검색
Article Search

JMB Journal of Microbiolog and Biotechnology

OPEN ACCESS eISSN 0454-8124
pISSN 1225-6951
QR Code

Article

Kyungpook Mathematical Journal 2021; 61(3): 645-660

Published online September 30, 2021

Copyright © Kyungpook Mathematical Journal.

3-Dimensional Trans-Sasakian Manifolds with Gradient Generalized Quasi-Yamabe and Quasi-Yamabe Metrics

Mohammed Danish Siddiqi, Sudhakar Kumar Chaubey, Ghodratallah Fasihi Ramandi*

Department of Mathematics, Faculty of Science, Jazan University, Jazan, Kingdom of Saudi Arabia
e-mail : msiddiqi@jazanu.edu.sa

Section of Mathematics, Department of Information Technology, University of Technology and Applied Sciences, Shinas, P. O. box 77, Postal Code 324, Oman
e-mail : sudhakar.chaubey@shct.edu.om

Department of Pure Mathematics, Faculty of Science, Imam Khomeini International University, Qazvin, Iran
e-mail : fasihi@sci.ikiu.ac.ir

Received: January 25, 2021; Revised: June 21, 2021; Accepted: July 6, 2021

This paper examines the behavior of a 3-dimensional trans-Sasakian manifold equipped with a gradient generalized quasi-Yamabe soliton. In particular, It is shown that α-Sasakian, β-Kenmotsu and cosymplectic manifolds satisfy the gradient generalized quasi-Yamabe soliton equation. Furthermore, in the particular case when the potential vector field ζ of the quasi-Yamabe soliton is of gradient type ζ = grad(ψ), we derive a Pois-son’s equation from the quasi-Yamabe soliton equation. Also, we study harmonic aspects of quasi-Yamabe solitons on 3-dimensional trans-Sasakian manifolds sharing a harmonic potential function ψ. Finally, we observe that 3-dimensional compact trans-Sasakian mani-fold admits the gradient generalized almost quasi-Yamabe soliton with Hodge-de Rham po-tential ψ. This research ends with few examples of quasi-Yamabe solitons on 3-dimensional trans-Sasakian manifolds.

In the past twenty years, geometric flows have emerged as versatile tools for describing geometric structures in Riemannian geometry. A specific class of solutions on which the metric evolves by dilation and diffeomorphisms plays a vital part in the study of singularities of the flows as they appear as possible singularity models. They are often called soliton solutions.

The theory of Yamabe flow was popularized by Hamilton in his prime research work [12] as a tool for constructing metrics of constant scalar curvature on an n-dimensional Riemannian manifold (Mn,g), n ≥ 3. The Yamabe flow is an evolution equation for metrics on Riemannian manifolds. It is given by

tg(t)=rg(t),g(0)=g0,

where r is the scalar curvature corresponding to Riemannian metric g and t is time. It is used to deform a metric by smoothing out its singularities.

A Yamabe soliton is a special solution of the Yamabe flow that moves by one parameter a family of diffeomorphisms generated by a fixed vector field E on Mn with a real constant λ satisfying the following equation

12LEg=(rλ)g.

Here LEg is the Lie derivative of the metric g along the vector field E, called the soliton vector field of the Yamabe soliton [12]. If λ<0, λ>0, or λ=0, then the (Mn,g) is called a Yamabe shrinker, Yamabe expander, or Yamabe steady soliton, respectively.

When the vector field E is the gradient of a smooth function ψ:Mn&#_10230;, the manifold will be called a gradient Yamabe soliton. The function ψ is called the potential function of the gradient Yamabe soliton. In this case equation (1.2) becomes

Hessψ=(rλ)g,

where Hess ψ stands for the Hessian of the potential function ψ. The gradient Yamabe soliton equation (1.3) links geometric information about the curvature of the manifold to the scalar curvature tensor and the geometry of the level sets of the potential function by means of their second fundamental form. This makes gradient Yamabe solitons under some curvature conditions an interesting topic of study.

An Einstein manifold [2] with a constant potential function is called a trivial gradient Ricci soliton. Gradient Yamabe solitons [14] play an important role in Yamabe flow as they correspond to self-similar solutions, and often arise as singularity models [18].

Introduced by Chen and Desahmukh in [4], a Riemannian manifold (Mn, g) is called a quasi-Yamabe solitonif it admits a vector field E such that

LEg+2(λr)g=2μEE,

for some real constant λ and smooth function µ, where E is the dual 1-form of E. The vector field E is also called a soliton vector field for the quasi-Yamabe soliton. We denote the quasi-Yamabe soliton satisfying (1.4) by (Mn,g,E,λ,μ). If E=ψ, then (1.4) becomes

2ψ=(rλ)g+μdψdψ,

which is the gradient generalized quasi-Yamabe soliton studied by Huang et al. [10] and Leandro et al. [13], where 2ψ denotes Hessian of ψ. This class of closely related Yamabe solitons hase be extensively studied; for further details see ([1], [5], [8], [19], [20], [22], [23], [25]).

According to Pigola et al. [17], if we replace the constant λ in (1.4) and (1.5) with a smooth function λC(M), called soliton function, then we can say that (Mn, g) is an almost quasi-Yamabe and gradient generalized almost quasi-Yamabe soliton, respectively.

On one hand, in 1985, Oubina [15] introduced a new class of almost contact metric manifolds, known as trans-Sasakian manifolds. This class consists the Sasakian, the Kenmotsu and the cosymplectic structures. The properties of trans-Sasakian manifolds have been studied by several authors, like Blair [3] and Marrero [14]. %The trans-Sasakian manifolds arose in a natural way from the classification of almost contact metric structures and they appear as a natural generalization of both Sasakian and Kenmotsu manifolds The main goal of this paper is to characterize the three-dimensional trans-Sasakian manifolds equipped with gradient generalized quasi-Yamabe solitons, quasi-Yamabe metrics, and gradient generalized almost quasi-Yamabe metrics.

Let M be a connected almost contact metric manifold equipped with almost contact metric structure (φ,ζ,η,g) consisting of a (1,1) tensor field ⊎, a vector field ζ, a 1-form η and a positive definite metric g such that

φ2=I+ηζ,η(ζ)=1,η°φ=0,φζ=0, g(φE,ϕF)=g(E,F)η(E)η(F),η(E)=g(E,ζ)

for all E,Fχ(M), where χ(M) denotes the collection of all smooth vector fields of M and dimM=2m+1.

In the Grey and Harvella [9] classification of almost Hermitian manifolds, there appears a class W4 of Hermitian manifolds which are closely related to the conformal Kaehler manifolds. In their classification, the class C6C5 (see [3], [6], [15], [16]) coincides with the class of trans-Sasakian structure of type (α,β). In fact, the local nature of two sub classes, namely C6 and C5 of trans-Sasakian structures are characterized completely. An almost contact metric structure (ϕ, ξ, η, g) on M is called a trans-Sasakian [21] if (M×,J,G) belongs to the class W4, where J is an almost complex structure on M× defined by

JE,fddt=φEfζ,η(E)ddt

for all vector fields X on M and smooth functions f on M×. Here G is the product metric on M× and denotes the set of real numbers. This may be expressed by the condition

(Eφ)F=α(g(E,F)ζη(F)E)+β(g(φE,F)ξη(F)φE)

where α and β are some scalars functions on M and ∇ denotes the Levi-Civita connection with respect to g. We note that the trans-Sasakian structures of type (0,0), (α,0) and (0,β) are the cosymplectic, α-Sasakian and β-Kenmotsu structures, respectively. In particular, if α=1,β=0, α=0,β=1 and α=0,β=0, then the trans-Sasakian manifold reduces to Sasakian, Kenmotsu and cosymplectic manifolds, respectively. From 1.3, it follows that

Eζ=αφE+β[Eη(E)ζ],

equivalent to

(Eη)F=αg(φE,F)+β[g(E,F)η(E)η(F)],E,Fχ(M).

In a 3-dimensional trans-Sasakian manifold M, we have the following relations [7]

R(E,F)ζ=(α2β2)[η(F)Eη(E)F]+2αβ[η(F)φEη(E)φF]    +[(Eα)φE(Xα)φF+(Fβ)φ2E(Eβ)φ2F], S(E,ζ)=[(2(α2β2)(ζβ)]η(E)+((φE)α)+(Eβ), Qζ=(2(α2β2)(ζβ))ζ+φ(gradα)(gradβ),

where R, S and Q denote the curvature tensor, Ricci tensor and Ricci operator of g, respectively. Also grad stands for gradient. Further, in a three-dimensional trans-Sasakian manifold we have

φ(gradα)=gradβ,

and

2αβ+(ζα)=0.

Using (1.9) and (1.10), for constants α and β, we have

R(ζ,E)F=(α2β2)[g(E,F)ζη(F)E], R(E,F)ζ=(α2β2)[η(F)Eη(E)F], S(E,ζ)=[2(α2β2)]η(E).

For a smooth function ψ on M, the gradient and Hessian of ψ are, respectively, defined by

g(gradψ,E)=E(ψ)and(Hessψ)(E,F)=g(Egradψ,F),E,FΓ(TM).

For EΓ(TM), we define EΓ(T¯M) by

E(F)=g(E,F).

The generalized quasi-Yamabe soliton equation [4] in a Riemannian manifold M is defined by

12LEg=μEE+(rλ)g.

Equation (1.3) is a generalization of Einstein manifold [10]. Note that if E=gradψ, where ψC(M), the gradient generalized quasi-Yamabe soliton equation is given by [10]:

Hessψ=μdψdψ+(rλ)g.

Main Result:

Theorem 3.1. Let M be a three-dimensional trans-Sasakian manifold satisfy the gradient generalized quasi-Yamabe soliton equation (1.4) with condition μ[λ+6(α2β2)]=0, then ψ is a constant function. Furthermore, if µ≠ 0, then λ=6(α2β2) is negative, that is, a three-dimensional trans-Sasakian manifold admits a shrinking gradient generalized quasi-Yamabe soliton.

From Theorem 3.1, we get the following remarks:

Remark. Let a three-dimensional trans-Sasakian manifold M satisfy the gradient generalized quasi-Yamabe soliton equation Hessψ=(rλ)g, then ψ is constant and M is η-Einstein.

Remark. In a three-dimensional trans-Sasakian manifold M, there is no non-constant smooth function ψ such that Hessψ=λg for some constant λ.

To prove the Theorem 3.1, we have to demonstrate the following lemmas.

Lemma 3.2. Let M be a three-dimensional trans-Sasakian manifold. Then we have

(Lζ(LEg))(F,ζ)=(α2β2){g(E,F)η(E)η(F)}+g(ζζE,F)+Fg(ζE,ξ),

where E,FΓ(TM).

Proof. From the property of Lie-derivative we note that

(Lζ(LEg))(E,ζ)=ζ((LEg)(F,ζ))(LEg)(LζF,ζ)(LEg)(F,Lζζ).

Since LζF=[ζ,F] and Lζζ=[ζ,ζ], therefore the above equation can be written as

(Lζ(LEg))(F,ζ)=ζg(FE,ζ)+ζg(ζE,F)g([ζ,F]E,ζ)g(ζE,[ζ,F])    =g(ζFE,ζ)+g(FE,ζζ)+g(ζζE,F)    +g(ζE,ζF)g(ζE,ζF)g([ζ,F]E,ζ)+g(ζE,Fζ).

From (1.4) we get ζζ=0, so the last equation gives

(Lζ(LEg))(F,ζ)=g(ζFE,ζ)+g(ζζE,F)g([ζ,F]E,ζ)      +Fg(ζE,ζ)g(FζE,ζ),

which gives

(Lζ(LEg))(F,ζ)=g(R(ζ,F)E,ζ)+g(ζζE,F)+Yg(ζE,ζ).

From (1.12), we lead

g(R(ζ,F)E,ζ)=g(R(F,ζ)ζ,E)=(α2β2){g(E,F)η(E)η(F)}.

The Lemma 3.2 follows from the last two equations. Particularly, if Y is orthogonal to ζ then equation (1.5) assumes the form

(Lζ(LEg))(E,ζ)=(α2β2)g(E,F)+g(ζζE,F)+Fg(ζE,ζ)

for all Eχ(M) and F orthogonal to ζ.

Lemma 3.3. Let M be a Riemannian manifold, and let ψC(M). Then we have

(Lζ(dψdψ))(F,ζ)=F(ζ(ψ))ζ(ψ)+F(ψ)ζ(ζ(ψ)).

Proof. We calculate:

(Lζ(dψdψ))(F,ζ)=ζ(F(ψ)ζ(ψ))[ζ,F](ψ)ζ(ψ)F(ψ)[ζ,ζ](ψ)          =ζ(F(ψ))ζ(ψ)+F(ψ)ζ(ζ(ψ))[ζ,F](ψ)ζ(ψ).

Since [ζ,F](ψ)=ζ(F(ψ))F(ζ(ψ)), therefore the above equation becomes

(Lζ(dψdψ))(F,ζ)=[ζ,F](ψ)ζ(ψ)+F(ζ(ψ))ζ(ψ)+F(ψ)ζ(ζ(ψ))[ζ,F](ψ)ζ(ψ)          =F(ζ(ψ))ζ(ψ)+F(ψ)ζ(ζ(ψ)).

Hence the statement of Lemma 3.3 is proved.

Lemma 3.4. Let a three-dimensional trans-Sasakian manifold M satisfy the gradient generalized quasi-Yamabe soliton equation 1.4. Then we have

ζgradψ=[λ6(α2β2)]ζ+μζ(ψ)gradψ.

Proof. Let FΓ(TM), then form the definition of Ricci tensor S, scalar curvature r and the curvature condition 1.12, we have

S(E,F)=i=13g(R(ζ,ei)ei,F)=i=13g(R(ei,F)ζ,ei)=2(α2β2),          r=6(α2β2),

where e1,e2,e3 is an orthonormal frame on M. From the above equations, we infer

λg(ζ,F)+rg(ζ,F)=[λ+6(α2β2)]g(ζ,F).

From (1.4) and (1.9), we obtain

(Hessψ)(ζ,F)=μζ(ψ)F(ψ)+[6(α2β2)λ]g(ζ,F)      =μζ(ψ)g(gradψ,F)+[6(α2β2)λ]g(ζ,F).

The Lemma 3.3 follows from equation (1.10) and the definition of Hessian (see (1.1)).

Now, we are going to prove our main Theorem 3.1 by using Lemma 3.2, Lemma 3.3 and Lemma 3.4.

Proof of Theorem 3.1. Let us suppose that the three-dimensional trans-Sasakian manifold satisfying the gradient generalized quasi-Yamabe soliton equation (1.10) and λ,μ. Let YΓ(TM), then Lemma together with E=grad ψ leads to

2(Lζ(Hessψ))(F,ζ)=(α2β2){F(ψ)ζ(ψ)η(F)}        +g(ζζ gradψ,F)+Fg(ζ gradψ,ζ).

From Lemma 3.4 and equations (1.1), (1.2), (1.4), (1.11), we get

2(Lζ(Hessψ))(F,ζ)=F(ψ)[(α2β2)+μ(ζ(ζ(ψ)))+(μ(ζ(ψ)))2]  +{μζ(ψ)[6(α2β2)λ]ζ(ψ)(α2β2)}η(F)  +F[6(α2β2)λ+μ(ζ(ψ))2]

for all FΓ(TM). Taking F orthogonal to ζ and therefore the above equation becomes

2(Lζ(Hessψ))(F,ζ)=F[6(α2β2)λ+μ(ζ(ψ))2]  +F(ψ)[(α2β2)+μ(ζ(ζ(ψ)))+(μ(ζ(ψ)))2].

Next, the Lie derivative of the gradient generalized quasi-Yamabe soliton equation (1.4) along the vector field ζ yields

2(Lζ(Hessψ))(F,ζ)=μ(Lζ(dψdψ))(F,ζ).

The last two equations together with Lemma infer

F(ψ)μζ(ζ(ψ))F(ψ)+μ2ζ(ψ)2F(ψ)2μζ(ψ)F(ζ(ψ))    =2μF(ζ(ψ))ζ(ψ)2μF(ψ)ζ(ζ(ψ)),

which is equivalent to

F(ψ)[1+μζ(ζ(ψ))+μ2ζ(ψ)2]=0.

According to Lemma 4.3, we have

μζ(ζ(ψ))=μζg(ζ,grad ψ)    =ag(ζ,ζ gradψ)      =μ[λ+6(α2β2)]μ2ζ(ψ)2,

by equations (1.15) and (1.16), we obtain

F(ψ)[λ+6(α2β2)]=0,

since [λ+6(α2β2)]0, we find that F(ψ)=0, i.e., gradψ is parallel to ζ. Hence grad ψ=0 as D=kerη is not integrable any where, which means ψ is a constant function.

Now, for particular values of α and β we turn up the following cases:

Case: For α=0, (β=1) and (α=β=0) we can state the following results:

Corollary 3.5. Let M be a 3-dimensional β-Kenmotsu (or Kenmotsu) manifold satisfies the gradient generalized quasi-Yamabe soliton 1.4 condition μ[λ6β2)]0, then ψ is a constant function. Furthermore, if µ≠ 0, implies λ=6β2), then M is expanding.

Case: For β=0, or (α=1) we can state:

Corollary 3.6. Let M be a 3-dimensional α-Sasakian (or Sasakian) manifold satisfies the gradient generalized quasi-Yamabe soliton1.4 condition μ[λ+6α2)]0, then ψ is a constant function. Furthermore, if µ≠ 0, implies λ=6α2), then M is shrinking.

Case: For α=β=0, we can state:

Corollary 3.7. Let M be a 3-dimensional cosymplectic manifold satisfies the gradient generalized quasi-Yamabe soliton 1.4 condition μ[λ]0, then ψ is a constant function. Furthermore, if µ≠ 0, implies λ=0, then M is steady.

Again, assume the equation

Lζg+(λR)g+μEE=0

where g is a Riemannian metric and R is the scalar curvature, ζ is vector field, E is a 1-form and λ and µ are real constant. The data (g,ζ,λ,µ) satisfies the equation (4.1) is called the quasi-Yamabe soliton. In particular, if µ=0, (g,ζ,λ) is a Yamabe soliton.

Using the definition of Lie derivative and (4.1), we obtain

(Rλ)g(F,G)=μE(F)E(G)12[g(Fζ,G)+g(F,Gζ)],

for any F,Gχ(M).

Contracting (4.2) we get

3λμ=3Rdiv(ζ).

Let (M,g,⊎,η,ζ) be a 3-dimensional trans-Sasakian manifold and (g,ζ,λ,µ) be a quasi-Yamabe soliton on M. Writing 4.2) for F=G=ζ, we obtain

λμ=6(α2β2).

Therefore

λ=6(α2β2)+div(ζ)2μ=12(α2β2)+div(ζ)2

Using (4.5) we can state the following results.

Theorem 4.1. Let (M,η,φ,ζ,g) be a 3-dimensional trans-Sasakian manifold and E be the g-dual 1-form of the gradient vector field ζ=grad(ψ). If (4.1) define a quasi-Yamabe soliton with non vanishing µ in M, then the Poisson equation satisfied by ψ becomes

Δ(ψ)=2[μ+12(α2β2)].

Once again, considering the equation (4.5) we can also obtain

Δ(ψ)=2[λ+6(α2β2)].

Remark.([24]) A C function f:M&#_10230; is said to be harmonic if Δf=0 , where Δ is the Laplacian operator in M.

Now, from equation (4.7) and using above remark, we obtain the following results:

Theorem 4.2. Let (M,η,φ,ζ,g) be a 3-dimensional trans-Sasakian manifold and E be the g-dual 1-form of the gradient potential vector field ζ=grad(ψ) . If the potential function ψ is harmonic, then quasi-Yamabe soliton is shrinking for the value of λ=3(α2β2).

Corollary 4.3. Let (M,η,φ,ζ,g) be a 3-dimensional α-Sasakian (or Sasakian) manifold and E be the g-dual 1-form of the gradient potential vector field ζ=grad(ψ) . If the potential function ψ is harmonic, then quasi-Yamabe soliton is shrinking for the value of λ=3α2.

Corollary 4.4. Let (M,η,φ,ζ,g) be a 3-dimensional β-Kenmotsu (or Kenmotsu) manifold and E be the g-dual 1-form of the gradient potential vector field ζ=grad(ψ). If the potential function ψ is harmonic, then quasi-Yamabe soliton is expanding for the value of λ=3β2.

Corollary 4.5. Let (M,η,φ,ζ,g) be a 3-dimensional cosymplectic manifold and E be the g-dual 1-form of the gradient potential vector field ζ=grad(ψ) . If the potential function ψ is harmonic, then quasi-Yamabe soliton is steady for the value of λ=0.

Example 5.1. Let M=(x,y,z)3:z0, where (x,y,z) is the standard coordinates of 3.

The vector fields are

e1=zyx,e2=y,e3=2x

Let g be the Riemannian metric defined by

g(e1,e1)=g(e2,e2)=g(e3,e3)=1,g(e1,e3)=g(e2,e3)=g(e1,e2)=0

that is, the form of the metric becomes Let η be the 1-form defiend by η(Z)=g(Z,e3) for any Zχ(M).

Also, let ⊎ be the (1,1) tensor field defined by

φ(e1)=e2,φ(e2)=e1,φ(e3)=0.

Thus, using the linearity of ⊎ and g, we have

η(e3)=0,,η(e1)=0,η(e2)=0, [e1,e2]=12e3,[e2,e3]=0,[e1,e3]=0, φ2Z=Z+η(Z)e3, g(φZ,φW)=g(Z,W)η(Z)η(W)

for any Z,Wχ(M).

Then for e3=ξ, the structure (φ,ξ,η,g) defines an almost contact metric structure on M.

Let ∇ be the Levi-Civita connection with respect to the metric g, then we have

2g(XY,Z)=Xg(Y,Z)+Yg(Z,X)Zg(X,Y)g(X,[Y,Z])g(Y,[X,Z])+g(Z,[X,Y]),

which is known as Koszul's formula.

Using Koszul's formula we have

e1e1=0,e1e2=14e3,e1e3=14e3, e2e1=14e3,e2e2=0,e2e3=14e1, e3e1=14e2,e3e2=14e1,e3e0=0.

From (5.1) we find that the structure (φ,ξ,η,g) satisfies the formula (4.5) for α=14 and ξ=e3. Hence the manifold is a 3-dimensional trans-Sasakian manifold of type (α,0) with the constant structure function α=14 and β=0.

Then the Riemannian and Ricci curvature tensor fields are given by:

R(e1,e2)e3=0,R(e2,e3)e3=116e2,R(e1,e3)e3=116e1, R(e1,e2)e2=316e1,R(e2,e3)e2=116e3,R(e1,e3)e2=0, R(e1,e2)e1=316e2,R(e2,e3)e1=0,R(e1,e3)e=116e3.

From the above expressions of the curvature tensor we obtain

S(e1,e1)=g(R(e1,e2)e2,e1)+g(R(e1,e3)e3,e1)=18

similarly we have

S(e1,e1)=S(e2,e2)=18,andS(e3,e3)=18.

Now, we have constant scalar curvature as follows,

R=S(e1,e1)+S(e2,e2)+S(e3,e3)=18.

By the definition of quasi-Yamabe soliton and using (1.4), we obtain

2β[g(ei,ei)+η(ei)η(ei)]+2(λR)g(ei,ei)+2μX(ei)X(ei)=0

for all i1,2,3, and we have

2(1+δi3)+2(λR)+2μδi3=0

for all i1,2,3.

Therefore λ=12 and μ=38 the data (g,ξ,λ,μ) admitting the shrinking quasi-Yamabe soliton on 3-dimensional trans-Sasakian manifolds with λ<0.

In [7] De and Sarkar proved that if a 3-dimensional trans-Sasakian

manifold is of constant curvature is compact and connected.

On the other hand, The classical theorem of de-Rham-Hodge asserts that the cohomology of an oriented closed Riemannian manifold can be represented by harmonic forms. The similar one still holds for an oriented compact Riemannian manifold with boundary by imposing certain boundary conditions, such as absolute and relative ones.

We consider M as a compact orientable trans-Sasakian manifold and X∈χ(M). Then Hodge-de Rham decomposition theorem [11] implies that E can be expressed as

E=h+F,

where hC(M) and div(F)=0. The function h is called the Hodge-de Rham potential [11].

Theorem 6.1. If (g,E,λ,μ) is a compact gradient almost quasi-Yamabe soliton on trans-Sasakian manifold M. If M is also a gradient almost quasi-Yamabe soliton with potential function ψ, then up to a constant, f equals to the Hodge-de Rham potential.

Proof. Since (g,E,λ,μ) is a compact almost quasi-Yamabe soliton, now taking the trace of (1.4), we find

div(E)=(Rλ)n+trce(μEE),

Hodge-de Rham decomposition implies that div(E)=Δh, hence the above equation, we get

R=λΔhn+13trce(μEE).

Again since M is generalized gradient almost quasi-Yamabe soliton with Perelman potential f, hence taking trace of (1.5), we have

R=λΔψ3+13μE2.

Now, equating the equations (5.3) and (5.4), we find 13Δ(ψh)=0. Hence ψ-h is a harmonic function in compact trans-Sasakian manifold. Hence f=h+c, for some constant c.

Let M=(x,y,z)3:z0 where (x,y,z) are the standard coordinates of 3.

The vector fields are

e1=zx,e2=zye3=zz

Let g be the Riemannian metric defined by

g(e1,e1)=g(e2,e2)=g(e3,e3)=1,g(e1,e3)=g(e2,e3)=g(e1,e2)=0

that is, the form of the metric becomes

g=dx2+dy2+dz2z2.

Let η be the 1-form defined by η(Z)=g(Z,e3) for any Zχ(M).

Also, let φ be the (1,1) tensor field defined by

φ(e1)=e2,φ(e2)=e1,φ(e3)=0.

Thus, using the linearity of φ and g, we have

η(e3)=0,,η(e1)=0,η(e2)=0, [e1,e2]=0,[e2,e3]=e2,[e1,e3]=e1, φ2Z=Z+η(Z)e3, g(φZ,φW)=g(Z,W)η(Z)η(W)

for any Z,Wχ(M).

Then for e3=ξ, the structure (φ,ξ,η,g) defines an almost contact metric structure on M.

Let ∇ be the Levi-Civita connection with respect to the metric g, then we have

2g(XY,Z)=Xg(Y,Z)+Yg(Z,X)Zg(X,Y)g(X,[Y,Z])g(Y,[X,Z])+g(Z,[X,Y]),

which is known as Koszul's formula.

Using Koszul's formula we have

e1e1=e3,e1e2=0,e1e3=e1, e2e1=0,e2e2=e3,e2e3=e2, e3e1=0,e3e2=0,e3e3=0.

From (5.1) we find that the manifold satisfies (1.4) for α=0 and β=-1 and ξ=e3. Hence the manifold is a 3-dimensional trans-Sasakian manifold of type (0,β) with the constant structure function α=0 and β=-1 [7].

Then the Riemannian and Ricci curvature tensor fields are given by:

R(e1,e2)e3=0,R(e2,e3)e3=e3,R(e1,e3)e3=e1, R(e1,e2)e2=e1,R(e2,e3)e2=e3,R(e1,e3)e2=0, R(e1,e2)e1=e2,R(e2,e3)e1=0,R(e1,e3)e=e3.

From the above expressions of the curvature tensor we obtain

S(e1,e1)=g(R(e1,e2)e2,e1)+g(R(e1,e3)e3,e1)=2

similarly, we have

S(e1,e1)=S(e2,e2)=S(e3,e3)=2.

Now, the scalar curvature

R=S(e1,e1)+S(e2,e2)+S(e3,e3)=6.

Because of scalar curvature r = 6, from Theorem (), we can conclude that M is an Einstein manifold.

By the definition of quasi-Yamabe soliton and using (1.4), we obtain

2β[g(ei,ei)+η(ei)η(ei)]+2(λR)g(ei,ei)+2μX(ei)X(ei)=0

for all i1,2,3, and we have

2(1+δi3)+2(λR)+2μδi3=0

for all i1,2,3.

Therefore λ =-1 and μ=38 the data (g,ξ,λ,µ) admitting the shrinking quasi-Yamabe soliton on 3-dimensional trans-Sasakian manifolds with λ<0.

  1. A. M. Blaga, A note on warped product almost quasi-Yamabe solitons, Filomat, 33(2019), 2009-2016.
    CrossRef
  2. A. L. Besse, Einstein manifolds, Ergebnisse der Mathematik und ihrer Grenzgebiete (3), vol. 10, Springer-Verlag, Berlin, 1987.
  3. D. E. Blair and J. A. Oubina, Conformal and related changes of metric on the product of two almost contact metric manifolds, Publ. Mat., 34(1990), 199-207.
    CrossRef
  4. B. Y. Chen and S. Desahmukh, Yamabe and quasi-Yamabe soliton on euclidean submanifolds, Mediterranean Journal of Mathematics, August 2018, DOI: 10.1007/s00009-018-1237-2.
    CrossRef
  5. S. Desahmukh and B. Y. Chen, A note on Yamabe solitons, Balk. J. Geom. Appl., 23(1)(2018), 37-43.
  6. D. Chinea and C. Gonzales, A classification of almost contact metric manifolds, Ann. Mat. Pura Appl., 156(1990) 15-30.
    CrossRef
  7. U. C. De, and A. Sarkar, On three-dimensional Trans-Sasakian Manifolds, Extracta Math., 23(2008) 265-277.
  8. C. Dey and U. C. De, A note on quasi-Yambe soliton on contact metric manifolds, J. of Geom., 111(11)(2020).
    CrossRef
  9. A. Gray and L. M. Harvella, The sixteen classes of almost Hermitian manifolds and their linear invariants, Ann. Mat. Pura Appl., 123(1980), 35-58.
    CrossRef
  10. G. Huang and H. Li, On a classification of the quasi Yamabe gradient solitons, Methods Appl. Anal., 21(3)(2014) 379-389.
    CrossRef
  11. C. Aquino, A. Barros and E. jr. Riberio, Some applications of Hodge-de Rham decomposition to Ricci solitons, Results. Math., 60(2011), 235-246.
    CrossRef
  12. R. S. Hamilton, The Ricci flow on surfaces, Mathematics and general relativity, Contemp. Math. Amer. Math. Soc., 71(1988), 237-262.
    CrossRef
  13. B. Leandro and H. Pina, Generalized quasi Yamabe gradient solitons, Differential Geom. Appl., 49(2016), 167-175.
    CrossRef
  14. K. Kenmotsu, A class of almost contact Riemannian manifolds, Tohoku Math. J., 24(1972), 93-103.
    CrossRef
  15. J. C. Marrero, The local structure of Trans-Sasakian manifolds, Annali di Mat. Pura ed Appl., 162(1992), 77-86.
    CrossRef
  16. J. A. Oubina, New classes of almost contact metric structures, Publ. Math. Debrecen, 32(1985), 187-193.
  17. S. Pigola, M. Rigoli, M. Rimoldi and A. Setti, Ricci almost solitons, Ann. Sc. Norm. Super. Pisa Cl. Sci., 10(5)(2011) 757-799.
    CrossRef
  18. A. A. Shaikh, M. H. Shahid and S. K. Hui, On weakly conformally symmetric manifolds. Matematicki Vesnik, 60(2008), 269-284.
  19. M. D. Siddiqi, Generalized Yamabe solitons on Trans-Sasakian manifolds, Matematika Instituti Byulleteni Bulletin of Institute of Mathematics, 3(2020), 77-85.
  20. J. B. Jun and M. D. Siddiqi, Almost Quasi-Yamabe Solitons on Lorentzian concircular structure manifolds-[(LCS)n], Honam Mathematical Journal, 42(3)(2020), 521-536.
  21. M. D. Siddiqi, Generalized Ricci Solitons on Trans-Sasakian Manifolds, Khayyam Journal of Math., 4(2)(2018), 178-186.
  22. M. D. Siddiqi, η-Ricci soliton in 3-diamensional normal almost contact metric manifolds, Bull. Transilvania Univ. Brasov, Series III: Math, Informatics, Physics, 11(60)(2018) 215-234.
  23. M. D. Siddiqi, η -Ricci soliton in (ε, δ)-trans-Sasakian manifolds, Facta. Univ. (Nis), Math. Inform., 34(1)(2019), 4556.
    CrossRef
  24. S. T. Yau, Harmonic functions on complete Riemannian manifolds, Commu. Pure. Appl. Math., 28(1975), 201-228.
    CrossRef
  25. M. El A. Mekki and A. M. Cherif, Generalised Ricci solitons on Sasakian folds, Kyungpook Math. J., 57(2017), 677-682.