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(4): 813-822

Published online December 31, 2021

Copyright © Kyungpook Mathematical Journal.

Ricci-Yamabe Solitons and Gradient Ricci-Yamabe Solitons on Kenmotsu 3-manifolds

Arpan Sardar and Avijit Sarkar*

Department of Mathematics, University of Kalyani, Kalyani 741235, West Bengal, India
e-mail : arpansardar51@gmail.com and avjaj@yahoo.co.in

Received: February 14, 2021; Revised: April 24, 2021; Accepted: June 14, 2021

The aim of this paper is to characterize a Kenmotsu 3-manifold whose metric is either a Ricci-Yamabe soliton or gradient Ricci-Yamabe soliton. Finally, we verify the obtained results by an example.

Keywords: Ricci-Yamabe soliton, Gradient Yamabe soliton, Kenmotsu manifolds, η-paralle Ricci tensor, scalar curvature

Hamilton introduced Yamabe flow in 1982 [8] at the same time he introduced Ricci flow. Ricci solitons and Yamabe solitons are the limit of the solutions of Ricci flow and Yamabe flow respectively. In dimension n = 2 the Yamabe soliton is equivalent to the Ricci soliton. However, in dimension n>2, the Yamabe and Ricci solitons do not agree as the Yamabe soliton preserves the conformal class of the metric but the Ricci soliton does not in general.

Over the past twenty years the theory of geometric flows, such as Ricci flow and Yamabe flow has been the focus of attention of many geometers. Recently, in 2019, Guler and Crasmareanu [7] introduced the study of a new geometric flow which is a scalar combination of Ricci and Yamabe flows under the name Ricci-Yamabe map. This is also called the Ricci-Yamabe flow of type (α,β). The Ricci-Yamabe flow is an evolution equation for the metrics on the Riemannian or semi-Riemannian manifolds defined as [7]

tg(t)=2αRic(t)+βr(t)g(t),g0=g(0).

A solution to the Ricci-Yamabe flow is called Ricci-Yamabe soliton if it depends only on one parameter group of diffeomorphism and scaling. A Ricci-Yamabe soliton on a Riemannian manifold (M,g) consists of data (g,V,λ,α,β) satisfying

£Vg+2αS+(2λβr)g=0,

where £V is the Lie-derivative, S is the Ricci tensor, r is the scalar curvature and λ,α,β. If V is the gradient of a smooth function f on M, then the above soliton is called the gradient Ricci-Yamabe soliton and then equation (1.2) reduces to

2f+αS=(λ12βr)g,

where 2f is the Hessian of f.

The Ricci-Yamabe soliton is said to be expanding, steady, or shrinking according to whether λ is negative, zero, or positive. It is called an almost Ricci-Yamabe soliton if α,β and λ are smooth functions on M. A Ricci-Yamabe soliton is said to be a [5]

  • • Ricci soliton [8] if α =1, β=0.

  • • Yamabe soliton[9] if α =0, β = 1.

  • • Einstein soliton [2] if α =1, β = -1.

  • • ρ-Einstein soliton [3] if α =1, β = -2ρ.

When α0,1, a Ricci-Yamabe soliton is proper. All of these classifications apply to gradient Ricci-Yamabe solitons as well.

The paper is organized as follows. After preliminaries in Section 2, we consider Ricci-Yamabe solitons on Kenmotsu 3-manifolds in Section 3. In Section 4 we study Ricci-Yamabe solitons on Kenmotsu 3-manifolds with η-parallel Ricci tensors. Section 5 is devoted to studying gradient Ricci-Yamabe solitons on Kenmotsu 3-manifolds. Finally, in Section 6 we construct an example of a 3-dimensional Kenmotsu manifold admitting a Ricci-Yamabe soliton.

An almost contact structure [1] on a (2n+1)-dimensional smooth manifold M2n+1 is a triplet (ϕ,ξ,η), where ϕ is a (1,1)-type tensor, ξ a global vector field and η a 1-form, such that

ϕ2=id+ηξ,η(ξ)=1,

where 'id' denotes the identity mapping and relation (2.1) implies that ϕ(ξ)=0, ηϕ=0 and rank(ϕ)=2n. The almost contact structure induces a natural almost complex structure J on the product manifold M× defined by J(U,λd/dt)=(ϕUλξ,η(U)d/dt), where U is tangent to M, t the coordinate of and λ a smooth function on M×. The almost contact structure is said to be normal [12] if the almost complex structure J is integrable or equivalently [ϕ,ϕ]+2dηξ vanishes, where [ϕ,ϕ] is the Nijenhuis torsion of ϕ. Let g be a compatible Riemannian metric with (ϕ,ξ,η), that is, let

g(ϕU,ϕV)=g(U,V)η(U)η(V)

or equivalently, Φ(U,V)=g(U,ϕV) along with g(U,ξ)=η(U) for all U,Vχ(M). With this, M is an almost contact metric manifold equipped with an almost contact metric structure (ϕ,ξ,η,g). An almost contact metric manifold is called a Kenmotsu manifold if it satisfies

(Uϕ)V=g(ϕU,V)ξη(V)ϕU

for all U,Vχ(M), where ∇ is Levi-Civita connection of the Riemannian metric. A Kenmotsu manifold is normal but not Sasakian. Moreover, it is also not compact since from the formula (2.3) we get

Uξ=Uη(U)ξ,

which gives divξ=2n. A conformal change g* of a Riemannian metric g is called a concircular transformation [14] if geodesic circles of g are transformed into geodesic circles of g*. Here a geodesic circle means a curve whose first curvature is constant and whose second curvature is identically zero. A cosymplectic structure is defined to be a normal almost contact metric structure (ϕ,ξ,η,g) with both the fundamental 2-form 𝚽 and the 1-form η is closed. An almost contact metric structure is cosymplectic if and only if ϕ=0. In [11], Kirichenko obtained the class of Kenmotsu manifolds from cosymplectic manifolds by the canonical concircular transformations. A Kenmotsu manifold is of constant curvature -1 if and only if it is canonically concircular to Cn× [11].

For a (2n+1)-dimensional Kenmotsu manifold, the following formulas hold:

R(U,V)ξ=η(U)Vη(V)U,
(Uη)V=g(U,V)η(U)η(V),
S(ξ,ξ)=g(Qξ,ξ)=2n

for any U,Vχ(M), where S is the Ricci tensor and Q is the Ricci operator.

From [4], we know that for a 3-dimensional Kenmotsu manifold

R(U,V)W=r+42[g(V,W)Ug(U,W)V]    r+62[g(V,W)η(U)ξg(U,W)η(V)ξ    +η(V)η(W)Uη(U)η(W)V],
QU=12[(r+2)U(r+6)η(U)ξ],
S(U,V)=12[(r+2)g(U,V)(r+6)η(U)η(V)],

where R is the curvature tensor and r is the scalar curvature of the manifold M.

An almost contact metric manifold is said to be η-Einstein if the Ricci tensor S satisfies

S(V,W)=ag(V,W)+bη(V)η(W)

for any vector field V,W on M and arbitrary functions a, b on M. An η-Einstein manifold with b vanishing and a constant is obviously an Einstein manifold. An η-Einstein manifold is said to be proper η-Einstein if b≠0.

Three-dimensional Kenmotsu manifold have been studied in the papers ([4], [13]).

Lemma 2.1.([13]) On any three-dimensional Kenmotsu manifold (M3,ϕ,ξ,η,g) we have

ξr=2(r+6).

Lemma 2.2.([4]) A 3-dimensional Riemannian manifold is a manifold of constant sectional curvature -1 if and only if the scalar curvature r=-6.

Lemma 2.3.(Proposition 8 of ([10]) Let M be an η-Einstein Kenmotsu manifold S=ag+bηη for scalars a and b. If either a or b is constant then the manifold becomes an Einstein manifold.

Assume that the Kenmotsu 3-manifold admits a proper Ricci-Yamabe soliton (g,ξ,λ,α,β). Then the relation (1.2) yields

(£ξg)(U,V)+2αS(U,V)+(2λβr)g(U,V)=0.

In Kenmotsu 3-manifolds

(£ξg)(U,V)=g(Uξ,V)+g(U,Vξ)    =2[g(U,V)η(U)η(V)].

Using (3.2) in (3.1), we get

S(U,V)=1α[(β2rλ1)g(U,V)+η(U)η(V)],

which implies

QU=1α[(β2rλ1)U+η(U)ξ].

Hence we can state the following theorem using Lemma 2.3 :

Theorem 3.1. A proper Ricci-Yamabe soliton on a 3-dimensional Kenmotsu manifold is an Einstein manifold.

In this section we study Ricci-Yamabe solitons on Kenmotsu 3-manifolds with η-parallel Ricci tensor. A Kenmotsu manifold is said to have η-parallel Ricci tensor if [6]

g((VQ)U,W)=0

for all smooth vector fields U,V,W.

In Kenmotsu 3-manifolds

(VQ)U=VQUQ(VU)    =1α[β2(Vr)U+((Vη)U)ξ+η(U)Vξ]    =1α[β2(Vr)U+g(U,V)ξ+η(U)V2η(U)η(V)ξ].

Using (4.2) in (4.1), we get

1α[β2(Vr)g(U,W)+g(U,V)η(W)+g(V,W)η(U)2η(U)η(V)η(W)]=0.

Putting V=ξ in (4.3) and using Lemma 2.1 we obtain

βα(r+6)g(U,W)=0.

If α=1, then (4.4) implies either β = 0 or, r=-6.

Case I: If β = 0 and α=1, then Ricci-Yamabe soliton becomes Ricci soliton.

Case II: If r=-6, then from Lemma 2.2, the manifold becomes a manifold of constant sectional curvature -1.

Hence we conclude the following:

Theorem 4.1. If a 3-dimensional Kenmotsu manifold admits a Ricci-Yamabe soliton with η-parallel Ricci tensor, then either it is a Ricci soliton or it is a manifold of constant sectional curvature -1, provided α=1.

Suppose a Kenmotsu 3-manifold admits the gradient Ricci-Yamabe soliton. Then from equation (1.3), we get

UDf=(λ12βr)UαQU.

Differentiating (5.1) covariantly along any vector field V, we get

VUDf=(λ12βr)VUβ2(Vr)UαVQU.

Interchanging U and V in the above equation, we infer

UVDf=(λ12βr)UVβ2(Ur)VαUQV.

Hence from the above equations, we get

R(U,V)Df=β2[(Vr)U(Ur)V]α[(UQ)V(VQ)U].

Now, in 3-dimension Kenmotsu manifolds

(UQ)V(VQ)U=12[(Ur)V(Ur)η(V)ξ(Vr)U+(Vr)η(U)ξ]        (r2+3)[η(Y)Xη(X)Y].

Using (5.5) in (5.4), we get

R(U,V)Df=β2[(Vr)U(Ur)V]      α2[(Ur)V(Ur)η(V)ξ(Vr)U+(Vr)η(U)ξ      (r+6){η(V)Uη(U)V}].

Contracting (5.6) and using Lemma 2.1, we get

S(V,Df)=(β+α2)(Vr).

Again, replacing U by Df in (2.8), we get

S(V,Df)=(r2+1)(Vf)(r2+3)η(V)(ξf).

In view of (5.7) and (5.8) we infer

(r2+1)(Vf)(r2+3)η(V)(ξf)=(β+α2)(Vr).

Putting V = ξ in the above equation, we get

ξf=(r+6)(β+α2).

Taking inner product of (5.7) with the vector field ξ, we get

η(V)(Uf)η(U)(Vf)=β2[(Vr)η(U)(Ur)η(V)].

Putting U = ξ in (5.11) and using Lemma 2.1, we get

Vf=β2(Vr)+α2(r+6)η(V).

Using (5.10) and (5.12) in (5.9), we obtain

(β2r+3β+α)[(Vr)+2(r+6)η(V)]=0.

Hence above equation implies either, β2r+3β+α=0 or, Vr+2(r+6)η(V)=0.

Case I: If β2r+3β+α=0, then r is constant.

Case II: If Vr=2(r+6)η(V). Using this in (5.12), we get

Df=(β+α2)(r+6)ξ,

which implies

ξDf=2(β+α2)(r+6)ξ.

Using the above equation in (5.1), we get

(32β+α)r=(12β+8α+λ),

which implies r is constant.

Thus, we state the following:

Theorem 5.1. If the metric of a Kenmotsu 3-manifold M is a gradient Ricci-Yamabe soliton, then the scalar curvature is constant.

If we take β =1 and α=0, then both cases implies r=-6. Then from Lemma 2.2, we say that it is a manifold of constant sectional curvature -1.

We consider the 3-dimensional manifold M={(x,y,z)3,z0}, where (x,y,z) are the standard coordinate of 3. Let {e1,e2,e3} be a linearly independent global frame on M given by

e1=zx,e2=zy,e3=zz.

Let g be the Riemannian metric defined by

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

Let η be the 1-form defined by η(W)=g(W,e3), for all Wχ(M) and ϕ be the (1,1)-tensor defined by

ϕe1=e2,ϕe2=e1,ϕe3=0.

Then using the linearity of ϕ and g, we get

ϕ2W=W+η(W)e3,η(e3)=1,
g(ϕV,ϕW)=g(V,W)η(V)η(W)

for any V,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

[e1,e2]=0,[e2,e3]=e2,[e1,e3]=e1.

The Riemannian connection ∇ of the metric g and using Koszul's formula , we have

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

For the above we see that Wξ=Wη(W)ξ for all Wχ(M). Hence the manifold is a Kenmotsu manifold.

Now, we have

R(U,V)W=UVWVUW[U,V]W.

With the help of the above results and using (6.1), we obtain

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

From the above, we can easily calculate the non-vanishing components of the Ricci tensor S as follows:

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

Hence from the above, we get

r=S(e1,e1)+S(e2,e2)+S(e3,e3)=6,

where r is the scalar curvature.

From (3.3) we obtain

S(e1,e1)=S(e2,e2)=S(e3,e3)=1α(3β+λ).

Therefore λ=2α3β. Hence it is Ricci-Yamabe soliton on Kenmotsu 3-manifolds.

  1. D. E. Blair, Riemannian Geometry of contact and symplectic manifolds, Progress in Mathematics, 203(2010), Birkhäuser, New work.
    CrossRef
  2. G. Catino and L. Mazzieri, Gradient Einstein solitons, Nonlinear Anal., 132(2016), 66-94.
    CrossRef
  3. G. Catino, L. Cremaschi, Z. Djadli, C. Mantegazza and L. Mazzieri, The Ricci-Bourguignon flow, Pacific J. Math., 28(2017), 337-370.
    CrossRef
  4. U. C. De and G. Pathak, On 3-dimensional Kenmotsu manifolds, Indian J. Pure Appl. Math., 35(2)(2004), 159-165.
  5. D. Dey, Almost Kenmotsu metric as Ricci-Yamabe soliton, arXiv: 2005.02322v1[math.DG], (2020 May 5).
    CrossRef
  6. 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
  7. S. Guler and M. Crasmareanu, Ricci-Yamabe maps for Riemannian flow and their volume variation and volume entropy, Turkish J. Math., 43(5)(2019), 2631-2641.
    CrossRef
  8. R. S. Hamilton, The Ricci flow on surfaces, Mathematics and general relativity, Contemp. Math., 71(1998), 237-262.
    CrossRef
  9. R. S. Hamilton, Lectures on geometric flows, 1989 (unpublished).
  10. K. Kenmotsu, A class of almost contact Riemannian manifolds, Tohoku Math. J., 24(1972), 93-103.
    CrossRef
  11. V. F. Kirichenko, On the geometry of Kenmotsu manifolds, Dokl. Akad. Nauk, 380(5)(2001), 585-587.
  12. S. Sasaki and Y. Hatakeyama, On differentiable manifolds with certain structures which are closely related to almost contact structure, II, Tohoku Math. J., 13(1961), 281-294.
    CrossRef
  13. Y. Wang, Yamabe soliton on 3-dimensional Kenmotsu manifolds, Bull. Belg. Math. Soc. Simon Stevin, 23(2016), 345-355.
    CrossRef
  14. K. Yano, Concircular geometry. I. Concircular transformations, Proc. Imp. Acad. Tokyo, 16(1940), 195-200.
    CrossRef