Next Article in Journal
Where Are We Going with Statistical Computing? From Mathematical Statistics to Collaborative Data Science
Previous Article in Journal
CISA: Context Substitution for Image Semantics Augmentation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

On Translation Curves and Geodesics in Sol14

Faculty of Organization and Informatics, University of Zagreb, HR-42000 Varaždin, Croatia
*
Author to whom correspondence should be addressed.
Mathematics 2023, 11(8), 1820; https://doi.org/10.3390/math11081820
Submission received: 23 March 2023 / Revised: 7 April 2023 / Accepted: 10 April 2023 / Published: 11 April 2023
(This article belongs to the Section Algebra, Geometry and Topology)

Abstract

:
A translation curve in a homogeneous space is a curve such that for a given unit vector at the origin, translation of this vector is tangent to the curve in its every point. Translation curves coincide with geodesics in most Thurston spaces, but not in twisted product Thurston spaces. Moreover, translation curves often seem more intuitive and simpler than geodesics. In this paper, we determine translation curves in Sol 1 4 space. Their curvature properties are discussed and translation spheres are presented. Finally, characterization of geodesics in Sol 1 4 space is given.
MSC:
53C30; 53B20; 53C22

1. Introduction

A homogeneous geometry is a pair ( G , X ) consisting of a smooth manifold X, equipped with the transitive action of a Lie group G. The manifold X defines the underlying homogeneous space, and the group G defines the set of allowable motions.
A Thurston geometry  ( G , X ) is a homogeneous space such that X is connected and simply connected, G acts transitively on X with compact point stabilizers, G is not contained in any larger group of diffeomorphisms of X, and there is at least one compact manifold modeled on ( G , X ) . More about eight three-dimensional Thurston spaces can be found in [1,2].
There are 19 homogeneous four-dimensional model spaces according to Filipkiewicz [3]. Our ambient space Sol 1 4 , is among them. Furthermore, it is one of the 14 spaces admitting a complex structure compatible with the geometric structure (cf. [4]). Sol 1 4 possesses a locally conformal Kahler (LCK) structure. This structure in [5] facilitated a study of minimal invariant, totally real, and CR-submanifolds of Sol 1 4 . On the other hand, J-trajectories, which represent a generalization of geodesics, are investigated in [6].
Some homogeneous spaces allow a specific translation different from the geodesic translation. This specific translation carries a unit vector given at the origin to any point in space by its tangent mapping. The corresponding curve is called the translation curve. Molnár and Szilágyi in [7] initiated the study of translation curves along with the investigation of translation spheres in three-dimensional product and twisted product Thurston geometries. In four-dimensional Thurston spaces, translation curves have first been considered in [8]. More about geodesics and their generalizations in some other four-dimensional Thurston spaces can be found in [9,10,11].
Here, we classify translation curves and investigate geodesics in Sol 1 4 space.
This paper is organized as follows. First, we recall basic properties of Sol 1 4 space; then, we determine its translation curves and discuss their curvature properties. Finally, we consider geodesics in Sol 1 4 .

2. The Model Space Sol 1 4

2.1. Lie Group

The underlying manifold of the model space Sol 1 4 is the simply connected solvable Lie group G 8 described in [3] (p. 101).
The Lie group multiplication is given by:
( x 1 , y 1 , z 1 , t 1 ) ( x 2 , y 2 , z 2 , t 2 ) = ( x 1 + e t 1 x 2 , y 1 + e t 1 y 2 , z 1 + z 2 + e t 1 x 1 y 2 , t 1 + t 2 ) .
This operation can be derived from the following matrix multiplication:
1 0 e t 1 x 1 z 1 0 e t 1 0 x 1 0 0 e t 1 y 1 0 0 0 1 1 0 e t 2 x 2 z 2 0 e t 2 0 x 2 0 0 e t 2 y 2 0 0 0 1 = = 1 0 e ( t 1 + t 2 ) ( x 1 + e t 1 x 2 ) z 1 + z 2 + e t 1 x 1 y 2 0 e t 1 + t 2 0 x 1 + e t 1 x 2 0 0 e ( t 1 + t 2 ) y 1 + e t 1 y 2 0 0 0 1
by the identification:
( x , y , z , t ) : = 1 0 e t x z 0 e t 0 x 0 0 e t y 0 0 0 1 .
Multiplication (1) can be interpreted as a translation T of ( x 2 , y 2 , z 2 , t 2 ) to ( x 1 , y 1 , z 1 , t 1 ) . The neutral element of (1) is ( 0 , 0 , 0 , 0 ) . The inverse element of ( x , y , z , t ) is given by:
( x , y , z , t ) 1 = ( e t x , e t y , z + x y , t ) .

2.2. Metric and Basis

Using the inverse translation (3), by pullback of coordinate differentials,
1 0 x x y z 0 e t 0 e t x 0 0 e t e t y 0 0 0 1 0 0 e t ( d x x d t ) d z 0 e t d t 0 d x 0 0 e t d t d y 0 0 0 0 = 0 0 e t d x d z x d y 0 d t 0 e t d x 0 0 d t e t d y 0 0 0 0
we obtain the left invariant Riemannian metric g of Sol 1 4 :
g = e 2 t d x 2 + e 2 t d y 2 + ( d z x d y ) 2 + d t 2 .
Hence, the orthonormal coframe { ϑ 1 , ϑ 2 , ϑ 3 , ϑ 4 } is given by:
ϑ 1 = e t d x , ϑ 2 = e t d y , ϑ 3 = d z x d y , ϑ 4 = d t .
Thus, the metrically dual left invariant basis vector fields are:
e 1 = e t x , e 2 = e t y + x z , e 3 = z , e 4 = t .
Remark 1.
If we use different representation than (2), i.e.,
( x , y , z , t ) = 1 0 e t x 0 z 0 e t 0 0 x 0 0 e t 0 y 0 0 0 1 t 0 0 0 0 1 ,
then by matrix multiplication we (again) obtain the multiplication rule (1), but also obtain a straightforward presentation of the metric:
1 0 x 0 x y z 0 e t 0 0 e t x 0 0 e t 0 e t y 0 0 0 1 t 0 0 0 0 1 d z d x d y d t 0 = d z x d y e t d x e t d y d t 0 .

2.3. Levi–Civita Connection

The Levi–Civita connection is given by:
e 1 e 1 = e 4 , e 1 e 2 = 1 2 e 3 , e 1 e 3 = 1 2 e 2 , e 1 e 4 = e 1 , e 2 e 1 = 1 2 e 3 , e 2 e 2 = e 4 , e 2 e 3 = 1 2 e 1 , e 2 e 4 = e 2 , e 3 e 1 = 1 2 e 2 , e 3 e 2 = 1 2 e 1 , e 3 e 3 = 0 , e 3 e 4 = 0 , e 4 e 1 = 0 , e 4 e 2 = 0 , e 4 e 3 = 0 , e 4 e 4 = 0 .
Basis vector fields satisfy the following commutation relations:
[ e 1 , e 3 ] = [ e 2 , e 3 ] = [ e 3 , e 4 ] = 0 , [ e 1 , e 2 ] = e 3 , [ e 1 , e 4 ] = e 1 , [ e 2 , e 4 ] = e 2 .

2.4. Lie algebra

The Lie algebra g of G 8 is given by:
0 0 t 1 t 3 0 t 4 0 t 1 0 0 t 4 t 2 0 0 0 0 : t 1 , t 2 , t 3 , t 4 R .
If we take the following basis:
e 1 = 0 0 1 0 0 0 0 1 0 0 0 0 0 0 0 0 , e 2 = 0 0 0 0 0 0 0 0 0 0 0 1 0 0 0 0 , e 3 = 0 0 0 1 0 0 0 0 0 0 0 0 0 0 0 0 , e 4 = 0 0 0 0 0 1 0 0 0 0 1 0 0 0 0 0 ,
we can verify commutation relations (5).

2.5. Riemannian and Sectional Curvatures

The non-vanishing components of the Riemannian tensor are:
R ( e 1 , e 2 ) e 1 = 1 4 e 2 , R ( e 1 , e 2 ) e 3 = 1 2 e 4 , R ( e 1 , e 2 ) e 4 = 1 2 e 3 , R ( e 3 , e 1 ) e 1 = 1 4 e 3 , R ( e 4 , e 1 ) e 1 = e 4 , R ( e 2 , e 3 ) e 3 = 1 4 e 2 , R ( e 2 , e 4 ) e 4 = e 2 , R ( e 3 , e 4 ) e 1 = 1 2 e 2 .
Hence, we obtain the following sectional curvatures:
K 12 = K 13 = K 23 = 1 4 , K 24 = K 14 = 1 , K 34 = 0 .
and the scalar curvature K = 5 2 .

3. Translation Curves in Sol 1 4

3.1. Translation Curves in Sol 1 4

For a given starting unit vector ( α , β , γ , δ ) = ( x ˙ ( 0 ) , y ˙ ( 0 ) , z ˙ ( 0 ) , t ˙ ( 0 ) ) at the origin ( x ( 0 ) , y ( 0 ) , z ( 0 ) , t ( 0 ) ) = ( 0 , 0 , 0 , 0 ) , we define its image in a point ( x ( s ) , y ( s ) , z ( s ) , t ( s ) ) by the translation T such that:
1 0 e t x z 0 e t 0 x 0 0 e t y 0 0 0 1 0 0 α γ 0 δ 0 α 0 0 δ β 0 0 0 0 = 0 0 α δ x e t γ + β e t x 0 δ e t 0 α e t 0 0 δ e t β e t 0 0 0 0 .
Hence, we obtain differential equations for the curve starting at the origin in direction ( α , β , γ , δ ) :
x ˙ ( s ) = α e t ( s ) y ˙ ( s ) = β e t ( s ) z ˙ ( s ) = γ + β x ( s ) e t ( s ) , t ˙ ( s ) = δ .
Remark 2.
Using the representation from Remark 1, we could explicitly obtain components of the translated initial vector:
1 0 e t x 0 z 0 e t 0 0 x 0 0 e t 0 y 0 0 0 1 t 0 0 0 0 1 γ α β δ 0 = γ + β x e t α e t β e t δ 0 .
Next, we solve the system (6). From the fourth equation we have t ( s ) = δ s . Substituting this into other equations of (6), after integration we obtain the following result.
Theorem 1.
Translation curves in Sol 1 4 space starting at the origin are given by the following parametric equations:
(a) 
x ( s ) = α δ e δ s 1 , y ( s ) = β δ e δ s 1 , z ( s ) = α β + γ δ δ s + α β δ 2 e δ s 1 , t ( s ) = δ s , for δ 0 R ,
(b) 
x ( s ) = α s , y ( s ) = β s , z ( s ) = γ s + α β 2 s 2 , t ( s ) = 0 , for δ = 0 ,
where α , β , γ R , such that α 2 + β 2 + γ 2 + δ 2 = 1 .
Remark 3.
Note that in the second case ( δ = 0 ) we consider translation curves in hyperspace isometric to the Nil 3 space. The obtained translation curves in this case coincide with the translation curves described in [7].

3.2. Curvature Properties of Translation Curves

Definition 1.
For a curve c in Sol 1 4 space parameterized by arc length s, we say that c is a Frenet curve of osculating order r ( r = 1 , , 4 ) if there exist orthonormal vector fields E 1 , E 2 , E 3 , and E 4 along c, such that:
c ˙ = E 1 , c ˙ E 1 = κ E 2 , c ˙ E 2 = κ E 1 + τ E 3 , c ˙ E 3 = τ E 2 + σ E 4 , c ˙ E 4 = σ E 3 ,
where κ, τ, and σ are positive C functions of s.
Vector fields E 1 , E 2 , E 3 , and E 4 are called the tangent, the normal, the binormal, and the trinormal vector field of the curve c, respectively. Functions κ ( s ) , τ ( s ) , and σ ( s ) are called the first, the second and the third curvature of c, respectively.
A geodesic is regarded as a Frenet curve of osculating order 1.
A helix of order 2 is a Frenet curve of osculating order 2 with constant curvature κ , i.e., it is a circle.
A helix of order 3 is a Frenet curve of osculating order 3 with constant curvatures κ and τ , i.e., it is a circular helix.
A helix of order 4 is a Frenet curve of osculating order 4 such that all curvatures κ , τ , and σ are constant (e.g., see [12]).
Next, we determine the curvatures of translation curves.
We start with the unit velocity vector E 1 = c ˙ = α e 1 + β e 2 + γ e 3 + δ e 4 . Using (4), we have:
c ˙ E 1 = ( β γ α δ ) e 1 + ( β δ α γ ) e 2 + ( α 2 β 2 ) e 4 .
By c ˙ E 1 = κ E 2 , we have:
κ 2 = ( α 2 β 2 ) 2 + ( α 2 + β 2 ) ( γ 2 + δ 2 ) 4 α β γ δ = c o n s t .
and:
E 2 = 1 κ ( β γ α δ ) e 1 + ( β δ α γ ) e 2 + ( α 2 β 2 ) e 4 .
Remark 4.
Notice that κ is zero in at least two cases. The first case is α = β = 0 and the second case is α = ± β and γ = ± δ . The translation curves obtained for α = β = 0 coincide with geodesics we will describe later in Case 1 of the proof of Theorem 3.
Next, using α 2 + β 2 + γ 2 + δ 2 = 1 , we can rewrite the first curvature as:
κ 2 = ( α 2 + β 2 ) 4 α β ( α β + γ δ ) .
Furthermore, using (4) and (7), we calculate:
c ˙ E 2 = 1 κ { γ 2 β δ α γ α ( α 2 β 2 ) e 1 + β ( α 2 β 2 ) γ 2 ( β γ α δ ) e 2 + + α β δ γ 2 ( α 2 + β 2 ) e 3 + 2 α β γ δ ( α 2 + β 2 ) e 4 } .
Then, from c ˙ E 2 = κ E 1 + τ E 3 , we have:
τ E 3 = 1 κ { γ 2 ( β δ α γ ) α ( α 2 β 2 ) + κ 2 α e 1 + β ( α 2 β 2 ) γ 2 ( β γ α δ ) + κ 2 β e 2 + + α β δ γ 2 ( α 2 + β 2 ) + κ 2 γ e 3 + 2 α β γ δ ( α 2 + β 2 ) + κ 2 δ e 4 } = 1 κ { γ 2 ( β δ α γ ) + 2 α β 2 4 α 2 β ( α β + γ δ ) e 1 + + γ 2 ( β γ α δ ) + 2 α 2 β 4 α β 2 ( α β + γ δ ) e 2 + + α β δ + γ 2 ( α 2 + β 2 ) 4 α β γ ( α β + γ δ ) e 3 + + 2 α β γ 4 α β δ ( α β + γ δ ) e 4 } .
Hence, we obtain that the second curvature is constant.
τ 2 = 1 κ 2 ( α 2 + β 2 ) 4 α 2 β 2 + c 2 4 + 3 α β γ δ 16 α 2 β 2 ( α β + γ δ ) 2 + α β δ ( α β δ γ 3 ) .
Next, we find:
c ˙ E 3 = 1 κ τ { [ 4 α β ( α β + γ δ ) ( α δ β γ ) + 1 2 α β ( β δ 2 α γ ) + + γ 4 α ( α β + γ δ ) + β ( β 2 γ 2 ) ] e 1 + + 4 α β ( α β + γ δ ) ( α γ β δ ) + 1 2 α β ( 2 β γ α δ ) γ 4 ( β ( α β + γ δ ) + α ( α 2 γ 2 ) ) e 2 + + ( α 2 β 2 ) α β + 1 4 γ δ e 3 + + ( β 2 α 2 ) 4 α β ( α β + γ δ ) + 1 2 γ 2 e 4 } .
Then, from c ˙ E 3 = τ E 2 + σ E 4 we have:
σ E 4 = 1 κ τ { 4 α β ( α β + γ δ ) ( α δ β γ ) + 1 2 α β ( β δ 2 α γ ) + γ 4 α ( α β + γ δ ) + β ( β 2 γ 2 ) τ 2 ( β γ α δ ) e 1 + + 4 α β ( α β + γ δ ) ( α γ β δ ) + 1 2 α β ( 2 β γ α δ ) γ 4 β ( α β + γ δ ) + α ( α 2 γ 2 ) τ 2 ( β δ α γ ) e 2 + + ( α 2 β 2 ) α β + 1 4 γ δ e 3 + + ( β 2 α 2 ) 4 α β ( α β + γ δ ) + 1 2 γ 2 τ 2 ( α 2 β 2 ) e 4 } .
Even though we are unable to express σ E 4 without κ and τ , we can conclude that σ = c o n s t . From (8) we obtain that σ 2 can be expressed in the following form:
σ 2 = 1 κ 2 τ 2 2 + 2 + ( α 2 β 2 ) ( α β + 1 4 γ δ ) 2 + 2 .
Thus, we conclude that the third squared term in the braces in general is not zero. Hence, σ = c o n s t 0 .
Thus, we have proved the following theorem.
Theorem 2.
Translation curves in Sol 1 4 space are helices of order 4.
Remark 5.
Note that translation curves in Sol 0 4 space are helices of order 3, i.e., circular helices [8].

3.3. Translation Spheres in Sol 1 4

Let us assume that the initial unit vector of translation curve (7) is given by:
α = sin ϑ cos φ cos ψ , β = sin ϑ cos φ sin ψ , γ = sin ϑ sin φ , δ = cos ϑ .
Then, we can define the sphere of radius R with its center at the origin. Namely, the unit velocity translation curves ending in parameter R describe the translation spheres.
Proposition 1.
The translation sphere of radius R in Sol 1 4 space is given by the following parametric equations:
x ( ϑ , φ , ψ ) = tan ϑ cos φ cos ψ e R cos ϑ 1 , y ( ϑ , φ , ψ ) = tan ϑ cos φ sin ψ e R cos ϑ 1 , z ( ϑ , φ , ψ ) = R tan ϑ 1 2 sin ϑ cos 2 φ sin 2 ψ + cos ϑ sin φ + 1 2 tan 2 ϑ cos 2 φ sin 2 ψ e R cos ϑ 1 , t ( ϑ , φ , ψ ) = R cos ϑ ,
where ϑ , φ , ψ [ 0 , 2 π ) , and R R + .

4. Geodesics in Sol 1 4

Here we consider geodesics in Sol 1 4 . Let c ( s ) = ( x ( s ) , y ( s ) , z ( s ) , t ( s ) ) be a curve in Sol 1 4 parameterized by arc length. Its unit tangent vector field is:
c ˙ ( s ) = x ¨ ( s ) x + y ˙ ( s ) y + z ˙ ( s ) z + t ˙ ( s ) t = e t ( s ) x ˙ ( s ) e 1 + e t ( s ) y ˙ ( s ) e 2 + ( z ˙ ( s ) x ( s ) y ˙ ( s ) ) e 3 + t ˙ ( s ) e 4 .
The arc-length condition is:
e 2 t ( s ) x ˙ ( s ) 2 + e 2 t ( s ) y ˙ ( s ) 2 + ( z ˙ ( s ) x ( s ) y ˙ ( s ) ) 2 + t ˙ ( s ) 2 = 1 .
Next, using (4), we find the acceleration:
c ˙ ( s ) c ˙ ( s ) = x ¨ ( s ) 2 x ˙ ( s ) t ˙ ( s ) e t + y ˙ ( s ) ( z ˙ ( s ) x ( s ) y ˙ ( s ) ) e t ( s ) e 1 + + y ¨ ( s ) + 2 y ˙ ( s ) t ˙ ( s ) e t x ˙ ( s ) ( z ˙ ( s ) x ( s ) y ˙ ( s ) ) e t ( s ) e 2 + + z ¨ ( s ) x ˙ ( s ) y ˙ ( s ) x ( s ) y ¨ ( s ) e 3 + + t ¨ ( s ) y ˙ ( s ) 2 e 2 t ( s ) x ˙ ( s ) 2 e 2 t ( s ) e 4 .
From the geodesic equation c ˙ ( s ) c ˙ ( s ) = 0 , we obtain the following system:
x ¨ ( s ) 2 x ˙ ( s ) t ˙ ( s ) = y ˙ ( s ) ( z ˙ ( s ) x ( s ) y ˙ ( s ) ) e 2 t ( s ) , y ¨ ( s ) + 2 y ˙ ( s ) t ˙ ( s ) = x ˙ ( s ) ( z ˙ ( s ) x ( s ) y ˙ ( s ) ) e 2 t ( s ) , z ¨ ( s ) = x ˙ ( s ) y ˙ ( s ) + x ( s ) y ¨ ( s ) , t ¨ ( s ) = y ˙ ( s ) 2 e 2 t ( s ) x ˙ ( s ) 2 e 2 t ( s ) .
Let the initial conditions be:
( x ( 0 ) , y ( 0 ) , z ( 0 ) , t ( 0 ) ) = ( x 0 , y 0 , z 0 , t 0 ) and ( x ˙ ( 0 ) , y ˙ ( 0 ) , z ˙ ( 0 ) , t ˙ ( 0 ) ) = ( α , β , γ , δ ) ,
where α , β , γ , δ R .
We state the following theorem.
Theorem 3.
Geodesics in Sol 1 4 geometry are solutions of the system of differential equations (9). In particular, some analytical solutions of System (9) with respect to initial conditions (10) are curves:
1. 
c 1 ( s ) = ( x 0 , y 0 , γ s + z 0 , δ s + t 0 ) ;
2. 
c 2 ( s ) = ( tanh ( α s ) , y 0 , z 0 , ln cosh ( α s ) ) ;
3. 
c 3 ( s ) = ( β e 2 t 0 s + x 0 , β s + y 0 , x 0 β s + 1 2 e 2 t 0 β 2 s 2 , t 0 ) ;
4. 
c 4 ( s ) = β C e C s + 2 t 0 , β C e C s , C s β 2 2 C 2 e 2 ( C s + t 0 ) , t 0 , for C = γ x 0 β 0 , where x 0 , y 0 , z 0 , t 0 , α , β , γ , δ R .
Proof. 
Integrating the third equation of (9), with respect to (10), we have:
z ˙ x y ˙ = C where C = γ x 0 β .
Substituting (11) into the first and the second equations of (9), we obtain:
x ¨ 2 x ˙ t ˙ = C y ˙ e 2 t , y ¨ + 2 y ˙ t ˙ = C x ˙ e 2 t .
Adding these equations after respective multiplication by y ˙ , x ˙ gives:
x ¨ y ˙ + x ˙ y ¨ = C x ˙ 2 e 2 t y ˙ 2 e 2 t ,
which through the fourth equation of (9) reduces to:
1 1 s ( x ˙ y ˙ ) = C t ¨ .
Unfortunately, solving this system presents a considerable challenge in the general case. We will look for geodesic lines in characteristic hypersurfaces of Sol 1 4 space.
Case 1: x = x 0
From the first equation of (12), for C 0 , we have y = y 0 (and hence β = 0 ) . Then, (11) and (14) imply z ( s ) = γ s + z 0 and t ( s ) = δ s + t 0 . Hence,
c 1 ( s ) = ( x 0 , y 0 , γ s + z 0 , δ s + t 0 ) .
If C = 0 , then from the second equation of (12) and the fourth equation of (9) we obtain the system y ¨ + 2 y ˙ t ˙ = 0 , t ¨ = y ˙ 2 e 2 t . Integrating the first equation, we have y ˙ = β e 2 t . Substituting this in the second equation, we obtain t ¨ β 2 · e 2 t = 0 . This equation only has a complex solution.
Case 2: y = y 0
From the second equation of (12), for C 0 , we have x = x 0 . Then, (11) and (14) again imply z ( s ) = γ s + z 0 , and t ( s ) = δ s + t 0 . If C = 0 , then from the first equation of (12) and the fourth equation of (9) we obtain the system x ¨ 2 x ˙ t ˙ = 0 , t ¨ = x ˙ 2 e 2 t . Integrating the first equation, we have x ˙ = α e 2 t . Substituting in the second equation we obtain t ¨ + α 2 · e 2 t = 0 . The solution of this equation is t ( s ) = ln cosh ( α s ) . Hence, x ˙ ( s ) = α cosh 2 ( α s ) and x ( s ) = tanh ( α s ) . From the second equation of (9) we have z ( s ) = z 0 . Thus, we obtain:
c 2 ( s ) = ( tanh ( α s ) , y 0 , z 0 , ln cosh ( α s ) ) .
Case 3: z = z 0
From (11) it follows that x y ˙ = c o n s t . = C . If C = 0 , then we have x = 0 or y ˙ = 0 . Both cases are already considered in Case 1 and Case 2, respectively. If C 0 we differ two subcases. In the first subcase we assume that both factors are constants, i.e., x = x 0 and y ˙ = c o n s t . = β . In this subcase, the first equation of (9) leads to a contradiction. In the second subcase we assume that both factors are not constants. Then we can repeat the procedure from (12) and obtain (13). Unfortunately, we could not find a general solution for the obtained system.
Case 4: t = t 0
Once more, we consider two cases regarding C = γ x 0 β . For C = 0 , the system (9) is rather simple:
x ¨ = 0 , y ¨ = 0 , z ˙ = x y ˙ , e 2 t 0 y ˙ 2 e 2 t 0 x ˙ 2 = 0 .
Its solution is the following curve:
c 3 ( s ) = α s + x 0 , β s + y 0 , γ s + α β 2 s 2 + z 0 , t 0 , where α = β e 2 t 0 and γ = x 0 β .
If C 0 , then from the fourth equation of (9), we obtain y ˙ = e 2 t 0 x ˙ . Substitution of this in the first equation of (9), followed by integration, obtains:
x ( s ) = α C e C s and hence , y ( s ) = β C e C s , where β = α e 2 t 0 .
From the third equation of (9), we have z ( s ) = C s + α β 2 C 2 e 2 C s . Hence,
c 4 ( s ) = α C e C s , β C e C s , C s + α β 2 C 2 e 2 C s , t 0 , where C = γ x 0 β , a = β e 2 t 0 .
Note that for t = 0 , the system (9) is:
x ¨ = C y ˙ , y ¨ = C x ˙ , C = z ˙ x y ˙ , y ˙ 2 = x ˙ 2 .
The first two equations suggest a solution in sine and cosine functions, but then the fourth equation leads to a contradiction. Therefore, the only possible solution, for C = 0 , has been presented already. □

Author Contributions

Writing—original draft, Z.E. and M.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The authors would like to thank the referees for their valuable comments and suggestions.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Scott, P. The geometries of 3-manifolds. Bull. London Math. Soc. 1983, 15, 401–487. [Google Scholar] [CrossRef]
  2. Thurston, W.M. Three-Dimensional Geometry and Topology, Volume I; Levy, S., Ed.; Princeton Mathematical Series; Princeton University Press: Princeton, NJ, USA, 1997. [Google Scholar]
  3. Filipkiewicz, R. Four Dimensional Geometries. Ph.D. Thesis, University of Warwick, Coventry, UK, 1983. [Google Scholar]
  4. Wall, C.T.C. Geometric structures on compact complex analytic surfaces. Topology 1986, 25, 119–153. [Google Scholar] [CrossRef] [Green Version]
  5. Erjavec, Z.; Inoguchi, J. Minimal submanifolds in Sol 1 4 . RACSAM 2022. submitted. [Google Scholar]
  6. Erjavec, Z.; Inoguchi, J. J-trajectories in 4-dimensional solvable Lie groups Sol 1 4 . J. Nonlinear Sci. 2022. submitted. [Google Scholar]
  7. Molnár, E.; Szilágyi, B. Translation curves and their spheres in homogeneous geometries. Publ. Math. Debrecen 2011, 78, 327–346. [Google Scholar] [CrossRef]
  8. Erjavec, Z. Geodesics and translation curves in Sol 0 4 . Mathematics 2023, 11, 1533. [Google Scholar] [CrossRef]
  9. De Leo, B.; Marinosci, R.A. Homogeneous geodesics of four-dimensional generalized symmetric pseudo-Riemannian spaces. Publ. Math. Debrecen 2008, 73, 341–360. [Google Scholar] [CrossRef]
  10. Erjavec, Z.; Inoguchi, J. J-trajectories in 4-dimensional solvable Lie groups Sol 0 4 . Math. Phys. Anal. Geom. 2022, 25, 8. [Google Scholar] [CrossRef]
  11. Inoguchi, J. J-trajectories in locally conformal Kähler manifolds with parallel anti Lee field. Int. Electron. J. Geom. 2020, 13, 30–44. [Google Scholar] [CrossRef]
  12. Maeda, S.; Adachi, T. Holomorphic helices in a complex space form. Proc. Amer. Math. Soc. 1997, 125, 1197–1202. [Google Scholar] [CrossRef] [Green Version]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Erjavec, Z.; Maretić, M. On Translation Curves and Geodesics in Sol14. Mathematics 2023, 11, 1820. https://doi.org/10.3390/math11081820

AMA Style

Erjavec Z, Maretić M. On Translation Curves and Geodesics in Sol14. Mathematics. 2023; 11(8):1820. https://doi.org/10.3390/math11081820

Chicago/Turabian Style

Erjavec, Zlatko, and Marcel Maretić. 2023. "On Translation Curves and Geodesics in Sol14" Mathematics 11, no. 8: 1820. https://doi.org/10.3390/math11081820

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

Article Metrics

Back to TopTop