δ-Ricci-Yamabe Almost Solitons on Paracontact Metric Manifolds">
검색
Article Search

JMB Journal of Microbiolog and Biotechnology

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

Article

Kyungpook Mathematical Journal 2023; 63(4): 623-638

Published online December 31, 2023

Copyright © Kyungpook Mathematical Journal.

The Geometry of δ-Ricci-Yamabe Almost Solitons on Paracontact Metric Manifolds

Somnath Mondal, Santu Dey, Young Jin Suh*, Arindam Bhattacharyya

Department of Mathematics, Jadavpur University, Kolkata-700032, India
e-mail : somnathmondal.math@gmail.com

Department of Mathematics, Bidhan Chandra College, Asansol-4, West Bengal-713304, India
e-mail : santu.mathju@gmail.com or santu@bccollegeasansol.ac.in

Department of Mathematics, Kyungpook National University, Daegu 41566, Republic of Korea
e-mail : yjsuh@knu.ac.kr

Department of Mathematics, Jadavpur University, Kolkata-700032, India
e-mail : bhattachar1968@yahoo.co.in

Received: February 15, 2023; Revised: March 30, 2023; Accepted: June 27, 2023

In this article we study a δ-Ricci-Yamabe almost soliton within the framework of paracontact metric manifolds. In particular we study δ-Ricci-Yamabe almost soliton and gradient δ-Ricci-Yamabe almost soliton on K-paracontact and para-Sasakian manifolds. We prove that if a K-paracontact metric g represents a δ-Ricci-Yamabe almost soliton with the non-zero potential vector field V parallel to ξ, then g is Einstein with Einstein constant -2n. We also show that there are no para-Sasakian manifolds that admit a gradient δ-Ricci-Yamabe almost soliton. We demonstrate a δ-Ricci-Yamabe almost soliton on a (κ,μ)-paracontact manifold.

Keywords: Para-Sasakian manifold, (κ, μ)-paracontact manifold, Paracontact metric manifolds, δ-Ricci-Yamabe almost soliton, Einstein manifold, Harmonic vector field

Paracontact geometry methods plays an important role in modern mathematics. In the sam way that almost contact manifolds extend almost Hermitian manifolds, the geometry of almost paracontact manifolds is a natural extension of almost paraHermitian geometry. Over the last few years, the study of paracontact geometry has evolved from the mathematical formalism of classical mechanics (see [13, 21]). The concept of Ricci flow, is an evolution equation for metrics defined on connected almost contact metric manifolds whose automorphism groups have maximal dimensions.

Very recently, in [14], Güler and Crasmareanu studied Ricci-Yamabe flow of the type (α, β). A soliton to the Ricci-Yamabe flow is called Ricci-Yamabe soliton (abbreviated to RYS) if it moves only by one parameter group of diffeomorphism and scaling. The metric of the Riemannain manifold (Mn,g), n>2 is said to admit a (α,β)-Ricci-Yamabe soliton or simply a Ricci-Yamabe soliton (g,V,λ,α,β) if it satisfies the equation

LVg+2αRicg+(2λβr)g=0,

where LVg denotes the Lie derivative of the metric g along the vector field V, Ricg is the Ricci tensor, r is the scalar curvature and λ, α, β are real scalars.

In [8], Dey et al. defined a δ-Ricci-Yamabe soliton (in short δ-RYS). A complete Riemannian manifold (Mn,g) is said to be a δ-Ricci-Yamabe almost soliton, denoted by (Mn,g,V,δ,λ), if there exists smooth vector field V on Mn, a soliton function λC(Mn) and a non-zero real valued function δ on Mn such that

δLVg+2αRicg+(2λβr)g=0.

This soliton is called shrinking, steady or expanding according as λ is negative, zero or positive respectively. If the potential vector field V can be written as a gradient of a smooth function u on Mn, then the δ-Ricci-Yamabe almost soliton is called a gradient δ-Ricci-Yamabe almost soliton. In this case, (1.1) can be expressed as

δ2u+αRicg+(λ12βr)g=0,

where 2u be the Hessian of u. We denote this as (Mn,g,Du,λ). Now, the identity (1.2) can be written as

δHessf+αRicg+(λ12βr)g=0.

There are many papers that prove the existence of Ricci solitons and gradient Ricci solitons on paracontact manifolds. In particular, Calvaruso et al. [3] exhibited Ricci solitons on 3-dimensional almost paracontact manifolds. Ricci solitons and their generalizations have been well studied within the framework of contact and paracontact metric manifolds. See [2] for fundamental background, and say, [9] which list many recent related papers in it extensive references. Recently, Erken [11] demonstrated Yamabe solitons on 3-dimensional para-cosymplectic manifold and proved, for example, that the manifold is either η-Einstein or Ricci flat.

In [17, 19], Patra gave answers to the following important questions associated to almost Ricci, and almost Ricci–Bourguignon, solitons: Under which conditions is a (gradient) Ricci almost soliton Einstein? ...trivial? and Under which conditions is a (gradient) Ricci–Bourguignon almost soliton Einstein (trivial) on a paracontact metric manifold?. It is natural to ask the same questions about more general solitons.

Question. Under which conditions is a (gradient) δ-Ricci-Yamabe almost soliton on a paracontact metric manifold Einstein (trivial)?

We find sufficient conditions under which a paracontact metric manifold admitting a δ-Ricci-Yamabe almost soliton or a gradient δ-Ricci-Yamabe almost soliton is Einstein (trivial). We prove the following.

Lemma 1. If a K-paracontact metric g is a δ-Ricci-Yamabe almost soliton, then

(LVη)(ξ)=η(LVξ)=1δ{4nα(2λβr)}.

Patra [17] proved that "if a paracontact metric manifold endows a Ricci soliton with nonzero potential vector field V parallel to the Reeb vector field ξ and the Ricci operator commutes with paracontact structure ϕ, then the manifold is Einstein with Einstein constant -2n". Here, we generalize this result for δ-Ricci-Yamabe almost soliton and, removing the commutativity condition, prove that the potential vector field V being parallel to ξ is a sufficient condition under which a K-paracontact manifold admitting a δ-Ricci-Yamabe almost soliton is Einstein (trivial). So, we have the following.

Theorem 1. If K-paracontact metric g endows a δ-Ricci-Yamabe almost soliton with the non-zero potential vector field V is parallel to ξ, then g is Einstein with Einstein constant -2n. Moreover, V is a constant multiple of ξ.

After Theorem 1, we prove the following result.

Proposition 1. Let M2n+1(ϕ,ξ,η,g) be a para-Sasakian manifold. If the metric g represents a δ-Ricci-Yamabe almost soliton with the potential vector field V, then the following relation holds:

(ξLV)(ξ,ξ)=(βr2λ+4nα)η(ξDδ)+β{η(ξDr)+ξ(ξ(r))ξ      ξ(r)DδξDr}2{η(ξDλ)+ξ(ξ(λ))ξξDλ      ξ(λ)Dδ}+(2λβr4nα)ξDδ.

Next, we get results on K-paracontact manifold and para-Sasakian manifold whose metric endows a gradient δ-Ricci-Yamabe almost soliton. We state this as follows.

Theorem 2. Let M2n+1(ϕ,ξ,η,g) be a K-paracontact manifold. If the metric g represents a gradient δ-Ricci-Yamabe almost soliton, then M2n+1 satisfies either

(ξQ)V1+2ϕQV1+4nϕV1=0

or α=0, that is, it becomes a gradient δ-Yamabe almost soliton, provided β=2.

In [12], Ghosh proved that if a K-contact manifold endows a gradient Ricci almost soliton, then it is of constant scalar curvature. Recently, Patra [18] generalized this result and proved that if a K-contact manifold admits a non-trivial gradient Ricci almost soliton, then the manifold becomes an Einstein metric with constant scalar curvature 2n(2n+1). Here, we prove the nonexistence of a para-Sasakian metric g admitting a gradient Ricci–Yamabe almost soliton with a Ricci operator Q which commutes with a paracontact metric structure ϕ.

Theorem 3. There does not exist a para-Sasakian manifold M2n+1(ϕ,ξ,η,g) with gradient δ-Ricci-Yamabe almost soliton.

As every para-Sasakian manifold is always K-paracontacto, this theorem also holds for K-paracontact manifolds.

Now, we turn our attention to a gradient δ-Ricci-Yamabe almost soliton on a (κ,μ)-paracontact manifold, and state the following results.

Lemma 5. If a (κ,μ)-paracontact manifold (dimension (2n+1)) with κ>1 endows a gradient δ-Ricci-Yamabe almost soliton, then we have

κ(2μ)=μ(n+1).

By virtue of Lemma 5 and Theorem 3, we can assert the following:

Theorem 4. If a (κ,μ)-paracontact manifold (dimension (2n+1)) with κ>1 admits a gradient δ-Ricci-Yamabe almost soliton, then the manifold is locally isometric to the product of a flat (n+1)-dimensional manifold and an n-dimensional manifold of negative constant curvature -4.

The structure of this paper is the following. In Section 2, after a brief introduction, we discuss some preliminaries of paracontact metric manifolds. In Section 3, we examine δ-Ricci-Yamabe almost solitons on K-paracontact and para-Sasakian manifolds. Also, we show that if K-paracontact metric g represents δ-Ricci-Yamabe almost soliton with the non-zero potential vector field V is parallel to ξ, then g is Einstein with Einstein constant -2n. Section 4 deals with a gradient δ-Ricci-Yamabe almost solitons on K-paracontact and para-Sasakian manifolds, and proves that there does not exist such a manifold. In the last section, we study δ-Ricci-Yamabe almost solitons within the framework of (κ,μ)-paracontact manifold. Here, we prove that if a (κ,μ)-paracontact manifold with κ>1 admits a gradient δ-Ricci-Yamabe almost soliton, then the manifold is locally isometric to the product of a flat (n+1)-dimensional manifold and a n-dimensional manifold of negative constant curvature -4.

In this section, we discuss some definitions and identities of paracontact metric manifolds (for more details see [4, 5, 15, 23]). A dimensional smooth manifold M is said to be an almost paracontact structure (ϕ,ξ,η) if it endows a (1,1)-tensor field ϕ, a vector field ξ and a 1-form η such that

ϕ2(V1)=V1η(V1)ξ,η(ξ)=1,ϕξ=0,ηϕ=0

and there is a paracontact distribution D:qMDqTqM:Dq=Ker(η)={xTqM:η(x)=0} generated by η. If an almost paracontact manifold admits a pseudo-Riemannian metric g such that

g(ϕV1,ϕV2)=g(V1,V2)+η(V1)η(V2)

for all V1,V2 on M, then M has an almost paracontact metric structure (ϕ,ξ,η,g) and g is called a compatible metric. Notice that, since Eq. (2.2) holds any compatible metric g has signature (n+1,n). The fundamental 2-form 𝚽 of an almost paracontact metric structure (ϕ,ξ,η,g) is defined by Φ(V1,V2)=g(V1,ϕV2) for all vector fields V1, V2 on M. The manifold M2n+1(ϕ,ξ,η,g) is called paracontact metric manifold, if 𝚽=dη. Here, η is a contact form, i.e., η(dη)n0, ξ is its Reeb vector field and M is a contact manifold (see [5]). We define self-adjoint operators h=12Lξϕ and l=R(.,ξ)ξ, where Lξ is the Lie-derivative along ξ and R is the Riemannian curvature tensor of g on a paracontact metric manifold. The operators h and l satisfy [23]:

Trgh=0,Trg(hϕ)=0,hξ=0,lξ=0,hϕ=ϕh.

The following results hold on a paracontact metric manifold [23]:

V1ξ=ϕV1+ϕhV1,ξξ=0,V1χ(M),
ξh=ϕ+ϕh2ϕl,
Ricg(ξ,ξ)=g(Qξ,ξ)=Trl=Tr(h2)2n
(ϕV1ϕ)ϕV2(V1ϕ)V2=2g(V1,V2)η(V2)(V1hV1+η(V1)ξ)

for all V1, V2 on M2n+1, where ∇ is the operator of covariant differentiation of g and Q denotes the Ricci operator given by Ricg(V1,V2)=g(ϕV1,V2)V1,V2 on M2n+1. M is said to be a K-paracontact manifold if the vector field ξ is a killing (equivalently h=0). On a K-paracontact manifold the following formula holds [23]:

V1ξ=ϕV1,(V1η)V2=g(V1,V2)η(V1)η(V2),
R(V1,ξ)ξ=V1+η(V1)ξ,
Qξ=2nξ

for any vector fields V1,V2 on M2n+1. Moreover, from [23] we have (Lξg)(V1,V2)=2g(V1,ϕhV2) and therefore, M is K-paracontact if and only if ϕ h=0.

A paracontact metric structure on M is said to be normal if the almost paracomplex structure on M×R defined by

J(V1,fd/dt)=(ϕV1+fξ,η(V1)d/dt),

where f is a real function on M×R, is integrable. A normal paracontact metric manifold is said to be para-Sasakian. A para-Sasakian manifold is always a K-paracontact manifold. A 3-dimensional K-paracontact manifold is a para-Sasakian manifold [3], which may not be true in higher dimensions [16]. Equivalently, a paracontact metric manifold is said to para-Sasakian if [23]:

(V1ϕ)V2=g(V1,V2)ξ+η(V2)V1

for any vector fields V1,V2 on M2n+1. Further, on any para-Sasakian manifold [23]:

R(V1,V2)ξ=η(V1)V2η(V2)V1,
R(V1,ξ)ξ=V1+η(V1)ξ

for any vector fields V1,V2 on M2n+1.

We recall the following commutation formula from [22]

(LVV3gV3LVg[V,V3]g)(V1,V2)=g((LV)(V3,V1),V2)              g((LV)(V3,V2),V1)

for all vector fields V1,V2 on M2n+1. By virtue of parallelism of the pseudo-Riemannian metric g, this formula yields

(V3LVg)(V1,V2)=g((LV)(V3,V1),V2)+g((LV)(V3,V2),V1)

for all vector fields V1,V2 on M2n+1. We also recall the following from [10, p. 39]

(LV)(V1,V2)=V1V2VV1 V2V+R(V,V1)V2

for any vector fields V1, V2, V on M2n+1.

Let R be the Riemannian curvature tensor of the Levi-Civita connection ∇ of g, given by

R(V1,V2)=V1V2V2V1[V1,V2],V1,V2χ(M).

where χ(M) is the set of all vectors fields on M. On a paracontact metric manifold the following formula holds

V1ξ=V1η(V1)ξϕhV1(ξξ=0)

for any V1, V2χ(M),

R(V1,V2)ξ=η(V1)(V2ϕhV2)η(V2)(V1ϕhV1)+(V2ϕh)V1(V1ϕh)V2

for any vector fields V1, V2χ(M).

The reading of nullity conditions on paracontact geometry is an attractive topic in paracontact geometry. In [6], Cappelletti-Montano et al. initiated the notion of (κ,μ)-paracontact structure. They defined a (κ,μ)-paracontact manifold as a paracontact metric manifold M2n+1(ϕ,ξ,η,g) whose curvature tensor satisfies

R(V1,V2)ξ=κ{η(V2)V1η(V1)V2}+μ{η(V2)hV1η(V1)hV2}

for some real numbers (κ,μ). Many geometers have studied (κ,μ)-paracontact manifolds and attained several significant properties of these manifold (see [7, 20]). On a (κ,μ)-paracontact manifold one has [6]

h=0h=0, h 2V1=(k+1)ϕ2V1,
h2(V1)=(κ+1)[V1η(V1)ξ]

for V1χ(M). And also we have the following

R(ξ,V1)V2=κ{g(V1,V2)ξη(V2)V1}2{g(hV1,V2)ξη(V2)hV1},
QV1=2nV1+2n(κ+1)η(V1)ξ2nh(V1),
r=2n(κ2n),
(V1η)V2=g(V1,V2)η(V1)η(V2)+g(hV1,V2),

where V1 and V2 are any vector fields on M.

In this section, we prove the results we stated about δ-Ricci-Yamabe almost solitons on K-paracontact and para-Sasakian manifolds. We begin with the following.

Proof of the Lemma 1. In light of identity (2.10), the soliton equation (1.1) gives

(LVg)(V1,ξ)=1δ{4nα(2λβr)}η(V1).

Taking the Lie differentiation of η(V1)=g(V1,ξ) by the vector field V, we achieve (LVη)(V1)g(LVξ,V1)=(LVg)(V1,ξ). By using (3.1), we acquire

(LVη)(V1)g(LVξ,V1)=1δ{4nα(2λβr)}η(V1).

The result then follows using (3.2) with g(ξ,ξ)=1.

Lemma 2. [17] On a K-paracontact manifold M2n+1(ϕ,ξ,η,g), we have

(i)(V1Q)ξ=QϕV1+2nϕV1,
(ii)(ξQ)V1=QϕV1ϕQV1

for all vector fields V1 on M2n+1(ϕ,ξ,η,g).

Proof of the Theorem 1. Since the potential vector field V is parallel to ξ, i.e., V=σξ for a non-zero smooth function σ on M, we acquire V1V=V1(σ)ξσ(ϕV1) by the derivative of V=σξ covariantly by V1χ(M) and using the identity (2.8). Thus, the equation (1.1) reduces to

δ{V1(σ)η(V2)+V2(σ)η(V1)}+2αRicg(V1,V2)+(2λβr)g(V1,V2)=0

for all V1,V2χ(M). Now, we insert V1=V2=ξ into (3.3) and use fact (2.10) to infer ξ(σ)=12δ{4nα(2λβr)}. Setting V2 in (3.3) and recalling (2.10), we get

V1(σ)=ξ(σ)η(V1),  V1χ(M)

and therefore, by (2.8), get

Hessσ(V1,V2)=V1(ξ(σ))η(V2)ξ(σ)g(ϕV1,V2),V1,V2χ(M).

Since Hessσ is symmetric and ϕ is skew-symmetric, by (2.1) and (3.4), we get

ξ(σ)dη(V1,V2)=0  V1,V2ξ,

as dη(V1,V2)=g(V1,ϕV2). This exposes that ξ(σ)=0, as dη is a non-zero

on M, hence, σ=0. Hence, σ is constant on M. This simplifies the equation (3.4) to

2αRicg(V1,V2)=(2λβr)g(V1,V2)=4nαg(V1,V2),V1,V2χ(M),

using Qξ=2nξ and hence (M,g) is an Einstein with Einstein constant -2n. This finishes the proof.

Proof of the Proposition 1.

Now, we use identities (1.1) and (2.14) to acquire

g((LV)(V3,V1),V2)+g((LV)(V3,V2),V1)=1δ[V3(δ)(LVg)(V1,V2)+2α(V3Ricg)(V1,V2){2V3(λ)βV3(r)}g(V1,V2)]

for all vector fields V1,V2,V3 on M2n+1. Interchanging cyclicly the roles of V1, V2 and V3 in the upstairs equalization and with the straight enumeration we gain

g((LV)(V1,V2),V3)=1δ[2α{(V1Ricg)(V2,V3)+(V2Ricg)(V1,V3)        (V3Ricg)(V1,V2)}+V1(δ)(LVg)(V2,V3)        +V2(δ)(LVg)(V1,V3)V3(δ)(LVg)(V1,V2)        +{2V3(λ)βV3(r)}g(V1,V2){2V1(λ)        βV1(r)}g(V2,V3){2V2(λ)βV2(r)}g(V1,V3)]

V1,V2,V3 on M2n+1. Recall the following from [23, Lemma 3.15]:

(V3Ricg)(V1,V2)=(V1Ricg)(V2,V3)(ϕV2Ricg)(ϕV1,V3)      η(V1)Ricg(V2,V3)2η(V2)Ricg(ϕV1,V3)      2nη(V1)g(ϕV2,V3)4nη(V2)g(ϕV1,V3).

Using (V3Ricg)(V1,V2)=g((V3Q)V1,V2) and the identity (2.1) of Lemma 2, we find ξQ=QϕϕQ=2nηξ after putting V3 into (3.5). With this, Lemma 2 and substituting V2 by ξ in (3.5) we can get

(LV)(V1,ξ)=2αδ(2nη(V1)+4nϕV1)+{βV1(r)2V1(λ)}ξ    {(2λβr)ξ4nα}V1(δ)+{βξ(r)2ξ(λ)}V1    {2αQV1+(2λβr)V1}ξ(δ)+{(2λβr4nα)Dδ    +2DλβDr}η(V1)

for all V1 on M. Now, we taking the covariant differentiation of (3.6) by a vector field V2 on M2n+1 and applying (2.8), (2.10) and (2.11) we obtain

(V2LV)(V1,ξ)+(LV)(V1,V2)η(V2)(LV)(V1,ξ)      =2αδ{2n(V2η)V1+4n(V2ϕ)V1}      +2αV2(δ)(2QϕV1+4nϕV1)+{βg(V1,V2Dr)      2g(V1,V2Dλ)}{βV1(r)2V1(λ)}ϕ2V2      +V1(δ){2λβr4nα}ϕ2V2(2λβr4nα)      g(V1,V2Dδ)+{βV2(ξ(r))2V2(ξ(λ))}V1      +{(2V2(λ)βV2(r))Dδ+(2λβr4nα)V2Dδ      +2V2DλβV2Dr}η(V1)+{(2λβr4nα)Dδ      +2DλβDr}(g(V1,V2)η(V1)η(V2)).

Now, we plug V1=ξ, V2=ξ in the equation (3.7) and using (2.1), (2.8) and (2.11) to achieve

(ξLV)(ξ,ξ)=(βr2λ+4nα)η(ξDδ)+β{η(ξDr)+ξ(ξ(r))ξ      ξ(r)DδξDr}2{η(ξDλ)+ξ(ξ(λ))ξξDλ      ξ(λ)Dδ}+(2λβr4nα)ξDδ.

This completes the proof.

In this section, we take into account the gradient almost δ-Ricci-Yamabe solitons on paracontact metric manifolds.

Let a paracontact metric manifold M2n+1(ϕ,ξ,η,g) admits a gradient almost δ-Ricci-Yamabe soliton. Then the soliton equation (1.2) can be demonstrated as

δV1u+αQV1+(λβr2)V1=0

for all V1χ(M) and hence the curvature tensor gained from (4.1) and (2.16) satisfies

δR(V1,V2)u=α{(V2Q)V1(V1Q)V2}+V1(λ)V2    V2(λ)V1β2{V1(r)V2V2(r)V1}.

Proof of the Theorem 2. First, we take the covariant differentiation of 2.10 through the vector field V1χ(M), and apply (2.8) to yield

(V1Q)ξ=QϕV1+2nϕV1.

Since ξ is killing, we have

0=(LξQ)V1=Lξ(QV1)Q(LξV1)=[ξ,QV1]Q([ξ,V1])=ξ(QV1)QV1ξQ(ξV1V1ξ)=(ξQ)V1QV1ξ+Q(V1ξ)

It follows from (2.8) that ξQ=QϕϕQ. Now, we replace V1 by ξ into identity (4.2) and then replace the scalar product with V1χ(M) to yield

δg(R(ξ,V2)u,V1)=α{g(ϕQV2,V1)+2ng(ϕV2,V1)+2nη(V1)η(V2)}        +{ξ(λ)β2ξ(r)}g(V1,V2)        {V2(λ)β2V2(r)}η(V1).

Now, by identity (4.3) and equation (2.8) we get

g((V1ϕ)V2,V3)g((V2ϕ)V1,V3)=g(R(V1,V2)V3,ξ).

Using the Bianchis's first identity, we achieve

g(R(ξ,V3)V2,V1)=g((V3ϕ)V2,V1),V1,V2,V3χ(M).

We insert the above identity into (4.4) and ξ(r)=0 (which holds since ξQ=QϕϕQ) to get

δg((V2ϕ)V1,u)+α{g(ϕQV2,V1)+2ng(ϕV2,V1)+2nη(V1)η(V2)}+ξ(λ)g(V1,V2){V2(λ)β2V2(r)}η(V1)=0.

Now setting V1=ϕV1, and V2=ϕV2 in (4.5) and eliminating (4.5) from the results expression, we get

δ{g((ϕV2ϕ)ϕV1,u)g((V2ϕ)V1,u)}αg(QϕV2+ϕQV2,V1)2ξ(λ)g(V1,V2)+V2(λβ2r)η(V1)+ξ(λ)η(V1)η(V2)4nαg(ϕV2,V1)=0.

Herre we also used (2.10). The following formula for paracontact metric manifolds is from [23, Lemma 2.7]:

(ϕV2ϕ)ϕV1(V2ϕ)V1=2g(V1,V2)ξη(V1){V2hV2+η(V2)ξ}.

Using (4.6) and (4.7), we can infer that

2ξ(δuλ)g(V1,V2)+V2(λβr2δu)η(V1)ξ(δuλ)η(V1)η(V2)=αg(QϕV2+ϕQV2,V1)+4nαg(ϕV2,V1),

since h=0 for K-paracontact manifold. At this point, placing V2 by ξ in (4.2), we get

δR(V1,ξ)u=α{(ξQ)V1(V1Q)ξ}+V1(λ)ξ      ξ(λ)V1β2{V1(r)ξξ(r)V1},

replacing the scalar product in the above result with ξ and using (2.9) and (4.3) we get

V1(λuδβr2)=ξ(λuδ)η(V1),

by ξQ=QϕϕQ. Let σ=λuδβr2. Equation (4.9) becomes V1(σ)=ξ(σ)η(V1), for V1χ(M) as ξ(r)=0. In this manner, by the argument in Section 3, we get that σ=λuδβr2 is constant on M. Using ξQ=QϕϕQ which follows from (4.8), we get

α{g((ξQ)V2,V1)+2g(ϕQV2,V1)+4ng(ϕV2,V1)}=0.

This implies that α{g((ξQ)V2+2ϕQV2+4nϕV2,V1)}=0. So either α=0 or g((ξQ)V2+2ϕQV2+4nϕV2,V1)=0, which completes the proof.

Proof of the Theorem 3. On a para-Saskian manifold, a Ricci operator satisfies the following (see [23, Lemma 3.15])

QV1=ϕQϕV12nη(V1)ξ.V1χ(M)

With this and (2.1) we see that the Ricci operator Q deflects the paracontact structure ϕ.

On the other hand, para-Sasakian manifolds are K-paracontact. So one has ξQ=QϕϕQ=2nηξ, which it implies the contradiction ηξ=0. This finishes the proof.

By virtue of this formula ξQ=QϕϕQ=2nηξ, Theorem 3 gives a non-existence theorem.

In this last section, we discuss the nullity agreements on paracontact geometry. In [5], Cappelletti-Montano et al. introduced the notion of (κ,μ)-paracontact structures. According to them a (κ,μ)-paracontact manifold is a paracontact metric manifold M2n+1(ϕ,ξ,η,g) whose curvature tensor satisfies

R(V1,V2)ξ=κ{η(V2)V1η(V1)V2}+μ{η(V2)hV1η(V1)hV2}

for all V1,V2χ(M) and for some real numbers (κ,μ). Equivalently, this equation can be written as

R(V1,ξ)V2=κ{η(V2)V1g(V1,V2)ξ}+μ{η(V2)hV1g(hV1,V2)ξ}

for all V1,V2χ(M). On a (κ,μ)-paracontact manifold one has [5]:

h2=(κ+1)ϕ2
Qξ=2nκξ

Lemma 3. (see [5]). In any (κ,μ)-paracontact manifold M2n+1(ϕ,ξ,η,g), the Ricci operator Q of M can be written as

QV1=[2(1n)+nμ]V1+[2(n1)+μ]hV1  +[2(n1)+n(2κμ)]η(V1)ξ,forκ>1

for any vector field V1 on M2n+1. Moreover, the scalar curvature of M is 2n(2(1n)+κ+nμ).

Lemma 4. (see [5]). On a (κ,μ)-paracontact manifold M2n+1(ϕ,ξ,η,g), we have

(ξh)V1=μϕhV1

for any vector field V1 in M2n+1.

Proof of the Lemma 5. First, taking the covariant derivative of (5.4) through an arbitrary vector field V1 on M and applying (2.4) we get

(V1Q)ξ=Q(ϕϕh)V12nκ(ϕϕh)V1.

With (1.2) this can be written as

δV2Du+αQV2+(λβr2)g=0

for all vector fields V2 on M2n+1. Using the foregoing equation in the famous manifestation of the curvature tensor R(V1,V2)=[V1,V2][V1,V2], we can easily derive

δR(V1,V2)Du=α{(V2Q)V1(V2Q)V1}

for all V1 and V2 on M2n+1. In this manner, taking the scalar product of (5.9) along ξ and making use of (5.3) and (5.7) leaves that

δg(R(V1,V2)Du,ξ)=α{g((Qϕ+ϕQ)V2,V1)      g((Qϕh+hϕQ)V2,V1)4nκg(ϕV2,V1)}.

Replacing V1 by ϕ V1 and V2 by ϕ V2 in (5.10) and noting that R(ϕV1,ϕV2)ξ=0 (from (5.1)) and (2.1), we get

QϕV1+ϕQV1+ϕQhV1+hQϕV14nκϕV1=0.

Now, we put V1=ϕV1 into (5.5) and use ϕξ=0 to obtain

QϕV1=[2(1n)+nμ]ϕV1+[2(n1)+μ]hϕV1.

Dy acting h on the last equation and making use of (2.1), (5.3) and hξ=0 leaves

hQϕV1=[2(1n)+nμ]hϕV1+(κ+1)[2(n1)+μ]ϕV1.

In addition, operating ϕ on (5.5) and using ϕξ=0, we get

ϕQV1=[2(1n)+nμ]ϕV1+[2(n1)+μ]ϕhV1.

Now, we replace V1 by hV1 in the foregoing equation and use (5.3) to yield

ϕQhV1=[2(1n)+nμ]ϕhV1+(κ+1)[2(n1)+μ]ϕV1.

Applying the last four equations in (5.9) and also using ϕh=hϕ we obtain 5.7. This completes the proof.

Proof of the Theorem 4. First, we substitute ξ for V1 in (5.10) and use the identity (5.4) and hξ=ϕξ=0 to acquire δg(R(ξ,V2)ξ,Du)=0. With this and (5.3) we get

κ{Du(ξu)ξ}+μhDu=0,

where we have used g(V1,Du)=V1u. Now, we take a covariant differentiation of the equation (5.12) by ξ and use the relation (5.6), ξξ=0 to achieve

κ{ξDuξ(ξu)ξ}+μ{μhϕDu+h(ξDu)}=0.

By equations (5.5) and (5.8), we have

δξD(λβr2)=(2nκα+λβr2)ξ

and also

δξ(ξu)=2nκα+(λβr2).

Making use of (5.14) and (5.15) in (5.13) and using ϕξ=0, we have μ2hϕDu=0. Applying this to ϕ and using (2.1) we get μ2hDu=0. By the operation of h and the use of (2.1) and (5.3) we get

μ2(κ+1)(Du(ξu)ξ)=0.

For κ>1, either (i)μ=0 or (ii)μ0.

Case (i). In this case, as κ>1 it follows from (1.6) that κ=0. Hence R(V1,V2)ξ=0 for any vector fields V1,V1χ(M), and therefore M is the product of a flat (n+1)-dimensional manifold of negative constant curvature -4 (see [24, Theorem 3.3]).

Case (ii). This case yields Du=(ξu)ξ. We differentiate this along with an arbitary vector field V1 together with (2.1) to acquire

V1Du=V1(ξu)ξ(ξu)(ϕV1ϕhV1).

As g(V1D(λβr2),V2)=g(V2D(λβr2),V1), the last equation gives

V1(ξu)η(V2)V2(ξu)η(V1)+(ξu)dη(V1,V2)=0.

Replacing V1 by ϕ V1 and V2 by ϕ V2 and using ϕ ξ=0 we find ξ u=0. We apply dη0 on M2n+1. Then, Du=0, i.e., u is constant and consequently (5.8), (5.14) and (5.15) yield Ricg=(λβr2)g=2nκα, i.e., M2n+1 is an Einstein. This gives r=2nκα(2n+1). In addition Lemma 3 yields r=2n{2(1n)+κα+nμ}. Combining both we have

nμ=2(nκα+n1).

Now, using (5.16) and Ricg=2nκαg in (5.5) we get 2(n1)+μ=0. Thus (1.6) yields κ=1n2n, a contradiciton. This finishes the proof.

  1. D. E. Blair. Riemannian geometry of contact and sympletic manifolds. Progress in Mathematics. MA: Birkhäuser Bostom; 2022. pp. xii+260, DOI:10.1007/978-1-4757-3604-5.
    CrossRef
  2. G. Calvaruso, Homogeneous paracontact metric three-manifolds, Illinois J. Math., 55(2011), 697-718.
    CrossRef
  3. G. Calvaruso and A. Perrone, Ricci solitons in three-dimensional paracontact geome-try, J. Geom. Phys., 98(2015), 1-12.
    CrossRef
  4. B. Cappelletti-Montano, A. Carriazo, V. Martin-Molina and Sasaki-Einstein, Sasaki-Einstein and paraSasaki-Einstein metrics from (κ, µ)-structures, J. Geom. Phys., 73(2013), 20-36.
    CrossRef
  5. B. Cappelletti-Montano, I. K. Erken and C. Murathan, Nullity conditions in para-contact geometry, Differential Geom. Appl., 30(2012), 665-693.
    CrossRef
  6. J. T. Cho and R. Sharma, Contact geometry and Ricci solitons, Int. J. Geom. Methods Mod. phys., 7(6)(2010), 951-960.
    CrossRef
  7. U. C. De, K. Mandal and Ricci almost solitons, Ricci almost solitons and gradient Ricci almost solitons in (κ, µ) paracontact geometry, Bol. Soc. Parana. Mat., 37(3)(2019), 119-130.
    CrossRef
  8. S. Dey, P. Laurian-Ioan and S. Roy, Geometry of ∗-κ-Ricci-Yamabe soliton and gradient ∗-κ-Ricci-Yamabe soliton on Kenmotsu manifolds, Hacet. J. Math. Stat., 52(4)(2023), 907-922.
    CrossRef
  9. S. Dey and S. Roy, Characterization of general relativistic spacetime equipped with η-Ricci-Bourguignon soliton, J. Geom. Phys., 178(2022).
    CrossRef
  10. K. L. Duggal and R. Sharma. Symmetries of spacetimes and Riemannian manifolds. Math. Appl. 487. Dordrecht: Kluwer Academic Publishers; 2020. x+214 pp.
    CrossRef
  11. I. K. Erken, Yamabe solitons on three-dimensional normal almost paracontact metric manifolds, Period. Math. Hungar., 80(2)(2020), 172-184.
    CrossRef
  12. A. Ghosh, Certain contact metrics as Ricci almost solitons, Results Math., 65(2014), 81-94.
    CrossRef
  13. A. Ghosh, R. Sharma and J. T. Cho, Contact metric manifolds with η-parallel torsion tensor, Ann. Global Anal. Geom., 34(3)(2008), 287-299.
    CrossRef
  14. S. G¨uler and M. Crasmareanu, Ricci-Yamabe maps for Riemannian flows and their volume variation and volume entropy, Turkish J. Math., 43(5)(2019), 2631-2641.
    CrossRef
  15. X. Liu and Q. Pan, Second order parallel tensors on some paracontact metric mani-folds, Quaest. Math., 40(7)(2017), 849-860.
    CrossRef
  16. V. Martín-Mólina, Paracontact metric manifolds without a contact metric counter-part, Taiwanese J. Math., 19(1)(2015), 175-191.
    CrossRef
  17. D. S. Patra, Ricci solitons and paracontact Geometry, Mediterr. J. Math., 16(6)(2019).
    CrossRef
  18. D. S. Patra, K-Contact metrics as Ricci almost solitons, Beitr. Algebra Geom., 62(3)(2021), 737-744.
    CrossRef
  19. D. S. Patra, A. Ali and F. Mofarreh, Characterizations of Ricci-Bourguignon almost solitons on Pseudo-Riemannian man-ifolds, Mediterr J. Math., 19(4)(2022).
    CrossRef
  20. D. G. Prakasha, L. M. Fernández and K. Mirji, The M-projective curvature tensor field on generalized (κ, µ)-paracontact metric manifolds, Georgian Math. J., 27(1)(2020), 141-147.
  21. R. Sharma and A. Ghosh, Sasakian 3-metric as a Ricci soliton represents the Heisen-berg group, Int. J. Geom. Methods Mod. phys., 8(1)(2011), 149-154.
    CrossRef
  22. K. Yano. Integral Formulas in Riemannian geometry. Pure Appl. Math. New York: Marcel Dekker, Inc.; 1970.
  23. S. Zamkovoy, Canonical connections on paracontact manifolds, Ann. Global Anal. Geom., 36(1)(2009), 37-60.
    CrossRef
  24. S. Zomkovoy and V. Tzanov, Non-existence of flat paracontact metric structures in dimensional greater than or equal to five, Annuaire Univ. Sofia Fac. Math. Inform., 100(2011), 27-34.