Next Article in Journal
The Barium Odd Isotope Fractions in Seven Ba Stars
Next Article in Special Issue
Optical Solitons and Conservation Laws for the Concatenation Model: Undetermined Coefficients and Multipliers Approach
Previous Article in Journal
The Primordial Particle Accelerator of the Cosmos
Previous Article in Special Issue
Construction of Exact Solutions for Gilson–Pickering Model Using Two Different Approaches
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Ricci Soliton and Certain Related Metrics on a Three-Dimensional Trans-Sasakian Manifold

1
School of Statistics, Jilin University of Finance and Economics, Changchun 130117, China
2
School of Mathematics, Hangzhou Normal University, Hangzhou 311121, China
3
Department of Mathematics, Jadavpur University, Kolkata 700032, India
4
Department of Mathematics, Bidhan Chandra College, Asansol 713304, India
*
Author to whom correspondence should be addressed.
Universe 2022, 8(11), 595; https://doi.org/10.3390/universe8110595
Submission received: 17 August 2022 / Revised: 28 September 2022 / Accepted: 8 November 2022 / Published: 11 November 2022
(This article belongs to the Special Issue Research on Optical Soliton Perturbation)

Abstract

:
In this article, a Ricci soliton and ∗-conformal Ricci soliton are examined in the framework of trans-Sasakian three-manifold. In the beginning of the paper, it is shown that a three-dimensional trans-Sasakian manifold of type ( α , β ) admits a Ricci soliton where the covariant derivative of potential vector field V in the direction of unit vector field ξ is orthogonal to ξ . It is also demonstrated that if the structure functions meet α 2 = β 2 , then the covariant derivative of V in the direction of ξ is a constant multiple of ξ . Furthermore, the nature of scalar curvature is evolved when the manifold of type ( α , β ) satisfies ∗-conformal Ricci soliton, provided α 0 . Finally, an example is presented to verify the findings.

1. Introduction

Richard S. Hamilton introduced the concept of Ricci flow (for details see [1]) which was named after the great Italian mathematician Gregorio Ricci-Curbastro. Later, Grigori Perelman [2,3,4] found it very useful to solve Poincare conjecture. If we take a smooth closed (compact without boundary) Riemannian manifold M equipped with a smooth Riemannian metric g then the Ricci flow is defined by the geometric evolution equation,
g ( t ) t = 2 S ( g ( t ) ) ,
where S is the Ricci curvature tensor of the manifold and g ( t ) is a one-parameter family of metrics on M.
A Riemannian manifold ( M , g ) is called a Ricci soliton if there exists a vector field V and a constant λ such that the following equation holds, [5]
1 2 L V g + S + λ g = 0 ,
where L V denotes Lie derivative along the direction of V. The vector field V is called the potential vector field and λ is called the soliton constant. The Ricci soliton, which is a natural extension of the Einstein manifold, is a self-similar solution of Ricci flow. When establishing the characteristics of the soliton, the potential vector field V and the soliton constant λ are crucial factors. According to whether λ < 0 , λ = 0 or λ > 0 , the soliton is said to be shrinking, steady or expanding. The Ricci soliton reduces to Einstein manifold if V is Killing vector field. Compact Ricci solitons are the fixed points of the Ricci flow (1) projected from the space of metrics onto its quotient modulo diffeomorphisms and scalings, and often arise as blow-up limits for the Ricci flow on compact manifolds.
By simply generalizing the classical Ricci flow equation and changing the unit volume constraint to a scalar curvature constraint, in 2005 A. E. Fischer [6] introduced conformal Ricci flow. The conformal Ricci flow equation was given by
g t + 2 ( S + g n ) = p g , r ( g ) = 1 ,
where g is a dynamically evolving metric, r is the scalar curvature, p is the scalar non-dynamical field and n is the dimension of the manifold. The conformal Ricci soliton, which is an extension of the Ricci soliton, was introduced by N. Basu and A. Bhattacharyya [7] in 2015 in relation to the conformal Ricci flow equation. The conformal Ricci soliton equation was given by,
L V g + 2 S + [ 2 λ ( p + 2 n ) ] g = 0 .
The definition of the Ricci soliton was modified in 2014 by G. Kaimakamis and K. Panagiotidou [8] who substituted the ∗-Ricci tensor S , introduced by S. Tachibana [9] and T. Hamada [10], respectively, for the Ricci tensor S. The ∗-Ricci tensor S is defined by
S ( X , Y ) = 1 2 ( t r a c e { ϕ . R ( X , ϕ Y ) } ) ,
for arbitrary vector fields X and Y on M, where R is the Riemannian curvature tensor and ϕ is a ( 1 , 1 ) tensor field. Within the context of real hypersurfaces of a complex space form, the ∗-Ricci soliton notion has been applied. A pseudo-Riemannian metric g is called a ∗-Ricci soliton if there exists a constant λ and a vector field V such that,
L V g + 2 S + 2 λ g = 0 .
Note that, ∗-Ricci soliton is trivial if the vector field V is Killing, and in this case, the manifold becomes ∗-Einstein. By a ∗-Einstein manifold we mean that the ∗-Ricci tensor ( S ) is proportional to the metric g. Thus, it is considered a natural generalization of ∗-Einstein metric. A ∗-Ricci soliton is said to be almost ∗-Ricci soliton if λ is a smooth function on M. Moreover, an almost ∗-Ricci soliton is called shrinking, steady, and expanding according to as λ is negative, zero, and positive, respectively. It was demonstrated by G. Kaimakamis et al. [8] that a real hypersurface in a complex projective space does not admit a ∗-Ricci soliton by studying real hypersurfaces of a non-flat complex space that admit a ∗-Ricci soliton whose potential vector field is the structure vector field. They also proved that a real hypersurface of complex hyperbolic space admitting a ∗-Ricci soliton is locally congruent to a geodesic hypersphere.
With the aid of (3), P. Majhi and D. Dey [11] further modified the aforementioned definition of ∗-Ricci soliton in 2020 and defined ∗-conformal Ricci soliton as follows,
L V g + 2 S + [ 2 λ ( p + 2 n ) ] g = 0 .
Ricci solitons have been studied in many contexts: on Kähler manifolds [12], on contact and Lorentzian manifolds [13,14], on K-contact manifolds [15], etc. by many authors. Later, H. G. Nagaraja and C. R. Premalatha [16] studied the nature of Ricci soliton on a three-dimensional trans-Sasakian manifold; C. Cǎlin and M. Crasmareanu [17] on f-Kenmotsu manifold; C. He and M. Zhu [18] on Sasakian manifold and G. Ingalahalli and C. S. Bagewadi [19] on α -Sasakian manifold. Recently, in 2017, Y. Wang [20] proved that if a three-dimensional cosymplectic manifold M 3 admits a Ricci soliton, then either M 3 is locally flat or the potential vector field is an infinitesimal contact transformation. Furthermore, S. Pahan and A. Bhattacharyya gave some insight into the trans-Sasakian manifold [21]. In 2016, T. Dutta et al. studied conformal Ricci soliton on a three-dimensional trans-Sasakian manifold [22].
As shown in the literature, ∗-Ricci soliton on contact geometry was studied by many authors: on Sasakian and ( κ , μ ) -contact manifold by A. Ghosh and D. S. Patra [23], on ( κ , μ ) -almost Kenmotsu manifolds by X. Dai, Y. Zhao and U. C. De [24], on contact 3-manifolds by Y. Wang [25], etc. It is worthy to mention that in [26], D. Dey and P. Majhi considered ∗-Ricci soliton on a three-dimensional trans-Sasakian manifold and proved that if the metric of the manifold represents ∗-Ricci soliton and if it satisfies a certain condition then the manifold reduces to a β -Kenmotsu manifold. Furthermore, very recently generalizations of ∗-Ricci soliton on contact geometry were studied by [5,27,28,29,30,31]. Moreover, some of the latest connected studies can be seen in [32,33,34,35,36,37,38,39,40,41,42,43,44,45,46,47,48,49,50,51,52,53,54,55,56,57,58,59,60,61,62,63,64,65,66,67,68,69,70,71,72,73,74,75,76,77,78,79,80,81,82,83,84,85,86,87,88,89,90,91].
Motivated from the above-mentioned well-praised works we studied the behaviour of Ricci soliton and ∗-conformal Ricci soliton on a three-dimensional trans-Sasakian manifold. In the later sections, we revisit some definitions and important properties of three-dimensional trans-Sasakian manifold and after that the main result of this paper, containing two theorems are described. We also provide an example to justify our findings.

Physical Motivation

The Ricci soliton has extensive applications, not only in mathematical physics but also in quantum cosmology, quantum gravity, and black holes as well. The Ricci soliton can be considered as a kinematic solution in fluid space–time, whose profile develops a characterization of spaces of constant curvature along with the locally symmetric spaces. It also expresses geometrical and physical applications with relativistic viscous fluid space–time admitting heat flux and stress, dark and dust fluid general relativistic space–time, and radiation era in general relativistic space–time. Ricci soliton has applications in the renormalization group (RG) flow for the bosonic nonlinear sigma model from two-dimensional space–time to a curved Riemannian tangent manifold. A two-dimensional Ricci soliton can be used to discuss the behavior of mass under Ricci flow. Ricci soliton is important as it can help in understanding the concepts of energy or entropy in general relativity. This property is the same as that of the heat equation due to which an isolated system loses the heat for thermal equilibrium.
As an application to cosmology and general relativity by investigating the kinetic and potential nature of relativistic space–time, we can present a physical model of three classes, namely, shrinking, steady, and expanding of perfect and dust fluid solutions of Ricci solitons space–time. The first case shrinking ( λ < 0 ) which exists on a minimal time interval 1 < t < b where b < 1 , steady ( λ = 0 ) which exists for all time or expanding ( λ > 0 ) which exists on maximal time interval a < t < 1 , a > 1 . These three classes give examples of ancient, eternal, and immortal solutions, respectively. From [92,93] (briefly discussed in the above section), we can think more about the physical applications of Ricci soliton.

2. Preliminaries

According to D. E. Blair [94], a differentiable manifold M of dimension ( 2 n + 1 ) is said to have an almost contact structure or ( ϕ , ξ , η ) structure if M permits a ( 1 , 1 ) tensor field ϕ , a vector field ξ , an 1-form η satisfying
ϕ 2 = I + η ξ ,
η ( ξ ) = 1 ,
where I is the identity mapping. A Riemannian metric g is said to be a compatible metric if it satisfies,
g ( ϕ X , ϕ Y ) = g ( X , Y ) η ( X ) η ( Y ) ,
for any vector fields X and Y on M. A manifold having almost contact structure along with compatible Riemannian metric is called almost contact metric manifold.
In an almost contact metric manifold the following conditions are satisfied for arbitrary X , Y χ ( M ) , where χ ( M ) denotes the set of all vector fields on M [94]:
ϕ ξ = 0 ,
η ϕ = 0 ,
g ( X , ξ ) = η ( X ) ,
g ( ϕ X , Y ) = g ( X , ϕ Y ) .
Let M be a ( 2 n + 1 ) -dimensional almost contact manifold. Then we define an almost complex structure J on M × R by J ( X , f d d t ) = ( ϕ X f ξ , η ( X ) d d t ) , where X is a tangent to M, t is the coordinate on R , and f a C function on M × R . Clearly, J 2 = I . If J is integrable then the almost contact structure is said to be normal. The normality of an almost contact metric manifold is equivalent to the vanishing of the tensor field [ ϕ , ϕ ] + 2 d η ξ , where [ ϕ , ϕ ] is the Nijenhuis torsion tensor of ϕ (for more details see [94]).
In 1985, J. A. Oubiña [95] introduced a new class of almost contact metric manifolds known as trans-Sasakian manifolds. Trans-Sasakian manifolds arose naturally from the classification of almost contact metric structures and they appear as a natural generalization of both Sasakian and Kenmotsu manifolds. An almost contact metric manifold M is called a trans-Sasakian manifold if ( M × R , J , G ) , where G is the product metric on M × R , belongs to the class W 4 (see [96]). If there are smooth functions α , β on an almost contact metric manifold ( M , ϕ , ξ , η , g ) satisfying [97]
( X ϕ ) Y = α [ g ( X , Y ) ξ η ( Y ) X ] + β [ g ( ϕ X , Y ) ξ η ( Y ) ϕ X ] ,
where X , Y χ ( M ) are arbitrary and ∇ is the Levi-Civita connection of g on M, then the manifold is called trans-Sasakian manifold of type ( α , β ) . α , β are called structure functions of the manifold. Trans-Sasakian manifolds of type ( 0 , 0 ) , ( α , 0 ) , ( 0 , β ) are called cosymplectic, α -Sasakian, and β -Kenmotsu manifolds, respectively. Then from (12), we can deduce that,
( X η ) ( Y ) = α g ( ϕ X , Y ) + β g ( ϕ X , ϕ Y ) ,
X ξ = α ϕ X + β ( X η ( X ) ξ ) .
J. C. Marrero [98] showed that a trans-Sasakian manifold of dimension 5 is either cosymplectic or α -Sasakian or β -Kenmotsu. Therefore, proper trans-Sasakian manifold exists only for dimension 3. In a 3-dimensional trans-Sasakian manifold the following relations hold, [21]
R ( X , Y ) Z = r 2 + 2 ξ β 2 ( α 2 β 2 ) [ g ( Y , Z ) X g ( X , Z ) Y ] g ( Y , Z ) [ r 2 + ξ β 3 ( α 2 β 2 ) η ( X ) ξ η ( X ) ( ϕ D α D β ) + ( ( X β ) + ( ϕ X ) α ) ξ ] + g ( X , Z ) [ r 2 + ξ β 3 ( α 2 β 2 ) η ( Y ) ξ η ( Y ) ( ϕ D α D β ) + ( ( Y β ) + ( ϕ Y ) α ) ξ ) ] [ ( ( Z β ) + ( ϕ Z ) α ) η ( Y ) + ( ( Y β ) + ( ϕ Y ) α ) η ( Z ) + r 2 + ξ β 3 ( α 2 β 2 ) η ( Y ) η ( Z ) ] X + [ ( ( Z β ) + ( ϕ Z ) α ) η ( X ) + ( ( X β ) + ( ϕ X ) α ) η ( Z ) + r 2 + ξ β 3 ( α 2 β 2 ) η ( X ) η ( Z ) ] Y ,
S ( X , Y ) = r 2 + ( ξ β ) ( α 2 β 2 ) g ( X , Y ) r 2 + ( ξ β ) 3 ( α 2 β 2 ) η ( X ) η ( Y ) ( ( Y β ) + ( ϕ Y ) α ) η ( X ) ( ( X β ) + ( ϕ X ) α ) η ( Y ) ,
S ( X , ξ ) = ( 2 ( α 2 β 2 ) ( ξ β ) ) η ( X ) ( X β ) ( ϕ X ) α ,
where D f denotes the gradient of the smooth function f defined on M and R, S, r are the Riemannian curvature tensor, Ricci tensor of type ( 0 , 2 ) , and scalar curvature of the manifold, respectively, and α , β are smooth functions on the manifold.
Here in this paper, we restricted the smooth functions α , β to be constant functions. Then we obtained some special relations compatible with our restrictions,
R ( X , Y ) ξ = ( α 2 β 2 ) [ η ( Y ) X η ( X ) Y ] ,
S ( X , Y ) = r 2 ( α 2 β 2 ) g ( X , Y ) r 2 3 ( α 2 β 2 ) η ( X ) η ( Y ) ,
S ( X , ξ ) = 2 ( α 2 β 2 ) η ( X ) ,
Q X = r 2 ( α 2 β 2 ) X r 2 3 ( α 2 β 2 ) η ( X ) ξ ,
where Q is the Ricci operator given by S ( X , Y ) = g ( Q X , Y ) . The expression of ∗-Ricci tensor (for details see Lemma 3.1 of [26]) on a three-dimensional trans-Sasakian manifold for arbitrary vector fields X and Y of χ ( M ) is given by,
S ( X , Y ) = r 2 2 ( α 2 β 2 ) [ g ( X , Y ) η ( X ) η ( Y ) ] .

3. Results

In this section, we consider the metric of a three-dimensional trans-Sasakian manifold as a Ricci soliton and a ∗-conformal Ricci soliton and prove the following two results.
Theorem 1.
Let M be a three-dimensional trans-Sasakian manifold of type ( α , β ) admitting a Ricci soliton where the structure functions α and β are non-zero constant. Then the following relations are satisfied,
1.
If ξ V is orthogonal to ξ, then the soliton is shrinking for α 2 < β 2 , steady for α 2 = β 2 , and expanding for α 2 > β 2 .
2.
If α 2 = β 2 , then the covariant derivative of the potential vector field V in the direction of ξ is a constant multiple of ξ.
Proof. 
In a three-dimensional trans-Sasakian manifold where α and β are non-zero constant, we know from (21) that the Ricci operator can be written as,
Q X = r 2 ( α 2 β 2 ) X r 2 3 ( α 2 β 2 ) η ( X ) ξ ,
where X χ ( M ) is any vector field. The aforementioned equation implies that it is an η -Einstein manifold. Now taking the covariant derivative of (23) along an arbitrary Y χ ( M ) , we have
( Y Q ) X = 1 2 ( Y r ) X 1 2 ( Y r ) η ( X ) ξ r 2 3 ( α 2 β 2 ) [ α g ( ϕ Y , X ) ξ + β g ( X , Y ) ξ α η ( X ) ( ϕ Y ) + β η ( X ) Y 2 β η ( X ) η ( Y ) ξ ] .
Contracting X and using the well-known formula t r a c e { X ( X Q ) Y } = 1 2 ( Y r ) in (24), we obtain
ξ r = 2 r β + 12 ( α 2 β 2 ) β .
Using (19) in the definition of Ricci soliton (2), we acquire
( L V g ) ( Y , Z ) = ( 2 λ r + 2 ( α 2 β 2 ) ) g ( Y , Z ) + ( r 6 ( α 2 β 2 ) ) η ( Y ) η ( Z ) ,
for any vector fields Y , Z χ ( M ) . Now taking the covariant derivative of (26) along an arbitrary vector field X χ ( M ) ,
( X L V g ) ( Y , Z ) = ( X r ) g ( Y , Z ) + ( X r ) η ( Y ) η ( Z ) + ( r 6 ( α 2 β 2 ) ) [ α g ( ϕ X , Y ) η ( Z ) + β g ( ϕ X , ϕ Y ) η ( Z ) α g ( ϕ X , Z ) η ( Y ) + β g ( ϕ X , ϕ Z ) η ( Y ) ] .
Again for any vector fields X , Y , Z χ ( M ) , we know [99]
( L V X g X L V g [ V , X ] g ) ( Y , Z ) = g ( ( L V ) ( X , Y ) , Z ) g ( ( L V ) ( X , Z ) , Y ) .
Since ∇ is Riemannian metric connection, g = 0 . So the above equation reduces to, ( X L V g ) ( Y , Z ) = g ( ( L V ) ( X , Y ) , Z ) + g ( ( L V ) ( X , Z ) , Y ) . Again, using symmetry of ( L V ) , i.e., ( L V ) ( X , Y ) = ( L V ) ( Y , X ) , we rewrite the last relation as
2 g ( ( L V ) ( X , Y ) , Z ) = ( X L V g ) ( Y , Z ) + ( Y L V g ) ( Z , X ) ( Z L V g ) ( X , Y ) .
Using (27) in the above equation, we obtain
( L V ) ( X , Y ) = 1 2 ( X r ) Y 1 2 ( Y r ) X + 1 2 g ( ϕ X , ϕ Y ) D r + 1 2 ( X r ) η ( Y ) ξ + 1 2 ( Y r ) η ( X ) ξ + ( r 6 ( α 2 β 2 ) ) [ α η ( Y ) ϕ X α η ( X ) ϕ Y + β g ( ϕ X , ϕ Y ) ξ ] ,
for all vector fields X and Y on M. Covariant derivative of (29) along an arbitrary vector field yields,
( X L V ) ( Y , Z ) = 1 2 g ( Z , X D r ) Y 1 2 g ( Y , X D r ) Z + 1 2 g ( ϕ Y , ϕ Z ) ( X D r ) α η ( Z ) ( X r ) ϕ Y α η ( Y ) ( X r ) ϕ Z + 1 2 [ ( Z r ) η ( Y ) + ( Y r ) η ( Z ) ] ( X ξ ) + 1 2 [ g ( Y , X D r ) η ( Z ) α ( Y r ) g ( ϕ X , Z ) + β g ( ϕ X , ϕ Z ) ( Y r ) + g ( Z , X D r ) η ( Y ) α ( Z r ) g ( ϕ X , Y ) + β g ( ϕ X , ϕ Y ) ( Z r ) + 2 β g ( ϕ Y , ϕ Z ) ( X r ) ] ξ + 1 2 [ α g ( ϕ X , Y ) η ( Z ) β g ( ϕ X , ϕ Y ) η ( Z ) + α g ( ϕ X , Z ) η ( Y ) β g ( ϕ X , ϕ Z ) η ( Y ) ] D r + ( r 6 ( α 2 β 2 ) ) [ { α 2 g ( ϕ X , Z ) α β g ( ϕ X , ϕ Z ) } ϕ Y + { α 2 g ( ϕ X , Y ) α β g ( ϕ X , ϕ Y ) } ϕ Z α η ( Z ) ( ( X ϕ ) Y ) α η ( Y ) ( ( X ϕ ) Z ) + β g ( ϕ Y , ϕ Z ) ( X ξ ) + { α β g ( ϕ X , Y ) η ( Z ) β 2 g ( ϕ X , ϕ Y ) η ( Z ) + α β g ( ϕ X , Z ) η ( Y ) β 2 g ( ϕ X , ϕ Z ) η ( Y ) } ξ ] .
From K. Yano [99], we know ( L V R ) ( X , Y ) Z = ( X L V ) ( Y , Z ) ( Y L V ) ( X , Z ) . Using this formula in the above equation we obtain,
( L V R ) ( X , Y ) Z = 1 2 g ( Z , Y D r ) X 1 2 g ( Z , X D r ) Y 1 2 [ α ( Y r ) g ( ϕ X , Z ) + β ( Y r ) g ( ϕ X , ϕ Z ) g ( Z , X D r ) η ( Y ) + α ( Z r ) g ( ϕ X , Y ) α ( X r ) g ( ϕ Y , Z ) β g ( ϕ Y , ϕ Z ) ( X r ) + g ( Z , Y D r ) η ( X ) α ( Z r ) g ( ϕ Y , X ) ] ξ + α η ( Z ) ( Y r ) ϕ X α η ( Z ) ( X r ) ϕ Y + α { η ( X ) ( Y r ) η ( Y ) ( X r ) } ϕ Z + 1 2 { α g ( ϕ X , Y ) η ( Z ) + α g ( X , ϕ Y ) η ( Z ) α g ( ϕ X , Z ) η ( Y ) β g ( ϕ X , ϕ Z ) η ( Y ) α g ( ϕ Y , Z ) η ( X ) + β g ( ϕ Y , ϕ Z ) η ( X ) } D r + 1 2 { ( Y r ) η ( Z ) + ( Z r ) η ( Y ) } ( X ξ ) 1 2 { ( X r ) η ( Z ) + ( Z r ) η ( X ) } ( Y ξ ) + 1 2 g ( ϕ Y , ϕ Z ) ( X D r ) 1 2 g ( ϕ X , ϕ Z ) ( Y D r ) + ( r 6 ( α 2 β 2 ) ) [ { α β g ( ϕ Y , ϕ Z ) α 2 g ( ϕ Y , Z ) } ϕ X { α β g ( ϕ X , ϕ Z ) α 2 g ( ϕ X , Z ) } ϕ Y + 2 α 2 g ( ϕ X , Y ) ϕ Z + { 2 α β g ( ϕ X , Y ) η ( Z ) + α β g ( ϕ X , Z ) η ( Y ) β 2 g ( ϕ X , ϕ Z ) η ( Y ) α β g ( ϕ Y , Z ) η ( X ) + β 2 g ( ϕ Y , ϕ Z ) η ( X ) } ξ + β g ( ϕ Y , ϕ Z ) ( X ξ ) β ( Y ξ ) g ( ϕ X , ϕ Z ) α η ( Z ) ( ( X ϕ ) Y ) α η ( Y ) ( ( X ϕ ) Z ) + α η ( Z ) ( ( Y ϕ ) X ) + α η ( X ) ( ( Y ϕ ) Z ) .
The above equation holds for any X , Y , Z χ ( M ) . Contracting X in (30), we achieve
( L V S ) ( Y , Z ) = Δ r 2 6 α 4 + 12 α 2 β 2 6 β 4 + r α 2 r β 2 g ( ϕ Y , ϕ Z ) ,
for any Y , Z χ ( M ) . Again, from (19), we obtain
( L V S ) ( Y , Z ) = 1 2 g ( ϕ Y , ϕ Z ) ( V r ) + r 2 ( α 2 β 2 ) { g ( Y V , Z ) + g ( Y , Z V ) } r 2 3 ( α 2 β 2 ) { η ( Z ) ( ( V η ) Y ) + η ( Y ) ( ( V η ) Z ) + η ( Z ) η ( Y V ) + η ( Y ) η ( Z V ) } .
Comparison of (31) with (32) yields,
Δ r 2 6 α 4 + 12 α 2 β 2 6 β 4 + r α 2 r β 2 g ( ϕ Y , ϕ Z ) = 1 2 { g ( ϕ Y , ϕ Z ) ( V r ) + r 2 ( α 2 β 2 ) { g ( Y V , Z ) + g ( Y , Z V ) } r 2 3 ( α 2 β 2 ) { η ( Z ) ( ( V η ) Y ) + η ( Y ) ( ( V η ) Z ) + η ( Z ) η ( Y V ) + η ( Y ) η ( Z V ) } .
Now, letting Y = Z = ξ gives rise to ( α 2 β 2 ) η ( ξ V ) = 0 . From here, two cases arise, either η ( ξ V ) = 0 or ( α 2 β 2 ) = 0 . From the definition of Ricci soliton (2), we have
1 2 ( g ( X V , Y ) + g ( Y V , X ) ) + S ( X , Y ) = λ g ( X , Y ) ,
for any vector fields X and Y. In first case, η ( ξ V ) = 0 which implies ξ V is orthogonal to ξ , putting X = Y = ξ in (34) gives 2 ( α 2 β 2 ) = λ . It directly implies that the soliton is shrinking if α 2 < β 2 , steady if α 2 = β 2 and expanding if α 2 > β 2 .
For the second case where α 2 = β 2 , then it follows directly from (34) that ξ V = λ ξ , i.e., the covariant derivative of the potential vector field V in the direction of ξ is λ -multiple of ξ . □
Theorem 2.
Let M be a three-dimensional trans-Sasakian manifold of type ( α , β ) where the structure functions α and β are constant with α 0 . If the metric g represents a ∗-conformal Ricci soliton then the scalar curvature of the manifold is given by r = 1 β 2 α 2 p 2 + 1 3 λ + 4 α 2 .
Proof. 
Since the metric g represents a ∗-conformal Ricci soliton, using (22) in the definition of ∗-conformal Ricci soliton (4), we obtain
( L V g ) ( X , Y ) = p + 2 3 + 4 ( α 2 β 2 ) r 2 λ g ( X , Y ) + ( r 4 ( α 2 β 2 ) ) η ( X ) η ( Y ) ,
for all vector fields X and Y on M. If we consider covariant derivative with respect to arbitrary vector field Z, then (35) reduces to
( Z L V g ) ( X , Y ) = ( Z r ) [ η ( X ) η ( Y ) g ( X , Y ) ] ( r 4 ( α 2 β 2 ) ) [ α g ( ϕ Z , X ) η ( Y ) β g ( ϕ X , ϕ Z ) η ( Y ) + α g ( ϕ Z , Y ) η ( X ) β g ( ϕ Y , ϕ Z ) η ( X ) ] ,
for all X , Y χ ( M ) . Using (11) and (28) in (36), we obtain
( L V ) ( X , Y ) = 1 2 ( D r ) [ g ( X , Y ) η ( X ) η ( Y ) ] 1 2 ( X r ) [ Y η ( Y ) ξ ] 1 2 ( Y r ) [ X η ( X ) ξ ] + ( r 4 ( α 2 β 2 ) ) [ β g ( ϕ X , ϕ Y ) ξ α η ( Y ) ( ϕ X ) α η ( X ) ( ϕ Y ) ] ,
for arbitrary vector fields X and Y on M. Setting Y = ξ in (37), we have
( L V ) ( X , ξ ) = 1 2 ( ξ r ) [ X η ( X ) ξ ] α ( r 4 ( α 2 β 2 ) ) ( ϕ X ) .
Applying covariant derivative along an arbitrary vector field Y and making use of (12)–(14), we obtain
( Y L V ) ( X , ξ ) = α ( L V ) ( X , ϕ Y ) β ( L V ) ( X , Y ) 1 2 ( Y ( ξ r ) ) [ X η ( X ) ξ ] + 1 2 ( ξ r ) [ α g ( ϕ X , Y ) ξ + β g ( ϕ X , ϕ Y ) ξ α η ( X ) ( ϕ Y ) + β η ( X ) Y β η ( Y ) X ] α ( Y r ) ( ϕ X ) α ( r 4 ( α 2 β 2 ) ) [ α g ( X , Y ) ξ α η ( X ) Y + β g ( ϕ Y , X ) ξ β η ( X ) ( ϕ Y ) + β η ( Y ) ( ϕ X ) ] .
From K. Yano [99], we know ( L V R ) ( X , Y ) Z = ( X L V ) ( Y , Z ) ( Y L V ) ( X , Z ) . Using (39) in this formula, we obtain
( L V R ) ( X , Y ) ξ = α ( L V ) ( ϕ X , Y ) α ( L V ) ( X , ϕ Y ) 1 2 ( X ( ξ r ) ) [ Y η ( Y ) ξ ] + 1 2 ( Y ( ξ r ) ) [ X η ( X ) ξ ] + 1 2 ( ξ r ) [ 2 α g ( X , ϕ Y ) ξ α η ( Y ) ( ϕ X ) + α η ( X ) ( ϕ Y ) + 2 β η ( Y ) X 2 β η ( X ) Y ] α ( X r ) ( ϕ Y ) + α ( Y r ) ( ϕ X ) α ( r 4 ( α 2 β 2 ) ) [ α η ( X ) Y α η ( Y ) X + 2 β g ( ϕ X , Y ) ξ + 2 β η ( X ) ( ϕ Y ) 2 β η ( Y ) ( ϕ X ) ] .
Setting Y = ξ in the foregoing equation, we acquire
( L V R ) ( X , ξ ) ξ = 1 2 ( ξ ( ξ r ) ) [ X η ( X ) ξ ] + β ( ξ r ) [ X η ( X ) ξ ] 2 α ( r 4 ( α 2 β 2 ) ) [ α X + α η ( X ) ξ β ( ϕ X ) .
Again, Lie differentiation of Equation (18) along soliton vector field V and use of (15) and (18) leads to,
( L V R ) ( X , ξ ) ξ = ( α 2 β 2 ) [ g ( X , L V ξ ) ξ ( ( L V η ) X ) ξ 2 η ( L V ξ ) X ] ,
which holds for arbitrary vector field X on M. Setting Y = ξ in (35) implies,
( L V η ) X g ( X , L V ξ ) = p + 2 3 2 λ η ( X ) .
Taking (42) into account, Lie derivative of η ( ξ ) = 1 along the direction of V leads to
2 η ( L V ξ ) = p + 2 3 2 λ .
After using (42) and (43), the Equation (41) reduces to
( L V R ) ( X , ξ ) ξ = ( α 2 β 2 ) p + 2 3 2 λ [ X η ( X ) ξ ] ,
for all X χ ( M ) . Comparing (40) with (44) we acquire,
( α 2 β 2 ) p + 2 3 2 λ [ X η ( X ) ξ ] = 1 2 ( ξ ( ξ r ) ) [ X η ( X ) ξ ] + β ( ξ r ) [ X η ( X ) ξ ] 2 α ( r 4 ( α 2 β 2 ) ) [ α X + α η ( X ) ξ β ( ϕ X ) ,
for any X χ ( M ) . Inner product of the foregoing equation with arbitrary vector field Y gives,
1 2 ( ξ ( ξ r ) ) + β ( ξ r ) + 2 α 2 ( r 4 ( α 2 β 2 ) ) ( α 2 β 2 ) p + 2 3 2 λ [ g ( X , Y ) η ( X ) η ( Y ) ] + 2 α β ( r 4 ( α 2 β 2 ) ) g ( ϕ X , Y ) = 0 .
Anti-symmetrizing the foregoing equation yields,
1 2 ( ξ ( ξ r ) ) + β ( ξ r ) + 2 α 2 ( r 4 ( α 2 β 2 ) ) ( α 2 β 2 ) p + 2 3 2 λ g ( ϕ X , ϕ Y ) = 0 .
Since this equation holds for arbitrary vector fields ϕ X and ϕ Y and as we know from (25) that ξ r = 2 r β + 12 β ( α 2 β 2 ) holds in a three-dimensional trans-Sasakian manifold, we conclude that the scalar curvature of the manifold satisfies r = 1 β 2 α 2 p 2 + 1 3 λ + 4 α 2 . □

4. Example of a Three-Dimensional Trans-Sasakian Manifold Admitting Ricci Soliton

In this section, we provide an example to verify our outcomes.
Example 1.
We consider the manifold as M = { ( x , y , z ) R 3 : y 0 } , where ( x , y , z ) are the standard coordinates in R 3 . The vector fields as defined bellow,
e 1 : = e 2 z x , e 2 : = e 2 z y , e 3 : = z ,
are linearly independent at each point of M. We define the Riemannian metric g as,
g i j = g ( e i , e j ) : = 1 , i f   i = j a n d i , j { 1 , 2 , 3 } 0 , o t h e r w i s e .
Let ξ = e 3 . Then the 1-form η is defined by η ( Z ) : = g ( Z , e 3 ) , for arbitrary Z χ ( M ) , then we have the following relations,
η ( e 1 ) = 0 , η ( e 2 ) = 0 , η ( e 3 ) = 1 .
Let us define the (1,1)-tensor field ϕ as
ϕ e 1 : = e 2 , ϕ e 2 : = e 1 , ϕ e 3 : = 0 ,
then the relations (5), (6), and (7) are satisfied. Thus, ( ϕ , ξ , η , g ) defines an almost contact metric structure on M. We can now easily conclude,
[ e 1 , e 2 ] = 0 , [ e 2 , e 3 ] = 2 e 2 , [ e 1 , e 3 ] = 2 e 1 .
Let be the Levi-Civita connection of g. Then from Koszul’s formula, 2 g ( X Y , Z ) = X g ( Y , Z ) + Y g ( Z , X ) Z g ( X , Y ) g ( X , [ Y , Z ] ) g ( Y , [ X , Z ] ) + g ( Z , [ X , Y ] ) , we can have
e 1 e 1 = 2 e 3 , e 1 e 2 = 0 , e 1 e 3 = 2 e 1 , e 2 e 1 = 0 , e 2 e 2 = 2 e 3 , e 2 e 3 = 2 e 2 , e 3 e 1 = 0 , e 3 e 2 = 0 , e 3 e 3 = 0 .
From here we can easily verify that the relations (12) and (13) are satisfied. Hence M becomes trans-Sasakian manifold of type ( 0 , 2 ) . The components of Riemannian curvature tensor are given by,
R ( e 1 , e 2 ) e 1 = 4 e 3 , R ( e 1 , e 2 ) e 2 = 4 e 1 , R ( e 1 , e 2 ) e 3 = 0 , R ( e 1 , e 3 ) e 1 = 4 e 2 , R ( e 1 , e 3 ) e 2 = 0 , R ( e 1 , e 3 ) e 3 = 4 e 1 , R ( e 2 , e 3 ) e 1 = 0 , R ( e 2 , e 3 ) e 2 = 4 e 2 , R ( e 2 , e 3 ) e 3 = 4 e 2 .
And the components of Ricci tensor are given by,
S ( e 1 , e 1 ) = 0 , S ( e 2 , e 2 ) = 0 , S ( e 3 , e 3 ) = 8 .
From here we can easily deduce that the scalar curvature of the manifold r = i = 1 3 S ( e i , e i ) = 8 . Let us define a vector field V by, V : = ξ . Then we can obtain,
( L V g ) ( e 1 , e 1 ) = 4 , ( L V g ) ( e 2 , e 2 ) = 4 , ( L V g ) ( e 3 , e 3 ) = 0 .
Contracting (2) and using the result r = 8 , we deduce λ = 4 . So g defines a Ricci soliton on this trans-Sasakian manifold for λ = 4 .

5. Conclusions

In this article, we used the methods of local Riemannian geometry to interpret solutions of (2) and (4) and impregnate Einstein metrics in a large class of metrics of Ricci soliton and ∗-conformal Ricci soliton on a trans-Sasakian manifold of the third dimension. Our results will not only play an indispensable and incitement role in contact geometry but also make a significant and motivational contribution in the area of further research of complex geometry, especially on Kähler and para-Kähler manifolds, etc. Some questions arise from our article to study in further research:
(i)
What will be the outcomes if we consider the structure functions α and β to satisfy ϕ D α = D β ?
(ii)
Do the above results hold without assuming any restrictions on structure functions?
(iii)
How do the aforementioned outcomes differ for the ∗- η Ricci soliton and the ∗-conformal η -Ricci soliton?

Author Contributions

Conceptualization, Z.C., Y.L., S.S., S.D. and A.B.; methodology, Z.C., Y.L., S.S., S.D. and A.B.; investigation, Z.C., Y.L., S.S., S.D. and A.B.; writing—original draft preparation, Z.C., Y.L., S.S., S.D. and A.B.; writing, Z.C., Y.L., S.S., S.D. and A.B. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by National Natural Science Foundation of China (Grant No. 12101168) and Zhejiang Provincial Natural Science Foundation of China (Grant No. LQ22A010014) and the UGC Senior Research Fellowship of India (Sr. No.: 2061540940, Ref. No.: 21/06/2015(i)EU-V).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

We gratefully acknowledge the constructive comments from the editor and the anonymous referees.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Hamilton, R.S. The Ricci flow on surfaces. Contemp. Math. 1988, 71, 237–261. [Google Scholar]
  2. Perelman, G. The entropy formula for the Ricci flow and its geometric applications. arXiv 2002, arXiv:math.DG/0211159. [Google Scholar]
  3. Perelman, G. Ricci flow with surgery on three-manifolds. arXiv 2003, arXiv:math.DG/0303109. [Google Scholar]
  4. Perelman, G. Finite extinction time for the solutions to the Ricci flow on certain three-manifolds. arXiv 2003, arXiv:math.DG/0307245. [Google Scholar] [CrossRef]
  5. Sarkar, S.; Dey, S.; Chen, X. Certain results of conformal and ∗-conformal Ricci soliton on para-cosymplectic and para-Kenmotsu manifolds. Filomat 2021, 35, 5001–5015. [Google Scholar] [CrossRef]
  6. Fischer, A.E. An introduction to conformal Ricci flow. Class. Quantum Gravity 2004, 21, S171–S218. [Google Scholar] [CrossRef]
  7. Basu, N.; Bhattacharyya, A. Conformal Ricci soliton in Kenmotsu manifold. Glob. J. Adv. Res. Class. Mod. Geom. 2015, 4, 15–21. [Google Scholar]
  8. Kaimakamis, G.; Panagiotidou, K. ∗-Ricci solitons of real hypersurface in non-flat comlex space forms. J. Geom. Phys. 2014, 86, 408–413. [Google Scholar] [CrossRef]
  9. Tachibana, S. On almost-analytic vectors in almost Kählerian manifolds. Tohoku Math. J. 1959, 11, 247–265. [Google Scholar] [CrossRef]
  10. Hamada, T. Real hypersurfaces of complex space forms in terms of Ricci ∗-tensor. Tokyo J. Math. 2002, 25, 473–483. [Google Scholar] [CrossRef]
  11. Majhi, P.; Dey, D. ∗-Conformal Ricci soliton on a class of almost Kenmotsu manifolds. São Paulo J. Math. Sci. 2020, 15, 335–343. [Google Scholar]
  12. Chodosh, O.; Fong, F.T. Rotational symmetry of conical Kähler-Ricci solitons. Math. Ann. 2016, 364, 777–792. [Google Scholar] [CrossRef] [Green Version]
  13. Bagewadi, C.S.; Ingalahalli, G. Ricci solitons in Lorentzian α-Sasakian manifolds. Acta Math. Acad. Paedagog. Nyíregyháziensis 2012, 28, 59–68. [Google Scholar]
  14. Bagewadi, C.S.; Ingalahalli, G.; Ashoka, S.R. A Study on Ricci Solitons in Kenmotsu Manifolds. ISRN Geom. 2013, 2013, 412593. [Google Scholar] [CrossRef]
  15. Sharma, R. Certain results on K-contact and (κ,μ)-contact manifolds. J. Geom. 2008, 89, 138–147. [Google Scholar] [CrossRef]
  16. Nagaraja, H.G.; Premalatha, C.R. Ricci Solitons in f -Kenmotsu Manifolds and 3-Dimensional Trans-Sasakian Manifolds. Prog. Appl. Math. 2012, 3, 1–6. [Google Scholar]
  17. Cǎlin, C.; Crasmareanu, M. From the Eisenhart problem to Ricci solitons in f-Kenmotsu manifolds. Bull. Malays. Math. Soc. Ser. 2 2010, 33, 361–368. [Google Scholar]
  18. He, C.; Zhu, M. The Ricci solitons on Sasakian manifolds. arXiv 2011, arXiv:1109.4407. [Google Scholar]
  19. Ingalahalli, G.; Bagewadi, C.S. Ricci solitons in α-Sasakian manifolds. ISRN Geom. 2012, 2012, 421384. [Google Scholar] [CrossRef]
  20. Wang, Y. Ricci solitons on 3-dimensional cosymplectic manifolds. Math. Slovaca 2017, 67, 979–984. [Google Scholar] [CrossRef]
  21. Pahan, S.; Bhattacharyya, A. Some Properties of Three Dimensional trans-Sasakian Manifolds with a Semi-Symmetric Metric Connection. Lobachevskii J. Math. 2016, 37, 177–184. [Google Scholar] [CrossRef]
  22. Dutta, T.; Basu, N.; Bhattacharyya, A. Almost conformal Ricci soliton on 3-dimensional trans-Sasakian manifold. Hacet. J. Math. Stat. 2016, 45, 1379–1392. [Google Scholar] [CrossRef]
  23. Ghosh, A.; Patra, D.S. ∗-Ricci soliton within the frame-work of Sasakian and (κ,μ)-contact manifold. Int. J. Geom. Methods Mod. Phys. 2018, 15, 1850120. [Google Scholar] [CrossRef]
  24. Dai, X.; Zhao, Y.; De, U.C. ∗-Ricci soliton on (κ,μ)′-almost Kenmotsu manifolds. Open Math. 2019, 17, 874–882. [Google Scholar] [CrossRef]
  25. Wang, Y. Contact 3-manifolds and ∗-Ricci soliton. Kodai Math. J. 2020, 43, 256–267. [Google Scholar] [CrossRef]
  26. Dey, D.; Majhi, P. ∗-Ricci solitons and ∗-gradient Ricci solitons on 3-dimensional trans-Sasakian man-ifolds. Commun. Korean Math. Soc. 2020, 35, 625–637. [Google Scholar]
  27. Dey, S.; Sarkar, S.; Bhattacharyya, A. ∗-η Ricci soliton and contact geometry. Ric. Mat. 2021. [Google Scholar] [CrossRef]
  28. Dey, S.; Uddin, S. Conformal η-Ricci almost solitons on Kenmotsu manifolds. Int. J. Geom. Methods Mod. Phys. 2022, 19, 2250121. [Google Scholar] [CrossRef]
  29. Dey, S.; Roy, S. ∗-η-Ricci Soliton within the framework of Sasakian manifold. J. Dyn. Syst. Geom. Theor. 2020, 18, 163–181. [Google Scholar] [CrossRef]
  30. Ganguly, D.; Dey, S.; Ali, A.; Bhattacharyya, A. Conformal Ricci soliton and Quasi-Yamabe soliton on generalized Sasakian space form. J. Geom. Phys. 2021, 169, 104339. [Google Scholar] [CrossRef]
  31. Roy, S.; Dey, S.; Bhattacharyya, A.; Hui, S.K. ∗-Conformal η-Ricci Soliton on Sasakian manifold. Asian-Eur. J. Math. 2022, 15, 2250035. [Google Scholar] [CrossRef]
  32. Yang, Z.C.; Li, Y.; Erdoǧdub, M.; Zhu, Y.S. Evolving evolutoids and pedaloids from viewpoints of envelope and singularity theory in Minkowski plane. J. Geom. Phys. 2022, 104513, 1–23. [Google Scholar] [CrossRef]
  33. Li, Y.; Ganguly, D.; Dey, S.; Bhattacharyya, A. Conformal η-Ricci solitons within the framework of indefinite Kenmotsu manifolds. AIMS Math. 2022, 7, 5408–5430. [Google Scholar] [CrossRef]
  34. Li, Y.; Dey, S.; Pahan, S.; Ali, A. Geometry of conformal η-Ricci solitons and conformal η-Ricci almost solitons on Paracontact geometry. Open Math. 2022, 20, 1–20. [Google Scholar] [CrossRef]
  35. Li, Y.; Abolarinwa, A.; Azami, S.; Ali, A. Yamabe constant evolution and monotonicity along the conformal Ricci flow. AIMS Math. 2022, 7, 12077–12090. [Google Scholar] [CrossRef]
  36. Li, Y.; Khatri, M.; Singh, J.P.; Chaubey, S.K. Improved Chen’s Inequalities for Submanifolds of Generalized Sasakian-Space-Forms. Axioms 2022, 11, 324. [Google Scholar] [CrossRef]
  37. Li, Y.; Mofarreh, F.; Agrawal, R.P.; Ali, A. Reilly-type inequality for the ϕ-Laplace operator on semislant submanifolds of Sasakian space forms. J. Inequal. Appl. 2022, 1, 102. [Google Scholar] [CrossRef]
  38. Li, Y.; Mofarreh, F.; Dey, S.; Roy, S.; Ali, A. General Relativistic Space-Time with η1-Einstein Metrics. Mathematics 2022, 10, 2530. [Google Scholar] [CrossRef]
  39. Li, Y.; Uçum, A.; İlarslan, K.; Camcı, Ç. A New Class of Bertrand Curves in Euclidean 4-Space. Symmetry 2022, 14, 1191. [Google Scholar] [CrossRef]
  40. Li, Y.; Şenyurt, S.; Özduran, A.; Canlı, D. The Characterizations of Parallel q-Equidistant Ruled Surfaces. Symmetry 2022, 14, 1879. [Google Scholar] [CrossRef]
  41. Li, Y.; Haseeb, A.; Ali, M. LP-Kenmotsu manifolds admitting η-Ricci solitons and spacetime. J. Math. 2022, 2022, 6605127. [Google Scholar] [CrossRef]
  42. Li, Y.; Mofarreh, F.; Abdel-Baky, R.A. Timelike Circular Surfaces and Singularities in Minkowski 3-Space. Symmetry 2022, 14, 1914. [Google Scholar] [CrossRef]
  43. Li, Y.; Alluhaibi, N.; Abdel-Baky, R.A. One-Parameter Lorentzian Dual Spherical Movements and Invariants of the Axodes. Symmetry 2022, 14, 1930. [Google Scholar] [CrossRef]
  44. Li, Y.; Prasad, R.; Haseeb, A.; Kumar, S.; Kumar, S. A Study of Clairaut Semi-Invariant Riemannian Maps from Cosymplectic Manifolds. Axioms 2022, 11, 503. [Google Scholar] [CrossRef]
  45. Li, Y.; Nazra, S.H.; Abdel-Baky, R.A. Singularity Properties of Timelike Sweeping Surface in Minkowski 3-Space. Symmetry 2022, 14, 1996. [Google Scholar] [CrossRef]
  46. Li, Y.; Eren, K.; Ayvacı, K.H.; Ersoy, S. Simultaneous characterizations of partner ruled surfaces using Flc frame. AIMS Math. 2022, 7, 20213–20229. [Google Scholar] [CrossRef]
  47. Li, Y.; Gur, S.; Senyurt, S. The Darboux trihedrons of timelike surfaces in the Lorentzian 3-space. Int. J. Geom. Methods Mod. Phys. 2022, 1–35. [Google Scholar] [CrossRef]
  48. Li, Y.; Eren, K.; Ayvacı, K.H.; Ersoy, S. The developable surfaces with pointwise 1-type Gauss map of Frenet type framed base curves in Euclidean 3-space. AIMS Math. 2023, 8, 2226–2239. [Google Scholar] [CrossRef]
  49. Li, Y.; Erdoğdu, M.; Yavuz, A. Differential Geometric Approach of Betchow-Da Rios Soliton Equation. Hacet. J. Math. Stat. 2022, 1–12. [Google Scholar] [CrossRef]
  50. Li, Y.; Abdel-Salam, A.A.; Saad, M.K. Primitivoids of curves in Minkowski plane. AIMS Math. 2023, 8, 2386–2406. [Google Scholar] [CrossRef]
  51. Li, Y.; Erdoğdu, M.; Yavuz, A. Nonnull soliton surface associated with the Betchov-Da Rios equation. Rep. Math. Phys. 2022, 90, 241–255. [Google Scholar] [CrossRef]
  52. Li, Y.; Mondal, S.; Dey, S.; Bhattacharyya, A.; Ali, A. A Study of Conformal η-Einstein Solitons on Trans-Sasakian 3-Manifold. J. Nonlinear Math. Phy. 2022, 1–27. [Google Scholar] [CrossRef]
  53. Gür, S.; Şenyurt, S.; Grilli, L. The Dual Expression of Parallel Equidistant Ruled Surfaces in Euclidean 3-Space. Symmetry 2022, 14, 1062. [Google Scholar]
  54. Çalışkan, A.; Şenyurt, S. Curves and ruled surfaces according to alternative frame in dual space. Commun. Fac. Sci. Univ. 2020, 69, 684–698. [Google Scholar] [CrossRef]
  55. Çalışkan, A.; Şenyurt, S. The dual spatial quaternionic expression of ruled surfaces. Therm. Sci. 2019, 23, 403–411. [Google Scholar] [CrossRef] [Green Version]
  56. Şenyurt, S.; Çalışkan, A. The quaternionic expression of ruled surfaces. Filomat 2018, 32, 5753–5766. [Google Scholar] [CrossRef] [Green Version]
  57. Şenyurt, S.; Gür, S. Spacelike surface geometry. Int. J. Geom. Methods Mod. Phys. 2017, 14, 1750118. [Google Scholar] [CrossRef]
  58. As, E.; Şenyurt, S. Some Characteristic Properties of Parallel-Equidistant Ruled Surfaces. Math. Probl. Eng. 2013, 2013, 587289. [Google Scholar] [CrossRef] [Green Version]
  59. Özcan, B.; Şenyurt, S. On Some Characterizations of Ruled Surface of a Closed Timelike Curve in Dual Lorentzian Space. Adv. Appl. Clifford Algebr. 2012, 22, 939–953. [Google Scholar]
  60. Antić, M.; Moruz, M.; Van, J. H-Umbilical Lagrangian Submanifolds of the Nearly Kähler S 3 × S 3 . Mathematics 2020, 8, 1427. [Google Scholar] [CrossRef]
  61. Antić, M.; Djordje, K. Non-Existence of Real Hypersurfaces with Parallel Structure Jacobi Operator in S6(1). Mathematics 2022, 10, 2271. [Google Scholar] [CrossRef]
  62. Antić, M. Characterization of Warped Product Lagrangian Submanifolds in C n . Results Math. 2022, 77, 1–15. [Google Scholar] [CrossRef]
  63. Antić, M.; Vrancken, L. Conformally flat, minimal, Lagrangian submanifolds in complex space forms. Sci. China Math. 2022, 65, 1641–1660. [Google Scholar] [CrossRef]
  64. Antić, M.; Hu, Z.; Moruz, M.; Vrancken, L. Surfaces of the nearly Kähler S 3 × S 3 preserved by the almost product structure. Math. Nachr. 2021, 294, 2286–2301. [Google Scholar] [CrossRef]
  65. Antić, M. A class of four-dimensional CR submanifolds in six dimensional nearly Kähler manifolds. Math. Slovaca 2018, 68, 1129–1140. [Google Scholar] [CrossRef]
  66. Antić, M. A class of four dimensional CR submanifolds of the sphere S6(1). J. Geom. Phys. 2016, 110, 78–89. [Google Scholar] [CrossRef]
  67. Ali, A.T. Non-lightlike constant angle ruled surfaces in Minkowski 3-space. J. Geom. Phys. 2020, 157, 103833. [Google Scholar] [CrossRef]
  68. Ali, A.T. A constant angle ruled surfaces. Int. J. Geom. 2018, 7, 69–80. [Google Scholar]
  69. Ali, A.T. Non-lightlike ruled surfaces with constant curvatures in Minkowski 3-space. Int. J. Geom. Methods Mod. Phys. 2018, 15, 1850068. [Google Scholar] [CrossRef]
  70. Ali, A.T.; Hamdoon, F.M. Surfaces foliated by ellipses with constant Gaussian curvature in Euclidean 3-space. Korean J. Math. 2017, 25, 537–554. [Google Scholar]
  71. Ali, A.T.; Abdel Aziz, H.S.; Sorour, A.H. On some geometric properties of quadric surfaces in Euclidean space. Honam Math. J. 2016, 38, 593–611. [Google Scholar] [CrossRef] [Green Version]
  72. Ali, A.T.; Abdel Aziz, H.S.; Sorour, A.H. On curvatures and points of the translation surfaces in Euclidean 3-space. J. Egypt. Math. Soc. 2015, 23, 167–172. [Google Scholar] [CrossRef] [Green Version]
  73. Jäntschi, L. Introducing Structural Symmetry and Asymmetry Implications in Development of Recent Pharmacy and Medicine. Symmetry 2022, 14, 1674. [Google Scholar] [CrossRef]
  74. Jäntschi, L. Binomial Distributed Data Confidence Interval Calculation: Formulas, Algorithms and Examples. Symmetry 2022, 14, 1104. [Google Scholar] [CrossRef]
  75. Jäntschi, L. Formulas, Algorithms and Examples for Binomial Distributed Data Confidence Interval Calculation: Excess Risk, Relative Risk and Odds Ratio. Mathematics 2021, 9, 2506. [Google Scholar] [CrossRef]
  76. Donatella, B.; Jäntschi, L. Comparison of Molecular Geometry Optimization Methods Based on Molecular Descriptors. Mathematics 2021, 9, 2855. [Google Scholar]
  77. Mihaela, T.; Jäntschi, L.; Doina, R. Figures of Graph Partitioning by Counting, Sequence and Layer Matrices. Mathematics 2021, 9, 1419. [Google Scholar]
  78. Kumar, S.; Kumar, D.; Sharma, J.R.; Jäntschi, L. A Family of Derivative Free Optimal Fourth Order Methods for Computing Multiple Roots. Symmetry 2020, 12, 1969. [Google Scholar] [CrossRef]
  79. Deepak, K.; Janak, R.; Jäntschi, L. A Novel Family of Efficient Weighted-Newton Multiple Root Iterations. Symmetry 2020, 12, 1494. [Google Scholar]
  80. Janak, R.; Sunil, K.; Jäntschi, L. On Derivative Free Multiple-Root Finders with Optimal Fourth Order Convergence. Mathematics 2020, 8, 1091. [Google Scholar]
  81. Jäntschi, L. Detecting Extreme Values with Order Statistics in Samples from Continuous Distributions. Mathematics 2020, 8, 216. [Google Scholar] [CrossRef] [Green Version]
  82. Deepak, K.; Janak, R.; Jäntschi, L. Convergence Analysis and Complex Geometry of an Efficient Derivative-Free Iterative Method. Mathematics 2019, 7, 919. [Google Scholar]
  83. Todorčević, V. Harmonic Quasiconformal Mappings and Hyperbolic Type Metrics; Springer International Publishing: Berlin/Heidelberg, Germany, 2019. [Google Scholar]
  84. Todorčević, V. Subharmonic behavior and quasiconformal mappings. Anal. Math. Phys. 2019, 9, 1211–1225. [Google Scholar] [CrossRef]
  85. Kojić, V.; Pavlović, M. Subharmonicity of |f|p for quasiregular harmonic functions, with applications. J. Math. Anal. Appl. 2008, 342, 742–746. [Google Scholar] [CrossRef] [Green Version]
  86. Kojić, V. Quasi-nearly subharmonic functions and conformal mappings. Filomat 2007, 21, 243–249. [Google Scholar] [CrossRef] [Green Version]
  87. Manojlović, V.; Vuorinen, M. On quasiconformal maps with identity boundary values. Trans. Am. Math. Soc. 2011, 363, 2367–2479. [Google Scholar] [CrossRef] [Green Version]
  88. Manojlović, V. On bilipschicity of quasiconformal harmonic mappings. Novi Sad J. Math. 2015, 45, 105–109. [Google Scholar] [CrossRef]
  89. Manojlović, V. Bilipschitz mappings between sectors in planes and quasi-conformality. Funct. Anal. Approx. Comput. 2009, 1, 1–6. [Google Scholar]
  90. Manojlović, V. Bi-Lipschicity of quasiconformal harmonic mappings in the plane. Filomat 2009, 23, 85–89. [Google Scholar] [CrossRef]
  91. Manojlović, V. On conformally invariant extremal problems. Appl. Anal. Discret. Math. 2009, 3, 97–119. [Google Scholar] [CrossRef]
  92. Duggal, K.L. Almost Ricci solitons and physical applications. Int. Electron. J. Geom. 2017, 10, 1–10. [Google Scholar]
  93. Woolgar, E. Some applications of Ricci flow in physics. Can. J. Phys. 2008, 86, 645–651. [Google Scholar] [CrossRef] [Green Version]
  94. Blair, D.E. Riemannian Geometry of Contact and Symplectic Manifolds, 2nd ed.; Birkhäuser: Boston, MA, USA, 2010. [Google Scholar]
  95. Oubiña, J.A. New classes of almost contact metric structures. Publ. Math. 1985, 32, 187–193. [Google Scholar]
  96. Gray, A.; Hervella, L.M. The Sixteen Classes of Almost Hermitian Manifolds and Their Linear Invariance. Ann. Mat. Pura Appl. 1980, 123, 35–58. [Google Scholar] [CrossRef]
  97. Blair, D.E.; Oubiña, J.A. Conformal and Related Changes of Metric on the Product of Two Almost Contact Metric Manifolds. Publicacions Mat. 1990, 34, 199–207. [Google Scholar] [CrossRef] [Green Version]
  98. Marrero, J.C. The local structure of trans-Sasakian manifolds. Ann. Mat. Pura Appl. (IV) 1992, 162, 77–86. [Google Scholar] [CrossRef]
  99. Yano, K. Integral Formulas in Riemannian Geometry; Marcel Dekker: New York, NY, USA, 1970. [Google Scholar]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Chen, Z.; Li, Y.; Sarkar, S.; Dey, S.; Bhattacharyya, A. Ricci Soliton and Certain Related Metrics on a Three-Dimensional Trans-Sasakian Manifold. Universe 2022, 8, 595. https://doi.org/10.3390/universe8110595

AMA Style

Chen Z, Li Y, Sarkar S, Dey S, Bhattacharyya A. Ricci Soliton and Certain Related Metrics on a Three-Dimensional Trans-Sasakian Manifold. Universe. 2022; 8(11):595. https://doi.org/10.3390/universe8110595

Chicago/Turabian Style

Chen, Zhizhi, Yanlin Li, Sumanjit Sarkar, Santu Dey, and Arindam Bhattacharyya. 2022. "Ricci Soliton and Certain Related Metrics on a Three-Dimensional Trans-Sasakian Manifold" Universe 8, no. 11: 595. https://doi.org/10.3390/universe8110595

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