1. Introduction
The theory of gyrogroups and gyrovector spaces has been intensively developed over recent years. The structure of gyrovector subspaces and orthogonal gyrodecomposition are studied in [
1]. Topological gyrogroups are the subject of investigations in [
2]. Article [
3] is devoted to metric properties of gyrovector spaces. Several geometric inequalities in gyrovector spaces are established in [
4]. Algebraic properties of gyrogroups in Hilbert spaces are investigated in [
5]. An introduction to a theory of harmonic analysis on gyrogroups is presented in [
6]. A study of isometries in generalized gyrovector spaces is presented in [
7]. Gyrogroup actions are studied in [
8]. An application of Einstein bi-gyrogroups to quantum multi-particle entanglement is presented in [
9]. Several recent studies of gyrogroups and gyrovector spaces are presented in [
10,
11,
12]. A number of fundamental results concerning gyrovector spaces and bi-gyrovector spaces are presented in [
13,
14,
15,
16,
17,
18,
19,
20]. The main concrete examples of gyrogroups and gyrovector spaces are those induced by the Einstein addition and by Möbius addition. Interestingly, (i) Einstein gyrovector spaces are based on the Einstein addition, and they provide the algebraic setting for the Klein ball model of hyperbolic geometry. Similarly, (ii) Möbius gyrovector spaces are based on Möbius addition, and they provide the algebraic setting for the Poincaré ball model of hyperbolic geometry, just as (iii) vector spaces form the algebraic setting for the common model of Euclidean geometry.
Recently, we developed in [
21] a differential geometry approach to the theory of gyrogroups and gyrovector spaces based on local properties of underlying binary operations and, particularly, on properties of canonical metric tensors (see Definition 1) of corresponding Riemannian manifolds. It turned out to be possible to restore Einstein addition and Möbius addition from corresponding canonical metric tensors using standard tools of differential geometry. These are the parallel transport and the geodesics. Among important properties of the resulting Einstein and Möbius gyrogroups and gyrovector spaces are the left cancellation law, the existence of gyrations, the gyrocommutative law, and the left reduction law. These were proved using the differential geometry approach. Moreover, we found in [
21] a gyrogroup and a gyrovector space in the ball
, which turn out to be a group and a vector space isomorphic to the Euclidean group and space. Here we may note that any group and vector space is a gyrogroup and gyrovector space with trivial gyrations.
A gyration is a groupoid automorphism that emerges as a mathematical extension by abstraction of the special relativistic effect known as Thomas precession. It gives rise to the prefix “gyro” that we extensively use in the resulting gyroformalism. We, accordingly, prefix a gyro to any term that describes a concept in Euclidean geometry and in associative algebra to mean the analogous concept in hyperbolic geometry and in nonassociative algebra. Our gyroterminology thus conveys a world of meaning in an elegant and memorable fashion. Thus, for instance, the Einstein addition and Möbius addition in the ball are neither commutative nor associative. However, they are both gyrocommutative and gyroassociative, giving rise to gyrogroups and gyrovector spaces [
20].
The new results presented in this paper split up into three classes:
Class 1: Einstein addition and Möbius addition are isomorphic to each other, giving rise to an isomorphism between corresponding gyrogroups and gyrovector spaces. There exists a one-parameter set of binary operations that are isomorphic to the Einstein addition, and which generate gyrogroups and gyrovector spaces isomorphic to Einstein ones. Möbius addition is one of these operations. We consider the following problem. Are there operations that generate gyrogroups and gyrovector spaces isomorphic to Einstein ones, which are other than those belonging to the one parameter set? In this paper we show that there is a large class of such operations parametrized by a function satisfying some mild conditions. All such operations are described in terms of corresponding canonical metric tensors.
Class 2: Each binary operation in that we study in this paper defines sets of lines called gyrolines and cogyrolines. Gyrolines and cogyrolines are well studied for the cases of the Einstein addition and Möbius addition. We encounter here the following problem. Does the set of cogyrolines of an operation parametrized by a function coincide with the set of gyrolines of some other operations? If the answer is yes, then how can we get such operations? In this paper we prove that such operations exist, and find the canonical metric tensors of these operations.
Class 3: It is known that the Gaussian curvature of the gyrovector space generated by Einstein addition is , and by Möbius addition is . What can we say about the Gaussian curvature of the gyrovector spaces generated by the operations found in Class 2? We provide an answer to this question. We prove that the Gaussian curvature of corresponding gyrovector spaces is equal to zero.
In this paper we extend the study of the differential geometry of binary operations in the ball that we initiated in [
21]. The organization of the paper is the following. In
Section 2 we present a short description of important results in [
21], following which we introduce a set of operations isomorphic to Einstein addition. We, then, find the canonical metric tensors of these operations, enabling us to formulate an operation of scalar multiplication determined uniquely by these operations. We, thus, get the corresponding gyrovector spaces. In
Section 3 we establish important properties of these operations that correspond to similar properties of Einstein gyrogroups.
Section 4 is devoted to gyrolines and cogyrolines. We find the differential equations of the sets of gyrolines and the sets of cogyrolines for the cases of Einstein and Möbius additions. Remarkably, the operations, which we find using the sets of cogyrolines of Einstein and Möbius additions, are coincident. Moreover, they turn out to be exactly the operation that we have encountered in [
21]. We also find the corresponding operations for an arbitrary function
. In
Section 5 we employ Brioschi formula [
22] to calculate the Gaussian curvature of line elements in manifolds generated by the operations corresponding to cogyrolines. We prove that this curvature is always equal to zero. Finally, in
Section 6 we present an interesting open problem.
2. Main Definitions, Procedures and Assumptions
Let
be the open unit ball in the
n-dimensional Euclidean space
,
We seek binary operations ⊕ in
that are invariant under unitary transformations, that is, for every vectors
and a unitary
-matrix
UAssuming that the function
is differentiable, we introduce the matrix-function
given by
and
where ⊤ denotes transposition.
The matrix-function
G is viewed as a metric tensor in
. We assume that this function has the canonical form (
5) in the following formal definition.
Definition 1. (Canonical Metric Tensor).The matrix function ,, where and are scalar functions satisfying Assumptions 1 and 2 below, is said to be the canonical metric tensor in parametrized by and . Assumption 1. The functions and are differentiable, positive, and .
Assumption 2. The function satisfies the condition Then
G is also differentiable, and invariant under unitary transformations, that is, for all
and
-matrices
U such that
we have
Having such a matrix
G we can restore the binary operation ⊕ using the following procedure that we introduced in [
21]. Let
. If
, then
. If
, then
. Otherwise we perform the following four steps that lead to
.
Step 1. We calculate the vector
Step 3. We find a solution
of the differential equation
with the initial values
,
. Here
and
,
.
Step 4. Then .
For such binary operations ⊕ we defined in [
21] an operation of scalar multiplication ⊗ satisfying the following properties: for all
and numbers
we have
The operation ⊗ is unique and is defined in [
21] as follows. We introduce the following strictly increasing function
h:
,
and denote by
the function inverse to
h. Then for all
,
,
We pay special attention to the binary operation
in
of the Beltrami–Klein ball model of hyperbolic geometry, known as the Einstein addition. For all
It is shown in [
21] that this operation enjoys the following nice properties:
1. Left Cancellation Law:
2. Existence of Gyrations: for every
there exists a unitary matrix denoted by
such that for all
we have the following gyroassociative law:
3. Gyrocommutative Law:
such that
4. Left Reduction Property:
The operation
, along with the corresponding scalar multiplication
and gyrations
, forms a gyrocommutative gyrogroup and a gyrovector space, as shown in [
21].
In this paper we show that there exists a large class of binary operations ⊕ in
that satisfy properties (
15)–(
19). These operations are isomorphic to the Einstein addition
, and are parametrized by special functions
.
We now introduce the set of gyrolines
and the set of cogyrolines
We find a binary operation for which the set of gyrolines coincides with the set of cogyrolines of the Einstein addition. The same results are obtained for Möbius addition, which is isomorphic to Einstein addition.
Finally, for a curvature of the manifold with canonical metric tensors G generated by binary operations ⊕ we calculate the Gaussian curvature in terms of coefficients of G. We show that the Gaussian curvature of Einstein and Möbius additions are constant, and the Gaussian curvature of Einstein and Möbius coadditions are zero.
3. New Binary Operations that Give Rise to Gyrogroups
3.1. A Family of Binary Operations
Every binary operation for which Properties (
15)–(
19) hold, determines its gyrocommutative gyrogroup structure [
13], presented in [
21]. We extend the study of Einstein and Möbius addition as follows.
Einstein addition and Möbius addition are isomorphic to each other (in the sense of gyrogroups) since they are related by the identities in ([
15] Equations (6.325)),
Owing to the isomorphism between Einstein and Möbius addition, it is obvious that, like the Einstein addition, Möbius addition also satisfies Properties (
15)–(
19), and therefore forms a gyrogroup.
Instead of the number 2 in (
22) it is possible to place any positive number
t, thus obtaining from the Einstein addition
a new binary operation
in
, given by the equation
When , the binary operation descends to the Einstein addition and when , the binary operation descends to Möbius addition.
It seems natural that for every
the ball
with the binary operation
forms a gyrocommutative gyrogroup (that is, it satisfies Properties (
15)–(
19)). We prove below that this is, indeed, the case.
More generally, we construct in this section a large family of binary operations (parametrized by a function
) that satisfy Properties (
15)–(
19).
3.2. Operations Parametrized by Functions
Let us consider an arbitrary bijection : , which is differentiable, strictly increasing, and satisfies . Since is an increasing bijection, we have .
Then there exists an inverse smooth bijection : , and .
We now define a function
:
as follows. We set
and for every
we set
The function
is differentiable everywhere in
including zero, since
is differentiable, and Equation (
24) holds. Moreover, the function
is a smooth bijection
, there exists an inverse bijection
:
and, as it may be checked directly,
Now we introduce a new operation determined by the function
. For every
we define
where
is the Einstein addition. Obviously, this operation is isomorphic to the Einstein addition. Still, it is necessary to prove that the gyration operator is actually an operator of multiplication by a unitary matrix. We also prove below that the operation
is a special case of the operation
.
3.3. The Canonical Metric Tensor
Let us find the canonical metric tensor determined by the operation
. We have
Denoting the coefficient of
in (
30) by
, we obtain
and the canonical metric tensor is
Hence, we have in (
32) the canonical metric tensor (
5), parametrized by
and
, given by
Noticing that
and
, the second equation in (
33) may be solved for
. We further notice that
. Hence,
Now we consider a set of functions
such that functions
are equal to the same function, which we denote by
. Due to (
34) all such functions
may be parametrized by a number
. We denote such functions by
:
As we show below in
Section 3.7, for each
and the same
we can find a function
and a corresponding function
such that a binary operation generated by the canonical metric tensor
is a gyrogroup operation, satisfying Properties (
15)–(
19).
Example 1. Then, by (35),and by (36), For we getwhich is the canonical metric tensor for Einstein addition as shown in [21]. For we getwhich is the canonical metric tensor for Möbius addition as shown in [21]. 3.4. Multiplication of Vectors by Numbers
The function
h in (
12) has the form
3.5. Relations between the Functions and
In this subsection we explore the relations between the functions and for which the corresponding tensor G determines a gyrocommutative gyrogroup operation.
Let us fix a smooth positive function
such that
and
is a bijection
. We choose an arbitrary positive number
t and define
and
The pair of functions
determines a canonical metric tensor (
37) and a binary operation of a gyrogroup satisfying properties (
15)–(
19). Then
and
Let a pair of smooth functions
and
satisfy (
48), and such that the function
is increasing and
. Then this pair determines a gyrocommutative gyrogroup operation in
, as we will show in
Section 3.7.
3.6. Unitary Gyration Operator
For every binary operation isomorphic to Einstein addition ⊕ it is possible to introduce the gyration operator : . In general this operator need not be linear. Remarkably, however, the gyration operator for the operation turns out to be linear, as we will see in Lemma 1.
Lemma 1. For every function introduced in (45), the gyration operator ,is a linear operator . The matrix of this operator is unitary. Moreover,for all . Proof. We use (
26), and the fact that the matrix of the gyration operator
for Einstein addition ⊕ is unitary. For every function
described in
Section 3.2, and vectors
we have
Hence, the operator is linear, and its matrix representation is the same as the matrix representation of the operator for the Einstein addition. This matrix is unitary. Therefore, the matrix of the linear operator is also unitary. The proof of the Lemma is, thus, complete. □
The gyrolinearity of the operation follows from the fact that the matrix is unitary.
3.7. Special Properties of Operations Parametrized by
Functions
Theorem 1. The operation has the same properties as those of Einstein addition:
1. Left cancellation law: 2. Existence of gyrations: for every there exists a unitary matrix denoted by such that for all we have the following gyroassociative law: 3. For all we have the following gyrocommutative law:implying 5. Linearity of gyrations with respect to addition and multiplication:for all and . Proof. The proof follows straightforwardly from the definition of the operation
given in (
27).
Hence, Property 1 is satisfied.
Hence, Property 2 holds.
Thus, Property 3 is valid.
Thus, Property 4 is valid.
Thus, Property 5 is valid, and the proof of the Theorem is complete. □
We now check properties of gyrocommutative gyrogroups for the groupoid .
1. From the following three results, (i) identity (
52) of Theorem 1, (ii)
(see (
24)), and (iii)
for all
, we obtain the existence of a left identity, that is, for all
2. From identity (
52) of Theorem 1 with
and (
63) we obtain the existence of a left inverse, that is, for all
3. Identity (
53) of Theorem 1 implies that the binary operation
obeys the left gyroassociative law, that is, for all
4. From statement 2 of Theorem 1 we see that
is a unitary matrix for all
. Therefore, the mapping
is invertible. Identity (
57) with
shows that this mapping is an automorphism of the groupoid
.
5. Identity (
56) of Theorem 1 implies that the operator
possesses the left reduction property.
Hence, as shown in [
21], the groupoid
is a gyrogroup.
Finally, identity (
54) of Theorem 1 implies that the groupoid
is gyrocommutative so that, by [
21], it is a gyrocommutative gyrogroup.
3.8. The Canonical Metric Tensor For Coaddition
Let
be the binary operation such that for every
the solution
x of the equation
is given by
The binary operation
turns out to be
Noticing that
we see that the canonical metric tensor in the space with the binary operation
is given by
Thus, we have the canonical metric tensor
G with
In particular, for the trivial case with linear
(i.e., when
) we have
Noticing that if
for some positive number
t, as for the cases of Einstein and Möbius additions, then we see that
4. Gyrolines and Cogyrolines
Let us consider a Riemannian manifold with a canonical metric tensor
G in
,
The geodesics in this manifold are solutions of the second order differential Equation (
10), that is,
We denote by ⊕ the binary operation introduced in (
14) and assume that Assumption 2 for the function
holds, that is,
Then, the product
in (
13) is well defined and belongs to
for every
,
, and
is also well defined and belongs to
for all
.
Definition 2. For every such that the curveis called a gyroline. For the Einstein addition
the gyrolines are Euclidean intervals in
. For Möbius addition the gyrolines are circular arcs that intersect the boundary of
orthogonally. Every gyroline is a geodesic in a Riemannian manifold
with a canonical metric tensor
G. Notice that in order to get gyrolines from a binary operation
we multiply the second vector
b by numbers
t, as in (
78).
Definition 3. For every such that the curveis called a cogyroline. In this section, we face the following problem. Is it possible to find a canonical metric tensor such that cogyrolines are geodesics in the Riemannian manifold with a canonical metric tensor ?
If b is parallel to a, then cogyrolines coincide with gyrolines and are segments of Euclidean lines: . We, therefore, assume that b is not parallel to a.
4.1. Einstein Cogyrolines
In this section, we consider the Einstein addition and Einstein multiplication .
4.1.1. Elliptic Curves
For every
not parallel to
we define
We notice that every cogyroline lies in the two-dimensional plane that contains a and b.
Theorem 2. For every and b not parallel to a the cogyroline lies in an ellipse: Proof. By (
13) with
we have
where
For
we have, by (
84), (
85) and (
14),
Therefore
and for all
,
Setting
, we have
We now set
. Then
and
The proof of the Theorem is, thus, complete. □
We notice that the derivative
is not equal to zero for all
t. Therefore for every
the function
tends to
as
, and the corresponding cogyroline is a half of ellipse (semi-ellipse), represented in (
82).
4.1.2. The Canonical Metric Tensor for Cogyrolines
Let us find the canonical metric tensor for which every geodesic lies on some ellipse (
82).
We consider the following second order differential equation
Theorem 3. Every solution of Equation (92) lies on a cogyroline. Every cogyroline is a set of all points of a solution of (92) defined on . Proof. Let
x be a solution of (
92). If the vectors
and
are parallel for any point
, then
belongs to the ray
for all
, and the set of points
coincides with the interval with endpoints
, which in turn coincides with a cogyroline.
We assume that the vectors
and
are not parallel for all
, and denote by
P the two-dimensional plane that contains the vectors
and
for some
. Then
for all
. Introducing an orthogonal basis of
P, let
,
be coordinates of
in this basis. We denote by
the 2-vector function
. Then
satisfies Equation (
92).
Let us consider the functions
Since the vectors
x and
are not parallel, at least one of
and
is not equal to zero. Hence, there exists a number
such that
The value
is not equal to zero for all
t since the vectors
and
are not parallel. We define a function
:
so that
Therefore, the function is constant, .
We now define a unitary matrix
U,
Then,
, and the equation
is equivalent to
We denote by
a a vector in
parallel to
, and denote
Therefore the whole solution lies on the same cogyroline determined by the vectors and the number .
Since as , the set of points is a cogyroline. The proof of the Theorem is, thus, complete. □
In order to find a canonical metric tensor for which the solutions of Equation (
92) are geodesics, we compare Equation (
92) with
We need to find functions
and
such that for all numbers
The elegant solution to the equations in (
104) for the unknowns
and
is
Let us consider the canonical metric tensor parametrized by the functions
and
in (
105),
The geodesics of the Riemannian manifold with the canonical metric tensor (
106) satisfy Equation (
92). Hence, every geodesic in this manifold is a cogyroline, and every cogyroline is a geodesic in the Riemannian manifold with the canonical metric tensor
.
The canonical metric tensor
for the Einstein addition
is parametrized by the functions
and
given by [
21]
The functions are the same for the tensors and . The distinction lies in the function .
4.1.3. A Binary Operation for Einstein Cogyrolines
The functions
and
satisfy Assumptions 1 and 2. Hence, we follow the four steps that lead to a binary operation
in
for which the canonical metric tensor is
. We assume
. If
a and
b are parallel, then
can be defined using multiplication of vectors by numbers in (
13). We now assume that the vectors
a and
b are not parallel. In particular, they are not zero vectors. We now follow the four steps that lead to the binary operation
in
when
a and
b are not parallel.
Step 1. We evaluate the integral
Step 2. We perform a parallel transport of the vector
along the interval
. The vector parallel to
at the point
a is
Step 3. We find a solution
of Equation (
92),
with initial data
,
.
We seek a solution having the form
where
is an orthonormal basis in the plane containing the vectors
and
,
d being a number,
, and
is a scalar function to be determined. Then
and
Therefore Equation (
92) takes the form
This equation is obviously equivalent to
the general solution of which is
where
,
are arbitrary constants. Equation (
111) shows that a general solution of Equation (
92) is
The initial conditions are
where
According to Theorem 3, the function x determines a cogyroline for Einstein addition.
If we use the standard notation
for all vectors
, then (
121) may be written in the symmetric form
This operation is obviously commutative. It has been studied in [
21].
The binary operation determines the canonical metric tensor .
To define an operation of multiplication of a vector by a number, as it is shown in [
21], we have to calculate the function
,
where
is given by (
105).
In particular, if , then .
For every
,
, the cogyroline
of the Einstein addition is given by
This curve is also a gyroline for the addition
,
4.1.4. Distance and Norm For Cogyrolines
We can define a cogyronorm as a norm
as it is described in [
21]. In particular, if
, then for the function
we have
and the equalities in each line is attained if and only if there exists a non negative number
such that
or
. Here
is the Euclidean norm.
The distance between points
a and
b is given by
For arbitrary three points
we have the triangle inequality
where equality is attained if and only if these points lie on the same cogyroline, and
b is between
a and
c. Hence, we can define the cogyronorm as follows.
For this norm we have
for all
,
, and
for all
.
4.2. M öbius Cogyrolines
In this section, we consider cogyrolines for the Möbius addition
:
4.2.1. Circular Arcs
For every
if
is a two-dimensional plane that contains both
a and
b, then the cogyroline
lies in
P. If
, then the cogyroline is a point
b. If
b is parallel to
a, that is, there exists a number
such that
, then the cogyroline is a segment
.
Theorem 4. For every and not parallel to a the cogyroline is an arc of a circle that intersects the unit circle at centrally symmetric points, that is, for the vectorsandwe have Proof. Since the right-hand side of (
138) is a curve in
that does not intersect itself, and that connects the points
a and
, it is sufficient to prove that for every point
we have
We notice that (
139) is equivalent to the equation
which determines a circle with radius
.
To verify (
139) we assume that
. Then, for some
we have
Then,
, and
Let us now calculate
,
Noticing that
, we have
Adding Equations (
144) and (
145) yields (
139), that is,
and the proof of the Theorem is complete. □
4.2.2. The Canonical Metric Tensor for Cogyrolines
As in
Section 4.1.2, we consider the second order differential equation
Theorem 5. For every solution of Equation (147) the set of its points is a cogyroline. Every cogyroline is a set of points of a solution of (147) defined on . Proof. Let
and
be arbitrary non parallel vectors. Below we prove that a solution
of (
147) with initial conditions
and
lies on a circular arc
defined in (
138) for any
. Since there are no stationary points of Equation (
147) in
, and every solution
with initial conditions
can’t reach the points
, these would imply that the set of all points of the solution
coincides with
, and the statement of the theorem holds.
For a curve
we use representation (
143) to get the initial conditions
We consider a solution
of (
147) with initial conditions
and
, and choose vectors
such that
Since
where
, we can define a vector
w,
Let a number
be such that
Set , . Then for the solution we have , and is parallel to the vector tangent to the curve .
Now we can define vectors
and
. Noticing that
and
, we have
and
According to Theorem 4, a solution
x lies on the circular arc
if and only if
for all
. To prove the theorem it is sufficient to show that (
155) holds, which we accomplish in (
164).
The vector
y belongs to the two dimensional plane that contains
and
. Hence, there exist numbers
and
such that
. Solving Equations (
153) and (
154) for these numbers yields
We now introduce the functions
If
for all
, then
for all
. Furthermore, we drop for clarity the argument
. We denote by
p and
q the numerator and denominator of the sum
,
We need to show that
. We have
Since
x is a solution of Equation (
147), we have
Straightforwardly, with the value of
in (
162), we have
Hence,
for all
, and the proof of Theorem 5 is complete. □
Equation (
147) has the same form as (
5) with
,
if for all
the following equations hold
The equations in (
165) possess the solution
Let us introduce the canonical metric tensor parametrized by
and
in (
166),
The geodesics of the Riemannian manifold with metric tensor (
167) satisfy Equation (
147). Hence, every geodesic in this manifold is a cogyroline, and every cogyroline is a geodesic in the Riemannian manifold with the metric tensor
.
We notice that the canonical metric tensor
for Möbius addition is parametrized by the functions
and
given by
where, remarkably,
, as shown in [
21].
The functions are the same for the tensors and . Again, as for the case of the Einstein addition, the difference lies in the function .
4.2.3. A Binary Operation for M öbius cogyrolines
In this section, we introduce a new binary operation
such that every Möbius cogyroline is a gyroline for this operation, and vice versa, every gyroline for this operation is a Möbius cogyroline. According to Theorem 5 it is sufficient to find a smooth binary operation satisfying Condition (
2) with a canonical metric tensor (
4) equal to
given in (
167).
Introduce an operation
:
,
where
for all
. This operation is well defined, smooth, and satisfies the invariance condition (
2).
Theorem 6. The canonical metric tensor (4) of the operation coincides with . Proof. In order to use the formula (
4) we consider the first two terms of the Taylor series at a point
x of the following function:
Therefore, the matrix
in (
3) is equal to
and we get the canonical metric tensor
which coincides with the metric tensor
in (
167). The theorem is thus proved. □
Notice that the operation is commutative.
To define an operation of multiplication of a vector by a number we have to calculate the function
In particular, if with a real number , then .
For every
,
, the cogyroline
of Möbius addition is given by
This curve is also a gyroline for the addition
,
4.2.4. Distance and Norm For Cogyrolines
We can define a cogyronorm as a norm
as it is described in [
21]. For the function
h in (
173), we have,
Equalities in (
177) are attained if and only if there exists a non negative number
such that
or
. Here
is the Euclidean norm.
4.3. Cogyrolines in Spaces Parametrized by Functions
Consider again an arbitrary bijection
:
, which is differentiable, strictly increasing, and
. Following (
25) for every
we define
In
Section 3, we introduced a canonical metric tensor
, corresponding functions
,
, and a binary operation
, which has the same properties as Einstein operation
. In this section we consider cogyrolines in the space with canonical metric tensor
. The cogyrolines with parallel
and
b are intervals of the form
. Hence further in this section we assume that
a and
b are not parallel. Recall that we have the Einstein addition if
, and the Möbius addition if
.
4.3.1. A Relation with Gyrolines in the Space with Einstein Addition
The cogyroline in the space with canonical metric tensor
and corresponding binary operation
and scalar multiplication
(see
Section 3) is defined as the set
where
a and
b are arbitrary points in the open ball
.
According to the definition of the operations with subindex
, the set (
179) coincides with the set
As we have seen above for non parallel points
a and
b, the set (
181),
is the following semiellipse in the plane
containing
a and
b:
where
, and
The set of such semiellipses parametrized by points and numbers coincides with the set of all cogyrolines in the space with the binary operation .
In
Section 4.1.3 we proved (see (
126)) that there exists a binary operation
with scalar multiplication
such that every cogyroline is a gyroline in the space with the binary operation
:
Every such a line is the set of all the points of some geodesic in the space with the canonical metric tensor (
106) parametrized by the functions
Such geodesics satisfy the second order differential Equation (
92),
In this section, we are going to find a canonical metric tensor parametrized by functions and , and an equation for geodesics such that every cogyroline in the space with a canonical metric tensor is a gyroline in the space with the canonical metric tensor .
4.3.2. Description of the Set of Cogyrolines
From (
180) and (
181), it follows that every cogyroline is an image of a semiellipse
under the mapping
.
Let us assume
and
. Then
, and
The vectors
,
form an orthonormal basis in the plane
. Therefore
and (
187) is equivalent to
Introduce a vector
. Assume
is a cogyroline which belongs to the set
. Denote
Moreover, Equation (
191) parametrizes all the cogyrolines by a non zero
n-vector
d as follows. For every cogyroline there exists
d such that (
191) holds for all
, and for every
for every cogyroline for which (
191) holds for at least one number
t, it holds for all
.
4.3.3. Differential Equations for Geodesics
For the sake of clarity we drop the argument
t. Differentiating the equation
we have
and
The curve
z lies in the plane
, and the vectors
z and
are non parallel. Therefore there exist functions
,
such that
Multiplying this equation by
d from the right yields
Noticing that by (
193) the coefficient of
in (
198) is equal to zero, and solving (
198) for
yields
Equation (
10) has the form
Equations (
195) and (
200) coincide with
,
if
and
and
We add Equations (
202)–(
203), and solve Equation (
203) for
. Then we get the following system of equations for
and
as functions of
f:
The solution of the system (
204) is given by
where
C is a positive constant.
Let
be the canonical metric tensor parametrized by the functions
and let
be the binary operation in the space with this canonical metric tensor. Then the set of cogyrolines
in (
179) coincides with the set of gyrolines in the space with the canonical metric tensor
.
Furthermore, let us normalize the functions
and
such that their values at zero are equal to one. To this end we need to choose
. Then
We now recall the values of these functions in (
33) for the space with the binary operation
:
Obviously, the functions and coincide, while the functions and are different.
In particular, if
, we get the functions for Einstein cogyrolines:
If
, we get the functions for Möbius cogyrolines: