1. Introduction
Let
be the unit, open ball in the Euclidean
n-space
,
, where
is the Euclidean norm.
Einstein addition is a binary operation,
, in the ball
that stems from his velocity composition law in the ball
of relativistically admissible velocities. Seemingly structureless, Einstein addition is neither commutative nor associative. However, Einstein addition turns out to be both
gyrocommutative and
gyroassociative, thus giving rise to the rich algebraic structures that became known as a (gyrocommutative) gyrogroup and a gyrovector space, the definitions of which are presented in Definitions 1–3,
Section 2.
Einstein addition, , and its isomorphic copy, Möbius addition, , are studied in the literature algebraically, along with applications to the hyperbolic geometry of Lobachevsky and Bolyai. Naturally, one may expect that the rich algebraic structure of Einstein addition can find home in differential geometry, giving rise to a novel branch called Binary Operations in the Ball.
Accordingly, the aim of this paper is to develop a differential geometry approach to Einstein addition and, hence, to discover the resulting novel branch of differential geometry that involves binary operations in the ball. We thus begin with the study of an arbitrary binary operation in the ball that satisfies some general conditions.
A binary operation in
is a function
f:
. We consider functions
f of class
, that is, functions
f having continuous second derivatives. This operation determines a metric tensor
in
given by
, where
denotes transposition, and
is the space of all real
-matrices. Then,
is a Riemannian manifold with a metric tensor
G.
We pay special attention to the following three binary operations in the ball, along with their associated scalar multiplication:
Einstein addition
in the ball, presented in (
136), and the scalar multiplication
that it admits, presented in (
138), are recovered in
Section 5 within the framework of differential geometry. The triple
is an Einstein gyrovector space that forms the algebraic setting for the Beltrami-Klein ball model of hyperbolic geometry.
Möbius addition
in the ball, presented in (
153), and the scalar multiplication
that it admits, presented in (
154), are recovered in
Section 6 within the framework of differential geometry. The triple
is a Möbius gyrovector space that forms the algebraic setting for the Beltrami–Poincaré ball model of hyperbolic geometry.
A novel, interesting binary operation ⊕ in the ball, presented in (
199), and the scalar multiplication ⊗ that it admits, presented in (
205), are discovered in
Section 7 within the framework of differential geometry. Remarkably, the triple
is a vector space isomorphic to the Euclidean vector space
. As such, the binary operation ⊕ in
is commutative, associative and distributive. Accordingly, the vector space
forms the algebraic setting for a novel
n-dimensional Euclidean geometry ball model.
A procedure that we present in this paper enables binary operations in the ball, like , and ⊕, to be obtained from corresponding metric tensors G. Interestingly, a metric tensor G is determined by the behavior of the function f in a neighborhood of the set , rather than in the whole of the space . Hence, the global operations turn out to be determined by local properties of the functions f and their second derivatives in the set .
The procedure is formulated in terms of geodesics and a parallel transport. It is applied to a wide class of metric tensors that satisfy three properties: (i) smoothness, (ii) rotation invariance, and (iii) plane invariance.
For Einstein addition
the operation of scalar multiplication
with
,
is well defined [
1]. It gives rise to a structure called a gyrovector space. For each metric tensor considered in this paper we define an operation
of scalar multiplication, which leads to corresponding gyrovector spaces.
For every the set is called a gyroline. The set of gyrolines defines a metric tensor G. The sets are called co-gyrolines. It is proved in the paper that for operations f isomorphic to Einstein addition, the co-gyrolines are gyrolines for other binary operations, , with metric tensors that we denote by . The Gaussian curvatures of spaces with metric tensors are equal to zero.
Various algebraic and geometric properties of the operations
and
have been intensively studied in recent papers and monographs; see, for instance, References [
1,
2,
3,
4,
5,
6,
7,
8,
9,
10]).
Einstein addition and Möbius addition are neither commutative nor associative. As such, they do not form group operations. Yet, the lack of the commutative and associative laws is compensated by the
gyrocommutative and
gyroassociative laws that these binary operations obey. As such, these binary operations give rise to the algebraic objects known as
gyrogroups and
gyrovector spaces. Remarkably, gyrovector spaces form the algebraic setting for various models of analytic hyperbolic geometry, just as the standard vector spaces form the algebraic setting for analytic Euclidean geometry; see, for instance, References [
4,
5,
9,
11,
12,
13].
The special interest of our study of both Einstein addition and Möbius addition within the framework of differential geometry stems from the result that they are
gyrocommutative gyrogroup operations. Indeed they give rise to (i) Einstein gyrogroups
and Einstein gyrovector spaces
; and to (ii) Möbius gyrogroups
and Möbius gyrovector spaces
. The definitions of the abstract (gyrocommutative) gyrogroup and gyrovector space are presented in
Section 2.
The organization of the paper is as follows. In
Section 2 we present the definitions of the abstract (gyrocommutative) gyrogroup and gyrovector space. In
Section 3 we introduce metric tensors satisfying three properties: (i) smoothness; (ii) rotation invariance; and (iii) plane invariance. Then we derive the equations for geodesics and parallel transport. In
Section 4 we (i) introduce binary operations defined by metric tensors
G; (ii) define corresponding operations of scalar multiplication; and (iii) introduce distances and gyronorms. In
Section 5 we prove that the operation
f, defined in terms of the metric tensor of Einstein addition, coincides with Einstein addition. In
Section 6 and
Section 7 similar results are obtained for Möbius addition and for a novel addition.
Section 8 is devoted to properties of binary operations similar to those of Einstein addition: (i) left cancellation law; (ii) existence and properties of unitary gyrators; and (iii) a gyrocommutative law.
In the paper, we use the following notation—
is the set of real numbers.
I is the identity matrix. A real square matrix
U is called unitary [
14] if
. The function
is the signum function in
. Components of an
n-vector
x are denoted by
,
.
2. Gyrogroups and Gyrovector Spaces
Definition 1. (
(Gyrogroup), [
4] Definition 2.5)
. A groupoid is a nonempty set S with a binary operation. A groupoid is a gyrogroup if its binary operation satisfies the following axioms. In S there is at least one element, 0, called a left identity, satisfying(G1) ,
for all . There is an element satisfying axiom such that for each there is an element , called a left inverse of a, satisfying
(G2)
Moreover, for any there exists a unique element such that the binary operation obeys the left gyroassociative law
(G3)
The map given by is an automorphism of the groupoid , that is,
(G4)
and the automorphism of S is called the gyroautomorphism, or the gyration, of S generated by . The operator is called the gyrator of S. Finally, the gyroautomorphism generated by any possesses the left reduction property
(G5)
called the reduction axiom.
The gyrogroup axioms ()–() in Definition 1 split up into three classes:
The first pair of axioms, and , is a reminiscent of the group axioms.
The last pair of axioms, and , presents the gyrator axioms.
The middle axiom, , is a hybrid axiom linking the two pairs of axioms in (1) and (2).
As in group theory, we use the notation in gyrogroup theory as well.
In full analogy with groups, gyrogroups split up into gyrocommutative and non-gyrocommutative ones.
Definition 2. (
(Gyrocommutative Gyrogroup), [
4] Definition 2.6)
. A gyrogroup is gyrocommutative if its binary operation obeys the gyrocommutative law(G6) ,
for all .
The first concrete example of a gyrogroup was discovered in 1988 [
15]. It became known as an Einstein gyrogroup. Einstein gyrogroups are employed, for instance, in References [
5,
16,
17,
18,
19,
20,
21]. Möbius gyrogroups are employed as well, for instance, in References [
4,
22,
23,
24]. In full analogy with groups, there are topological gyrogroups, studied in References [
25,
26]. Gyrogroups share remarkable analogies with groups studied, for instance, in References [
27,
28,
29,
30,
31,
32]. Applications of gyrogroups in harmonic analysis are found in References [
17,
23,
33]. For other interesting studies of gyrogroups see References [
34,
35,
36,
37,
38,
39]. Einstein gyrogroups and gyrovector spaces are extended to Einstein bi-gyrogroups and bi-gyrovector spaces in References [
10,
40], along with an application to relativistic quantum entanglement of multi-particle systems.
The gyrocommutative gyrogroups that we study in this paper admit scalar multiplication, turning themselves into gyrovector spaces, the formal definition of which follows.
Definition 3. (
(Gyrovector space), [
4] Definition 6.2)
. A real inner product gyrovector space (gyrovector space in short) is a gyrocommutative gyrogroup that obeys the following axioms:(1) G is a subset of a real inner product vector space V called the carrier of G, , from which it inherits its inner product, ·, and norm, , which are invariant under gyroautomorphisms, that is,
(V0) for all .
(2) G admits a scalar multiplication, ⊗, possessing the following properties. For all real numbers and all points :
(V1) ,
(V2) ,
(V3) ,
(V4) ,
(V5) ,
(V6) ,
(V7) ,
(V8) .
Like gyrogroups, also gyrovector spaces are studied in the literature. The papers [
41,
42,
43,
44,
45] are devoted to various aspects of Möbius gyrovector spaces. Einstein and Möbius gyrovector spaces in the context of a gyrovector space approach to hyperbolic geometry are the subject of Reference [
6]. Generalized gyrovector spaces are studied in References [
46,
47,
48]. Interesting results about the differential geometry of some gyrovector spaces may be found in Reference [
49]. Other interesting studies of gyrovector spaces are found in References [
50,
51,
52,
53,
54,
55].
In this paper we introduce, within the framework of differential geometry, a large number of gyrocommutative gyrogroup operations, which enjoy key properties of Einstein addition and Möbius addition. Furthermore, we present corresponding scalar multiplications that turn these gyrocommutative gyrogroups into gyrovector spaces.
4. Binary Operations
We are now in the position to define a binary operation ⊕ in
that results from the metric tensor
G in (
6).
4.1. Vector Addition
In this subsection we define a binary operation ⊕ in the ball . Let vectors be given. If , then . If , then . For the case , we perform the following four steps that lead to the definition of .
Step 1. We calculate the value
using Formula (
35) with
a replaced by
b,
Step 2. We calculate the value
using Formula (
40) with initial condition
,
Step 3. We solve the second order differential Equation (
25),
with initial conditions
,
.
Step 4. A binary operation
is defined by the equation
We say that the binary operation ⊕ in
is generated by the metric tensor
G given by (
2).
4.2. Elementary Properties of the Binary Operations ⊕
Proof. The first equation stems from the definition of ⊕. Assume
,
,
. Applying steps 1–4, we have
Consider a solution
x of Equation (
43) with initial conditions
,
. Since the vectors
,
are parallel to
v, the solution
x has the form
, where
g is a scalar function to be determined. Equation (
43) shows that
g satisfies the conditions
The solution
g of this initial value problem satisfies the equation
Obviously, . Hence, . □
4.3. Metric Tensors Associated with Binary Operations
In this section we show that the metric tensor G that generates the operation ⊕ can be recovered from the operation ⊕.
Theorem 1. Let G be given by (6) and let ⊕ be the binary operation (44) generated by G. Furthermore, for every let be the -matrix such that Proof. The matrix
exists since the functions
and
are smooth. It is sufficient to prove that
when
, and
when
.
Let’s assume that the vectors
and
are parallel, that is, for some number
we have
. We use the procedure described in
Section 4.1 with
and
. Then, following (
40),
The vectors
and
are parallel. Hence (
25) takes the form
with initial conditions
,
. The unique solution of this initial value problem has the form
, where the scalar function
q satisfies the second order differential equation
and the initial conditions
Integrating Equation (
57) yields
Equation (
59) is separable. Integrating it over the interval
yields
Since
, we get
Thus, for
, and
we have
so that (
53) is proved.
We now assume that
. Then
, and
Let us consider the solution
x of (
25) with initial conditions
and
. For every
the vector
belongs to a two-dimensional plane that contains
and
. We introduce an orthonormal basis with the first unit vector
, and the second unit vector
, and use the notation
and
.
Then
for all
. Owing to (
25), the functions
and
satisfy the same second order differential equation
with initial conditions
where
Since the functions
and
are bounded on
and
,
, we have
uniformly for
. Denote by
a solution of Equation (
25) with initial data
Then
,
for all
,
is increasing, and
Denote by
the value
such that
. Then
. On the interval
we have
Since
,
, and
, Equations (
69) and (
70) imply
On the other hand for every differentiable scalar function
w we have
Therefore, there exist constants
and
that are independent of bounded
, and such that
Since
x and
are solutions of Equation (
66), there exist constants
,
and
independent on bounded
and such that
Taking into account that
,
, we get
Since
, for
we have
In order to prove (
54) we consider the Wronskian
First, we calculate the exponent. Since
and
, we have
Then, we calculate the values of the Wronskian:
Taking into account that
and
, we get
and
This completes the proof of the theorem. □
Theorem 1 shows that there is a one-to-one correspondence between metric tensors
G in the form (
6), and binary operations ⊕ that
G generates, as defined in
Section 4.1.
4.4. Multiplication of Vectors by Numbers
In this subsection we define a function ⊗:
(multiplication of vectors by numbers) that is compatible with the binary operation ⊕ in the sense that
for all
. Notice that
and
in (
91) are the common addition and multiplication of
and
in
. The first identity in (
91) is called the
scalar distributive law, and the second identity in (
91) is called the
scalar associative law.
Let
be two nonzero parallel vectors (that is, belong to the same line passing through the origin). We calculate
using the four steps described in
Section 4.1.
Step 2. We compute
using (
40) with
, noting that
Step 3. We integrate Equation (
25) with initial data
and
. We parametrize the solution
x as follows:
with scalar function
p. Then
,
,
. Equation (
34) implies
Following obvious simplifications we get
Step 4. We have .
Therefore
. We define a function
h as follows. For every number
p we set
Then,
and
The function h is monotonically increasing. Therefore, h is invertible, the inverse of which is denoted by .
Under this assumption the function h is a bijection . In particular, , and for every and there exists .
Now for every number
t we define
Then
and
. We assume
,
. Then
Owing to the property , the cases with and of arbitrary signs are considered similarly. Accordingly, the operation of multiplication by a number is well defined.
From the definition of scalar product (
101) it is easy to see that
and for all
,
4.5. Distances and Norms
We introduce the standard definition of distance between points a and b of the unit ball .
Definition 4. The distance between points is the minimal length of a curve connecting a and b,where the minimum is taken over all smooth functions x: with boundary conditions and . Obviously,
for all
, and
iff
since
for all
. Besides, we have the triangle inequality: for all
The value is called the norm of a, denoted by .
Lemma 2. For every there exists such that .
Proof. Fix arbitrary points
. Since
is convex, there exist curves in
connecting
a and
b. The minimum of the lengths of such curves does exist since
as
. Let this minimum be attained at a curve
q:
. Then
q is a geodesic which connects
a and
b:
,
. Consider the vector
. Make a parallel transport of this vector along the interval connecting
a and the origin. Denote the vector at the origin by
y. Consider a geodesic
w:
with initial conditions
,
. Denote
. Then according to the steps 1,2,3 described in the
Section 4.1 we have
. The lemma is proved. □
Consider a geodesic
x:
with boundary conditions
and
. It is known that
is equal to the length of this geodesic, and that along the geodesic the integrand on the right-hand side of (
107) is constant [
56]. Therefore
Let us consider a geodesic
y:
with boundary points
and
. The value of
is constant over
. Hence,
But the vector
is a parallel transport of the vector
along the curve
. Therefore,
In view of the equation
we have
Interestingly, the norm does not depend on the function . In particular, if we have two spaces with metric tensors that have the same function , then the distances between points (and hence the norms) in these two spaces coincide, and the operations of multiplication by a number also coincide. We’ll see this result in several examples, including the spaces with Einstein and Möbius additions.
Now let us consider a binary operation ⊕ applied to numbers. That is, given functions
,
, we consider the tensor
G given by (
6), and its resulting binary operation ⊕ introduced in
Section 4.1. Also, let’s consider a tensor with the same functions
and
in the one dimensional space. The corresponding operation between numbers is denoted by the same symbol, ⊕.
For arbitrary numbers
p,
q the value of
is defined by the four-steps procedure presented in
Section 4.1. If
, then
. Assume
.
Step 2. Since
, we have
Step 3. If
x is a scalar then Equation (
25) for geodesics takes the form
A solution
x of (
118) with the initial values
,
satisfies the condition
for all
.
Step 4. Since
, we have
Thus, taking into account the definition of the function
h, we get
Similarly, the operation ⊗ for numbers is defined by (
101). In particular, for arbitrary numbers
r and
p we have
Thus, for every number
r and vector
Let
. The triangle inequality implies
and equality is attained only if the vectors
a and
b belong to the same ray, that is, if there exists a positive number
such that
.
In terms of the function
h, inequality (
124) yields
We have thus proved the following theorem.
Theorem 2. For all vectors we have the gyrotriangle inequalityand equality is attained iff the vectors lie on the same ray, that is, for some nonnegative number λ we have or . Inequality (
127) may be considered as a triangle inequality in the spaces of vectors and numbers with the same addition operation ⊕.
It is shown below that for Einstein addition and all additions isomorphic to Einstein addition (for example, Möbius addition) a solution of the equation is given by . Hence, we have the following result.
Theorem 3. If a solution x of the equation is given by , then