Home Yamabe Solitons and τ-Quasi Yamabe Gradient Solitons on Riemannian Manifolds Admitting Concurrent-Recurrent Vector Fields
Article Open Access

Yamabe Solitons and τ-Quasi Yamabe Gradient Solitons on Riemannian Manifolds Admitting Concurrent-Recurrent Vector Fields

  • Devaraja Mallesha Naik , Ghodratallah Fasihi-Ramandi EMAIL logo , H. Aruna Kumara and Venkatesha Venkatesha
Published/Copyright: March 31, 2023
Become an author with De Gruyter Brill

Abstract

We consider a Riemannian manifold (M,g) admitting a concurrent-recurrent vector field for which the metric g is a Yamabe soliton or a τ-quasi Yamabe gradient soliton. We show that if the metric of a Riemannian three-manifold (M, g) admitting a concurrent-recurrent vector field is a Yamabe soliton, then M is of constant negative curvature –α2. In this case, we see that the potential vector field is Killing. Next, we show that if the metric of a Riemannian manifold M admitting concurrent-recurrent vector field is a non-trivial r-quasi Yamabe gradient soliton with potential function f, then M has constant scalar curvature and is equal to –n(n – 1)α2. Finally, an illustrative example is presented.

2020 Mathematics Subject Classification: Primary 06B05; 06D35; Secondary 35R30; 39A10

1 Introduction

Let (M, g) be a compact Riemannian manifold with s as its scalar curvature. Then, the Yamabe problem concerns the existence of a Riemannian metric g′ conformal to g, for which the scalar curvature s′ of the metric g′ is constant. As an effort to solve the Yamabe problem, Hamilton in [4] came up with the concept of Yamabe flow. Given a conformal class of Riemannian metrics, Yamabe flow can be used for constructing metrics whose scalar curvature s is constant. The Yamabe flow is an evolving metric family (g(t)) satisfying

1.1 tg(t)=s(t)g(t)

with the initial data g(0) = g. Geometric flows like Yamabe flows, Ricci flows, mean and the inverse mean curvature flows, and Kaehler-Ricci flows have been applied to a variety of topological, geometric, and physical problems. It is interesting to note that in dimension two the Yamabe flow and Ricci flow are equivalent, but in higher dimension they are non-identical.

The Yamabe solitons are special solutions of the Yamabe flows, that is, there exist scalars σ(t) and diffeomorphisms ϕt in such manner that g(t)=σ(t)ϕt(g0) is the solution of the Yamabe flow (1.1), with σ(0) = 1 and ϕ0 = I. In other words, a Riemannian metric g is called a Yamabe soliton if there exist a smooth vector field V∈ 𝔛 (M) (called a potential vector field) and a scalar λ∈ ℝ such that

1.2 12LVg=(sλ)g,

where V stands for the Lie derivative operator along V. In the case which V is Killing, we say that the Yamabe soliton is trivial. In particular, if the potential vector field V= Df, where D is the gradient operator and f is a smooth scalar function, then we say that the metric g is Yamabe gradient soliton and in this case (1.2) turns into

Hessf=(sλ)g.

where Hessf is Hessian of f. If f is constant, then the above soliton called trivial Yamabe gradient soliton.

It is interesting to notice that there is a nice connection between warped product structure and Yamabe soliton. In [6], Ma and Cheng showed that a non-compact complete Riemannian manifold admitting a Yamabe gradient soliton has a warped product structure. In [10], Tokura et al. studied Yamabe gradient soliton on warped product manifold with compact Riemannian base and in that case it has been shown that the soliton is trivial. Yamabe solitons have been studied by many geometers in many different contexts (see [13,8,9,11,13]). Recently, in[7], the first author introduced a special vector field ν which satisfies the relation

1.3 Xν=α{Xν(X)ν},

where α∈ ℝ and ν is the 1-form equivalent to ν in a Reimannian manifold (M, g). A unit non-parallel (i.e., α ≠ 0) vector field ν satisfying the preceding equation is called a concurrent-recurrent vector field. It is shown that an n-dimensional connected Riemannian manifold (M, g) admits concurrent-recurrent vector field ν, if and only if, (M, g) is the warped product I × f(t) F, where I is an open interval and f(t) = eαt (see [7: Theorem 3]). In this paper, we consider a Yamabe soliton on a Riemannian manifold admitting concurrent-recurrent vector field and prove:

Theorem 1.1

Let (M, g) be Riemannian three-manifold admitting concurrent-recurrent vector field. If the metric g is a Yamabe soliton with potential vector field V, then the manifold is of constant negative curvature −α2 and V is Killing.

Due to the Theorem 3 of [7], we have:

Corollary 1.1.1

Let M = I × f(t)) F with the warping function f(t) = eαt, where α∈ ℝ, I is an open interval inand F is a Riemannian 2-manifold. If the metric of M is a Yamabe soliton, then the manifold is of constant negative curvatureα2.

In [5], Huang and Li introduced the notion of a τ-quasi Yamabe gradient soliton which naturally extends the concept of Yamabe gradient soliton. According to Huang and Li[5], a τ-quasi Yamabe gradient soliton is a Riemannian metric g satisfying

1.4 Hessf=1τdfdf+(sλ)g,

where f is a smooth scalar function and τ > 0 is a constant. Notice that a Yamabe gradient soliton is nothing but an ∞-quasi Yamabe gradient soliton. For λ < 0 the Yamabe soliton (or τ-quasi Yamabe gradient soliton) is said to be shrinking, for λ > 0 is said to be expanding, and for λ = 0 is said to be steady. In [12], Wang showed that a non-compact complete Riemannian manifold admitting a τ-quasi Yamabe gradient soliton has warped product structure. In this direction, we consider a τ-quasi Yamabe gradient soliton on a Riemannian manifold admitting concurrent-recurrent vector field and prove the following theorem.

Theorem 1.2

Let M be Riemannian manifold admitting concurrent-recurrent vector field. If the metric of M is non-trivial τ-quasi Yamabe gradient soliton with potential function f, then the scalar curvature of M is constant and is equal ton(n−1)α2.

From Theorem 3 of [7], we immediately have the following.

Corollary 1.2.1

Let M = I × f(t) F with the warping function f(t) = eαt, where α∈ ℝ, I is an open interval inand F is a Riemannian n-manifold. If the metric of M is a non-trivial τ-quasi Yamabe gradient soliton, then the scalar curvature of M is constant and is equal ton(n− 1)α2.

2 Background and key lemmas

A unit vector field ν on a Reimannian manifold (M, g) is said to be a concurrent-recurrent vector field if it satisfies

2.1 Xν=α{Xν(X)ν},

where ∇ is the Levi-Civita connection of g and α is a non-zero constant. In [7], the author constructed certain examples of n-dimensional Riemannian manifolds admitting such vector fields. An interesting property of this vector field is that it is an eigenvector of the Ricci operator of the Riemannian manifold (M, g) on which this vector field is defined. Moreover, the defining equation (2.1) dictates that integral curves of ν are geodesics. The following result has been proved in [7].

Theorem 2.1

A Riemannian n-manifold admitting a concurrent-recurrent vector field is locally isometric to the warped product I × f(t) F, where I⊆ ℝ is an open interval and F is a Riemannian (n−1)-manifold. Conversely, the warped product I × f(t) F with the warping function f(t) =eαt admits a concurrent-recurrent vector field .

2.1 Key lemmas

In this subsection, we give some lemmas that are needed to prove our main results.

Lemma 2.1

A Riemannian manifold equipped with a concurrent-recurrent vector field ν satisfies

2.2 ν(s)=2α(s+n(n1)α2).

Proof. Using (1.3) in the definition of Riemann curvature tensor, we obtain

2.3 R(X,Y)ν=α2{ν(Y)Xν(X)Y}.

Contracting the above equation gives

2.4 Ric(X,ν)=(n1)α2ν(X),

which yields = −(n−1)α2ν, where Q is Ricci operator defined by g(QX,Y) = Ric(X,Y). Differentiating =−(n−1)α2ν along X implies that

2.5 (XQ)ν=(n1)α3XαQX.

Taking the g-trace of the above equation gives (2.2).

Let (M, g) be a Riemannian manifold. If there exists ρC(M), called the potential function, such that

£Vg=2ρg

then we say that the vector field V is a conformal vector field (see Yano [14]). Moreover, V is homothetic when ρ is constant, whereas Killing when ρ= 0. Now, we recall the following result from Yano [14].

Lemma 2.2

A conformal vector field V on a Riemannian n-manifold (Mn,g) satisfies

(£VRic)(X,Y)=(n2)g(XDρ,Y)(Δρ)g(Y,X),£Vs=2ρs2(n1)Δρ,

where Δ = div D is the Laplacian operator of g.

Lemma 2.3

If the metric of Riemannian three-manifold M equipped with a concurrent-recurrent vector field is a Yamabe soliton, then the scalar curvature of M is harmonic and the Yamabe soliton is shrinking with λ = –6α2.

Proof. First, we take Lie-derivative to g (ν,ν) = 1 along V, and employ the equations [1.2] and [1.3] to get

2.6 (LVν)ν=ν(LVν)=(sλ).

Now, taking n= 3 and ρ= sλ in Lemma 2.2, we find

2.7 (VRic)(X,Y)=g(XDs,Y)(Δs)g(X,Y),

2.8 Vs=2s(sλ)4Δs.

In Riemannian three-manifolds, the curvature tensor is given by

2.9 R(X,Y)Z=g(Z,Y)QXg(Z,X)QY+g(Z,QY)Xg(Z,QX)Ys2{g(Z,Y)Xg(Z,X)Y}.

Taking Z = ν in the above equation and using (2.3), we easily deduce

2.10 Ric(X,Y)=s2+α2g(Y,X)3α2+s2ν(X)ν(Y).

Lie-differentiating of (2.10) along V and employing the equations (2.8) and(1.2), we find

(LVRic)(X,Y)=(2α2(sλ)2Δs)g(X,Y)+(s(sλ)+2Δs)ν(X)ν(Y)s2+3α2{(LVν)(X)ν(Y)+ν(X)(LVν)(Y)}.

Comparing the above equation with (2.7),we obtain

2.11 g(XDs,Y)=(Δs2α2(sλ))g(X,Y)(s(sλ)+2Δs)ν(X)ν(Y)+s2+3α2{(LVν)(X)ν(Y)+ν(X)(LVν)(Y)}.

Replacing X and Y in the above equation by ν and utilizing(2.6), we see that

ν(νs)=Δs+4α2(sλ.

Now, we use (2.2) in the above equation in order to obtain

2.12 Δs=4α2(λ+6α2).

We replace Y by ν in (2.11) and use the equations (2.6) and (2.12) to deduce

g(XDs,ν)=(4α2(λ+6α2)+((α2s2)sλ))ν(X)+(s2+3α2)(LVν)X.

On the other hand, differentiating (2.2) along X and utilizing (1.3) we obtain

g(XDs,ν)=3α(Xs)2α2(s+6α2)ν(X).

Using the preceding equation in (2.11), we get

s2+3α2(LVν)(X)={(sλ)(s2α2)2α2(s+6α2)4α2(λ+6α2)}ν(X)3α(Xs).

Substituting the above equation in (2.11) and using(2.12), we find

g(XDs,Y)=2α2(λ+s+12α2)g(X,Y)+2α2(λ3s12α2)ν(X)ν(Y)3α(Xs)ν(Y)3α(Ys)ν(X),

which further leads to

2.13 XDs=2α2(λ+s+12α2)X+2α2(λ3s12α2)ν(X)ν3α(Xs)ν3αν(X)Ds.

Replacing X in the previous equation by ν and applying(2.2), we get

νDs=2α2(s+6α2)ν3αDs.

Operating the above equation by ∇X gives us

2.14 XνDs=2α2(Xs)ν3αXDs2α2(s+6α2)Xν.

On the other hand, differentiating (2.13) along ν, we get

νXDs=2α2(νs)X2α2(λ+s+12α2)νX6α2(νs)ν(X)ν+2α2(λ3s12α2)ν(νX)ν3αν(Xs)ν3αν(νX)Ds3αν(X)νDs.

Again from (2.13)we immediately get

2.15 [X,ν]Ds=2α2(λ+s+12α2)(XννX)3αg(XννX,Ds)ν+2α2(λ3s12α2)ν(XννX)ν3αν(XννX)Ds.

Now, employing the equations (2.14)–(2.15), one can easily get

R(X,ν)Ds=2α2(Xs)ν3αXDs2α2(s+6α2)Xν+2α2(νs)X+2α2(λ+s+12α2)νX+6α2(νs)ν(X)ν+3αν(X)νDs2α2(λ3s12α2)ν(νX)ν+3αν(Xs)ν+3αν(νX)Ds+2α2(λ+s+12α2)(XννX)+3αg(XννX,Ds)ν+3αν(XννX)Ds2α2(λ3s12α2)ν(XννX)ν.

Contracting the above equation, we find

Ric(ν,Ds)=10α2(νs)3αΔs4α3(s+6α2)+6αν(νs)+4α3(λ+s+12α2).

Making use of (2.2) together with (2.12)in the preceding equation, we obtain

Ric(ν,Ds)=4α3(3λ+s+30α2).

Now, we employ (2.4) and (2.2) in the above equation to deduce the value of soliton constant λ=−6α2, which means Yamabe soliton is expanding. Using this in (2.12), we have Δs= 0, that is, the scalar curvature is harmonic. □

3 Proof of the main results

Proof of Theorem 1.1. first, differentiate (1.2) along Z to achieve

3.1 (ZVg)=2(Zs)g(X,Y).

In [14], Yano reveals the following relation:

(LVXgXLVg[V,X]g)(Y,Z)=g((LV)(X,Z),Y)g((LV)(X,Y),Z).

Due to ∇g = 0, it appears from the preceding equation that

(XLVg)(X,Y)=g((LV)(X,Z),Y)+g((LV)(X,Y),Z).

Utilizing the symmetric property of LV∇, the foregoing equation brings into view

2g((LV)(X,Y),Z)=(YLVg)(Z,X)+(XLVg)(Y,Z)(ZLVg)(X,Y).

Utilizing (3.1) in the previous equation, we find

3.2 (LV)(X,Y)=(Xs)Y+(Ys)Xg(Y,X)Ds.

Taking differentiation of (3.2) covariantly along Z yields

(ZL)(X,Y)=g(ZDs,X)Y+g(ZDs,Y)Xg(Y,X)ZDs.

Applying the above relation on the following well known formula

(LVR)(X,Y)Z=(XLV)(Y,Z)(YLV)(X,Z),

appears that

(LVR)(X,Y)Z=g(XDs,Z)Yg(Z,Y)XDsg(YDs,Z)X+g(X,Z)YDs.

Replacing Z in the previous equation by ν and calling back λ = –6α2 and the equation (2.13), we infer

3.3 (LVR)(X,Y)ν=(4α2(s+6α2)ν(Y)+3α(Ys))X3α(Ys)ν(X)ν(4α2(s+6α2)ν(X)+3α(Xs))Y+3α(Xs)ν(Y)ν.

At this point, we differentiate (2.3)and apply (1.2) to ensure

3.4 (LVR)(X,Y)ν=α2{(g(Y,LVν)+2(s+6α2)ν(Y))X(g(X,LVν)+2(s+6α2)ν(X))Y}R(X,Y)LVν.

Subtracting (3.3) from (3.4), we see that

R(X,Y)LVν=(6α2(s+6α2)ν(X)+3α(Xs)+α2g(LVν,X))Y+3α(Ys)ν(X)ν(6α2(s+6α2)ν(Y)+3α(Ys)+α2g(Y,LVν))X3α(Xs)ν(Y)ν.

Contracting the above equation, we may obtain that

Ric(Y,LVν)=6α2(s+6α2)3α(Ys)2α2g(Y,LVν).

By the support of (2.10) and (2.6), the above equation shows that

3.5 (s+6α2)LVν=(s+18α2)(s+6α2)ν6αDs.

If possible, we suppose that on an open subset O of M there holds s ≠ –6α2. Then it appears from equation (3.5) that

3.6 LVν=(s+18α2)ν6αs+6α2Ds.

Replacing Y with ν in the well-known formula (see [14]):

(LV)(X,Y)=LVXYXLVY[V,X]Y,

and then utilizing λ = –6α2,(1.3) and (3.6)we obtain the following equality

3.7 (LV)(X,ν)=α(s+30α2)ν(X)ν+s6α2s+6α2(Xs)ν12α2s+6α2ν(X)Ds+α(s+6α2)X6α(s+6α2)2(Xs)Ds.

On the other hand, replacing Y in (3.2) by ν and using (2.2) we have

3.8 (LV)(X,ν)=(Xs)ν2α(s+6α2)Xν(X)Ds.

Comparing (3.7) with (3.8) implies that

3.9 1s6α2s+6α2(Xs)ν3α(s+6α2)X+12α2s+6α2ν(X)Ds+α(s+30α2)ν(X)ν+6α(s+6α2)2(Xs)Ds=0.

Finally, replacing X by ν in (3.9)and applying (2.2) we get

3.10 Ds=2α(s+6α2)ν.

Using the previous equation in (3.6) gives us LVν = –(s + 6α2)ν. This relation together with (1.2) gives

3.11 (LVν)(X)=(LVg)(X,ν)+g(X,LVν)=(s+6α2)ν(X).

With the help of Lemma (2.3), one can easily derived from (2.8) that

3.12 LVs=2s(s+6α2).

Now we write (3.10) as: ds = –2α(s + 6α2)ν. Now, we take Lie derivative to this equation in order to deduce

LVds=2α(LVs)ν2α(s+6α2)LVν.

From (3.12), the preceding equation transforms into

3.13 LVds=4αs(s+6α2)ν2α(s+6α2)LVν.

Operating the equation (3.12) by d and using ds = –2α(s+6α2)ν, we have

3.14 LVds=4α(s+6α2)2ν+4αs(s+6α2)ν,

where used the fact that Lie-derivative commutes with the exterior derivative. Now, comparing (3.13) with (3.14), leads us to the following formula

(s+6α2){LVν+2(s+6α2)}ν=0.

As we know s ≠ –6α2 on O, we must have

LVν=2(s+6α2)ν.

Comparing the previous equation with (3.11) shows that the scalar curvature s = –6α2 on O and this is a contradiction. So, we must have s = –6α2 on M. Substituting this in (2.10)we see that Ric = (1–n)α2g which along with(2.9) shows that the manifold is of constant negative curvature –α2, and consequently, V is Killing. This concludes the proof of the theorem.

Proof of Theorem 1.2. We may write the equation (1.4) as

3.15 XDf=1τg(X,Df)Df+(sλ)Df.

Using the previous equation in the definition of curvature tensor, we find

3.16 R(X,Y)Df=(sλ)τ{(Yf)X(Xf)Y}+(Xs)Y(Ys)X.

Replacing X by ν in the previous equation and comparing the obtained equation with R(ν, X)Df = –α2{g(X, Df)ν – (νf)X} (which follows from(2.3)), we have

3.17 α2(Yf)ν+α2(νf)Y=sλτ(Yf)νsλτ(νf)Y+(νs)Y(Ys)ν.

On the other hand, we contract the equation (3.16) over X to find the expression of Ricci tensor as

3.18 Ric(Y,Df)=(n1)(sλ)τ(Yf)(n1)(Ys).

Replacing Y in the above equation by ν and utilizing (2.4) imply

3.19 (νs)α2(νf)=sλτ(νf),

which is a relation involving the scalar curvature and the potential function of the τ-quasi Yamabe gradient soliton. Feeding (3.19) in (3.17), we obtain

3.20 τDs=(sλ+τα2)Df.

Differentiating (3.20) with respect to X and using (3.15), we reach at

3.21 τXDs=(Xs)Df+(Xf)Df+(sλ+τα2)(sλ)X.

Taking scalar product of (3.16) with D f, we have (Xs)Df= (Xf) Ds. Employing this in (3.21), we find

3.22 τXDs=2g(X,Df)Ds+(sλ+τα2)(sλ)X.

At this stage, we use (3.20) in (3.28) in order to ensure

QDf=α2(n1)Df.

Differentiating the previous equation with respect to X and calling back (3.15) give

3.23 (XQ)Df+(sλ)QX+α2(n1)(sλ)X=0.

On the other hand, from second Bianchi identity one can find

traceg{X(XQ)Y}=(div Q)(Y)=12Y(s).

Now, we contract (3.23) over X and utilize the above identity to deduce

3.24 g(Ds,Df)+2(sλ)(s+nα2(n2))=0.

Combining the equations (3.15),(3.20), and (3.22) one can easily find

3τg(X,Df)g(Ds,Df)+2(sλ+τα2)(sλ)τg(X,Df)=2(2s2λ+τα2)(sλ+τα2)τg(X,Df)=0.

Setting X = Df in the above equation and using the fact that |Df| ≠ 0 (as the soliton is non-trivial), we have

3τg(Ds,Df)+2(sλ+τα2)(sλ)τ+2(2s2λ+τα2)(sλ+τα2)τ=0.

Now, we employ (3.24) in the preceding equation to deduce

(sλ+τα2)(3s3λ+τα2)3(sλ)(s+nα2(n1))=0.

The above equation shows that s is constant. So that we have ν(s)= 0, which together tracing of (2.5) gives that s = –2(n–1). This completes the proof.

4 Example

In this section, we construct a Riemannian manifold (M, g) admitting a concurrent-recurrent vector field for which the metric g is a Yamabe soliton.

Consider a manifold M = {(u, v, w) ∈ ℝ3:v > 0,w ≠ 0} with a coordinate system (u, v, w). Let us define a Riemannian metric on M as

g=2α2w2(du)2+α2w22v(du dv+dvdu)+α2w24v2(dv)2+1α2w2(dw)2, 

where α( ≠ 0) ∈ ℝ. From Koszul’s formula, one can easily compute

uu =2α4w3w,uv=α4w32vw, uw=1wu,vu=α4w32vw,vv=1vvα4w34v2w,vw=1wv,wu=1wu,wv=1wv,ww=1ww. 

Let us take ν=αww. Then, from above we can verify that

Xiν=α{Xiν(Xi)ν},

for all 1≤ i≤ 3, where X1=u, X2=v and X3=w. Thus, the vector field ν=αww is concurrent-recurrent vector field. Now we use Levi-Civita connection to find the non-zero components of curvature tensor as given below

R(u,v)u=α4w22vu+2α4w2v, R(v,w)v =α4w24v2w,R(u,w)w=1w2u, R(u,w)v=α4w22vw,R(u,v)v=α4w22vvα4w24v2u, R(v,w)w=1w2v,R(u,w)u=2α4w2w, R(v,w)u=α4w22vw,

From the curvature tensor we find the scalar curvature as s = –6α2. Also, it is not hard to verify that

R(Xi,Xj)Xk=α2(g(Xj,Xk)Xig(Xi,Xk)Xj),

for all 1≤ i, j, k ≤ 3. This shows that (M, g) is of constant curvature –α2.

Now we shall show that the metric g is a Yamabe soliton on M. Let

V=ln(v)2u2v(ln(v)+2u)v+ww

be a vector field on M. It is not hard to see that LVg = 0, and so V is Killing. Thus, we see that

LVg=(sλ)g

for λ = –6α2. Hence, g is a Yamabe soliton having the potential vector field V=ln(v)2u2v(ln(v)+2u)v+ww and the soliton constant λ = –6α2. Also, we see that this example verifies our Theorem 1.1.


(Communicated by Július Korbaš)


References

[1] DESHMUKH, S.—CHEN, B. Y.: A note on Yamabe solitons, Balkan J. Geom. Appl. 23(1) (2018), 37—43.Search in Google Scholar

[2] ERKEN, I. K.: Yamabe solitons on three-dimensional normal almost paracontact metric manifolds, Period. Math. Hungar. 80 (2020), 172—184.10.1007/s10998-019-00303-3Search in Google Scholar

[3] GHOSH, A.: Yamabe soliton and quasi Yamabe soliton on Kenmotsu manifold, Math. Slovaca 70(1) (2020), 151—160.10.1515/ms-2017-0340Search in Google Scholar

[4] HAMILTON, R.: The Ricci flow on surfaces, Contemp. Math. 71 (1988), 237—262.10.1090/conm/071/954419Search in Google Scholar

[5] HUANG, G. Y.—LI, H. Z.: On a classification of the quasi Yamabe gradient solitons, Methods Appl. Anal. 21(3) (2014), 379—390.10.4310/MAA.2014.v21.n3.a7Search in Google Scholar

[6] MA, L.—CHENG, L.: Properties of complete non-compact Yamabe solitons, Ann. Global Anal. Geom. 40(3) (2011), 379—387.10.1007/s10455-011-9263-3Search in Google Scholar

[7] NAIK, D. M.: Ricci solitons on Riemannian manifolds admitting certain vector field, Rie. Mat. (2021), https://doi.org/10.1007/sll587-021-00622-z.10.1007/s11587-021-00622-zSearch in Google Scholar

[8] SUH, Y. J.—DE, U. C: Yamabe solitons and Ricci solitons on almost co-Kähler manifolds, Canad. Math. Bull. 62 (2019), 653—661.10.4153/S0008439518000693Search in Google Scholar

[9] SUH, Y. J.—DE, U. C: Yamabe solitons and gradient Yamabe solitons on three-dimensional N(κ)-contact manifolds, Int. J. Geom. Methods Mod. Phys. 17(12) (2020), Art. ID 2050177.10.1142/S0219887820501777Search in Google Scholar

[10] TOKURA, W.—ADRIANO, L.—PINA, R.—BARBOZA, M.: On warped product gradient Yamabe solitons, J. Math. Anal. Appl. 473(1) (2019), 201—214.10.1016/j.jmaa.2018.12.044Search in Google Scholar

[11] VENKATESHA, V.—NAIK, D. M.: Yamabe Solitons on 3-dimensional contact metric manifolds with Qφ = φQ, Int. J. Geom. Methods Mod. Phys. 16(3) (2019), Art. ID 1950039.10.1142/S0219887819500397Search in Google Scholar

[12] WANG, L. F.: On non compact quasi Yamabe gradient solitons, Differential Geom. Appl. 31 (2013), 337—347.10.1016/j.difgeo.2013.03.005Search in Google Scholar

[13] WANG, Y.: Yamabe soliton on three dimensional Kenmotsu manifolds, Bull. Belg. Math. Soc. Simon Stevin 23 (2016), 345—355.10.36045/bbms/1473186509Search in Google Scholar

[14] YANO, K.: Integral Formulas in Riemannian Geometry, Marcel Dekker, 1970.Search in Google Scholar

Received: 2022-01-11
Accepted: 2022-03-19
Published Online: 2023-03-31
Published in Print: 2023-04-01

© 2023 Mathematical Institute Slovak Academy of Sciences

This work is licensed under the Creative Commons Attribution 4.0 International License.

Downloaded on 19.9.2025 from https://www.degruyterbrill.com/document/doi/10.1515/ms-2023-0037/html
Scroll to top button