1,2), we can find a t = t(p), such that 3p 1 > g(t). As with the second condition, 6p < 2p + 2⋅4p ⇔ 3p < 1 + 2⋅2p, it holds for any such p ∈ (1,2), since the function h(p) = 1 + 2⋅2p − 3p is decreasing on (1,2) and positive.
(ii)
Case 2: p ∈ (2,∞). Now, we choose x = (1,−1,2), y = (1,2,−1), a = (−t,t,t) for some t ∈ (0,1/2). Therefore, x + y = (2,1,1), x − a = (1 + t,−1 − t,2 − t), y − a = (1 + t,2 − t,−1 − t), x + y − a = (2 + t,1 − t, 1 − t); hence, ║xp = ║yp = ║x + yp = (2p + 2),x − ap = ║y − ap = 2(1 + t)p+ (2 − t)p and ║x + yap = (2 + t)p + 2(1 − t)p. We claim that for every p > 2, there is t ∈ (0,1/2), such that ║x║ > ║xa║, ║y║ > ║ya║ but ║x + y║ < ║x + y − a║. Indeed, let p > 2 be fixed and let f,g be defined by f(t) = 2p + 2 − 2(1 + t)p − (2 − t)p, g(t) = (2 + t)p +2(1 − t)p 2p − 2. Note that f(0) = g(0) = 0 and that f′(0) = g′(0) = p(2p−1 − 2). If p > 2, then 2p−1 > 2; hence, the derivatives are positive. This means that for small t, we have f(t) > 0, g(t) > 0; this fact agrees with our claim. For p = ∞, it is even simpler, since now ║x║ = ║y║ =║x + y║ = 2,x − a║ = ║y − a║ = 2 − t and ║x + yt║ = 2 + t.
Thus, the two-dimensional space ℜ2 endowed with the norm lp, 1 < p < ∞ has the mediator property, while ℜ3 endowed with the same norm does not have it. □

4. Open Problems

  • The only three-dimensional spaces that satisfy the mediator property (B) are inner-product spaces. Prove or disprove that if dim(X) ≥ 3, then property (M) implies the fact that X has an inner product. In other problems connected with perpendicular bisectors, this was indeed the case [5,6,7,8,9].
  • The only examples of spaces possessing the mediator property also satisfy the property C. Prove or disprove that (M) ⇔ (C).
  • Prove or disprove that if M + MML, then X has the property (M).

Funding

The research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The author declares no conflict of interest.

Appendix A

Proof of Lemma 1.
Let f(t) = ║(1,t)║. Obviously f is convex, lim t f ( t ) t = lim t f ( t ) t = ( 0 , 1 ) : = m > 0 and if x ≠ 0 we can write ║(x,y)║=|x| ( 1 , y x ) . If x = 0 then ║(0,y)║= |y|║(0,1)║ = m|y|.
Conversely, let f: ℜ → (0,∞) be a convex function, such that lim t f ( t ) t = lim t f ( t ) t = m > 0 . We claim that the equality
p ( x , y ) = { | x | f ( y x ) i f x 0 m | y | i f x = 0
defines a norm: it is sub-additive, p(x,y) = 0 ⇔ x = y = 0 and p(tx,ty) = |t|p(x,y). As the last two assertions are obvious, we shall focus on the sub-additivity. One must prove that
| x + x | f ( y + y x + x )   | x | f ( y x ) + | x | f ( y x )   if   x , x   and   x + x     0
| x | f ( y + y x )   | x | f ( y x ) + m | y |   if   x     0 ,   x = 0
| x | f ( y + y x )   | x | f ( y x ) + m | y |   if   x     0 ,   x = 0
m | y + y |     | x | ( f ( y x ) + f ( y x ) )   if   x + x = 0 ,   x     0
Of course, the most important is (A2). If both x and x′ are positive, this is easy. Indeed, p(x + x′, y + y′) = ( x + x ) f ( y + y x + x ) = ( x + x ) f ( x x + x y x + x x + x y x ) | x | f ( y x ) + | x | f ( y x ) due to convexity. However, if one of them is negative, this argument no longer holds. We shall use another.
We shall prove that any convex function f: ℜ → (0,∞), such that lim t f ( t ) t = lim t f ( t ) t = m > 0 admits the following representation
f ( t ) = a + | t c | d ν ( c )
where ν is a measure on the real line, such that ν(ℜ) = m and μ1:= | c | d ν ( c ) < ∞
It is well known (see, for instance, in [7], pg 96, formula (3.3)) that any convex function g:[0,∞) → ℜ that is differentiable at 0 admits a representation of the form
g ( t ) = a + b t + ( t c ) + d ν ( c )
where ν is a Stieltjes measure concentrated on (0,∞). If b ≥ 0, we can replace ν with ν + bδ0 (here, δ0 is the Dirac measure concentrated at 0) to obtain a simpler relation
g ( t ) = a + ( t c ) + d ν ( c ) ,   Supp ( ν )     [ 0 , )
Now, let us observe our function f: ℜ → (0,∞). It is convex and non-increasing on some intervals (−∞,u) and non-decreasing on (u,∞). Suppose that u = 0. There are, then, two Stieltjes measures, ν1 and ν2, Supp(ν1) ⊆ (−∞,0), Supp(ν2) ⊆ (0,∞) such that
f ( t ) = { a + ( t + c ) + d ν 1 ( c ) i f t < 0 a + ( t c ) + d ν 2 ( c ) i f t 0
As lim t f ( t ) t = lim t ( 1 c t ) + d ν 2 ( t ) = ν 2 ( [ 0 , ) ) = ν2(ℜ) (we applied Beppo Levi’s theorem) and lim t f ( t ) t = lim t ( 1 + c t ) + d ν 1 ( t ) = ν 1 ( ( , 0 ] ) = ν1(ℜ), we obtained ν1(ℜ) = ν2(ℜ):= m
Now, use the relation 2x+ = |x| + x. Consequently, (A9) becomes
2 f ( t ) 2 a = { | t c | d ν 1 ( c ) t m + c d ν 1 ( c ) i f t < 0 | t c | d ν 2 ( c ) + t m c d ν 2 ( c ) i f t 0
Write (A10) as
2 f ( t ) 2 a = { | t c | d ν 1 ( c ) + | t | m | c | d ν 1 ( c ) i f t < 0 | t c | d ν 2 ( c ) + | t | m | c | d ν 2 ( c ) i f t 0 = { f 1 ( t ) i f t < 0 f 2 ( t ) i f t 0
Now, note that t ≥ 0 ⇒ f1(t) = 0 and t ≤ 0 ⇒ f2(t) = 0 (this should be obvious if we look at (A9)). It means that we can write
2f(t) − 2a = f1 + f2 = | t c | d ν ( c ) + 2 m | t | | c | d ν ( c ) .
Therefore, we arrived at a representation of the form
f ( t ) = a + m | t | + | t c | d μ ( c )
that holds for any convex function non-increasing on (−∞,0) and non-decreasing on (0,∞), such that lim t f ( t ) t = lim t f ( t ) t = m > 0 . If we replace the measure μ = ν/2 by μ + mδ0, we arrive at the formula (A6). Subsequently, we can replace our assumption that f is decreasing on (−∞,0) and increasing on (0,∞) by “non-increasing on (−∞,u), non-decreasing on (u,∞)” for some u ∈ ℜ; the proof remains the same.
Therefore, any convex function, as in Lemma 1, can be represented by a finite Stieltjes measure by formula (A6).
Let C = {f: ℜ → [0,∞): f satisfies conditions (A2)−(A5)}
Consequently, C contains all the positive constants (for them m = 0) and all the functions f(x) = |x − a| (indeed, now m = 1: | x + x | f ( y + y x + x ) = |(y − ax) + ( y′ − ax′)|, | x | f ( y x ) = |y − ax|, | x | f ( y x ) = |y′ − ax′|, the inequality | x | f ( y + y x ) | x | f ( y x ) +m|y′| becomes |y + y′ − ax| ≤ |yax| + |y′| hence C is a cone : f,gCaf + bgCa,b ≥ 0. Moreover, it is closed with respect to pointwise convergence. As any integral is a limit of finite sums, C contains all the functions of the form (A6). This ends the proof. □
Proof of Lemma 2.
Let f(ℜ)=(v,∞), and f1({v}) = (u1, u2).
Let f1: (−∞,u1) → (v,∞) and f2: (u2,∞) → (v,∞) be defined by fj(x) = f(x). These functions are invertible. We know that M = {(x,y) | f ( y x ) = f ( y 1 x ) }. There are two possibilities
-
Case 1. x > 0. Now y 1 x < y x and the equality f ( y x ) = f ( y 1 x ) is possible only if u1 y 1 x < y x u2 or if y 1 x < u 1 < y x < u 2 .
In the first case x 1 u 2 u 1 and 1 + xu1yxu2.
In the second case, f ( y x ) = f 2 ( y x ) and f ( y 1 x ) = f 1 ( y 1 x ) ; hence, (x,y) ∈ M means that f 1 ( y 1 x ) = f 2 ( y x ) . Let t > v, t = f ( y x ) . Therefore, y = x f 2 1 ( t ) and y 1 = x f 1 1 ( t ) . Consequently, the set M can be described by the parametric curve
x ( t ) = 1 f 2 1 ( t ) f 1 1 ( t ) , y ( t ) = f 2 1 ( t ) f 2 1 ( t ) f 1 1 ( t ) , t ( v , ) .
This is the graph of the function ψ2: ( 0 , 1 u 2 u 1 ] → ℜ with the property that
ψ 2 ( 1 u 2 u 1 ) = u 2 u 2 u 1   and   φ ( 0 + 0 ) = m 1 m 1 + m 2   .
Indeed, lim t f 2 1 ( t ) f 2 1 ( t ) f 1 1 ( t ) = lim t f 2 1 ( t ) / t f 2 1 ( t ) / t f 1 1 ( t ) / t
However, lim t f 2 1 ( t ) t = lim x x f 2 ( x ) = 1 m 2 and lim t f 1 1 ( t ) t = 1 m 1
It follows that φ ( 0 + 0 ) = 1 / m 2 1 / m 2 + 1 / m 1 = m 1 m 1 + m 2 and we obtain
-
Case 2. x < 0. Now, y 1 x > y x and the equality f ( y x ) = f ( y 1 x ) is possible only if u1 y x < y 1 x u2 or if y x < u 1 < y 1 x < u 2 .
In the first case, x 1 u 2 u 1 and 1 + xu2yxu1.
In the second case, (x,y) ∈ M f 1 ( y x ) = f 2 ( y 1 x ) = t for some t > v. Thus, y = 1 + x f 2 1 ( t ) = x f 1 1 ( t ) and its parametric description is
x * ( t ) = 1 f 2 1 ( t ) f 1 1 ( t ) , y * ( t ) = 1 f 2 1 ( t ) f 2 1 ( t ) f 1 1 ( t ) , t ( v , )
Comparing this to (A13), we see that the curve (x*(t),(y*(t))t > v is the graph of the function ψ1:[ 1 u 2 u 1 , 0) → ℜ which is the symmetric of the graph of ψ2 with respect to the point (0, ½):
ψ 1 ( x ) = 1 ψ 2 ( x )     x     ( 0 ,   1 u 2 u 1 ] .
Add all these facts together. Let φj: ℜ* → ℜ be defined by
φ 1 ( x ) = { 1 + x u 2 i f x < 1 u 2 u 1 1 ψ 2 ( x ) 1 u 2 u 1 x < 0 ψ 2 ( x ) i f 0 < x 1 u 2 u 1 1 + x u 1 i f x > 1 u 2 u 1 ,   φ 2 ( x ) = { x u 1 i f x < 1 u 2 u 1 1 ψ 2 ( x ) 1 u 2 u 1 x < 0 ψ 2 ( x ) i f 0 < x 1 u 2 u 1 1 + x u 1 i f x > 1 u 2 u 1 .
We know that M = {h = 0} = {(x,y) ∈ ℜ* × ℜ: φ1(x)∧φ2(x) ≤ y ≤ φ1(x)∨φ2(x)}
Note that if f u1 = u2, then φ1 = φ2. This happens, for instance, if f is strictly convex.
Now, let us determine the set L = {h > 0} = {(x,y): f ( y x ) > f ( y 1 x ) }. If x > 0, this set surely contains all the points from {(x,y) ∈ (0,∞) × ℜ: y x > y 1 x > u 2 } since on the interval (u2,∞), the function f is increasing. Thus, H includes the set {(x,y) ∈ (0,∞) × ℜ: y > 1 + xu2}. If x < 0, it contains the set {(x,y) ∈ (−∞,0) × ℜ: y x < y 1 x < u 1 } since on (−∞,u1), the function f is decreasing. To conclude, L ⊇ {(x,y) ∈ ℜ* × ℜ: y > max(1+ xu1, 1 + xu2)}.
Similarly, L ⊇ {(x,y) ∈ ℜ* × ℜ: y < min(xu1, xu2)}.
We found that some points from L are above the graph of φ1∨φ2 and some points from L are below it. This means that all the points from L are in the set {(x,y)∈ ℜ* × ℜ: y > φ1(x)∨φ2(x)} and all the points from H are in the set {(x,y) ∈ ℜ* × ℜ: y < φ1(x)∨φ2(x)}. Suppose ad absurdum that L contains a point (x0, y0) such that y0 < (φ1∧φ2)(x0). There are some instances of c > 0 such that y0c < min(xu1, xu2) ⇒ (x0, y0c) ∈ L. According to the Darboux theorem, on the segment joining the points (x0, y0) and (x0, y0c), there must be at least one, let us say, (x0, η), such that h(x0, η) = 0 ⇔ (x0, η) ∈ M, contradicting the fact that (x,η) ∈ M ⇔ φ1(x0)∧φ2(x0) ≤ η ≤ φ1(x0)∨φ2(x0).
To conclude, L = {(x,y): y > (φ1∨φ2) (x)} and H = {(x,y): y < (φ1∧φ2) (x)}.
The situation simplifies significantly if u1 = u2 (for instance, if f is strictly convex): in this case, φ1 = φ2 = φ, M is the graph of φ, L is the points above the graph, and H is the points below it.
Now, prove the claim (A16). Suppose that (x,y) ∈ M and (tx, ty) ∈ M for some t > 1. Thus f ( y x ) = f ( y 1 x ) = f ( t y 1 t x ) . However, the only way in which the equation f(x) = α can have more than two solutions is if t = v; hence, these solutions are included in (u1, u2). It follows either that x > 0 and y 1 x < t y 1 t x < y x , meaning ⇔ x > 1 u 2 u 1 , or that x < 0 and y 1 x > t y 1 t x > y x ; hence, x < 1 u 2 u 1 . □
Proof of Lemma 3.
This proof uses the following inequality, which is interesting in itself.
Lemma 4.
(An elementary inequality)
Let a,b,s,t ∈ ℜ such that st > 0. Consequently, min[a(1 + s t ), b(1 + t s )] ≤ a + b.
Proof. 
Let λ = s t > 0. We want to check that a(1 + λ) ≤ a + b or b( 1 + 1 λ ) ≤ a + b aλ ≤ b or b λ a ⇔ aλ ≤ b or b ≤ aλ. This is obvious. □
The proof of Lemma 3 is as follows:
(i)
Note that g(0) = g(λ0) ≤ λg(0) ∀ λ ≥ 1; hence, g(0) ≥ 0, and if g satisfies the stronger condition λ > 1 ⇒ gx) > λg(x), then g(0) > 0.
Let s,t ∈ ℜ, such that st ≥ 0. If s = 0, the claim becomes g(t) ≤ g(t) + g(0) ⇔ g(0) ≥ 0, true.
If g satisfies the stronger condition gs) < λg(s) ∀ λ > 1, then this implies g(0) > 0, true. Therefore, if s = 0 we have proven that g(s+t) ≤ g(s) + g(t). If t = 0, the proof is the same. Suppose now that st > 0.
Consequently, g(s + t) = g ( ( 1 + t s ) s ) ( 1 + t s ) g ( s ) and g(s + t) = g ( ( 1 + s t ) t ) ( 1 + s t ) g ( t ) .
Thus, by (i), g(s + t) ≤ min( ( 1 + t s ) g ( s ) , ( 1 + t t ) g ( t ) ) ≤ g(s) + g(t).
Obviously, the stronger condition λ > 1 ⇒ gs) < λg(s) ∀ s implies the strict inequality g(s + t) < g(s) + g(t).
(ii)
It is obvious that if g is a sub-additive function, then the function h(s) = αgs) + β is again sub-additive for any λ ∈ ℜ, α,β > 0. If we write s = ax, the symmetry condition (S) becomes
g(ax) + g(a(1 − x)) = b g ( a x ) b + g ( a ( 1 x ) ) b = 1 .
It follows that the function g is sub-additive if, and only if, the function φ(x) = g ( a x ) b is sub-additive. This means that we can include the condition (S) a = b = 1.
If st ≥ 0, the sub-additivity has been already proven at (ii).
Suppose s and t have opposite signs, for example, s > 0 and t < 0. We shall use the symmetry property (S).
Write t = 1 − v. Consequently, v > 1. The claimed inequality becomes
φ(s + 1 − v) < φ(s) + φ( 1 − v) ⇔ 1 − φ(v − s) < φ(s) + 1 − φ(v) ⇔ φ(v) < φ(s) + φ(vs).
This is true if v − s ≥ 0, since v = s + (v − s) and s and v-s are positive. This holds if s + t ≤ 1. If s + t > 1, we write s = 1 − u; note that u − t < 0 and write the claimed inequality as φ(1 − u + t) < φ(1 − u) +φ(t) ⇔ 1 − φ(ut) < 1 − φ(u) + φ(t) ⇔ φ(ut) + φ(t) > φ(u), which is true again according to Lemma 1. This ends the proof. □

References

  1. Rădulescu, S.; Rădulescu, M. On the Maksa-Volkmann conjecture. In Proceedings of the 7th Congress of Romanian Mathematicians, Brasov, Romania, 29 June–5 July 2011. [Google Scholar]
  2. Horvath, A.G. On Bisectors in Minkowski Normed Spaces. Acta Math. Hungarica 2000, 89, 233–246. [Google Scholar] [CrossRef]
  3. Horvăth, A.G. Bisectors in Minkowski 3-space. Beiträge Zur Algebra Und Geom. Contrib. Algebra Geom. 2004, 45, 225–238. [Google Scholar]
  4. Tamas, M.S. Circumcenters in Real Normed Spaces. Boll. Dell’unione Mat. Ital. 2005, 15, 421–430. [Google Scholar]
  5. Guijarro, P.; Tomas, M.S. Perpendicular bisectors and orthogonality. Arch. Math. 1997, 69, 491–496. [Google Scholar] [CrossRef]
  6. Maksa, G.; Volkmann, P. Characterization of group homomorphisms having values in an inner product space. Publ. Math. Debr. 2000, 56, 197–200. [Google Scholar]
  7. Cohen, J.E.; Kemperman, J.H.B.; Zbăganu, G. Comparisons of Stochastic Matrices. with Applications in Information Theory, Statistics, Economics, and Population Sciences; Birkhäuser Boston, Inc.: Boston, MA, USA, 1998. [Google Scholar]
  8. Leopold, U.; Martini, H. Geometry of simplices in Minkowski Spaces. Results Math. 2018, 73, 1–17. [Google Scholar] [CrossRef] [Green Version]
  9. Jahn, T.; Spirova, M. On Bisectors in Normed Spaces. arXiv 2014, arXiv:1409.1833. [Google Scholar]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Zbăganu, G. On the Semi-Group Property of the Perpendicular Bisector in a Normed Space. Axioms 2022, 11, 125. https://doi.org/10.3390/axioms11030125

AMA Style

Zbăganu G. On the Semi-Group Property of the Perpendicular Bisector in a Normed Space. Axioms. 2022; 11(3):125. https://doi.org/10.3390/axioms11030125

Chicago/Turabian Style

Zbăganu, Gheorghiță. 2022. "On the Semi-Group Property of the Perpendicular Bisector in a Normed Space" Axioms 11, no. 3: 125. https://doi.org/10.3390/axioms11030125

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop