This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

00footnotetext: 𝟐𝟎𝟐𝟎𝐌𝐚𝐭𝐡𝐞𝐦𝐚𝐭𝐢𝐜𝐬𝐒𝐮𝐛𝐣𝐞𝐜𝐭𝐂𝐥𝐚𝐬𝐬𝐢𝐟𝐢𝐜𝐚𝐭𝐢𝐨𝐧.\bf{2020\ Mathematics\ Subject\ Classification\>}. 53D50, 53C25, 53C80.
Key words and phrases: (k,μ)(k,\mu)^{\prime}- almost Kenmotsu manifolds, Kenmotsu manifolds, (m,ρ)(m,\rho)-quasi Einstein solitons.
thanks: Corresponding author

A note on gradient Solitons on two classes of almost Kenmotsu Manifolds

Krishnendu De and Uday Chand De Department of Mathematics, Kabi Sukanta Mahavidyalaya, The University of Burdwan. Bhadreswar, P.O.-Angus, Hooghly, Pin 712221, West Bengal, India. ORCID iD: https://orcid.org/0000-0001-6520-4520 [email protected] Department of Mathematics, University of Calcutta, West Bengal, India. ORCID iD: https://orcid.org/0000-0002-8990-4609 uc-[email protected]
Abstract

The purpose of the article is to characterize gradient (m,ρ)(m,\rho)-quasi Einstein solitons within the framework of two classes of almost Kenmotsu Manifolds. Finally, we consider an example to justify a result of our paper.

1 Introduction

The techniques of contact geometry carried out a significant role in contemporary mathematics and consequently it have become well-known among the eminent researchers. Contact geometry has manifested from the mathematical formalism of classical mechanics. This subject matter has more than one attachments with the alternative regions of differential geometry and outstanding applications in applied fields such as phase space of dynamical systems, mechanics, optic and thermodynamics. In the existing paper, we have a look at the gradient solitons with nullity distribution which play a functional role in coeval mathematics.

In the study of Riemannian manifolds (M,g)(M,g), Gray [7] and Tanno [11] introduced the concept of kk-nullity distribution (k)(k\in\mathbb{R}) and is defined for any pMp\in M and kk\in\mathbb{R} as follows:

Np(k)={WTpM:R(U,V)W=k[g(V,W)Ug(U,W)V]},N_{p}(k)=\{W\in T_{p}M:R(U,V)W=k[g(V,W)U-g(U,W)V]\}, (1.1)

for any U,VTpMU,V\in T_{p}M, where RR indicates the Riemannian curvature tensor of type (1,3)(1,3).

Recently, the generalized idea of the kk-nullity distribution, named as (k,μ)(k,\mu)-nullity distribution on a contact metric manifold (M2n+1,η,ξ,ϕ,g)(M^{2n+1},\eta,\xi,\phi,g) intro by Blair, Koufogiorgos and Papantoniou [1] and defined for any pM2n+1p\in M^{2n+1} and k,μk,\mu\in\mathbb{R} as :

Np(k,μ)={WTpM2n+1:R(U,V)W\displaystyle N_{p}(k,\mu)=\{W\in T_{p}M^{2n+1}:R(U,V)W =\displaystyle= k[g(V,W)Ug(U,W)V]\displaystyle k[g(V,W)U-g(U,W)V] (1.2)
+μ[g(V,W)hUg(U,W)hV]},\displaystyle+\mu[g(V,W)hU-g(U,W)hV]\},

for any U,VTpMU,V\in T_{p}M and h=12£ξϕh=\frac{1}{2}\pounds_{\xi}\phi, where £\pounds denotes the Lie differentiation.

In 20092009, Dileo and Pastore [5] introduced a further generalized concept of the (k,μ)(k,\mu)-nullity distribution, called the (k,μ)(k,\mu)^{\prime}-nullity distribution on an almost Kenmotsu manifold (M2n+1,η,ξ,ϕ,g)(M^{2n+1},\eta,\xi,\phi,g) and is defined for any pM2n+1p\in M^{2n+1} and k,μk,\mu\in\mathbb{R} as :

Np(k,μ)={WTpM2n+1:R(U,V)W\displaystyle N_{p}(k,\mu)^{\prime}=\{W\in T_{p}M^{2n+1}:R(U,V)W =\displaystyle= k[g(V,W)Ug(U,W)V]\displaystyle k[g(V,W)U-g(U,W)V] (1.3)
+μ[g(V,W)hUg(U,W)hV]},\displaystyle+\mu[g(V,W)h^{\prime}U-g(U,W)h^{\prime}V]\},

for any U,VTpMU,V\in T_{p}M and h=hϕh^{\prime}=h\circ\phi.

A Riemannian metric gg on a Riemannian manifold MM is called a gradient (m,ρ)(m,\rho)-quasi Einstein soliton if there exists a smooth function f:Mnf:M^{n}\rightarrow\mathbb{R} and three real constants ρ\rho, λ\lambda and mm (0<m)(0<m\leq\infty) such that

S+2f1mdfdf=βg=(ρr+λ)g,S+\nabla^{2}f-\frac{1}{m}df\otimes df=\beta g=(\rho r+\lambda)g, (1.4)

where 2\nabla^{2} and \otimes indicate the Hessian of gg and tensor product, respectively. The expression S+2f1mdfdfS+\nabla^{2}f-\frac{1}{m}df\otimes df is the mm-Bakry-Emery Ricci tensor, which is proportional to the metric gg and λ=constant\lambda=constant [20]. The soliton becomes trivial if the potential function ff is constant and the triviality condition implies that the manifold is an Einstein manifold. Furthermore, when m=m=\infty, the foregoing equation reduces to gradient ρ\rho-Einstein soliton. This notion was introduced in [2] and recently, Venkatesha et al. studied ρ\rho- Einstein solitons [12] on almost Kenmotsu manifold. In this connection, the properties of (m,ρ)(m,\rho)-quasi Einstein solitons in different geometrical structures have been studied (in details) by ([6], [8]) and others. To know more about almost Kenmotsu manifold, here we may mention the work of Venkatesha et al. and Wang and his collaborator ([13], [15], [17], [19]). Recently, Wang [14] has studied Yamabe solitons and gradient Yamabe solitons within the context of almost Kenmotsu (k,μ)(k,\mu)^{\prime} manifolds.

Motivated from the above studies, we make the contribution to investigate gradient (m,ρ)(m,\rho)-quasi Einstein solitons in almost Kenmotsu Manifolds.

The current paper is constructed as : In section 2, we recall some basic facts and formulas of almost Kenmotsu manifolds which we will need throughout the paper. In sections 3, we characterize gradient (m,ρ)(m,\rho)-quasi Einstein solitons with nullity distribution and obtain some interesting results. In the next section we consider the gradient (m,ρ)(m,\rho)-quasi Einstein soliton in a 33-dimensional Kenmotsu manifold and proved that the manifold is of constant sectional curvature 1-1, provided r(λ+2)r\neq-(\lambda+2). Then we consider an example to verify the result of our paper.

2 Almost Kenmotsu manifolds

In this section we gather the formulas and results of almost Kenmotsu manifolds which will be required on later sections. A differentiable manifold M2n+1M^{2n+1} of dimension (2n+1)(2n+1) is called almost contact metric manifold if it admits a covariant vector field η\eta, a contravariant vector field ξ\xi, a (1,1)(1,1) tensor field ϕ\phi and a Riemannian metric gg such that

ϕ2=I+ηξ,η(ξ)=1,\phi^{2}=-I+\eta\otimes\xi,\;\eta(\xi)=1, (2.1)
g(ϕU,ϕV)=g(U,V)η(U)η(V),g(\phi U,\phi V)=g(U,V)-\eta(U)\eta(V),

where II indicates the identity endomorphism . Then also ϕξ=0\phi\xi=0 and ηϕ=0\eta\circ\phi=0; in a straight forward calculation both can be extracted from (2.1).

On an almost Kenmotsu manifold M2n+1M^{2n+1}, the two symmetric tensor fields h=12£ξϕh=\frac{1}{2}\pounds_{\xi}\phi and l=R(,ξ)ξl=R(\cdot,\xi)\xi, satisfy the following relations [5]

hξ=0,lξ=0,tr(h)=0,tr(h)=0,hϕ+ϕh=0,h\xi=0,\;l\xi=0,\;tr(h)=0,\;tr(h^{\prime})=0,\;h\phi+\phi h=0, (2.2)
Uξ=ϕ2U+hU(ξξ=0),\nabla_{U}\xi=-\phi^{2}U+h^{\prime}U(\Rightarrow\nabla_{\xi}\xi=0), (2.3)
ϕlϕl=2(h2ϕ2),\phi l\phi-l=2(h^{2}-\phi^{2}), (2.4)
R(U,V)ξ=η(U)(VϕhV)η(V)(UϕhU)+(Vϕh)U(Uϕh)V,R(U,V)\xi=\eta(U)(V-\phi hV)-\eta(V)(U-\phi hU)+(\nabla_{V}\phi h)U-(\nabla_{U}\phi h)V, (2.5)

for any vector fields U,VU,V.

Let the Reeb vector field ξ\xi of an almost Kenmotsu manifold belonging to the (k,μ)(k,\mu)^{\prime}-nullity distribution. Then the symmetric tensor field hh^{\prime} of type (1,1)(1,1) satisfies the relations hϕ+ϕh=0h^{\prime}\phi+\phi h^{\prime}=0 and hξ=0h^{\prime}\xi=0. Also, it is understandable that

h2=(k+1)ϕ2(h2=(k+1)ϕ2),h=0h=0.h^{\prime 2}=(k+1)\phi^{2}\,(\Leftrightarrow h^{2}=(k+1)\phi^{2}),\;\;h=0\Leftrightarrow h^{\prime}=0. (2.6)

From the equations (2.1) and (2.6), it follows that k1k\leq-1. Certainly, by (2.6), we conclude that hh^{\prime} vanishes if and only if k=1k=-1. In view of the Proposition 4.1 of [5] on a (k,μ)(k,\mu)^{\prime}-almost Kenmotsu manifold with k<1k<-1 leads to μ=2\mu=-2.

For an almost Kenmotsu manifold, we possess from (2.3)

R(U,V)ξ=k[η(V)Uη(U)V]+μ[η(V)hUη(U)hV],R(U,V)\xi=k[\eta(V)U-\eta(U)V]+\mu[\eta(V)h^{\prime}U-\eta(U)h^{\prime}V], (2.7)
R(ξ,U)V=k[g(U,V)ξη(V)U]+μ[g(hU,V)ξη(V)hU],R(\xi,U)V=k[g(U,V)\xi-\eta(V)U]+\mu[g(h^{\prime}U,V)\xi-\eta(V)h^{\prime}U], (2.8)

where k,μk,\mu\in\mathbb{R}. Contracting VV in (2.8) we have

S(U,ξ)=2kη(U).S(U,\xi)=2k\eta(U). (2.9)

Also, the distribution which is indicated by 𝒟\mathcal{D} (defined by 𝒟=\mathcal{D}=ker η\eta). Presume X𝒟X\in\mathcal{D} be the eigen vector of hh^{\prime} corresponding to the eigen value λ\lambda. Then λ2=(k+1)\lambda^{2}=-(k+1), a constant, which follows from (2.6). Therefore k1k\leq-1 and λ=±k1\lambda=\pm\sqrt{-k-1}. The non-zero eigen value λ\lambda and λ-\lambda are respectively indicated by [λ][\lambda]^{\prime} and [λ][-\lambda]^{\prime}, which are the corresponding eigen spaces associated with hh^{\prime}. Now before introducing the detailed proof of our main theorem, we first write the following lemma:

Lemma 2.1.

(Lemma. 3.2 of [18]) Let (M2n+1,η,ξ,ϕ,g)(M^{2n+1},\eta,\xi,\phi,g) be an almost Kenmotsu manifold such that ξ\xi belongs to the (k,μ)(k,\mu)^{\prime}-nullity distribution and h0h^{\prime}\neq 0. Then the Ricci operator QQ of M2n+1M^{2n+1} is given by

Q=2nid+2n(k+1)ηξ2nh,\displaystyle Q=-2nid+2n(k+1)\eta\otimes\xi-2nh^{\prime}, (2.10)

where k<1k<-1, moreover, the scalar curvature of M2n+1M^{2n+1} is 2n(k2n)2n(k-2n).

3 Gradient (m,ρ)(m,\rho)-quasi Einstein solitons on a (2n+1)(2n+1)-dimensional almost Kenmotsu manifold

Let us assume that the Riemannian metric of a (2n+1)(2n+1)-dimensional almost Kenmotsu manifold such that ξ\xi belongs to the (k,μ)(k,\mu)^{\prime}-nullity distribution and h0h^{\prime}\neq 0, is a (m,ρ)(m,\rho)-quasi Einstein soliton. Then the equation (1.4) may be expressed as

UDf+QU=1mg(U,Df)Df+βU.\nabla_{U}Df+QU=\frac{1}{m}g(U,Df)Df+\beta U. (3.1)

Taking covariant derivative of (3.1) along the vector field VV, we get

VUDf\displaystyle\nabla_{V}\nabla_{U}Df =\displaystyle= VQU+1mVg(U,Df)Df\displaystyle-\nabla_{V}QU+\frac{1}{m}\nabla_{V}g(U,Df)Df (3.2)
+1mg(U,Df)VDf+βVU.\displaystyle+\frac{1}{m}g(U,Df)\nabla_{V}Df+\beta\nabla_{V}U.

Interchanging UU and VV in (3.2), we lead

UVDf\displaystyle\nabla_{U}\nabla_{V}Df =\displaystyle= UQV+1mUg(V,Df)Df\displaystyle-\nabla_{U}QV+\frac{1}{m}\nabla_{U}g(V,Df)Df (3.3)
+1mg(V,Df)UDf+βUV\displaystyle+\frac{1}{m}g(V,Df)\nabla_{U}Df+\beta\nabla_{U}V

and

[U,V]Df=Q[U,V]+1mg([U,V],Df)Df+β[U,V].\displaystyle\nabla_{[U,V]}Df=-Q[U,V]+\frac{1}{m}g([U,V],Df)Df+\beta[U,V]. (3.4)

Equations (3.1)-(3.4) and the symmetric property of Levi-Civita connection together with R(U,V)Df=UVDfVUDf[U,V]DfR(U,V)Df=\nabla_{U}\nabla_{V}Df-\nabla_{V}\nabla_{U}Df-\nabla_{[U,V]}Df we lead

R(U,V)Df\displaystyle R(U,V)Df =\displaystyle= (VQ)U(UQ)V+βm{(Vf)U(Uf)V}\displaystyle(\nabla_{V}Q)U-(\nabla_{U}Q)V+\frac{\beta}{m}\{(Vf)U-(Uf)V\} (3.5)
+1m{(Uf)QV(Vf)QU}.\displaystyle+\frac{1}{m}\{(Uf)QV-(Vf)QU\}.

Taking inner product of (3.5) with ξ\xi, we have

g(R(U,V)Df,ξ)\displaystyle g(R(U,V)Df,\xi) =\displaystyle= βm{(Vf)η(U)(Uf)η(V)}\displaystyle\frac{\beta}{m}\{(Vf)\eta(U)-(Uf)\eta(V)\} (3.6)
+1m{(Uf)η(QV)(Vf)η(QU)}.\displaystyle+\frac{1}{m}\{(Uf)\eta(QV)-(Vf)\eta(QU)\}.

Again (2.7) implies that

g(R(U,V)ξ,Df)=k[η(V)(Uf)η(U)(Vf)]+μ[η(V)(hUf)η(U)(hVf)].g(R(U,V)\xi,Df)=k[\eta(V)(Uf)-\eta(U)(Vf)]+\mu[\eta(V)(h^{\prime}Uf)-\eta(U)(h^{\prime}Vf)]. (3.7)

Combining equation (3.6) and (3.7) reveal that

k[η(U)(Vf)η(V)(Uf)]+μ[η(U)(hVf)η(V)(hUf)]\displaystyle k[\eta(U)(Vf)-\eta(V)(Uf)]+\mu[\eta(U)(h^{\prime}Vf)-\eta(V)(h^{\prime}Uf)]
=βm{(Vf)η(U)(Uf)η(V)}\displaystyle=\frac{\beta}{m}\{(Vf)\eta(U)-(Uf)\eta(V)\}
+1m{(Uf)η(QV)(Vf)η(QU)}.\displaystyle+\frac{1}{m}\{(Uf)\eta(QV)-(Vf)\eta(QU)\}. (3.8)

Setting V=ξV=\xi in the foregoing equation yields

2mhDf=(β2nkmk){Df(ξf)ξ},2mh^{\prime}Df=(\beta-2nk-mk)\{Df-(\xi f)\xi\}, (3.9)

where we have used μ=2\mu=-2.

Letting β=(2nk+mk+2m)\beta=(2nk+mk+2m) and taking into account the equation (2.6) and operating hh^{\prime} on (3.9) produces that

(k+1){Df(ξf)ξ}=hDf,-(k+1)\{Df-(\xi f)\xi\}=h^{\prime}Df, (3.10)

Comparing the antecedent relation with (3.9) gives that

(k+2){Df(ξf)ξ}=0.(k+2)\{Df-(\xi f)\xi\}=0. (3.11)

This shows that either k=2k=-2 or {Df(ξf)ξ}=0\{Df-(\xi f)\xi\}=0.

Case i: If k=2k=-2, then the Proposition 4.1 and Corollary 4.2 of [5] state that M2n+1M^{2n+1} is locally isometric to the Riemannian product n+1(4)×n\mathbb{H}^{n+1}(-4)\times\mathbb{R}^{n}. In fact, from ([9],[10]) we can say that the product n+1(4)×n\mathbb{H}^{n+1}(-4)\times\mathbb{R}^{n} is a rigid gradient Ricci soliton.

Case ii:

Df=(ξf)ξ.Df=(\xi f)\xi. (3.12)

Executing the covariant differentiation of (3.12) along Uχ(M)U\in\chi(M) and utilizing (2.3) we get

UDf=U(ξf)ξ+(ξf)U(ξf)η(U)ξ+(ξf)hU.\nabla_{U}Df=U(\xi f)\xi+(\xi f)U-(\xi f)\eta(U)\xi+(\xi f)h^{\prime}U. (3.13)

Replacing the foregoing equation into (3.1) revels that

QX=βU(ξf)UU(ξf)ξ+(ξf)η(U)ξ(ξf)hU+1mg(U,Df)Df.QX=\beta U-(\xi f)U-U(\xi f)\xi+(\xi f)\eta(U)\xi-(\xi f)h^{\prime}U+\frac{1}{m}g(U,Df)Df. (3.14)

Comparing (2.10) and (3.14) give that

{β+2n(ξf)}U+{2n(ξf)}hU+{(ξf)η(U)\displaystyle\{\beta+2n-(\xi f)\}U+\{2n-(\xi f)\}h^{\prime}U+\{(\xi f)\eta(U)
U(ξf)2n(k+1)η(U)}ξ+1mg(U,Df)Df=0.\displaystyle-U(\xi f)-2n(k+1)\eta(U)\}\xi+\frac{1}{m}g(U,Df)Df=0. (3.15)

Now operating hh^{\prime} we get

{β+2n(ξf)}hU+(k+1){2n(ξf)}ϕ2U\displaystyle\{\beta+2n-(\xi f)\}h^{\prime}U+(k+1)\{2n-(\xi f)\}\phi^{2}U
+1mg(U,Df)hDf=0.\displaystyle+\frac{1}{m}g(U,Df)h^{\prime}Df=0. (3.16)

Contracting UU in the above equation we infer

2n(k+1){2n(ξf)}+1mg(hDf,Df)=0.2n(k+1)\{2n-(\xi f)\}+\frac{1}{m}g(h^{\prime}Df,Df)=0. (3.17)

Putting U=ξU=\xi in (3.16), we have

η(Df)hDf=0.\eta(Df)h^{\prime}Df=0. (3.18)

This shows that either η(Df)=0\eta(Df)=0 or hDf=0.h^{\prime}Df=0.

Case (i): If η(Df)=0\eta(Df)=0, then we get ξf=0\xi f=0 and hence from equation (3.12) it follows that f=constantf=constant. Then we get from (3.1) that the manifold is an Einstein manifold.

Case (ii): If hDf=0h^{\prime}Df=0, then from (3.17) we get (ξf)=2n(\xi f)=2n. Now putting this value in (3.16) we obtain βhU=0\beta h^{\prime}U=0. Hence either β=0\beta=0, or hU=0h^{\prime}U=0. Both contradicts our assumptions β=(2n+m)k+2m\beta=(2n+m)k+2m and k<1k<-1 respectively .

Thus, we can state the following theorem:

Theorem 3.1.

Let the Riemannian metric of a (2n+1)(2n+1)-dimensional (k,μ)(k,\mu)^{\prime} almost Kenmotsu manifold with h0h^{\prime}\neq 0, be the gradient (m,ρ)(m,\rho)-quasi Einstein soliton. Then either M2n+1M^{2n+1} is locally isometric to a rigid gradient Ricci soliton n+1(4)×n\mathbb{H}^{n+1}(-4)\times\mathbb{R}^{n} or M2n+1M^{2n+1} is an Einstein manifold, provided β=(2nk+mk+2m)\beta=(2nk+mk+2m).

We know that when m=m=\infty, the (m,ρ)(m,\rho)-quasi Einstein soliton becomes gradient ρ\rho-Einstein soliton. Putting the value m=m=\infty in (3) and by a straight forward calculation we find that the manifold M2n+1M^{2n+1} is locally isometric to a rigid gradient Ricci soliton n+1(4)×n\mathbb{H}^{n+1}(-4)\times\mathbb{R}^{n}. Thus, we can state:

Corollary 3.2.

Let the Riemannian metric of a (2n+1)(2n+1)-dimensional almost Kenmotsu manifold such that ξ\xi belongs to the (k,μ)(k,\mu)^{\prime}-nullity distribution and h0h^{\prime}\neq 0, be the gradient ρ\rho- Einstein soliton. Then M2n+1M^{2n+1} is locally isometric to a rigid gradient Ricci soliton n+1(4)×n\mathbb{H}^{n+1}(-4)\times\mathbb{R}^{n}.

Remark 3.3.

The above corollary have been proved by Venkatesha and Kumara in their paper [12]. They also prove that the potential vector field is tangential to the Euclidean factor Rn\emph{R}^{n}.

4 Gradient (m,ρ)(m,\rho)-quasi Einstein solitons on a 33-dimensional Kenmotsu manifold

In [5], Dileo and Pastore proved that the following conditions are equivalent:

(1) the (1,1)-type tensor field hh vanishes and foliations of the distribution 𝒟\mathcal{D} are Ka¨\ddot{a}hlerian.

(2) almost contact metric structure of an almost Kenmotsu manifold is normal.

As a outcome, we get instantly that a 3-dimensional almost Kenmotsu manifold reduces to a Kenmotsu manifold if and only if h=0h=0. In this section, we target to investigate the gradient (m,ρ)(m,\rho)-quasi Einstein soliton on a 33-dimensional Kenmotsu manifold. Making use of h=0h=0 in equation (2.3) we get ξ=ϕ2\nabla\xi=-\phi^{2}, and this revels that

R(U,V)ξ=η(V)U+η(U)VR(U,V)\xi=-\eta(V)U+\eta(U)V (4.1)

for any U,VU,V χ(M)\in\chi(M) and therefore by contracting VV in (4.1) we obtain Qξ=2ξQ\xi=-2\xi. From [4] we know that for a 3-dimensional Kenmotsu manifold

R(U,V)W\displaystyle R(U,V)W =\displaystyle= (r+42)[g(V,W)Ug(U,W)V]\displaystyle(\frac{r+4}{2})[g(V,W)U-g(U,W)V]
(r+62)[g(V,W)η(U)ξg(U,W)η(V)ξ\displaystyle-(\frac{r+6}{2})[g(V,W)\eta(U)\xi-g(U,W)\eta(V)\xi
+η(V)η(W)Uη(U)η(W)V],\displaystyle+\eta(V)\eta(W)U-\eta(U)\eta(W)V],
S(U,V)=12[(r+2)g(U,V)(r+6)η(U)η(V)].S(U,V)=\frac{1}{2}[(r+2)g(U,V)-(r+6)\eta(U)\eta(V)]. (4.3)

Now before introducing the detailed proof of our main theorem, we first state the following result:

Lemma 4.1.

(lemma. 4.1 of [16]) For a 3-dimensional Kenmotsu manifold (M3,ϕ,ξ,η,g)(M^{3},\phi,\xi,\eta,g), we have

ξr=2(r+6)\displaystyle\xi r=-2(r+6) (4.4)

where rr denotes the scalar curvature of MM.

Let us assume that the Riemannian metric of a 33-dimensional almost Kenmotsu manifold be a (m,ρ)(m,\rho)-quasi Einstein soliton. Then the equation (1.4) may be expressed as

UDf+QU=1mg(U,Df)Df+βU.\nabla_{U}Df+QU=\frac{1}{m}g(U,Df)Df+\beta U. (4.5)

Taking covariant derivative of (4.5) along the vector field VV, we get

VUDf\displaystyle\nabla_{V}\nabla_{U}Df =\displaystyle= VQU+1mVg(U,Df)Df\displaystyle-\nabla_{V}QU+\frac{1}{m}\nabla_{V}g(U,Df)Df (4.6)
+1mg(U,Df)VDf+βVU+(Vβ)U.\displaystyle+\frac{1}{m}g(U,Df)\nabla_{V}Df+\beta\nabla_{V}U+(V\beta)U.

Interchanging UU and VV in (4.6), we lead

UVDf\displaystyle\nabla_{U}\nabla_{V}Df =\displaystyle= UQV+1mUg(V,Df)Df\displaystyle-\nabla_{U}QV+\frac{1}{m}\nabla_{U}g(V,Df)Df (4.7)
+1mg(V,Df)UDf+βUV+(Uβ)V\displaystyle+\frac{1}{m}g(V,Df)\nabla_{U}Df+\beta\nabla_{U}V+(U\beta)V

and

[U,V]Df=Q[U,V]+1mg([U,V],Df)Df+β[U,V].\displaystyle\nabla_{[U,V]}Df=-Q[U,V]+\frac{1}{m}g([U,V],Df)Df+\beta[U,V]. (4.8)

Equations (4.5)-(4.8) and the symmetric property of Levi-Civita connection together with R(U,V)Df=UVDfVUDf[U,V]DfR(U,V)Df=\nabla_{U}\nabla_{V}Df-\nabla_{V}\nabla_{U}Df-\nabla_{[U,V]}Df we lead

R(U,V)Df\displaystyle R(U,V)Df =\displaystyle= (VQ)U(UQ)V+βm{(Vf)U(Uf)V}\displaystyle(\nabla_{V}Q)U-(\nabla_{U}Q)V+\frac{\beta}{m}\{(Vf)U-(Uf)V\} (4.9)
+1m{(Uf)QV(Vf)QU}+{(Uβ)V(Vβ)U}.\displaystyle+\frac{1}{m}\{(Uf)QV-(Vf)QU\}+\{(U\beta)V-(V\beta)U\}.

Taking inner product of (4.9) with ξ\xi and using (4.3) we have

g(R(U,V)Df,ξ)\displaystyle g(R(U,V)Df,\xi) =\displaystyle= βm{(Vf)η(U)(Uf)η(V)}\displaystyle\frac{\beta}{m}\{(Vf)\eta(U)-(Uf)\eta(V)\} (4.10)
+1m{(Uf)η(QV)(Vf)η(QU)}\displaystyle+\frac{1}{m}\{(Uf)\eta(QV)-(Vf)\eta(QU)\}
+{(Uβ)η(V)(Vβ)η(U)}.\displaystyle+\{(U\beta)\eta(V)-(V\beta)\eta(U)\}.

Again, from (4.1) we infer that

g(R(U,V)Df,ξ)={(Vf)η(U)(Uf)η(V)}.g(R(U,V)Df,\xi)=-\{(Vf)\eta(U)-(Uf)\eta(V)\}. (4.11)

Combining equation (4.10) and (4.11) reveal that

{(Vf)η(U)(Uf)η(V)}\displaystyle-\{(Vf)\eta(U)-(Uf)\eta(V)\} =\displaystyle= βm{(Vf)η(U)(Uf)η(V)}\displaystyle\frac{\beta}{m}\{(Vf)\eta(U)-(Uf)\eta(V)\} (4.12)
+1m{(Uf)η(QV)(Vf)η(QU)}\displaystyle+\frac{1}{m}\{(Uf)\eta(QV)-(Vf)\eta(QU)\}
+{(Uβ)η(V)(Vβ)η(U)}.\displaystyle+\{(U\beta)\eta(V)-(V\beta)\eta(U)\}.

Replacing VV by ξ\xi in the foregoing equation, we get

d(fβ)=ξ(fβ)η,d(f-\beta)=\xi(f-\beta)\eta, (4.13)

where dd stands for the exterior differentiation, provided β=2\beta=-2. In other word, fβf-\beta is invariant along 𝒟\mathcal{D}, i.e., X(fβ)=0X(f-\beta)=0 for any X𝒟X\in\mathcal{D}. Taking into account the above fact and using Lemma 6.1, we infer

(ξf)=(ξβ)=ρ(ξr)=2ρ(r+6).(\xi f)=(\xi\beta)=\rho(\xi r)=-2\rho(r+6). (4.14)

The contraction of the equation (4.9) along UU and applying Lemma 6.1, we get

S(V,Df)\displaystyle S(V,Df) =\displaystyle= 12(Vr)+2βm(Vf)\displaystyle\frac{1}{2}(Vr)+\frac{2\beta}{m}(Vf) (4.15)
1m{r(Vf)g(QV,Df)}2(Vβ).\displaystyle-\frac{1}{m}\{r(Vf)-g(QV,Df)\}-2(V\beta).

Clearly, comparing the above equation with (4.3) yields

(Vr)4βm(Vf)+2m{r(Vf)g(QV,Df)}\displaystyle-(Vr)-\frac{4\beta}{m}(Vf)+\frac{2}{m}\{r(Vf)-g(QV,Df)\}
+4(Vβ)+(r+2)(Vf)(r+6)η(V)ξf=0.\displaystyle+4(V\beta)+(r+2)(Vf)-(r+6)\eta(V)\xi f=0. (4.16)

By a straight forward calculation, replacing VV by ξ\xi in (4) and using (4.14), we can easily obtain

(ξf)=2m(r+6)4β2r4.(\xi f)=\frac{2m(r+6)}{4\beta-2r-4}. (4.17)

Comparing the antecedent equation with (4.14) reveals that

2(r+6){ρ+m4β2r4}=0.2(r+6)\{\rho+\frac{m}{4\beta-2r-4}\}=0. (4.18)

This shows that either r=6r=-6 or {ρ+m4β2r4}=0\{\rho+\frac{m}{4\beta-2r-4}\}=0. Next, we split our investigation as :

Case (i): If r=6r=-6, then from equation (4.3) we conclude that S=2gS=-2g. Therefore, by using equation (4) we sum up that the manifold is of constant sectional curvature 1-1.

Case (ii): If {ρ+m4β2r4}=0\{\rho+\frac{m}{4\beta-2r-4}\}=0, then by a simple calculation we get

r=4λ44ρλ4ρ+m2(2ρ23ρ+1)=constant,r=\frac{4\lambda-4-4\rho\lambda-4\rho+m}{2(2\rho^{2}-3\rho+1)}=constant, (4.19)

provided ρ1\rho\neq 1. Hence by applying Lemma 6.1 we can easily get r=6r=-6. Therefore, from Case (i) we see that the manifold is of constant sectional curvature 1-1. After combining the two conditions namely, β=2\beta=-2 and ρ1\rho\neq 1, we can write r(λ+2)r\neq-(\lambda+2). Thus we have the following theorem:

Theorem 4.2.

Let the Riemannian metric of a 33-dimensional Kenmotsu manifold be the gradient (m,ρ)(m,\rho)-quasi Einstein soliton. Then the manifold is of constant sectional curvature 1-1, provided r(λ+2)r\neq-(\lambda+2).

We know that when m=m=\infty, the (m,ρ)(m,\rho)-quasi Einstein soliton gives the so called gradient ρ\rho-Einstein soliton. Putting the value m=m=\infty in (4.12) and by a straight forward calculation we find that the manifold is of constant sectional curvature 1-1. Thus, we can state:

Corollary 4.3.

Let the Riemannian metric of a 33-dimensional Kenmotsu manifold be the gradient ρ\rho- Einstein soliton. Then the manifold is of constant sectional curvature 1-1.

5 Example

Here we consider an example cited in our paper [3].

We consider the 3-dimensional manifold M={(u,v,w)ε3,w0},M=\{(u,v,w)\varepsilon\mathbb{R}^{3},w\neq 0\}, where (u,v,w)(u,v,w) are standard coordinate of 3.\mathbb{R}^{3}.

The vector fields

E1=wu,E2=wv,E3=wwE_{1}=w\frac{\partial}{\partial u},\hskip 7.0ptE_{2}=w\frac{\partial}{\partial v},\hskip 7.0ptE_{3}=-w\frac{\partial}{\partial w}

are linearly independent at each point of M.M.

Let gg be the Riemannian metric defined by

g(E1,E3)=g(E1,E2)=g(E2,E3)=0,g(E_{1},E_{3})=g(E_{1},E_{2})=g(E_{2},E_{3})=0,
g(E1,E1)=g(E2,E2)=g(E3,E3)=1.g(E_{1},E_{1})=g(E_{2},E_{2})=g(E_{3},E_{3})=1.

Let η\eta be the 1-form defined by η(W)=g(W,E3)\eta(W)=g(W,E_{3}) for any Wεχ(M).W\varepsilon\chi(M).

Let ϕ\phi be the (1,1)(1,1) tensor field defined by

ϕ(E1)=E2,ϕ(E2)=E1,ϕ(E3)=0.\phi(E_{1})=-E_{2},\hskip 7.0pt\phi(E_{2})=E_{1},\hskip 7.0pt\phi(E_{3})=0.

Then for E3=ξE_{3}=\xi , the structure (ϕ,ξ,η,g)(\phi,\xi,\eta,g) defines an almost contact metric structure on MM. Further Koszul’s formula yields

E1E3=E1,E1E2=0,E1E1=E3,\nabla_{E_{1}}E_{3}=E_{1},\hskip 10.0pt\nabla_{E_{1}}E_{2}=0,\hskip 10.0pt\nabla_{E_{1}}E_{1}=-E_{3},
E2E3=E2,E2E2=E3,E2E1=0,\nabla_{E_{2}}E_{3}=E_{2},\hskip 10.0pt\nabla_{E_{2}}E_{2}=E_{3},\hskip 10.0pt\nabla_{E_{2}}E_{1}=0,
E3E3=0,E3E2=0,E3E1=0.\nabla_{E_{3}}E_{3}=0,\hskip 10.0pt\nabla_{E_{3}}E_{2}=0,\hskip 10.0pt\nabla_{E_{3}}E_{1}=0. (5.1)

From the above it follows that the manifold satisfies Uξ=Uη(U)ξ\nabla_{U}\xi=U-\eta(U)\xi, for ξ=E3\xi=E_{3}. Hence the manifold is a Kenmotsu manifold.
We verified that

R(E1,E2)E3=0,R(E2,E3)E3=E2,R(E1,E3)E3=E1,R(E_{1},E_{2})E_{3}=0,\hskip 10.0ptR(E_{2},E_{3})E_{3}=-E_{2},\hskip 10.0ptR(E_{1},E_{3})E_{3}=-E_{1},
R(E1,E2)E2=E1,R(E2,E3)E2=E3,R(E1,E3)E2=0,R(E_{1},E_{2})E_{2}=-E_{1},\hskip 10.0ptR(E_{2},E_{3})E_{2}=E_{3},\hskip 10.0ptR(E_{1},E_{3})E_{2}=0,
R(E1,E2)E1=E2,R(E2,E3)E1=0,R(E1,E3)E1=E3.R(E_{1},E_{2})E_{1}=E_{2},\hskip 10.0ptR(E_{2},E_{3})E_{1}=0,\hskip 10.0ptR(E_{1},E_{3})E_{1}=E_{3}.

From the above expressions of the curvature tensor RR we obtain

S(E1,E1)\displaystyle S(E_{1},E_{1}) =\displaystyle= g(R(E1,E2)E2,E1)+g(R(E1,E3)E3,E1)\displaystyle g(R(E_{1},E_{2})E_{2},E_{1})+g(R(E_{1},E_{3})E_{3},E_{1}) (5.2)
=\displaystyle= 2.\displaystyle-2.

Similarly, we have

S(E2,E2)=S(E3,E3)=2.S(E_{2},E_{2})=S(E_{3},E_{3})=-2.

Therefore,

r=S(E1,E1)+S(E2,E2)+S(E3,E3)=6.r=S(E_{1},E_{1})+S(E_{2},E_{2})+S(E_{3},E_{3})=-6.

Let f:M3f:M^{3}\rightarrow\mathbb{R} be a smooth function defined by f=w2f=-w^{2}. Then gradient of ff with respect to gg is given by

Df=2ww=2E3.Df=-2w\frac{\partial}{\partial w}=2E_{3}.

With the help of (5.1) we can easily get

Hessf(E3,E3)=0.Hessf(E_{3},E_{3})=0.

Thus gradient ρ\rho-Einstein soliton equation shows

Hessf(E3,E3)+S(E3,E3)+2g(E3,E3)=0.Hessf(E_{3},E_{3})+S(E_{3},E_{3})+2g(E_{3},E_{3})=0.

Similarly checking the other components we conclude that that M3M^{3} satisfies

Hessf(U,V)+S(U,V)+2g(U,V)=0.Hessf(U,V)+S(U,V)+2g(U,V)=0.

Thus gg is a gradient ρ\rho-Einstein soliton with f=w2f=-w^{2} and β=2\beta=-2. Hence the Corollary 4.3 is verified.

6 Declarations

6.1 Funding

Not applicable.

6.2 Conflicts of interest/Competing interests

The authors declare that they have no conflict of interest.

6.3 Availability of data and material

Not applicable.

6.4 Code availability

Not applicable.

References

  • [1] Blair, D.E., Koufogiorgos, T. and Papantoniou, B.J., Contact metric manifolds satisfying a nullity condition, Israel. J. Math. 91(1995), 189-214.
  • [2] Catino, G. and Mazzieri, L., Gradient Einstein-solitons, arXiv:1201.6620.
  • [3] De, U. C. and De, K., On ϕ\phi-Concircularly Symmetric Kenmotsu manifolds, Thai Journal of Mathematics 10 (2012), 1-11.
  • [4] De, U.C. and Pathak, G., On 3-dimensional Kenmotsu manifold, Indian J. Pure Appl. Math., 35(2004), 159-165.
  • [5] Dileo, G. and Pastore, A.M., Almost Kenmotsu manifolds and nullity distributions, J. Geom. 93(2009), 46-61.
  • [6] Ghosh, A., (m,ρ)(m,\rho)-quasi-Einstein metric within the framework of K-contact manifolds, J. Math. Phys. Anal. Geom. 17 (2014), 369-376.
  • [7] Gray, A., Spaces of constancy of curvature operators, Proc. Amer. Math. Soc., 17(1966), 897-902.
  • [8] Huang, G. and Wei, Y., The classification of (m,ρ)(m,\rho)-quasi-Einstein manifolds, Ann. Global Anal. Geom., 44 (2013), 269-282.
  • [9] Petersen, P. and Wylie, W., Rigidity of gradient Ricci solitons, Pacific J. Math. 241(2009), 329-345.
  • [10] Petersen, P. and Wylie, W., On gradient Ricci solitons with symmetry, Proc. Amer. Math. Soc. 137(2009), 2085-2092.
  • [11] Tanno, S., Some differential equations on Riemannian manifolds, J. Math. Soc. Japan, 30(1978), 509-531.
  • [12] Venkatesha, V. and Kumara, H. A., gradient ρ\rho-Einstein soliton on almost Kenmotsu manifolds, ANNALI DELL’UNIVERSITA’ DI FERRARA 65(2019), 375-388.
  • [13] Venkatesha, V., Naik, D. M. and Kumara, H. A., \ast-Ricci solitons and gradient almost \ast-Ricci solitons on Kenmotsu manifolds, Math. Slovaca 69 (2019), 1447-1458.
  • [14] Wang, Y., Almost Kenmotsu (k,μ)(k,\mu)^{\prime}-manifolds with Yamabe solitons , RACSAM 115, 14 (2021). https://doi.org/10.1007/s13398-020-00951-y
  • [15] Wang, Y., Conformally flat almost Kenmotsu 3-manifolds, Mediterr. J. Math. 14 (2017), DOI https://doi.org/10.1007/s00009-017-0984-9.
  • [16] Wang, Y., Gradient Ricci almost Solitons on two classes of almost Kenmotsu Manifolds, J. Korean Math. Soc. 23 (2016), 1101-1114.
  • [17] Wang, Y. and Dai, X., Cyclic-parallel Ricci tensors on a class of almost Kenmotsu 3-manifolds, Inter. J. Geom. Meth. Modern Phys. 16 (2019), 1950092.
  • [18] Wang, Y. and Liu, X., On almost Kenmotsu manifolds satisfying some nullity distributions, Proc. Natl. Acad. Sci., India,Sect. A Phys. Sci 86(2016), 347-353.
  • [19] Wang, Y. and Liu, X., A Schur-type theorem for CR-integrable almost Kenmotsu manifolds, Math. Slovaca 66 (2016), 1217-1226.
  • [20] Wei, G. and Wylie, W., Comparison geometry for the Bakry-Emery Ricci tensor, Journal of Differential Geometry, 83(2009), 337-405. doi:10.4310/jdg/1261495336.