Home Hyperbolic Geometry For Non-Differential Topologists
Article Open Access

Hyperbolic Geometry For Non-Differential Topologists

  • Piotr Niemiec EMAIL logo and Piotr Pikul EMAIL logo
Published/Copyright: February 16, 2022
Become an author with De Gruyter Brill

Abstract

A soft presentation of hyperbolic spaces (as metric spaces), free of differential apparatus, is offered. Fifth Euclid’s postulate in such spaces is overthrown and, among other things, it is proved that spheres (equipped with great-circle distances) and hyperbolic and Euclidean spaces are the only locally compact geodesic (i.e., convex) metric spaces that are three-point homogeneous.

2020 Mathematics Subject Classification: Primary 51F99; Secondary 51-01

1. Introduction

Hyperbolic geometry of a plane is known as a historically first example of a non-Euclidean geometry; that is, geometry in which all of Euclid’s postulates are satisfied, apart from the fifth, called parallel, which is false. Discovered in the first half of the 19th century, it is one of the greatest mathematical achievements of those times. Although it merits special attention and mathematicians all over the world hear about it sooner or later, there are plenty of them whose knowledge on hyperbolic geometry is superficial and far from formal details. This sad truth concerns also topologists which do not specialise in differential geometry. One of reasons for this state is concerned with the extent of differential apparatus that one needs to learn in order to get to know and understand hyperbolic spaces. It was one of our two main sakes to propose and prepare an introduction to hyperbolic spaces (and geometry) that will be self-contained, free of differential language and accessible to ‘everyone’ (e.g., to accomplished mathematicians as well as to students).

The other story about hyperbolic spaces concerns their free mobility (or, in other words, absolute metric homogeneity). A metric space is absolutely homogeneous if all its partial isometries (that is, isometries between its subspaces) extend to global (bijective) isometries. According to a deep result from the 50’s of the 20th century, known almost only by differential/Riemannian geometrists, hyperbolic spaces, beside Euclidean spaces and Euclidean spheres, are (in a very strong sense) the only connected locally compact metric spaces that have this property. One may even assume less about a connected locally compact metric space – that only partial isometries between 3-point and 2-point subspaces extend to global isometries. Then such a space is ‘equivalent’ to one of the aforementioned Riemannian manifolds and therefore is automatically absolutely homogeneous (the equivalence we speak here about does not imply that spaces are isometric, but is much stronger than a statement that they are homeomorphic). So, high level of metric homogeneity makes hyperbolic spaces highly exceptional. This is another reason for putting special attention on them. The main result of the present paper (Theorem 7.1 in Section 7 below) gives a full classification, up to isometry, of all connected locally compact metric spaces that are 3-point homogeneous. Although it is a consequence of the Freudenthal’s theorem [8] from the 1950’s, this classification has never been provided in the literature (according to the best knowledge of the authors). The present paper fills this astonishing lack in the theory of homogeneous metric spaces that have been intensively studied for many years.

The paper is organised as follows. In Section 2, we introduce (one of possible models of) hyperbolic spaces and prove that their metrics satisfy the triangle inequality and are equivalent to Euclidean metrics. In the next third section, we show that hyperbolic spaces are absolutely homogeneous and admit no dilations other than isometries (for dimension greater than 1). The fourth section is devoted to a counterpart of the Lebesgue measure in the hyperbolic realm, whereas the fifth discusses in details straight lines in hyperbolic geometry. We prove there that in a hyperbolic plane there are infinitely many straight lines passing through a given point and disjoint from a fixed straight line that does not pass through the given point. We also show that onedimensional hyperbolic space is isometric to the ordinary real line. Section 6 deals with Tarski’s axioms of the plane geometry in the context of the hyperbolic plane. In particular, we show that all the axioms of this system, except for the one that corresponds to the fifth Euclid’s postulate, are satisfied. The last, seventh section is devoted to the classification (up to isometry) of all connected locally compact metric spaces that are 3-point homogeneous. All proofs are included, apart from the proof of a deep and difficult result due to Freudenthal [8] (see Theorem 7.3 below). This result is applied only once in this paper – to classify 3-point homogeneous metric spaces described above.

The reader interested in Riemannian geometry may consult, e.g., [13] or [6]. An axiomatic approach to various kinds of geometries (including Klein’s projective model, Poincaré’s model of the hyperbolic plane, and others) can be found e.g. in [7] – this material is also free of differential geometry, but totally differs from ours (it has a geometric spirit, whereas ours has a metric one). The reader interested in a presentation of hyperbolic geometry that is closer to ours is referred to [2: Subsection 1.10] by Benz, from where we took the formula (2.1) (see Section 2). Benz skips some non-trivial details, e.g. the proof of the triangle inequality for the hyperbolic metric, which is the most difficult part of the basics of hyperbolic geometry. Besides, he introduces hyperbolic spaces in a geometric fashion – starting from the isometry groups (given a priori) of these spaces. Our approach is purely metric space theoretic and we give all details.

Notation and terminology

In this paper, the term metric means a function (in two variables) that assigns to a pair of points of a fixed set their distance (so, metric does not mean Riemannian metric, a common and important notion in differential geometry).

For a pair of vectors x = (x1, …, xn) and y = (y1, …, yn) in ℝn, we denote by 〈x, y〉 their standard inner product (that is, x,y=k=1nxkyk), whereas ||x||=x,x is the Euclidean norm of x. The Euclidean metric (induced by ∥ · ∥) will be denoted as de. To simplify further arguments and statements, we introduce the following notation:

(1.1)[x]=1+x2.

By an isometric map between metric spaces we mean any function that preserves the distances, that is, ƒ: (X, dX) → (Y, dY) is isometric if dY(ƒ (p), ƒ (q)) = dX (p, q) for any p, qX. The term isometry is reserved for surjective isometric maps. Additionally, a map ƒ is a dilation if there is a constant c > 0 such that dY(ƒ(p), ƒ(q)) = cdX(p, q) for any p, qX (we do not assume that dilations are surjective). If there exists a bijective dilation between two metric spaces, they are said to be homothetic. We will denote by Iso(X, dX) the full isometry group of (X, dX); that is,

IsoX,dX=u:X,dXX,dX,u isometry .

A metric space (X, dX) is said to be geodesic (or convex) if for any two distinct points a and b of X there exists a dilation γ: [0, 1] → (X, dX) such that γ(0) = a and γ(1) = b (we do not assume the uniqueness of such γ).

For the reader’s convenience, let us recall that the inverse hyperbolic cosine is defined as cosh1(t)=log(t+t21) for t ≥ 1 (throughout this paper ‘log’ stands for the natural logarithm).

2. Hyperbolic distance

Below we introduce one of many equivalent models of hyperbolic spaces – the one most convenient for us.

DEFINITION 2.1.

The n-dimensional (real) hyperbolic space is a metric space

Hn(),dh

, where Hn(ℝ) = ℝn and dh is a metric (called hyperbolic) given by

(2.1)dh(x,y)=cosh1([x][y]x,y)x,yn

(see (1.1)).

The above formula has its origin in differential/Riemannian geometry (most often it is defined as the length of a geodesic arc – so, to get it one needs to find geodesics and compute integrals related to them), see, e.g., [14] (consult also [1] and [2: Subsection 1.10], where (2.1) appears without derivation). We underline here – at the very beginning of our presentation – that establishing the triangle inequality for dh using elementary methods is undoubtedly the most difficult part in the whole of this approach.

Remark 2.2.

Our definition of Hn(ℝ) is actually a somewhat modified Minkowski model of hyperbolic spaces. Indeed, in the last mentioned model one introduces a bilinear form

B:n+1×n+1((x0,,xn),(y0,,yn))x0y0j=1nxjyj

and deals with an n-dimensional hyperboloid

H=x=x0,,xnn+1:x0>0 and B(x,x)=1

equipped with a metric d(x,y) = cosh−1 (B(x,y)). It is straightforward to check that the function (Hn(ℝ), dh) ∋ x ↦ ([x], x) ∈ (H, d) is a well defined bijective isometry.

Remark 2.3.

As we will see in Corollary 5.2, the metric space (H1(ℝ), dh) is isometric to (ℝ, de). This contrasts with all other cases, as for n > 1 the metric space (Hn(ℝ), dh) admits no dilations other than isometries (see Theorem 3.5 below). In particular, for any integer n > 1 and positive real r ≠ 1 the metric spaces (Hn(ℝ), dh) and (Hn(ℝ), rdh) are non-isometric but homothetic. Each of the metrics rdh (with fixed r > 0) may serve as a ‘standard’ hyperbolic metric. Actually, everything that will be proved in this paper about the metric spaces (Hn(ℝ), dh) remains true when the metric is replaced by rdh.

Note also that, similarly to Euclidean geometry, for j < k the space Hj(ℝ) can naturally be considered as the subspace ℝj × {0}k–j of Hk(ℝ). Under this identification, the hyperbolic metric of Hj(ℝ) coincides with the metric induced from the hyperbolic one of Hk(ℝ). This is the main reason why we ‘forget’ the dimension n in the notations ‘dh’ and ‘de’.

The aim of this section is to show that dh is a metric on ℝn equivalent to de.

LEMMA 2.4.

For anyx,yHn(ℝ), dh(x,y) is well defined, non-negative anddh(x,y) = dh(y,x). Moreover, dh(x,y) = 0 iff x = y.

Proof. It follows from the Schwarz inequality that

(2.2)||x||2||y||2+||xy||2x,y2.

Equivalently, ∥x2y2 + ∥x2 −2 ⟨x,y⟩ + ∥y2 ≥ 〈x, y2, which gives 1 + ||x||2 + ||y||2 + ||x||2||y||2 ≥ 1 + 2 〈x, y〉 + 〈x, y2 and hence (1 + ∥x2)(1 + ∥y2) ≥ (1 + 〈x, y〉)2. Taking square roots from both sides shows that [x][y] – 〈x, y〉 ≥ 1 and thus dh(x, y) is well defined (and, of course, non-negative). Further, we see from (2.2) that [x][y] – ⟨x, y⟩ = 1 iff x = y, which gives the last claim of the lemma. Symmetry is trivial. □

To establish the triangle inequality, first we show its special case, which will be used later in the proof of a general case.

LEMMA 2.5.

For anyx, yHn(ℝ), dh(x, y) ≤ dh(x, 0) + dh (0, y), and equality appears iff eitherx = ty or y = tx for some t ≤ 0.

Proof. Observe that

dh(x,0)+dh(0,y)=cosh1([x])+cosh1([y])=log([x]+||x||)+log([y]+||y||)=log([x][y]+[x]||y||+[y]||x||+||x||||y||).

On the other hand,

dh(x,y)=cosh1([x][y]x,y)cosh1([x][y]+||x||||y||)

and equality holds in the above iff ⟨x, y⟩ = −∥x∥ ∥y∥, or, equivalently, if either x = ty or y = tx for some t ≤ 0. So, to complete the whole proof, we only need to check that

[x][y]+[x]||y||+[y]||x||+||x||||y||=expcosh1([x][y]+||x||||y||),

which is left to the reader as an elementary exercise. □

To get the triangle inequality for general triples of elements of the hyperbolic space, we will use certain one-dimensional perturbations of the identity map, which turn out to be isometries (with respect to the hyperbolic distance). They are introduced in the following.

DEFINITION 2.6.

For any yHn(ℝ) let a map Ty : Hn(ℝ) → Hn(ℝ) be defined by

Ty(x)=x+[x]+x,y[y]+1yxHn().
LEMMA 2.7.

For anyyHn(ℝ) the mapTyis bijective and fulfills the equation:

(2.3)dh(Ty(a),Ty(b))=dh(a,b)a,bHn().

Moreover,Tyis the inverse ofTy.

Proof. We start from computing [Ty(u)] for uHn(ℝ):

[Ty(u)]2=1+u+[u]+u,y[y]+1y2=1+u2+2[u]+u,y[y]+1u,y+[u]+u,y[y]+12||y||2=[u]2+2[u]u,y+2u,y2[y]+1+[u]2||y||2+2[u]u,y+u,y2[y]+1[y]21[y]+1=[u]2[y]2+2[u]u,y[y]+u,y2,

and thus

(2.4)[Ty(u)]=[u][y]+u,y.

For simplicity, putα(x)=[x]+x,y[y]+1 (then Ty(x) = x + α(x)y). Observe that (2.3) is equivalent to

[Ty(a)][Ty(b)]Ty(a),Ty(b)=[a][b]a,b,

which can easily be transformed to an equivalent form:

[Ty(a)][Ty(b)]=[a][b]+α(b)a,y+α(a)b,y+α(a)α(b)y2.

The right-hand side expression of the above equation can be transformed as follows:

[a][b]+α(b)a,y+α(a)b,y+α(a)α(b)||y||2=[a][b]+[b]a,y+b,y[y]+1a,y+[a]b,y+a,y[y]+1b,y  +[a][b]([y]21)+([b]a,y+[a]b,y)[y]21[y]+1+a,yb,y([y]+1)2([y]21)=[a][b][y]2+[b]a,y[y]+[a]b,y[y]+a,yb,y=([a][y]+a,y)([b][y]+b,y),

which equals [Ty(a)] [Ty(b)], by (2.4). So, (2.3) is proved and, combined with Lemma 2.4, implies that Ty is one-to-one. To finish the whole proof, it suffices to check that T−yTy coincides with the identity map on Hn(ℝ) (because then, by symmetry, also TyT−y will coincide with the identity map). To this end, we fix xHn(ℝ), and may and do assume that y ≠ 0. Then there exist a unique vector z and a unique real number β such that ⟨z, y⟩ = 0 and x = z + βy. Note that Ty (x) = z + ([x] + β[y])y and, consequently,

Ty(Ty(x))=z+([x]+β[y])yTy(x)([x]+β[y])([y]1)y=zTy(x)y+([x]+β[y])[y]y.

Finally, an application of (2.4) enables us continuing the above calculations to obtain:

TyTy(x)=z([x][y]+x,y)y+[x][y]+β[y]2y=zβ||y||2[y]2y=x.

We are now able to prove the main result of this section.

THEOREM 2.8.

The functiondhis a metric inducing the Euclidean topology.

Proof. To show the triangle inequality, consider arbitrary three points x, y and z of Hn(ℝ). Since Ty(0) = y, it follows from Lemmas 2.7 and 2.5 that

dh(x,z)=dhTy(x),Ty(z)dhTy(x),0+dh0,Ty(z)=dh(x,y)+dh(y,z).

Further, to establish the equivalence of the metrics, observe that

(2.5)dh(x,0)=cosh1([x])

and that both Ty and Ty are continuous with respect to the Euclidean metric (and thus they are homeomorphisms in this metric). Thus, for an arbitrary sequence xnn=1 of elements of ℝn and any x ∈ ℝn, we have (note that Tx(x) = 0):

xndhxdhxn,x0dhTxxn,00Txxn1Txxnde0xndeTx(0)=x.

The following is an immediate consequence of Theorem 2.8 (and the fact that the collections of all closed balls around 0 with respect to dh and de, respectively, coincide – only radii change when switching between dh and de). We skip its simple proof.

COROLLARY 2.9.

For eachn, the hyperbolic spaceHn(ℝ) is locally compact and connected and the metric dh is proper (that is, all closed balls are compact) and, in particular, complete.

3. Absolute (metric) homogeneity

DEFINITION 3.1.

A metric space (X, d) is said to be absolutely (metrically) homogeneous if any isometric map ƒ0 : (X0, d) → (X, d) defined on a non-empty subset X0 of X extends to an isometry ƒ: (X, d) → (X, d).

In this section, we will show that hyperbolic spaces are absolutely homogeneous. According to Theorem 7.1 (see Section 7 below), this property makes them highly exceptional among all connected locally compact metric spaces.

The following is a reformulation of Lemma 2.7.

COROLLARY 3.2.

For anyyHn(ℝ), the mapTy : (Hn(ℝ), dh) → (Hn(ℝ), dh) is an isometry.

LEMMA 3.3.

For a mapu : AHn(ℝ), where A is a subset ofHn(ℝ) containing the zero vector, the following conditions are equivalent:

  1. uis isometric (with respect todh) and u(0) = 0;

  2. u(x), u(y)⟩ = ⟨x, yfor allx, yA.

Proof. Assume (i) holds and fix x, yA. Then [u(x)] = cosh(dh(u(x), u(0))) = cosh(dh(x, 0)) = [x]. Similarly, [u(y)] = [y] and thus 〈u(x), u(y)〉 = [u(x)] [u(y)] – cosh(dh(u(x), u(y))) = [x] [y] – cosh(dh(x, y)) = 〈x, y〉. Conversely, if (ii) holds and x, yA, then 〈u(x), u(x)〉 = 〈x, x〉 and hence [u(x)] = [x] (and, similarly, [u(y)] = [y]) and u(0) = 0. But then easily dh(u(x), u(y)) = dh(x, y) and we are done. □

THEOREM 3.4.

For anynthe hyperbolic space (Hn(ℝ), dh) is absolutely homogeneous.

Proof. Fix an isometric map v: AHn(ℝ) where A is a non-empty subset of Hn (ℝ). Take any aA, put B = T−a(A) and u = Tv(a)vTa : BHn(ℝ), and observe that 0 ∈ B, u(0) = 0 and u is isometric (with respect to dh). So, the map u satisfies condition (ii) from Lemma 3.3. It is well known (and easy to show) that each such a map extends to a linear map U: ℝn → ℝn that corresponds (in the canonical basis of ℝn) to an orthogonal matrix. The last property means precisely that also U fulfills the equation from condition (ii) of Lemma 3.3. Hence, U is isometric with respect to dh. Then Tv(a)UT−a is an isometry that extends v. □

The above proof enables us to describe all isometries of the hyperbolic space Hn(ℝ): all of them are of the form u = TaU where a is a vector and U: ℝn → ℝn is an orthogonal linear map (both a and U are uniquely determined by u). In particular, the isotropy groups stab(x) = {u ∈ Iso(Hn(ℝ), dh): u(x) = x} of elements xHn(ℝ) are pairwise isomorphic and stab(0) is precisely the group On of all orthogonal linear transformations of ℝn. The group On also coincides with the isotropy group of the zero vector with respect to the isometry group of the n-dimensional Euclidean space. So, hyperbolic spaces are quite similar to Euclidean. However, the last spaces have many (bijective) dilations, which contrasts with hyperbolic geometry, as shown by

THEOREM 3.5.

Forn > 1, every dilation on (Hn(ℝ), dh) is an isometry. More generally, ifu : Hn(ℝ) → Hm(ℝ) is a dilation (and n > 1), thenuis isometric.

Proof. Assume u: Hn(ℝ) → Hm(ℝ) satisfies dh (u(x), u(y)) = cdh(x, y) for all x, yHn(ℝ) and some constant c > 0. Our aim is to show that then c = 1. (That is all we need to prove, since any global isometric map on Hn(ℝ) is onto – its inverse map is extendable to an isometry which means that actually it is an isometry.)

Replacing u by Tu(0)u, we may and do assume that u(0) = 0. Then it follows from (2.5) that for any x, yHn(ℝ):

(3.1)||u(x)||=||u(y)||  ||x||=||y||.

Further, dh(x, 0) = dh(−x, 0) and dh(x, –x) = dh(x, 0) + dh(0, –x). Consequently, dh(u(x), 0) = dh(u(–x), 0) and dh(u(x), u(–x)) = dh(u(x), 0) + dh(0, u(–x)). So, we infer from Lemma 2.5 (and from (3.1)) that u(–x) = –u(x) for all xHn(ℝ).

Fix arbitrary two vectors x, yHn(ℝ) such that 〈x, y〉 = 0. Then

coshdh(x,y)=[x][y]=[x][y]=coshdh(x,y),

thus dh (u(x), u(y)) = dh(u(x), u(–y)) = dh(u(x), –u(y)). This implies that

u(x),u(y)=0.

Now for arbitrary t ≥ 1 choose two vectors x, yHn(ℝ) such that [x] = [y] = t and 〈x, y〉 = 0. Then also 〈u(x), u(y)〉 = 0 and (by (3.1)) [u(y)] = [u(x)] = cosh(dh(u(x),0)) = cosh(cdh(x, 0)) = cosh(c cosh−1(t)). It follows from the former property that c cosh−1([x] [y]) = cdh(x, y) = dh(u(x), u(y)) = cosh−1 ([u(x)] [u(y)]), which combined with the latter yields

(3.2)coshccosh1t2=coshccosh1(t)2.

The above equation is valid for every t ≥ 1 only if c = 1. Although this is a well-known fact, for the reader’s convenience we give its brief proof. Observe that

coshccosh1t2=12t2+t41c+t2+t41c

and

coshccosh1(t)2=14t+t21c+t+t21c2.

As a consequence, limtcoshccosh1t2t2c=2c, whereas

limtcoshccosh1(t)2t2c=22c1.

So, if (3.2) holds for all t ≥ 1, then 2c = 22c−1 and c = 1. □

We underline that the claim of Theorem 3.5 is false for n = 1 (see Corollary 5.2 below).

As an immediate consequence of the above result, we obtain:

COROLLARY 3.6.

For anyn > 1, the metric spaces (Hn(ℝ), dh) and (ℝn, de) are non-isometric.

4. “Hyperbolic” measure

For the purposes of this section, let us call a positive Borel measure on a metric space (M, dM) an iso-measure if it is finite on compact sets and invariant under each isometry of (M, dM). (Recall that a Borel measure μ on a topological space X is invariant under a map u: XX if μ(u−1 (A)) = μ(A) for any Borel set AX.)

The Lebesgue measure on ℝn can be characterised in many different ways. One of them is related to metric homogeneity, namely: the Lebesgue measure is a unique (up to a scalar multiple) iso-measure on (ℝn, de). Actually, this property is a very special case of an old result due to Loomis [11, 12], applicable also to hyperbolic spaces. It asserts that every (non-empty) metrically homogenenous metric space in which all closed balls are compact admits a unique (up to a multiple constant) iso-measure. The aim of this short section is to find an iso-measure on a hyperbolic space. It can be considered as a “hyperbolic” counterpart of the Lebesgue measure. Being unique, it is a natural “ingredient” of the hyperbolic “world.”

THEOREM 4.1.

The iso-measuremhon the hyperbolic spaceHn(ℝ) and the n-dimensional Lebesgue measure are mutually absolutely continuous. More precisely, for any Borel setA ⊂ ℝn:

(4.1)mh(A)=Adx[x],
where the right-hand side integral is computed with respect to the Lebesgue measure onn.

Proof. We only need to check that the measure mh defined by (4.1) is invariant under any isometry u of Hn(ℝ). To this end, write u = TaU where UOn and a ∈ Hn(ℝ). We will show separately the invariance under U and Ta, applying the change-of-variables theorem for the Lebesgue integral.

For the orthogonal matrix U this is quite easy: substituting y = U(x), we get:

mhU1(A)=U1(A)dx[x]=Ady[U1(y)]=mh(A),

since [U−1(y)] = [y]. For Ta, we set y = Ta (x) and use the following two properties:

  • Ta1=Ta;

  • the Jacobian of T−a at y equals 1y[y]a1+[a],a (and is positive).

The latter property easily follows from the fact that the derivative of Ta at y is a linear map of the form

zzz,y[y]a1+[a]a.

So, by (2.4):

mhTa1(A)=A1y[y]a1+[a],aTa(y)dy=A1+[a]+||a||21+[a]a,y[y][a][y]a,ydy=A[a]a,y[y][a][y]a,ydy=Ady[y]=mh(A).

Denoting by B¯ϱ(a,r) the closed ball around a of radius r with respect to a metric ϱ, we have (of course) mB¯de(a,r)=α0rtn1dt where m stands for the Lebesgue measure on ℝn and α is a constant that depends only on the dimension n of the space ℝn. A counterpart of this formula for hyperbolic spaces is formulated in the next result and reveals another connection with hyperbolic functions (first such a connection is exposed in the formula for dh).

PROPOSITION 4.2.

For anyn > 0, aHn(ℝ) andR > 0:

mhB¯dh(a,R)=γ0Rsinhn1(t)dt
whereγis a constant that depends only onn.

Proof. For simplicity, set F(R)=mhB¯dh(a,R). By metric homogeneity of Hn(ℝ) and invariance of mh under any isometry, we have F(R)=mhB¯dh(0,R). Observe that B¯dh(0,R)=B¯de(0,sinh(R)). So, F(R)=B¯de(0,sinh(R))dx[x]. When n > 1, a usage of the hyperspherical variables transforms this integral to

[0,π]n2×[0,2π]0sinh(R)rn1sinn2φ1sin0φn11+r2drdφ1dφn1,

and hence F(R)=γ0sinh(R)rn11+r2dr (this formula is correct also for n = 1, with γ = 2). Now to finish the proof it remains to substitute r = sinh(t) where 0 ≤ tR and notice that then dr = cosh(t) dt and cosh(t)=1+sinh2(t). □

The proof presented above shows that for n = 1, mhB¯dh(a,R)=2R and hence the one-dimensional measure mh coincides with the one-dimensional Lebesgue measure after an isometric identification of (H1(ℝ), dh) and (ℝ, de) (cf. Corollary 5.2 below).

5. Straight lines in hyperbolic spaces

In this section, we study in detail counterparts of straight lines in hyperbolic spaces. Our final result here (Theorem 5.4 below) is a negation of a possible interpretation of Euclid’s fifth postulate. The property formulated in that result made hyperbolic geometry iconic. We discuss all the axioms of a contemporary geometry in the context of the hyperbolic plane in the next section.

Recall that a straight line in a metric space is an isometric image of the real line, and a straight line segment is an isometric image of the compact interval in ℝ. Additionally, for simplicity, we call three points a, b, c in a metric space (X, d) metrically collinear if for some x, y, z with {x, y, z} = {a, b, c} we have d(x, z) = d(x, y) + d(y, z).

To avoid confusions, straight lines in (ℝn, de) (that is, one-dimensional affine subspaces) will be called Euclidean lines, whereas straight lines in (Hn(ℝ), dh) will be called hyperbolic lines. We will use analogous names for other geometric notions.

THEOREM 5.1.

Letaandbbe two distinct points ofHn(ℝ).

  1. The metric segmentI (a, b) = {xHn (ℝ): dh(a, b) = dh(a, x) + dh(x, b)} is a unique straight line segment inHn(ℝ) that joinsaandb.

  2. The set L(a, b) of allxHn(ℝ) such thatx, a, bare metrically collinear is a unique hyperbolic line passing througha and b.

  3. Every isometric mapγ: ℝ → Hn(ℝ) that sends 0 to a is of the formγ(t) = Ta (sinh(t)z) wherezHn(ℝ) is such thatz∥ = 1.

Proof. First assume a = 0. All the assertions of the theorem in this case will be shown in a few steps.

For any zHn(ℝ) with ∥z∥ = 1 denote by γz : ℝ → Hn(ℝ) a map given by γz(t) = sinh(t)z. This map is isometric, which can be shown by a straightforward calculation:

dh(γ(s),γ(t))=cosh1([sinh(s)z][sinh(t)z]sinh(s)sinh(t))=cosh1sinh(s)2+1sinh(t)2+1sinh(s)sinh(t)=cosh1(cosh(s)cosh(t)sinh(s)sinh(t))=cosh1(cosh(st))=|st|.

Observe that the image of γz coincides with the linear span of the vector z. Thus, every Euclidean line passing through the zero vector is also a hyperbolic line.

Now let x, y, zHn(ℝ) satisfy dh(x, z) = dh(x, y) + dh(y, z) and let 0 ∈ {x, y, z}. We claim that x, y, z lie on a Euclidean line. Indeed, if y = 0, it suffices to apply Lemma 2.5 and otherwise, we may assume, without loss of generality, that x = 0. In that case, we proceed as follows. The linear span L of y is a hyperbolic line (by the previous paragraph). So, there exists an isometry q: LL that sends y to 0. By absolute homogeneity established in Theorem 3.4, there exists an isometry Q: Hn(ℝ) → Hn(ℝ) that extends q. So, Q(L) = L, Q(y) = 0 and dh(Q(x), Q(z)) = dh(Q(x), Q(y)) + dh(Q(y), Q(z)). Again, Lemma 2.5 implies that Q(z) = tQ(x) for some t ≤ 0 (as Q(x) ≠ 0 = Q(y)). But Q(x) = Q(0) ∈ Q(L) = L and therefore also Q(z) ∈ L. So, x, y, zL and we are done.

Now we prove items (A)–(C) in the case a = 0. Let zHn(ℝ) be a vector such that ∥z∥ = 1 and b ∈ γz (ℝ). It follows from the last paragraph that each element xHn(ℝ) such that x, 0, b are metrically collinear belongs to γz (ℝ). We also know that γz (ℝ) is a hyperbolic line. This shows that L(0, b) = γz(ℝ) and proves (B), from which (A) easily follows. Finally, if γ: ℝ → Hn(ℝ) is an isometric map such that γ(0) = 0, then for any t ∈ ℝ, the points 0, γ(1) and γ(t) are metrically collinear, so γ(t) ∈ L(0, γ(1)). It follows from the above argument that L(0, γ(1)) = γz(ℝ) for some unit vector zHn(ℝ). Then v=γz1γ: is a Euclidean isometry sending 0 to 0. Thus v(t) = t or v(t) = – t. In the former case, we get γ = γz, whereas in the latter we have γ = γ−z, which finishes the proof of (C).

Now we consider a general case. When a is arbitrary, ba and γ: ℝ → Hn(ℝ) is isometric and sends 0 to a, it is easy to verify that T−a(I(a, b)) = I(0, T−a(b)), T−a (L(a, b)) = L(0, T−a(b)) and T−aγ is an isometric map from ℝ into Hn(ℝ) that sends 0 to 0. So, the first part of the proof implies that I (a, b) and L(a, b) are a unique straight line segment joining a and b, and – respectively – a unique hyperbolic line passing through a and b. Similarly, T−aγ = γz for some unit vector z. Then γ = Taγz and we are done.

It follows from the above result that two hyperbolic lines either are disjoint or have a single common point, or coincide.

As a consequence of Theorem 5.1, we obtain the following result, which is at least surprising when one compares the formulas for dh and de.

COROLLARY 5.2.

The map (ℝ, de) ∋ t ↦ sinh(t) ∈ (H1(ℝ), dh) is an isometry.

The above result implies, in particular, that H1(ℝ) admits many non-isometric dilations. Its assertion is the reason for considering (in almost whole existing literature) only hyperbolic spaces of dimension greater than one.

Another immediate consequence of Theorem 5.1 reads as follows.

COROLLARY 5.3.

Hyperbolic spaces are geodesic.

The next result may be named Failure of the Classical Euclid’s Fifth Postulate.

THEOREM 5.4.

Let n > 1, Lbe a hyperbolic line inHn(ℝ) and aL. There are infinitely many hyperbolic lines that pass through a and are disjoint fromL.

Proof. Thanks to the metric homogeneity of Hn(ℝ), we may and do assume that a = 0. So, L is a hyperbolic line that does not pass through 0. We claim that then there exist two linearly independent vectors a and b such that

(5.1)L={sinh(t)a+cosh(t)b:t}.

Indeed, we know that L = γ(ℝ) where γ = Tyγz, y ≠ 0, ∥z∥ = 1 and γz is given by γz(t) = sinh(t)z. Then γ(t)=sinh(t)z+[sinh(t)z]+sinh(t)z,y[y]+1y=sinh(t)z+z,y[y]+1y+cosh(t)y.. So, substituting a=z+z,y[y]+1y and b = y, it remains to check that a and b are linearly independent to get (5.1). If a and b were not such, then L would be a subset of the linear span Y of y. But Y is a hyperbolic line and therefore we would obtain that L = Y and hence 0 ∈ L. This proves (5.1).

Now let μ ∈ ℝ be such that |μ| > 1. Put c = μa + b and let Lμ be the linear span of c. We claim that Lμ is disjoint from L. Indeed, let s and t be real and assume, on the contrary, that sc = sinh(t)a + cosh(t)b (cf. (5.1)). It follows from the linear independence of a and b that sinh(t) = and cosh(t) = s. Consequently, |sinh(t)| = |cosh(t)μ| > |cosh(t)|, which is impossible.

So, for any real μ with |μ| > 1 the set Lμ is a hyperbolic line disjoint from L and passing through 0. Since a and b are linearly independent, these sets Lμ are all different. □

The following theorem shows one more non-Euclidean property of the discussed space.

THEOREM 5.5.

For any two hyperbolic lines inH2(ℝ) there exists a hyperbolic line disjoint from both of them.

Proof. Observe that for two orthogonal vectors a and b, the hyperbolic lines Ta(ℝb) and Ta(ℝb) are contained in different components of ℝ2 \ ℝb. This, by metric homogeneity, proves that for an arbitrary hyperbolic line, each component of its complement contains a hyperbolic line.

If the given two lines K and L are disjoint (or equal), we can pick a line from the component of ℝ \ K different than the one containing L.

Now proceed to the remaining case of the two intersecting lines. We may and do assume that they intersect at 0. Let them be the linear spans of independent unit vectors, a and b, respectively. We claim that there exist β > 0 such that Tβ(a+b)(ℝ(ab)) is a hyperbolic line disjoint from ℝa and ℝb.

For convenience, let us transform the whole setting by Tβ(a+b), obtaining hyperbolic lines {Tβ(a+b)(ta): t ∈ ℝ}, {Tβ(a+b)(tb): t ∈ ℝ} and {s(ab): s ∈ ℝ}. We want to prove that for sufficiently large β neither Tβ(a+b)(ta) = s(ab) nor Tβ(a+b)(tb) = s(ab), regardless of the values of real parameters s and t. Here we focus only on the former equation (the latter can be treated analogously). Using the explicit formula for the isometry, we obtain (in the former equation):

ta+[ta]+ta,β(a+b)[β(a+b)]+1β(a+b)=s(ab).

Using linear independence of a and b, we obtain a system of two equations:

t+βt2+1+tβ2a,a+b[β(a+b)]+1=sβt2+1+tβ2a,a+b[β(a+b)]+1=s.

Adding the equations, we obtain

(5.2)t+2βt2+1+t2β2a,a+b[β(a+b)]+1=0.

Since a and b are unit vectors, 2 〈a, a + b〉 = ∥a + b2. Expanding [β(a + b)] and simplifying, we arrive at

(5.3)2βt2+1=t[β(a+b)].

Observe that the above equation is fully symmetric with respect to a and b.

Consider the asymptotic behaviour of the [ · ] function:

limβ[β(a+b)]β||a+b||=1.

Since ∥a + b∥ < 2, for sufficiently large parameter β, 2β > [β(a + b)]. Then |t[β(a+b)]|<2β|t|<2βt2+1,, hence the equation (5.3) holds for no t.

The reader interested in establishing further geometric properties by means of the metric is referred to [4].

6. Tarski’s axioms of the plane geometry

While the Euclidean geometry for centuries used to be a canonical example of deductive approach, the foundations given by Euclid came out to be not fully sufficient from the viewpoint of mathematical logic. The first successful system of axioms for the geometry was completed by David Hilbert in 1899. We will discuss a different famous system, invented by Alfred Tarski, and prove how its statements are satisfied in our model. Since hyperbolic geometry is non-Euclidean not all the axioms would hold.

Tarski’s axioms of geometry base on a universe of points and two relations (there is also the relation “=” of identity, but it is more a part of the logic framework than an element of the constructed geometry): a quaternary relation of equidistance denoted by abcd and a ternary relation of betweenness which is usually denoted as B(abc). The narrow set of primitive notions is one of the most significant advantages of Tarski’s approach.

The initial list of axioms presented in 1926 consisted of twenty statements and one axiom schema. Over the years, several axioms were shown to be derivable. Below we present the list of eleven axioms (the axiom 11 is not a first-order sentence; it can be replaced by an axiom schema while it is out of our interest in this paper) sufficient for Tarski’s system of Euclidean geometry. The order coincides with presented in [15] and [16].

  1. (Reflexivity of equidistance) For any points a, b,

    abba.
  2. (Transitivity of equidistance) For any points a, b, c, d, e, f,

    (abcdabef)cdef.
  3. (Identity for equidistance) For any points a, b, c,

    abcca=b.
  4. (Segment Construction) For any points a, b, c, q,

    x(B(qax)axbc).
  5. (Five-Segment Axiom) For any points a, b, c, d, a′, b′, c′, d′,

    abababbcbcacac           bdbdB(abd)Babd  cdcd.
  6. (Identity for betweenness) For any points a, b,

    B(aba)  a=b.
  7. (Pasch Axiom) For any points a, b, c, p, q,

    B(apb)  B(aqc)  x B(cxp) B(bxq).
  8. (Lower bound for dimension) There exist three non-collinear points.

  9. (Upper bound for dimension) For any points a, b, p, q, r,

    (abappbaqqbarrb)(B(pqr)B(qrp)B(rpq)).
  10. (Euclid’s Axiom) For any points a, b, c, d, t,

    (ad  B(adt)  B(bdc))x,y(B(abx)B(acy)B(xty)).
  11. (Axiom of Continuity) For any sets X and Y of points,

    (axXyYB(axy)) (b xX yY B(xby)).

Axioms 8 and 9 are specific for two-dimensional geometry. To determine higher dimension different statements are required. Examples of such axioms can be found e.g. in [10: Remark 4] and [15: p. 23]. Note that the form of higher dimension axioms presented in [16] (and cited in several places) is incorrect. The four-dimensional Euclidean space satisfies ‘Ax. 8(n)’ (formulated therein) for every n ≥ 2.

For further remarks on the system, its history and importance the reader is referred to [16]. A systematic development of Euclidean geometry based on Tarski’s axioms can be found in the book [15].

DEFINITION 6.1.

In the hyperbolic space (Hn(ℝ), dh), we define relations of equidistance and betweenness in the following way (a, b, c, dHn(ℝ)):

abcddh(a,b)=dh(c,d),B(abc)dh(a,b)+dh(b,c)=dh(a,c)bI(a,c).

Observe that all the relations are strictly preserved by isometries, that means for any formula ϕ(p1,…, pk) of k free variables and fixed points x1,…, xk, the following are equivalent

  1. ϕ(x1 …, xk),

  2. Φ∈Iso(Hn (ℝ), deϕ(Φ(x1),…,Φ(xk),

  3. Φ∈Iso(Hn(ℝ), de)ϕ(Φ(x1),…, Φ(xk)).

We will show that with elementary relations introduced as above, our model of hyperbolic geometry satisfies all of the Tarski’s axioms except for the Euclid’s Axiom, what shall be expected. We focus on the two dimensional case (we have listed Tarski’s axioms particularly for dimension 2), while most of the proofs apply to the n-dimensional case.

Axioms 1–3 are obvious consequences of the general properties of a distance function.

Ad 4. As we mentioned before, we can reduce to the case when a = 0 (using the isometry T−a). Then it suffices to take x = –q/∥q∥ sinh(dh(b, c)), for q ≠ 0 (cf. item (C) in Theorem 5.1). If q = 0 = a, an x suitable for arbitrary q ≠ 0 would be appropriate.

Ad 5. The Fixe-Segment Axiom follows from the metric 3-homogeneity. Under its assumptions, we can find a global isometry Φ such that Φ(a) = a′, Φ(b) = b′ and Φ(c) = c′. If then Φ(d) = d′, the assertion of the axiom is satisfied. But since ab, the isometry Φ is uniquely determined on the line L(a, b), hence indeed Φ(d) = d′.

Ad 6. By the definition of betweenness, 2dh(a, b) = dh(a, a) = 0 and then a = b.

Ad 7. Since any line on a hyperbolic plane (H2(ℝ), dh) can be mapped onto a one-dimensional vector subspace by an isometry of the plane, its complement has two connected components.

LEMMA 6.2.

Let L be a line in (H2(ℝ), dh) and letC1andC2be the two connected components of2 \ L. Fix pointsxL, yC1andz ∈ ℝ2. Then

  1. zbelongs toC1Lprovided thatB(xyz),

  2. zdoes not belong toC1provided thatB(yxz).

Proof. By metric homogeneity, we may reduce to the case when x = (0,0) and L = {(t, 0): t ∈ ℝ}. Since hyperbolic lines passing through 0 are simply vector subspaces, both cases are obvious. □

To prove that the Pasch Axiom is satisfied in (H2(ℝ), dh), we observe that b is not in the same component of ℝ2 \ L(p, c) as a and q. (One can easily verify the degenerated cases when [some of] those points lie on L(p, c).) Therefore I(b, q) has to intersect L(p, c) in a (unique) point, say x.

The same reasoning applied to c, L(q, b), a and p leads to the conclusion that I(c, p) intersects L(b, q):

I(c,p)L(b,q)=L(c,p)L(b,q)=L(c,p)I(c,p)=I(c,p)I(b,q)={x}.

Observe that the setting of the Pasch Axiom is contained within two intersecting lines, and therefore can be assumed to be contained in ℝ2 × {0}n−2Hn(ℝ), which is naturally isometric with (H2(ℝ), dh). Hence the axiom is satisfied in higher dimensions as well.

Ad 8. Axiom 8 is witnessed e.g. by the triple (0, 0), (0, 1) and (1, 0).

Ad 9. By an isometric transformation, we can reduce the issue to the case when b = –a. Observe that 2 〈x, a〉 = cosh(dh(x, –a)) – cosh(dh(x, a)), for any x ∈ ℝn. Hence all the points equidistant from a and –a are vectors perpendicular to a. In H2(ℝ) the set of all such points is a straight line and the axiom 9 follows.

Ad 10. That the Euclid’s Axiom is violated in the hyperbolic plane, can be easily deduced from Theorem 5.4 or 5.5.

We pick two lines, K and L, intersecting in a point a. Then – thanks to Theorem 5.5 – we find a third line, M, disjoint from K and L. Then we pick an arbitrary point t on M and points b, c, d (with bK and cL close to a, and dI (a, t) ⋂ I (b, c)) to complete the setting of the axiom. Points x and y from the assertion cannot be found, since one of them has to belong to the other component of ℝ2 \ M than KL (by Lemma 6.2).

Ad 11. The continuity axiom is trivial if one of the sets consists of less than two points. In the other case both sets have to be contained in a hyperbolic line which is isometric to ℝ. On the real line the assertion is satisfied.

7. Absolute vs. 3-point homogeneity

Looking at a complicated formula for the hyperbolic metric, it is a natural temptation to search for a ‘simpler’ realisation of non-Euclidean (hyperbolic) geometry. Even if it is possible, a new model will not be as ‘perfect’ as the hyperbolic space described in this paper. More precisely, if a certain metric space (X, d) non-homothetic to Hn(ℝ) defined in Section 2 was used as a model for the hyperbolic geometry, then d would not be geodesic or else (X, d) would not be 3-point homogeneous. (Recall that, for a positive integer n, a metric space (X, d) is metrically n-point homogeneous if any isometric map u: (A, d) → (X, d) defined on a subset A of X that has at most n elements is extendable to an isometry v : (X, d) → (X, d).) The above statement is a consequence of deep achievements (which we neither discuss here in full details nor give their proofs) of the 50’s of the 20th century due to Wang [18], Tits [17] and Freudenthal [8]. They classified all connected locally compact metric spaces that are 2-point homogeneous. All that follows is based on Freudenthal’s work [8]; we also strongly recommend his paper [9] where main results of the former article are well discussed (see, e.g., Subsection 2.21 therein).

To formulate the main result on the classification, up to isometry, of all connected locally compact 3-point homogeneous metric spaces, let us introduce necessary notions.

For any positive integer n let Sn stand for the Euclidean unit sphere in ℝn+1:

Sn=xn+1:||x||=1.

We equip Sn with metric ds of the great-circle distance, given by

ds(x,y)=1πarccos(x,y)=1πarccos2de(x,y)22.

It is well-known (and easy to prove) that ds is a metric equivalent to de (restricted to Sn) such that the metric space (Sn, ds) has the following properties:

  • it is geodesic and absolutely homogeneous;

  • it has diameter 1;

  • any two points of Sn whose ds-distance is smaller than 1 (that is, which are not antipodal) can be joint by a unique straight line segment.

Further, for any subinterval I of [0, ∞) containing 0 let Ω(I) be the set of all continuous functions ω: I → [0, ∞) that vanish at 0 and satisfy the following two conditions for all x, yI:

(ω1) x < yω(x) < ω(y);

(ω2) ω(x + y) ≤ ω(x) + ω(y) provided x + yI.

Finally, let Ω = Ω = Ω([0, ∞)) and Ω1 = Ω([0,1]). For any ω ∈ Ω there exists limtω(t)[0,],, which will be denoted by ω(∞).

After all these preparations, we are ready to formulate the main result of this section.

THEOREM 7.1.

Each connected locally compact 3-point homogeneous metric space having more than one point is isometric to exactly one metric space (X, d) among all listed below (everywhere belowndenotes a positive integer).

  • X = ℝn, d = ωdewhereω ∈ Ω andω(1)=min(1,12ω()).

  • X=Sn, d = ωdswhereω ∈ Ω1.

  • X = Hn(ℝ) = d = ωdhwheren > 1 and ω ∈ Ω.

In particular:

  • all connected locally compact 3-point homogeneous metric spaces are absolutely homogeneous, and each of them is homeomorphic either to a Euclidean space or to a Euclidean sphere (unless it has at most one point);

  • each locally compact geodesic 3-point homogeneous metric space having more than one point is isometric to exactly one of the spaces: (ℝn, de), (Sn,rds)(where r > 0), (Hn(ℝ), rdh) (where n > 1 and r > 0).

Remark 7.2.

Busemann’s [5] and Birkhoff’s [3, 4] results have a similar spirit. However, both of them assume, besides metric convexity, a sort of uniqueness of straight line segments joining sufficiently close two points, and their statements are less general.

It seems that the assertion of Theorem 7.1 (in this form) has never appeared in the literature. Nevertheless, we consider this result as Freudenthal’s theorem – not ours. Such a thinking is justified by the fact that Theorem 7.1 easily follows from Freudenthal’s theorem, stated below, proved in [8].

Denoting by Pn(ℝ) the n-dimensional real projective space (realised as the quotient space of Sn obtained by gluing antipodal points) equipped with a geodesic metric dp(u¯,v¯)=2πarccos(|u,v|) (where u,vSn and u¯={u,u} and v¯={v,v} denote their equivalence classes belonging to Pn(ℝ)), we can formulate Freudenthal’s Hauptsatz IV from [8] as follows.

THEOREM 7.3.

Let (Z, λ) be a connected locally compact metric space that satisfies the following condition:

There are two positive realsγandγsuch that:

  1. γ < λ(v, w) for somev, wZ;

  2. γ′ < λ(a, b) for some a, bZfor which there iscZ with λ(a, c) = λ (b, c) = γ;

  3. any isometric mapu : (A, λ) → (Z, λ) defined on an arbitrary subsetAofZof the formml:

    • A = {x, y} where λ(x, y) = γ, or

    • A = {x, y, z} where λ(x, y) = λ(y, z) = γ and λ(x, z) = γis extendable to an isometry v : (Z, λ) → (Z, λ).

Then there is a unique space (M, ϱ) among (ℝn, de) (n > 0), (Sn,ds) (n > 0), (Hn(ℝ), dh) (n > 1), (Pn(ℝ), dp) (n > 1) and a homeomorphismh: ZMsuch that:

(7.1)huh1:uIso(Z,λ)=Iso(M,ϱ).

We prove Theorem 7.1 in few steps, each formulated as a separate lemma. All of them are already known but – for the reader’s convenience – we give their short proofs.

LEMMA 7.4.

Forn > 1 the space (Pn(ℝ), dp) is 2-point, but not 3-point, homogeneous.

Proof. We use here the notation introduced in the paragraph preceding the formulation of Theorem 7.3.

Let (u, v) and (x, y) be two pairs of points of Sn such that dp(u¯,v¯)=dp(x¯,y¯). This means that there is ε ∈ {1, –1} such that ds(u, v) = ds(x, εy). Consequently, there exists AOn+1 (which is automatically an isometry with respect to ds) such that A(u) = x and A(v) = εy. Every member B of On+1 naturally induces an isometry B¯:Pn()Pn() that satisfies B¯(z¯)=B(z)¯ for any zSn. Moreover, the assignment

(7.2)On+1BB¯Iso(Pn(),dp)

is surjective (this is classical, but non-trivial). So, A¯Iso(Pn(),dp) satisfies A¯(u¯)=x¯ and A¯(v¯)=y¯, which shows that Pn(ℝ) is 2-point homogeneous.

To convince oneself that Pn(ℝ) is not 3-point homogeneous for n > 1, it is enough to consider two triples (x, y, z1) and (x, y, z2) of points of Sn with x=(1,0,0,0), y=22,22,0,0, z1=14,14,144,0  and z2=14,34,64,0 where 0 denotes the zero vector in ℝn−2 (which has to be omitted when n = 2). Observe that ⟨x, z1⟩ = ⟨x, z2⟩ and ⟨y, z1) = – ⟨y, z2⟩. Consequently, dpw¯,z¯1=dpw¯,z¯2 for any w ∈ {x, y}. If Pn(ℝ) was 3-point homogeneous, there would exist an isometry of Pn(ℝ) that fixes x and y and sends z1 onto z2, which is impossible. To see this, assume – on the contrary – there exists such an isometry. It then follows from the surjectivity of (7.2) that there exists BOn+1 such that B(x) = ±x, B(y) = ±y and B(z1) = ±z2. Then, replacing if needed B by –B, we may and do assume B(x) = x and thus, since ⟨B(x), B(u)⟩ = ⟨x, u⟩ for any u ∈ ℝn+1, also B(y) = y and B(z1) = z2. But then ⟨y, z2⟩ = ⟨B(y), B(z1)⟩ = ⟨y, z1⟩ which is false. □

LEMMA 7.5.

Let (M, ϱ) be a 2-point homogeneous geodesic metric space having more than one point and I denoteϱ(M × M), and letƒ: I → [0, ∞) be a one-to-one function such thatϱƒ = ƒϱis a metric onMequivalent toϱ. Then:

  1. Iis an interval andƒ ∈ Ω(I);

  2. (M, ϱƒ) is 2-point homogeneous;

  3. Iso(M, ϱƒ) = Iso(M, ϱ);

  4. (M, ϱƒ) is absolutely homogeneous iff so is (M, ϱ);

  5. (M, ϱƒ) is geodesic iffƒ (t) = ctfor some positive constantc (and alltI).

Proof. Since ƒ is one-to-one, any map u: AM (where AM) is isometric with respect to ϱƒ if and only if it is isometric with respect to ϱ. This implies (v1), (v2) and (v3). It follows from 1-point homogeneity of (M, ϱ) that I = {ϱ(a, x): xM} where a is arbitrarily fixed element of M. But this, combined with continuity of ϱƒ with respect to ϱ, yields that ƒ is continuous at 0. Further, since M is geodesic, I is an interval and for any x, yI with x + yI there are points a, b, cM such that ϱ(a, c) = x + y, ϱ(a, b) = x and ϱ(b, c) = y. Then ƒ (x + y) = ϱƒ (a, c) ≤ ϱƒ (a, b) + ϱƒ (b, c) = ƒ (x) + ƒ (y). So, (ω2) holds. This implies that |ω(x) – ω(y)| ≤ ω(|xy|) for any x, yI. Thus f, being continuous at 0, is continuous (at each point of I). Finally, a continuous one-to-one function vanishing at 0 satisfies (ω1) and therefore f ∈ Ω(I).

It remains to show that if ϱf is geodesic, then f is linear. Since f ∈ Ω(I), it is a bijection between I and J = f (I), and J is an interval. Moreover, ϱ = f−1ϱf. So, assuming that ϱf is geodesic, it follows from the first part of the proof that also f−1 ∈ Ω(J). But if f ∈ Ω(I) and f−1 ∈ Ω(J), then f(x + y) = f(x) + f(y) for all x, yI with x + yI. The last equation implies that f is linear (for f is continuous). □

LEMMA 7.6.

Assume (Z, λ) and (M, ϱ) are two 2-point homogeneous metric spaces having more than one point and satisfying the following three conditions:

  • (M, ϱ) is geodesic;

  • there existsR ∈ {1, ∞} such thatϱ(M × M) = {r ∈ ℝ: 0 ≤ rR}

  • there exists a homeomorphismh: ZMsuch that (7.1) holds.

Then there exists a unique ω ∈ ΩRsuch that λ = ωϱ ○ (h × h). Moreover, (Z, λ) is 3-point or absolutely homogeneous iff so is (M, ϱ).

Proof. Recall that h × h: Z × ZM × M is given by (h × h)(x, y) = (h(x), h(y)). Further, for simplicity, denote I = ϱ(M × M). It follows from our assumptions that I = [0, 1] (if R = 1) or I = [0, ∞) (if R = ∞).

For any four points a, b, x and y in an arbitrary 2-point homogeneous metric space (Y, p) the following equivalence holds:

p(a,b)=p(x,y)ΦIso(Y,p):Φ(a)=x and Φ(b)=y.

The above condition, applied for both (Z, λ) and (M, ϱ), combined with (7.1) yields that

(7.3)λ(a,b)=λ(x,y)ϱ(h(a),h(b))=ϱ(h(x),h(y))(a,b,x,yZ).

We infer that there exists a one-to-one function ω: I → [0, ∞) such that λ = ωϱ ○ (h × h). Since ωϱ = λ ○ (h−1 × h−1) is a metric equivalent to ϱ, we conclude from Lemma 7.5 that ω ∈ ΩR. The uniqueness of ω is trivial.

Finally, the remainder of the lemma (about 3-point or absolute homogeneity) follows from (7.3). Indeed, this condition implies that a map u: (A, λ) → (Z, λ) (where AZ) is isometric iff the map huh−1|h(A): (h(A), ϱ) → (M, ϱ) is isometric. □

Proof of Theorem 7.1.

First of all, since each of the spaces

(7.4)(n,de)(n>0),(Sn,ds)(n>0),(Hn(),dh)(n>1)

is absolutely homogeneous, it follows from Lemma 7.5 that any space (X, d) listed in the statement of the theorem is also absolutely homogeneous (and, of course, connected, locally compact and has more than one point). Lemma 7.5 enables us recognizing those among them that are geodesic.

Further, if (Z, λ) is a connected locally compact 3-point homogeneous metric space having more than one point, we infer from Theorem 7.3 that there are a metric space (M, ϱ) and a homeomorphism h: ZM such that (7.1) holds and either (M, ϱ) is listed in (7.4) or it is (Pn(ℝ), dp) with n > 1. However, since (M, ϱ) is 2-point homogeneous (cf. Lemma 7.4), it follows from Lemma 7.6 that it is also 3-point homogeneous (because (Z, λ) is so). Thus, Lemma 7.4 implies that (M, ϱ) is listed in (7.4), and we conclude from Lemma 7.6 that there is ω˜Ω(I) (where I = ϱ(M × M)) for which

(7.5)λ=ω˜ϱ(h×h).

Now if (M, ϱ) is not a Euclidean space, put ω=ω˜ and observe that (X, d) = (M, ωϱ) is listed in the statement of the theorem and h: (Z, λ) → (X, d) is an isometry, thanks to (7.5).

In the remaining case, when (M, ϱ) = (ℝn, de), we proceed as follows. There is α > 0 such that ω˜(α)=min1,12ω˜(). We define ω ∊ Ω by ω(t)=ω˜(αt). Observe that then ω()=ω˜() and hence ω(1)=min1,12ω(). So, (X, d) = (ℝn, ωde) is listed in the statement of the theorem. What is more, the map

(Z,λ)z1αh(z)(X,d)

is an isometry, again by (7.5) (and the definition of ω).

The last thing we need to prove is that all the metric spaces (X, d) listed in the statement of the theorem are pairwise non-isometric. To this end, let (M1, ϱ1) and (M2, ϱ2) be two spaces among listed in (7.4), ω1 and ω2 be two functions such that (X, d) = (Mj, ωjϱj) for j ∈ {1, 2} is listed in the statement of the theorem; and let h: (M1, ω1ϱ1) → (M2, ω2ϱ2) be an isometry. Then

Iso(M1,ω1ϱ1)=Iso(M1,ϱ1)=h1uh:uIsoM2,ϱ2.

So, we infer from the uniqueness in Theorem 7.3 that (M1, ϱ1) = (M2, ϱ2). To simplify further arguments, we denote (M, ϱ) = (M1, ϱ1). Thus, h is an isometry from (M, ω1ϱ) onto (M, ω2ϱ). We conclude that:

ϱ(a,b)=ϱ(x,y)ϱ(h(a),h(b))=ϱ(h(x),h(y))(a,b,x,yM)

(because both ω1 and ω2 are one-to-one). The above condition implies that there is a one-to-one function f: II where I = ϱ(M × M) such that ϱ ○ (h × h) = fϱ. Since both ϱ and ϱ ○ (h × h) are geodesic, it follows from Lemma 7.5 that there is a constant c > 0 such that f(t) = ct. So, h is a dilation on (M, ϱ). Now we consider two cases. First assume that (M, ϱ) is not a Euclidean space. Then c = 1, which for hyperbolic spaces follows from Theorem 3.5 and for spheres is trivial (just compare diameters). Thus, h ∈ Iso(M, ϱ) and therefore ω1ϱ = ω2ϱ. Consequently, ω1 = ω2.

Finally, assume (M, ϱ) = (ℝn, de). We have already known that ϱ(h(x), h(y)) = cϱ(x, y) for all x, y ∈ ℝn. Simultaneously, ω2(ϱ(h(x), h(y))) = ω1(ϱ(x, y)) for any x, y ∈ ℝn. Both these equations imply that ω2(ct) = ω1(t) for any t ≤ 0. In particular, ω2(∞) = ω1(∞) and thus ω2(c)=ω1(1)=min1,12ω1()=ω2(1). Since ω2 is one-to-one, we get c = 1 and hence ω2 = ω1. □


(Communicated by L’ubica Holá)


REFERENCES

[1] BENZ, W.: A characterization of dimension-free hyperbolic geometry and the functional equation of 2-point invariants, Aequat. Math. 80 (2010), 5–11.10.1007/s00010-010-0050-1Search in Google Scholar

[2] BENZ, W.: Classical Geometries in Modern Contexts. Geometry of Real Inner Product Spaces, 3rd ed., BirkhaÜser/Springer, Basel, 2012.10.1007/978-3-0348-0420-2Search in Google Scholar

[3] BIRKHOFF, G.: Metric foundations of geometry, Proc. Natl. Acad. Sci. USA 27 (1941), 402–406.10.1073/pnas.27.8.402Search in Google Scholar

[4] BIRKHOFF, G.: Metric foundations of geometry I, Trans. Amer. Math. Soc. 55 (1944), 465–492.10.1090/S0002-9947-1944-0010393-5Search in Google Scholar

[5] BUSEMANN, H.: On Leibniz’s definition of planes, Amer. J. Math. 63 (1941), 101–111.10.2307/2371280Search in Google Scholar

[6] CHAVEL, I.: Riemannian Geometry. A Modern Introduction, 2nd ed., Cambridge University Press, New York, 2006.10.1017/CBO9780511616822Search in Google Scholar

[7] COXETER, H. S. M.: Non-Euclidean Geometry, 6th ed., AMS, Washington DC, 1998.10.5948/9781614445166Search in Google Scholar

[8] FREUDENTHAL, H.: Neuere Fassungen des Riemann-Helmholtz-Lieschen Raumproblems, Math. Z. 63 (1956), 374–405 (in German).10.1007/BF01187949Search in Google Scholar

[9] FREUDENTHAL, H.: Lie groups in the foundations of geometry, Adv. Math. 1 (1964), 145–190.10.1016/0001-8708(65)90038-1Search in Google Scholar

[10] KORDOS, M.: On the syntactic form of dimension axiom for affine geometry, Bull. Acad. Pol. Sci. Sér. Sci. Math. 18 (1969), 833–837.Search in Google Scholar

[11] LOOMIS, L. H.: Abstract congruence and the uniqueness of Haar measure, Ann. Math. 46 (1945), 348–355.10.2307/1969028Search in Google Scholar

[12] LOOMIS, L. H.: Haar measure in uniform structures, Duke Math. J. 16 (1949), 193–208.10.1215/S0012-7094-49-01620-8Search in Google Scholar

[13] MORGAN, F.: Riemannian Geometry. A Beginner’s Guide, Jones and Bartlett Publishers, Boston, 1993.10.1201/9781315275482Search in Google Scholar

[14] REYNOLDS, W. F.: Hyperbolic geometry on a hyperboloid, Amer. Math. Monthly 100 (1993), 442–455.10.1080/00029890.1993.11990430Search in Google Scholar

[15] SCHWABHÄUSER, W.—SZMIELEW, W.—TARSKI, A.: Metamathematische Methoden in der Geometrie, Springer-Verlag, Berlin, 1983 (in German).10.1007/978-3-642-69418-9Search in Google Scholar

[16] TARSKI, —GIVANT, S.: Tarski’s system of geometry, Bull Symb Logic. Logic 5 (1999), 175–214.10.2307/421089Search in Google Scholar

[17] TITS, J.: Sur certaines classes d’espaces homogènes de groupes de Lie, Mem. Acad. Royale Belgique, Classe des Sciences XXIX, Fasc. 3 (1955), 270 pp. (in French).Search in Google Scholar

[18] WANG, H.-C.: Two-point homogeneous spaces, Ann. Math. 55 (1952), 177–191.10.2307/1969427Search in Google Scholar

Received: 2020-09-21
Accepted: 2021-02-09
Published Online: 2022-02-16
Published in Print: 2022-02-16

© 2022 Mathematical Institute Slovak Academy of Sciences

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

Downloaded on 14.9.2025 from https://www.degruyterbrill.com/document/doi/10.1515/ms-2022-0012/html
Scroll to top button