Next Article in Journal
Minkowski–Sierpinski Fractal Structure-Inspired 2 × 2 Antenna Array for Use in Next-Generation Wireless Systems
Previous Article in Journal
Coefficients Inequalities for the Bi-Univalent Functions Related to q-Babalola Convolution Operator
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Characteristics of Sasakian Manifolds Admitting Almost ∗-Ricci Solitons

by
Vladimir Rovenski
1,* and
Dhriti Sundar Patra
2
1
Department of Mathematics, University of Haifa, Mount Carmel, Haifa 3498838, Israel
2
Department of Mathematics, Indian Institute of Technology, Hyderabad 502285, India
*
Author to whom correspondence should be addressed.
Fractal Fract. 2023, 7(2), 156; https://doi.org/10.3390/fractalfract7020156
Submission received: 14 December 2022 / Revised: 1 February 2023 / Accepted: 2 February 2023 / Published: 4 February 2023
(This article belongs to the Section Geometry)

Abstract

:
This article presents some results of a geometric classification of Sasakian manifolds (SM) that admit an almost ∗-Ricci soliton (RS) structure ( g , ω , X ) . First, we show that a complete SM equipped with an almost ∗-RS with ω const is a unit sphere. Then we prove that if an SM has an almost ∗-RS structure, whose potential vector is a Jacobi vector field on the integral curves of the characteristic vector field, then the manifold is a null or positive SM. Finally, we characterize those SM represented as almost ∗-RS, which are ∗-RS, ∗-Einstein or ∗-Ricci flat.

1. Introduction

Nonlinear systems that support solitons can generate self-similarity and fractals on successively smaller scales. Very often, the solitons of a given system are all self-similar to each other, see [1].
Several authors study Ricci solitons (RS) in contact metric geometry, where Sasakian manifolds (SM) are especially important. SM, which are odd-dimensional analogues of Kähler manifolds, are used to build geometrical examples, e.g., manifolds of special holonomy, Einstein manifolds (EM), Kähler manifolds and orbifolds. The geometry of SM (in particular, η -EM and Sasaki–Einstein manifolds) attracts much attention from mathematicians. We refer the reader to [2,3,4] for the latest developments in the theory of these manifolds.
Many tools for Kähler manifolds have analogues for contact manifolds, among them the ∗-Ricci tensor, introduced in [5] for almost Hermitian manifolds. In [6], the ∗-Ricci tensor was applied to a real hypersurface in a non-flat complex space form. Namely, for an almost contact manifold M 2 n + 1 ( φ , ζ , η , g ) and any vector fields Y 1 , Y 2 , Y 3 on M 2 n + 1 , we get
Ric g * ( Y 1 , Y 2 ) = 1 2 trace g Y 3 R ( Y 1 , φ Y 2 ) φ Y 3 .
If Ric g * vanishes identically then such M 2 n + 1 is called ∗-Ricci flat. A ∗-RS structure has been introduced in [7], using Ric g * instead of the Ricci tensor in the equation of RS,
£ X g + 2 Ric g * = 2 ω g .
Here, £ is the Lie derivative. If ω in (2) belongs to C ( M ) and is not a constant, then we get an almost ∗-RS, the triple ( g , ω , X ) . An almost ∗-RS is shrinking for ω > 0 , steady for ω = 0 and expanding for ω < 0 . Observe that a ∗-RS is ∗-EM (i.e., the ∗-Ricci tensor and the metric tensor are homothetic) if X is a Killing vector field (KVF), see [8]. Thus, it is a generalization of a ∗-EM. In particular, if X = f (the gradient of f C ( M ) ) in (2), then we obtain a gradient almost ∗-RS. Some generalizations of RS were studied by several authors, see [9,10,11,12,13,14,15].
Ghosh-Patra [16] first studied ∗-RS and gradient almost ∗-RS on a contact metric manifold, in particular, on an SM and ( k , μ ) -contact manifolds. Later on, ∗-RS were studied on almost contact manifolds in [15,17,18,19,20]. Wang [20] proved that “if the metric of a Kenmotsu 3-manifold represents a ∗-RS, then the manifold is locally the hyperbolic space H 3 ( 1 ) ”, and it was proved in [18] that “if a non-Kenmotsu ( k , μ ) -almost Kenmotsu manifold has a ∗-RS structure, then it is locally isometric to H n + 1 ( 4 ) × R under some restrictions”.
There is a natural question: “under what conditions a (gradient) almost RS is a unit sphere”? Several affirmative answers to this question on Riemannian and contact metric manifolds, are given in [9,10,12,13,14,21,22]. An almost ∗-RS is a generalization of ∗-RS and ∗-EM; thus we ask a question: “under what conditions a (gradient) almost ∗-RS is a unit sphere?”
In [16], they answered this question by showing that “if a complete SM admits a gradient almost ∗-RS, then it is a unit sphere”. Here, we consider non-gradient almost ∗-RS in the case of SM, and we are looking for conditions under which SM having an almost ∗-RS structure is a unit sphere. Our main achievement is the following.
Theorem 1.
If a complete SM M 2 n + 1 ( φ , ζ , η , g ) of dimension greater than three has an almost ∗-RS structure ( g , ω , X ) with ω const, then it is the unit sphere S 2 n + 1 .
The following question arises: “under what conditions an almost RS is an RS, or, EM”? Several answers to this question can be found in [14,22,23,24]. An almost ∗-RS generalizes ∗-RS and ∗-EM; so the following question arises:
“when an almost ∗-RS is a ∗-RS, for example, an ∗-EM ?”
In [16], they found such a condition on an SM, and this question on a Kenmotsu manifold was studied in [15]. In this regard, we are interested in characterizing those SM represented as almost ∗-RS, which are ∗-RS or are ∗-EM. In the following theorem, we assume that the potential vector field is a Jacobi vector field on the integral curves of the Reeb vector field.
Theorem 2.
If an SM M 2 n + 1 ( φ , ζ , η , g ) has an almost ∗-RS structure ( g , ω , X ) such that X is a Jacobi field on the ζ-integral curves, then ( g , ω , X ) is a ∗-RS on M.
Now, we recall some definitions, e.g., [3]. A SM M 2 n + 1 ( φ , ζ , η , g ) is an η-EM if
Ric g = α 1 g + α 2 η η ,
where functions α 1 and α 2 belong to C ( M ) . For K-contact manifolds (for example, SM) of dimension 4 , α 1 , α 2 are real constants, see [25], thus, the scalar curvature is constant. An η -Einstein SM is a null-SM when α 1 = 2 and α 2 = 2 n + 2 , and is a positive-SM when α 1 > 2 , see [3]. Using ([16], Theorem 8), we get the following consequence of Theorem 2.
Corollary 1.
If an SM M 2 n + 1 ( φ , ζ , η , g ) has an almost ∗-RS structure ( g , ω , X ) such that X is a Jacobi vector field on the ζ-integral curves, then M is positive SM and X is KVF (and g is ∗-EM), or M is null-SM and φ is invariant under X.
Remark 1.
Observe from [24] that if an SM has an almost RS structure, whose potential field is a Jacobi vector field on the ζ -integral curves, then the manifold is null-SM with the expanding RS.
A vector field Y on a contact manifold ( M , η ) that preserves the contact form η is called an infinitesimal contact transformation, see [26,27], i.e.,
£ Y η = ν η ,
for some ν C ( M ) ; and if ν = 0 , then Y is said to be strict. A vector field Y preserving φ , ζ , η and g on a contact metric manifold is called an infinitesimal automorphism. In [16], they addressed the question by showing that “if an SM represents an almost ∗-RS such that the potential vector field is an infinitesimal contact transformation, then it is a ∗-RS". We improve this result in the following statement.
Theorem 3.
Let an SM represent an almost ∗-RS ( g , ω , X ) such that X is an infinitesimal contact transformation. Then X leaves φ invariant, and the manifold is η-EM of scalar curvature ( ω + 4 n ) n .
Assuming compactness we get one more sufficient condition for g to be ∗-Ricci flat (e.g., ∗-EM).
Theorem 4.
Let a compact SM M 2 n + 1 ( φ , ζ , η , g ) admit an almost ∗-RS ( g , ω , X ) such that X is an infinitesimal contact transformation. Then the soliton is steady ( ω = 0 ) , X is an infinitesimal automorphism, and g is ∗-Ricci flat of scalar curvature 4 n 2 .
Finally, we show that the property “potential vector field and the characteristic vector field are parallel" provides “∗-Ricci flat" (e.g., ∗-EM).
Theorem 5.
Let an SM M 2 n + 1 ( φ , ζ , η , g ) admit an almost ∗-RS ( g , ω , X ) such that X is parallel to ζ. Then the soliton is steady ( ω = 0 ) , V is KVF, and g is ∗-Ricci flat of scalar curvature 4 n 2 .
Remark 2.
Observe from [14] that if a compact SM has an almost RS structure ( g , ω , X ) such that X is an infinitesimal contact transformation, then the soliton is shrinking ( ω = 2 n ), X is strict and leaves φ invariant, and the manifold is EM of scalar curvature 2 n ( 2 n + 1 ) . By Theorem 4, the soliton is steady ( ω = 0 ) , X is strict and leaves φ , ζ and η invariant, and the manifold is ∗-EM of scalar curvature 4 n 2 . In addition, note that an SM considered as an RS or an almost RS with X parallel to ζ , is an EM of scalar curvature 2 n ( 2 n + 1 ) , and X is a KVF, see [28].
Within the framework of Theorems 3–5, we not only find sufficient conditions for ∗-EM, but also characterize the structural tensor fields and the potential vector field.

2. Preliminaries

Here, we recall some properties of SM, see [2,26]. We suppose that all manifolds are smooth and connected. The curvature tensor R on a Riemannian manifold ( M , g ) is given by
R ( Y 1 , Y 2 ) = [ Y 1 , Y 2 ] [ Y 1 , Y 2 ] , Y 1 , Y 2 X ( M ) ,
where ∇ is the Levi-Civita connection and X ( M ) is the space of vector fields on M. They define the Ricci operator Q by
g ( Q Y 1 , Y 2 ) = Ric g ( Y 1 , Y 2 ) = trace g Y 3 R ( Y 3 , Y 1 ) Y 2 , Y 1 , Y 2 , Y 3 X ( M )
is a symmetric ( 1 , 1 ) -tensor. Then r = trace g Q is the scalar curvature. We get the following formula (follows from twice contracted second Bianchi identity, see, e.g., [29]):
1 2 g ( Y 1 , r ) = ( div Q ) ( Y 1 ) = i g ( ( E i Q ) Y 1 , E i ) , Y 1 X ( M ) ,
where { E i } is any local orthonormal basis on M.
For the Lie derivative of f C ( M ) along Y 1 X ( M ) we have
£ Y 1 f = Y 1 ( f ) = g ( Y 1 , f ) ,
where the Hessian of f is defined by
Hess f ( Y 1 , Y 2 ) = 2 f ( Y 1 , Y 2 ) = ( Y 1 f ) Y 2 = Y 1 Y 2 ( f ) ( Y 1 Y 2 ) ( f ) , Y 1 , Y 2 X ( M ) .
The hessian Hess f is symmetric and g ( Y 1 f , Y 2 ) = g ( Y 2 f , Y 1 ) is true, so it is a section of T * M T * M . A vector field Y on ( M , g ) is conformal if
£ Y g = f g
for some f C ( M ) . Such Y is called homothetic if f = c o n s t a n t , and Y is said to be KVF if f = 0 .
A manifold M 2 n + 1 is a contact manifold if η ( d η ) n 0 for some global 1-form η . In this case, ζ satisfying d η ( ζ , · ) = 0 and η ( ζ ) = 1 is called a characteristic vector field. Polarization of d η on D = ker η allows to find a ( 1 , 1 ) -tensor φ and a Riemannian structure g satisfying
φ 2 = i d T M + η ζ , η = g ( ζ , · ) ,
d η ( · , · ) = g ( · , φ · ) .
They call such M 2 n + 1 ( φ , ζ , η , g ) a contact metric manifold with associated metric g. By (6), we get
φ ( ζ ) = 0 , η φ = 0 , and rank ( φ ) = 2 n .
Further, a contact metric manifold is called SM if the following condition is valid, see [3,26]:
[ φ Y 1 , φ Y 2 ] = 2 d η ( Y 1 , Y 2 ) ζ , Y 1 , Y 2 X ( M ) .
The curvature tensor of an SM satisfies
R ( Y 1 , Y 2 ) ζ = η ( Y 2 ) Y 1 η ( Y 1 ) Y 2 , Y 1 , Y 2 X ( M ) .
Recall that if ζ is KVF, then M is called a K-contact manifold. A SM is K-contact, the converse is valid in dimension 3 only, see ([26], p. 87). SM satisfy the following, see ([26], p. 113):
Y 1 ζ = φ Y 1 , Y 1 X ( M ) ,
Q ζ = 2 n ζ .
According to ([16], Lemma 2.1):
ζ Q = Q φ φ Q ;
therefore, ζ Q = 0 , as φ and the Ricci operator commute on an SM, see [26]. Using (9) we find the derivative of (10) in the direction of Y 1 X ( M ) ,
( Y 1 Q ) ζ = Q φ Y 1 2 n φ Y 1 .
The ∗-Ricci tensor Ric g * on an SM has the view (see [16], Lemma 5):
Ric g * ( Y 1 , Y 2 ) = Ric g ( Y 1 , Y 2 ) ( 2 n 1 ) g ( Y 1 , Y 2 ) η ( Y 1 ) η ( Y 2 ) , Y 1 , Y 2 X ( M ) .
Thus, we can write the ∗-RS structure (2) on an SM as
( £ V g ) ( Y 1 , Y 2 ) + 2 Ric g ( Y 1 , Y 2 ) = 2 ( ω + 2 n 1 ) g ( Y 1 , Y 2 ) + 2 η ( Y 1 ) η ( Y 2 )
for all Y 1 , Y 2 X ( M ) .
Remark 3.
Contracting the derivative of (8) using a local orthonormal basis { E i } 1 i 2 n + 1 on M and using (9), we get the following:
( div R ) ( Y 1 , Y 2 ) ζ g ( R ( Y 1 , Y 2 ) φ E i , E i ) = 2 g ( φ Y 1 , Y 2 ) .
Contracting the second Bianchi identity, they reduce the above equation to the following:
g ( ( Y 1 Q ) Y 2 ( Y 2 Q ) Y 1 , ζ ) g ( R ( Y 1 , Y 2 ) φ E i , E i ) = 2 g ( φ Y 1 , Y 2 ) .
Using (11), from (14) we obtain
g ( R ( Y 1 , Y 2 ) φ E i , E i ) = g ( Q φ Y 1 , Y 2 ) + g ( φ Q Y 1 , Y 2 ) 2 ( 2 n 1 ) g ( φ Y 1 , Y 2 ) .
Replacing Y 2 by φ Y 2 in the preceding equation and using the definition (1), we represent the ∗-Ricci tensor of an SM in the form (12).
To prove Theorems 1 and 2, we need the following three results.
Proposition 1.
If an SM admits an almost ∗-RS structure, then we get the following:
( £ X ) ( Y 1 , ζ ) + 2 Q φ Y 1 = 2 ( 2 n 1 ) φ Y 1 + Y 1 ( ω ) ζ + ζ ( ω ) Y 1 η ( Y 1 ) ω , Y 1 X ( M ) .
Proof. 
Using the derivative of (13) for an arbitrary Y 3 X ( M ) and (9), gives
( Y 3 £ X g ) ( Y 1 , Y 2 ) + 2 ( Y 3 Ric g ) ( Y 1 , Y 2 ) = 2 Y 3 ( ω ) g ( Y 1 , Y 2 ) 2 η ( Y 1 ) g ( Y 2 , φ Y 3 ) 2 η ( Y 2 ) g ( Y 1 , φ Y 3 )
for Y 1 , Y 2 X ( M ) . Following Yano ([30] p. 23), for Y 1 , Y 2 , Y 3 X ( M ) we write
( £ X Y 3 g Y 3 £ X g [ X , Y 3 ] g ) ( Y 1 , Y 2 ) = g ( ( £ X ) ( Y 3 , Y 1 ) , Y 2 ) g ( ( £ X ) ( Y 3 , Y 2 ) , Y 1 ) .
Inserting (16) in the above relation and using g = 0 yields the following:
g ( ( £ X ) ( Y 3 , Y 1 ) , Y 2 ) + g ( ( £ X ) ( Y 3 , Y 2 ) , Y 1 ) + 2 ( Y 3 Ric g ) ( Y 1 , Y 2 ) = 2 Y 3 ( ω ) g ( Y 1 , Y 2 ) η ( Y 1 ) g ( Y 2 , φ Y 3 ) η ( Y 2 ) g ( Y 1 , φ Y 3 ) .
Due to the symmetry ( £ X ) ( Y 1 , Y 2 ) = ( £ X ) ( Y 2 , Y 1 ) , using cyclical permutations of Y 1 , Y 2 , Y 3 in the preceding equation, we get
g ( ( £ X ) ( Y 1 , Y 2 ) , Y 3 ) = ( Y 3 Ric g ) ( Y 1 , Y 2 ) ( Y 1 Ric g ) ( Y 2 , Y 3 ) ( Y 2 Ric g ) ( Y 3 , Y 1 ) + Y 1 ( ω ) g ( Y 2 , Y 3 ) + Y 2 ( ω ) g ( Y 3 , Y 1 ) Y 3 ( ω ) g ( Y 1 , Y 2 ) 2 η ( Y 2 ) g ( φ Y 1 , Y 3 ) 2 η ( Y 1 ) g ( φ Y 2 , Y 3 ) .
Since the Ricci operator Q is self-adjoint, using ζ Q = 0 and (11), and replacing Y 2 by ζ in (17), we get (15). □
Proposition 2.
Let an SM represent an almost ∗-RS and ζ leave ω invariant. Then we get
R ( Y 1 , Y 2 ) ω g ( R ( Y 1 , Y 2 ) ω , ζ ) ζ = 4 ( Y 1 Q ) Y 2 ( Y 2 Q ) Y 1 2 Y 1 ( ω ) Y 2 Y 2 ( ω ) Y 1 + 2 Y 1 ( ω ) η ( Y 2 ) ζ Y 2 ( ω ) η ( Y 1 ) ζ + 2 ( ω 2 ) { 2 g ( Y 1 , φ Y 2 ) ζ + η ( Y 1 ) φ Y 2 η ( Y 2 ) φ Y 1 } + η ( Y 2 ) Y 1 ζ ω η ( Y 1 ) Y 2 ζ ω + 2 g ( φ Y 2 , Y 1 ) ζ ω + g ( φ Y 2 , Y 1 ω ) ζ g ( φ Y 1 , Y 2 ω ) ζ + g ( ζ , Y 1 ω ) φ Y 2 g ( ζ , Y 2 ω ) φ Y 1 , Y 1 , Y 2 X ( M ) .
Proof. 
Applying the Lie derivative for R ( Y 1 , ζ ) ζ = Y 1 η ( Y 1 ) ζ for all Y 1 X ( M ) (follows from (8)) along X and using (8), yields
( £ X R ) ( Y 1 , ζ ) ζ + R ( Y 1 , ζ ) £ X ζ + η ( £ X ζ ) Y 1 + ( £ X g ) ( Y 1 , ζ ) ζ + g ( Y 1 , £ X ζ ) ζ = 0 .
Replacing ( Y 1 , Y 2 ) by ( Y 1 , ζ ) in (13) and using (10), we acquire the following:
( £ X g ) ( Y 1 , ζ ) = 2 ω η ( Y 1 ) .
Plugging it into the Lie derivative of η ( Y 1 ) = g ( Y 1 , ζ ) and η ( ζ ) = 1 , we find η ( £ X ζ ) = ω and ( £ X η ) ( ζ ) = ω . Thus, in view of (8), Equation (19) gives us
( £ X R ) ( Y 1 , ζ ) ζ = 2 ω Y 1 η ( Y 1 ) ζ , Y 1 X ( M ) .
By conditions, ζ ( ω ) = 0 is valid, thus (15) reduces to the following:
( £ X ) ( Y 1 , ζ ) + 2 Q φ Y 1 = 2 ( 2 n 1 ) φ Y 1 + Y 1 ( ω ) ζ η ( Y 1 ) ω , Y 1 X ( M ) .
Using ζ -derivative of (21), gives
( ζ £ X ) ( Y 1 , ζ ) = g ( Y 1 , ζ ω ) ζ η ( Y 1 ) ζ ω ,
where equalities ζ Q = ζ ζ = ζ φ = 0 for an SM were used. On the other hand, using Y 1 = ζ in (21), then differentiating along Y 1 and using (9) and the symmetry of £ X , we get
( Y 1 ( £ X ) ) ( ζ , ζ ) = 2 ( £ X ) ( φ Y 1 , ζ ) Y 1 ω .
Next, differentiating g ( ζ , ω ) = 0 along Y 1 X ( M ) and using (6), gives ( φ Y 1 ) ( ω ) = g ( ζ , Y 1 ω ) ; therefore, it suffices to combine (6), (10), (21) and (23) to arrive at the result
( Y 1 £ X ) ( ζ , ζ ) = 4 Q Y 1 4 ( 2 n 1 ) Y 1 4 η ( Y 1 ) ζ + 2 g ( ζ , Y 1 ω ) ζ Y 1 ω .
We need the following commutation result, see ([30], p. 23):
( £ X R ) ( Y 1 , Y 2 ) Y 3 = ( Y 1 £ X ) ( Y 2 , Y 3 ) ( Y 2 £ X ) ( Y 1 , Y 3 ) .
Next, replacing both Y 2 and Y 3 by ζ in (25) and then plugging the values of ( £ X R ) ( Y 1 , ζ ) ζ , ( ζ £ X ) ( Y 1 , ζ ) and ( Y 1 £ X ) ( ζ , ζ ) from (20), (22) and (24), respectively, we get
Y 1 ω = 4 Q Y 1 2 ω + 2 ( 2 n 1 ) Y 1 + 2 ( ω 2 ) η ( Y 1 ) ζ + g ( ζ , X ω ) ζ + η ( Y 1 ) ζ ω
for Y 1 X ( M ) . Using (9) in the Y 2 -derivative of (26), we acquire
Y 2 Y 1 ω = 4 ( Y 2 Q ) Y 1 + Q ( Y 2 Y 1 ) 2 Y 2 ( ω ) Y 1 η ( Y 1 ) ζ 2 ω + 2 ( 2 n 1 ) Y 2 Y 1 + 2 ( ω 2 ) η ( Y 2 Y 1 ) ζ g ( Y 1 , φ Y 2 ) ζ η ( Y 1 ) φ Y 2 + { η ( Y 2 Y 1 ) g ( Y 1 , φ Y 2 ) } ζ ω + η ( Y 1 ) Y 2 ζ ω g ( φ Y 2 , Y 1 ω ) ζ + g ( ζ , Y 2 Y 1 ω ) ζ g ( ζ , Y 1 ω ) φ Y 2 .
Since Hess ω is symmetric and φ is skew-symmetric, using (26) and the above equation in (4) completes the proof of (18). □
Recall that the contact metric structure commutes with the Ricci operator, i.e., Q φ = φ Q , e.g., [26]. Its covariant derivative and (5) provide the following.
Lemma 1
(see [12]). For an SM M and all Y 1 X ( M ) , we have
( i ) i = 1 2 n + 1 g ( ( φ Y 1 Q ) φ E i , E i ) = 0 , ( i i ) i = 1 2 n + 1 g ( ( φ E i Q ) φ Y 1 , E i ) = 1 2 Y 1 ( r ) ,
where { E i } 1 i 2 n + 1 is a local orthonormal basis on M.

3. Proof of Results

Proof of Theorem 1.
Since the characteristic vector field is KVF, £ ζ g = 0 = £ ζ Ric g is valid. Applying this to the Lie derivative of (13) along ζ , and using £ Y 1 £ Y 2 g £ Y 2 £ Y 1 g = £ [ Y 1 , Y 2 ] g , e.g., [30], we get
£ [ X , ζ ] g = 2 ζ ( ω ) g .
By (27), [ X , ζ ] is a conformal vector field. This gives us the following alternatives:
(I) [ X , ζ ] is non-homothetic, (II) [ X , ζ ] is homothetic.
Okumura [31] proved that “if a complete SM of dimension > 3 has a non-Killing conformal vector field, then it is a unit sphere”. Applying this theorem for (I), we conclude that ( M , g ) is a unit sphere.
We will finish the proof by showing a contradiction for case (II). Sharma [32] proved the following: “a homothetic vector field on an SM (more generally, K-contact manifold) is necessarily a KVF”. So, (27) implies that ζ leaves ω invariant. Thus, (18) holds. Contracting it over Y 1 and then using (5), (8), we obtain
Ric g ( Y 2 , ω ) = 4 g ( φ Y 2 , ζ ω ) + η ( Y 2 ) div ( ζ ω ) g ( ζ , Y 2 ζ ω ) 2 g ( Y 2 , r ) + 2 ( 2 n 1 ) g ( Y 2 , ω ) ,
where we used trace g φ = 0 = φ ζ , the skew-symmetry of φ and symmetry of Hess ω . Differentiating the equality g ( ζ , ζ ω ) = 0 along Y 2 X ( M ) and using (9), gives
g ( ζ , Y 2 ζ ω ) = g ( φ Y 2 , ζ ω ) .
Thus, (28) can be rewritten as
Ric g ( Y 2 , ω ) = 3 g ( φ Y 2 , ζ ω ) + η ( Y 2 ) div ( ζ ω ) 2 Y 2 ( r ) + 2 ( 2 n 1 ) Y 2 ( ω )
for all Y 2 X ( M ) . Next, recall the following result for SM, see [26]:
R ( φ Y 2 , φ Y 1 ) Y 3 = R ( Y 2 , Y 1 ) Y 3 + g ( Y 2 , Y 3 ) Y 1 g ( Y 1 , Y 3 ) Y 2 g ( Y 2 , Y 3 ) Y 1 + g ( Y 1 , Y 3 ) Y 2
for all Y 1 , Y 2 , Y 3 X ( M ) . Substituting Y 1 = φ Y 1 and Y 2 = φ Y 2 in (18) and using the last formula, in view of (6), φ ζ = 0 and the skew-symmetry of φ , we obtain
R ( Y 1 , Y 2 ) ω g ( R ( Y 1 , Y 2 ) ω , ζ ) ζ = 4 ( φ Y 1 Q ) φ Y 2 ( φ Y 2 Q ) φ Y 1 + Y 2 ( ω ) 2 Y 1 η ( Y 1 ) ζ Y 1 ( ω ) 2 Y 2 η ( Y 2 ) ζ + 4 ( ω 2 ) g ( Y 1 , Y 2 ) ζ + 2 g ( Y 1 , Y 2 ) ζ ω + 2 η ( Y 2 ) g ( ζ , Y 1 ω ) ζ 2 η ( Y 1 ) g ( ζ , Y 2 ω ) ζ g ( Y 2 , Y 1 ω ) ζ g ( Y 1 , Y 2 ω ) ζ + g ( ζ , Y 2 ω ) Y 1 g ( ζ , Y 1 ω ) Y 2 .
On the other hand, contracting (30) over Y 1 and applying (8), (5), (29) and Lemma 1, we find
η ( Y 2 ) div ( ζ ω ) + 2 ( n 1 ) Y 2 ( ω ) g ( φ Y 2 , ζ ω ) = 0 ,
where we have used the symmetry of Hess ω and that ζ leaves ω invariant. Next, replacing Y 2 by φ Y 2 in (31), noting that (6) and using g ( ζ , ζ ω ) = 0 , we acquire
2 ( n 1 ) g ( φ Y 2 , ω ) + g ( Y 2 , ζ ω ) = 0 , Y 2 X ( M ) .
Furthermore, differentiating g ( ζ , ω ) = 0 along Y 2 X ( M ) and using (9), we achieve
g ( ζ , Y 2 ω ) = g ( φ Y 2 , ω ) .
Thus, (32) for n > 1 gives us g ( φ Y 2 , ω ) = 0 ; consequently, ω = 0 . Hence, ω is constant – a contradiction with the conditions of the theorem. □
Proof of Theorem 2.
Applying Proposition 1 to the well-known formula:
Y 1 Y 2 X Y 1 Y 2 X R ( Y 1 , X ) Y 2 = ( £ X ) ( Y 1 , Y 2 ) ,
see ([30], p. 23), we acquire
Y 1 ζ X Y 1 ζ X R ( Y 1 , X ) ζ = 2 Q φ Y 1 + 2 ( 2 n 1 ) φ Y 1 + Y 1 ( ω ) ζ + ζ ( ω ) Y 1 η ( Y 1 ) ω .
By conditions, X is a Jacobi vector field on the ζ -integral curves, see [33], i.e.,
ζ ζ X + R ( X , ζ ) ζ = 0 .
Using Y 1 = ζ in (33) and ζ ζ = 0 (that is a consequence of (9)), we achieve the equality ω = 2 ζ ( ω ) ζ , or, using the exterior derivative,
d ω = 2 ζ ( ω ) η .
Applying exterior derivative, the Poincaré lemma ( d 2 = 0 ), and the wedge product with η , we acquire ζ ( ω ) η d η = 0 ; thus, ζ ( ω ) = 0 , as η d η is nowhere zero on a contact manifold. Thus, d ω = 0 , i.e., ω = c o n s t a n t . □
Corollary 4 follows from ([16], Theorem 8) and our Theorem 2.
Proof of Theorem 3.
Using d £ X = £ X d (d commutes with the Lie derivative) and applying the operator d to (3), gives
( £ X d η ) ( Y 1 , Y 2 ) = 1 2 Y 1 ( ν ) η ( Y 2 ) Y 2 ( ν ) η ( Y 1 ) + ν d η ( Y 1 , Y 2 )
for a function ν C ( M ) and any Y 1 , Y 2 X ( M ) . Applying the Lie derivative of (7) in the X-direction and using (2), (3) and (34), we obtain
2 ( £ X φ ) ( Y 1 ) + 2 ( 2 ω ν + 2 ( 2 n 1 ) ) φ Y 1 = 4 Q φ Y 1 + η ( Y 1 ) ν Y 1 ( ν ) ζ .
From the first equality of (6), for Y 1 X ( M ) we get
( £ X φ ) ( φ Y 1 ) + φ ( £ X φ ) ( Y 1 ) = ( £ X η ) ( Y 1 ) ζ + η ( Y 1 ) £ X ζ .
As a result of (10), ∗-RS Equation (2) gives us
( £ X g ) ( Y 1 , ζ ) = 2 ω η ( Y 1 ) .
Taking into account this, as well as (3), it suffices to show that
g ( £ X ζ , Y 1 ) = ( ν 2 ω ) η ( Y 1 ) .
By direct calculation using φ ζ = 0 we get ( £ X φ ) ( ζ ) = 0 ; hence, (35) gives ν = ζ ( ν ) ζ . Thus, by (9) we get
Hess ν ( Y 1 , Y 2 ) = Y 1 ( ζ ( ν ) ) η ( Y 2 ) ζ ( ν ) g ( φ Y 1 , Y 2 ) .
Since φ is skew-symmetric and Hess ν is symmetric, by (7) and (38), for Y 1 , Y 2 orthogonal to ζ we achieve
ζ ( ν ) d η ( Y 1 , Y 2 ) = 0 .
Thus, ζ ( ν ) = 0 , as d η is nonzero; hence, ν = 0 . Thus, ν is constant. Combining (2), (37) and η ( ζ ) = 1 , we get ν = ω . Substituting this and () in (35), gives
( £ X φ ) ( φ Y 1 ) = φ ( £ X φ ) ( Y 1 ) = 2 Q Y 1 + ( ω + 2 ( 2 n 1 ) ) Y 1 ( ω 2 ) η ( Y 1 ) ζ ,
where the equality Q φ = φ Q (for an SM) has been used, see ([26], p. 116). Now, by (3), (36), (37), (39) and ν = ω , we obtain the relation
2 Ric g = ( ω + 2 ( 2 n 1 ) ) g ( ω 2 ) η η .
Hence, ( M , g ) is an η -EM of scalar curvature n ( ω + 4 n ) . Substituting (40) into (35), we find £ X φ = 0 ; thus, X leaves φ invariant. □
Proof of Theorem 4.
The volume form Ω on a contact metric manifold satisfies Ω = η ( d η ) n 0 ; therefore, its Lie derivative in the X-direction and (3) give
£ X Ω = ( n + 1 ) ν Ω .
Proceeding, we obtain the equality £ X Ω = ( div X ) Ω , from which we deduce div X = ( n + 1 ) ν . Applying the Divergence theorem (for compact M), this gives ν = 0 ; consequently, ω = 0 by the proof of Theorem 3. Thus, (3) and (37), respectively, follows from the condition that X leaves η and ζ invariant. Moreover, Equation (40) becomes
Ric g = ( 2 n 1 ) g + η η .
Thus, from (12) we conclude that ( M , g ) is ∗-Ricci flat of zero ∗-scalar curvature r * . Using (2), we find that X is KVF. Applying Theorem 3, completes the proof. □
Proof of Theorem 5.
By conditions, X = σ ζ , where σ is a non-zero smooth function. By the skew-symmetry of φ and (9), we obtain
( £ X g ) ( Y 1 , Y 2 ) = g ( Y 1 X , Y 2 ) + g ( X Y 2 , Y 1 ) = Y 1 ( σ ) η ( Y 2 ) + Y 2 ( σ ) η ( Y 1 ) .
Thus, (13) becomes
Y 1 ( σ ) η ( Y 2 ) + Y 2 ( σ ) η ( Y 1 ) + 2 Ric g ( Y 1 , Y 2 ) = 2 ( ω + 2 n 1 ) g ( Y 1 , Y 2 ) + 2 η ( Y 1 ) η ( Y 2 ) .
Using Y 2 = ζ in (43) and (10), yields
Y 1 ( σ ) = ( 2 ω ζ ( σ ) ) η ( Y 1 ) , Y 1 X ( M ) .
Again, using Y 1 = ζ and Y 2 = ζ in (43) and applying (10), we get ζ ( σ ) = ω ; hence, from (44) it follows that Y 1 ( σ ) = ζ ( σ ) η ( Y 1 ) , for Y 1 X ( M ) . By the above argument, we get that σ is constant. By (42), X is KV, and using (44) we get ω = 0 . The above reduces (43) to (41), and therefore, g is ∗-Ricci flat (follows from (12)) and of constant scalar curvature 4 n 2 . □

4. Conclusions

A modern geometrical concept of an almost ∗-RS can be important in differential geometry and theoretical physics. We study the interaction of this structure on a smooth manifold with the well-known Sasakian structure and prove some results of geometric classification. Theorem 1 contains a condition for a complete SM equipped with an almost ∗-RS to be a unit sphere. Theorems 2–5 having local character, contain conditions, ensuring that an SM equipped with an almost ∗-RS structure is a ∗-RS, e.g., a ∗-EM. Using the fact that an almost ∗-Ricci tensor on an SM can be written as (12), an almost ∗-RS reduces to the form (13), which is less general than the almost η -RS equation for an SM:
1 2 £ X g + Ric g + ω g + δ η η = 0 .
In connection with the above, the following question arises: “are our Theorems 1–5 true under the condition (45) instead of (13), where ω and δ are arbitrary smooth functions on M?”

Author Contributions

Conceptualization, V.R. and D.S.P.; methodology, V.R. and D.S.P.; investigation, V.R. and D.S.P.; writing—review and editing, V.R. and D.S.P.; funding acquisition, D.S.P. All authors have read and agreed to the published version of the manuscript.

Funding

This research of Dhriti Sundar Patra was funded by Indian Institute of Technology Hyderabad grant number SG/IITH/F295/2022-23/SG-133.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Soljacic, M.; Sears, S.; Segev, M.; Krylov, D.; Bergman, K. Self-similarity and fractals driven by soliton dynamics. Invited Paper, Special Issue on Solitons. Photonics Sci. News 1999, 5, 3–12. [Google Scholar]
  2. Boyer, C.P.; Galicki, K. Sasakian Geometry; Oxford University Press: Oxford, UK, 2008. [Google Scholar]
  3. Boyer, C.P.; Galicki, K.; Matzeu, P. On η-Einstein Sasakian geometry. Comm. Math. Phys. 2006, 262, 177–208. [Google Scholar] [CrossRef]
  4. Sparks, J. Sasakian–Einstein manifolds. Surveys Diff. Geom. 2011, 16, 265–324. [Google Scholar] [CrossRef]
  5. Tachibana, S. On almost-analytic vectors in almost Kählerian manifolds. Tohoku Math. J. 1959, 11, 247–265. [Google Scholar] [CrossRef]
  6. Hamada, T. Real hypersurfaces of complex space forms in terms of Ricci *-tensor. Tokyo J. Math. 2002, 25, 473–483. [Google Scholar] [CrossRef]
  7. Kaimakamis, G.; Panagiotidou, K. ∗-Ricci solitons of real hypersurface in non-flat complex space forms. J. Geom. Phys. 2014, 76, 408–413. [Google Scholar] [CrossRef]
  8. Ivey, T.A.; Ryan, P.J. The ∗-Ricci tensor for hypersurfaces in CPn and CHn. Tokyo J. Math. 2011, 34, 445–471. [Google Scholar]
  9. Barros, A.; Ribeiro, E., Jr. Some characterizations for compact almost Ricci solitons. Proc. Amer. Math. Soc. 2012, 140, 1033–1040. [Google Scholar] [CrossRef]
  10. Barros, A.; Batista, R.; Ribeiro, E., Jr. Compact almost Ricci solitons with constant scalar curvature are gradient. Monatsh. Math. 2014, 174, 29–39. [Google Scholar] [CrossRef]
  11. Crasmareanu, M. A new approach to gradient Ricci solitons and generalizations. Filomat 2018, 32, 3337–3346. [Google Scholar] [CrossRef]
  12. Gangadharappa, N.H.; Sharma, R. D-homothetically deformed K-contact Ricci almost solitons. Results Math. 2020, 75, 124. [Google Scholar] [CrossRef]
  13. Ghosh, A.; Sharma, R. K-contact and Sasakian metrics as Ricci almost solitons. Int. J. Geom. Methods Mod. Phys. 2021, 18, 2150047. [Google Scholar] [CrossRef]
  14. Patra, D.S. K-contact metrics as Ricci almost solitons. Beitr. Algebra Geom. 2021, 62, 737–744. [Google Scholar] [CrossRef]
  15. Patra, D.S.; Ali, A.; Mofarreh, F. Geometry of almost contact metrics as almost ∗-Ricci solitons. arXiv 2021, arXiv:2101.01459. [Google Scholar]
  16. Ghosh, A.; Patra, D.S. ∗-Ricci Soliton within the frame-work of Sasakian and (κ,μ)-contact manifold. Int. J. Geom. Methods Mod. Phys. 2018, 15, 1850120. [Google Scholar] [CrossRef]
  17. Dai, X. Non-existence of ∗-Ricci solitons on (k,μ)-almost cosymplectic manifolds. J. Geom. 2019, 110, 30. [Google Scholar] [CrossRef]
  18. Dai, X.; Zhao, Y.; De, U.C. ∗-Ricci soliton on (k,μ)-almost Kenmotsu manifolds. Open Math. 2019, 17, 874–882. [Google Scholar] [CrossRef]
  19. Venkatesha, V.; Naik, D.M.; Kumara, H.A. ∗-Ricci solitons and gradient almost ∗-Ricci solitons on Kenmotsu manifolds. Math. Slovaca 2019, 69, 1447–1458. [Google Scholar] [CrossRef]
  20. Wang, Y. Contact 3-manifolds and ∗-Ricci soliton. Kodai Math. J. 2020, 43, 256–267. [Google Scholar] [CrossRef]
  21. Deshmukh, S. Almost Ricci solitons isometric to spheres. Int. J. Geom. Methods Mod. Phys. 2019, 16, 1950073. [Google Scholar] [CrossRef]
  22. Deshmukh, S.; Al-Sodais, H. A note on almost Ricci solitons. Anal. Math. Phys. 2020, 10, 75. [Google Scholar] [CrossRef]
  23. Ghosh, A. Certain contact metrics as Ricci almost solitons. Results Math. 2014, 65, 81–94. [Google Scholar] [CrossRef]
  24. Ghosh, A. Ricci almost solitons and contact geometry. Adv. Geom. 2021, 21, 169–178. [Google Scholar] [CrossRef]
  25. Yano, K.; Kon, M. Structures on Manifolds; Series in Pure Math. 3; World Scientific Pub. Co.: Singapore, 1984. [Google Scholar]
  26. Blair, D.E. Riemannian Geometry of Contact and Symplectic Manifolds; Birkhäuser: Boston, MA, USA, 2010. [Google Scholar]
  27. Tanno, S. Note on infinitesimal transformations over contact manifolds. Tohoku Math. J. 1962, 14, 416–430. [Google Scholar] [CrossRef]
  28. Sharma, R. Certain results on K-contact and (κ,μ)-contact manifolds. J. Geom. 2008, 89, 138–147. [Google Scholar] [CrossRef]
  29. Besse, A. Einstein Manifolds; Springer: New York, NY, USA, 2008. [Google Scholar]
  30. Yano, K. Integral Formulas in Riemannian Geometry; Marcel Dekker: New York, NY, USA, 1970. [Google Scholar]
  31. Okumura, M. On infinitesimal conformal and projective transformations of normal contact spaces. Tohoku Math. J. 1962, 14, 398–412. [Google Scholar] [CrossRef]
  32. Sharma, R. Addendum to our paper “Conformal motion of contact manifolds with characteristic vector field in the k-nullity distribution”. Ill. J. Math. 1998, 42, 673–677. [Google Scholar] [CrossRef]
  33. Blair, D.E.; Vanhecke, L. Jacobi vector fields and the volume of tubes about curves in a Sasakian space forms. Annali Mat. Pura App. 1987, 148, 41–49. [Google Scholar] [CrossRef]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Rovenski, V.; Patra, D.S. Characteristics of Sasakian Manifolds Admitting Almost ∗-Ricci Solitons. Fractal Fract. 2023, 7, 156. https://doi.org/10.3390/fractalfract7020156

AMA Style

Rovenski V, Patra DS. Characteristics of Sasakian Manifolds Admitting Almost ∗-Ricci Solitons. Fractal and Fractional. 2023; 7(2):156. https://doi.org/10.3390/fractalfract7020156

Chicago/Turabian Style

Rovenski, Vladimir, and Dhriti Sundar Patra. 2023. "Characteristics of Sasakian Manifolds Admitting Almost ∗-Ricci Solitons" Fractal and Fractional 7, no. 2: 156. https://doi.org/10.3390/fractalfract7020156

Article Metrics

Back to TopTop