Next Article in Journal
Influence of Homo- and Hetero-Junctions on the Propagation Characteristics of Radially Propagated Cylindrical Surface Acoustic Waves in a Piezoelectric Semiconductor Semi-Infinite Medium
Next Article in Special Issue
Correction: Alohali et al. Ricci Vector Fields Revisited. Mathematics 2024, 12, 144
Previous Article in Journal
An Approach to Multidimensional Discrete Generating Series
Previous Article in Special Issue
Tensor Decompositions and Their Properties
 
 
Correction published on 7 February 2024, see Mathematics 2024, 12(4), 512.
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Ricci Vector Fields Revisited

by
Hanan Alohali
1,†,
Sharief Deshmukh
1,† and
Gabriel-Eduard Vîlcu
2,*,†
1
Department of Mathematics, College of Science, King Saud University, P.O. Box 2455, Riyadh 11451, Saudi Arabia
2
Department of Mathematics and Informatics, National University of Science and Technology Politehnica Bucharest, 313 Splaiul Independenţei, 060042 Bucharest, Romania
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Mathematics 2024, 12(1), 144; https://doi.org/10.3390/math12010144
Submission received: 25 November 2023 / Accepted: 29 December 2023 / Published: 1 January 2024 / Corrected: 7 February 2024
(This article belongs to the Special Issue Special (Pseudo-) Riemannian Manifolds)

Abstract

:
We continue studying the σ -Ricci vector field u on a Riemannian manifold ( N m , g ) , which is not necessarily closed. A Riemannian manifold with Ricci operator T, a Coddazi-type tensor, is called a T-manifold. In the first result of this paper, we show that a complete and simply connected T-manifold ( N m , g ) , m > 1 , of positive scalar curvature τ , admits a closed σ -Ricci vector field u such that the vector u σ is an eigenvector of T with eigenvalue τ m 1 , if and only if it is isometric to the m-sphere S α m . In the second result, we show that if a compact and connected T-manifold ( N m , g ) , m > 2 , admits a σ -Ricci vector field u with σ 0 and is an eigenvector of a rough Laplace operator with the integral of the Ricci curvature R i c u , u that has a suitable lower bound, then ( N m , g ) is isometric to the m-sphere S α m , and the converse also holds. Finally, we show that a compact and connected Riemannian manifold ( N m , g ) admits a σ -Ricci vector field u with σ as a nontrivial solution of the static perfect fluid equation, and the integral of the Ricci curvature R i c u , u has a lower bound depending on a positive constant α , if and only if ( N m , g ) is isometric to the m-sphere S α m .

1. Introduction

In a recent paper, (cf. [1]), a σ -Ricci vector field (abbreviated as σ -RVF) u on a m-Riemannian manifold N m , g is introduced, being defined by
1 2 £ u g = σ R i c ,
where £ u g is the Lie derivative of the metric g with respect to u , σ is a smooth function and R i c is the Ricci tensor of N m , g . A σ -RVF is a generalization of conformal vector fields (known for their utility in studying geometry and relativity), on Einstein manifolds (see [1,2,3,4,5,6,7,8,9,10,11]). Moreover, it represents a Killing vector field, which is known to have a great influence on the geometry as well as topology on which it lives (see [12,13,14,15]). Apart from these generalizations, a σ -RVF is a particular form of potential field of generalized solitons considered in [16,17,18]. Note that a 1-RVF  u on a m-Riemannian manifold N m , g is a stable Ricci soliton N m , g , u , 0 (see [19]). Indeed, in [1], it has been observed that a σ -RVF on N m , g is a stable solution of the generalized Ricci flow (or a σ -Ricci flow),
t g = 2 σ R i c , g 0 = g ,
of the form g t = ρ ( t ) φ t * g , where φ t : N m N m is a 1-parameter family of diffeomorphisms generated by the vector fields U t and ρ ( t ) is a scale factor, under the initial conditions ρ ( 0 ) = 1 , ρ . 0 = 0 , U 0 = u and φ 0 = i d .
In [1], a closed σ -RVF u , with σ 0 , on a compact and connected m-Riemannian manifold N m , g , m > 2 , of nonzero scalar curvature is used with an appropriate lower bound on the integral of the Ricci curvature R i c u , u to find a characterization of the m-sphere S m ( c ) . Moreover, in [1], a closed σ -RVF u on a complete and simply connected m-Riemannian manifold N m , g , m > 2 , of positive scalar curvature, is used, where the function σ is a nontrivial solution of the Fischer–Marsden equation (cf. [20]) with an appropriate upper bound on the length u of the covariant derivative of u , to find another characterization of the sphere S m ( c ) .
The Ricci operator T of a Riemannian manifold N m , g is a symmetric operator defined by
R i c E , F = g T E , F , E , F Γ ( N m ) ,
where Γ ( N m ) is a space of vector fields on N m . A Riemannian manifold N m , g is said to be a T-manifold, if the Ricci operator T is a Codazzi tensor, i.e., it satisfies
D E T ( F ) = D F T ( E ) , E , F Γ ( N m ) ,
where D is the Riemannian connection on N m , g . It is worth noting that a T-manifold N m , g has a constant scalar curvature.
In this article, we are interested in studying the geometry of N m , g equipped with a σ -RVF u . In the first result, we consider a T-manifold N m , g that possesses a closed σ -RVF u and we observe that, in this case, the vector field u σ has a special role to play in shaping the geometry of the T-manifold N m , g . It is shown that if the scalar curvature τ of a compact T-manifold N m , g is positive (note that τ is a constant for a T-manifold) and the vector field u σ satisfies
T u σ = τ m u σ ,
then N m , g is isometric to the m-sphere S c m of constant curvature c, where τ = m ( m 1 ) c , and the converse also holds (cf. Theorem 1).
Then, we concentrate on a σ -RVF u on N m , g that is not necessarily closed. In this case, the 1-form β dual to u gives rise to a skew symmetric operator Ψ : Γ ( N m ) Γ ( N m ) defined by
g Ψ E , F = 1 2 d β E , F , E , F Γ ( N m ) ,
and we call the operator Ψ the associated operator of the σ -RVF u . In the second result of this paper, we consider a compact and connected T-manifold N m , g with scalar curvature τ = m ( m 1 ) c that possesses a σ -RVF u , σ 0 , with associated operator Ψ satisfying
Δ u = c u , N m R i c u , u N m m 1 m τ 2 σ 2 + Ψ 2 ,
which necessarily implies that N m , g is isometric to the m-sphere S c m of constant curvature c, and the converse is also true (cf. Theorem 2), where Δ is the rough Laplace operator acting on vector fields on N m , g .
Recall the differential equation on a Riemannian manifold N m , g considered by Obata (cf. [18,21]), namely
H e s s ( σ ) = c σ g ,
where σ is a non-constant smooth function, c is a positive constant and H e s s ( σ ) is the Hessian of σ defined by
H e s s ( σ ) ( E , F ) = g D E σ , F , E , F Γ ( N m ) .
It is known that a complete, simply connected N m , g admits a nontrivial solution of (4) if and only if N m , g is isometric to the sphere S c m (cf. [18,21]).
There is yet another important differential equation on a Riemannian manifold N m , g (cf. [7] and references therein), given by
σ R i c H e s s ( σ ) = 1 m τ σ Δ σ g ,
known as the static fluid equation, where Δ σ is the Laplacian of σ with respect to the metric g. A Riemannian manifold N m , g that admits a nontrivial solution of the static fluid equation is called a static space. Note that under the additional assumption
Δ σ = τ m 1 σ ,
the static fluid equation reduces to the Fischer–Marsden equation (cf. [20])
Δ σ g + σ R i c = H e s s σ .
In the last result of this paper, we show that a compact and connected Riemannian manifold N m , g with scalar curvature τ possessing a σ -RVF u with associated operator Ψ and the function σ is a nontrivial solution of the static perfect fluid Equation (5); furthermore, for a positive constant c, the following inequality holds:
N m R i c u , u N m m 1 m τ 2 σ 2 + 1 m Δ σ + n c σ 2 + Ψ 2 ,
which necessarily implies that N m , g is isometric to the sphere S c m , and the converse is also true (cf. Theorem 3).

2. Preliminaries

For a σ -RVF u on an m-dimensional Riemannian manifold N m , g , we let β be the 1-form dual to u , i.e.,
β E = g u , E , E Γ N m .
Then, we have the associated operator Ψ satisfying
d β E , F = 1 2 g Ψ E , F , E , F Γ N m ,
which shows that Ψ is a skew symmetric operator. Using Equations (1) and (8), we obtain the following expression for the covariant derivative E u
D E u = σ T E + Ψ E , E Γ N m .
where T is the Ricci operator defined by
R i c E , F = g T E , F , E , F Γ N m .
On employing the following expression for the curvature tensor field R of N m , g ,
R E , F G = D E , D F G D [ E , F ] G , E , F , G Γ N m ,
with Equation (9), we obtain
R E , F u = E σ T F F σ T E + ρ D E T ( F ) D F T ( E ) + D E Ψ ( F ) D F Ψ ( E ) ,
for any E , F Γ N m , where
D E T ( F ) = D E T F T D E F .
The scalar curvature τ of N m , g is given by
τ = α = 1 m g T E α , E α ,
where E 1 , , E m is a local frame on N m . The Ricci tensor is given by
R i c E , F = α = 1 m g R E α , E F , E α ,
and employing it in Equation (10), we conclude
R i c F , u = R i c F , σ τ F σ + σ g F , α = 1 m F α T ( F α ) ρ Y τ g F , α = 1 m F α Ψ ( F α ) ,
where σ is the gradient of σ and we have used the symmetry of the Ricci operator T and the skew symmetry of the associated operator Ψ . It is known that the gradient of scalar curvature τ satisfies (cf. [22])
1 2 τ = α = 1 m D F α T ( F α ) .
Thus, on using Equation (12) in (11), we arrive at
R i c F , u = R i c F , σ τ F σ 1 2 σ F τ g F , α = 1 m F α Ψ ( F α )
and, therefore,
T u = T σ τ σ 1 2 ρ τ α = 1 m F α Ψ ( F α ) .
Lemma 1. 
For a σ-RVF u on a T-manifold N m , g , the associated operator Ψ satisfies
D E Ψ ( F ) = R E , u F R i c E , F σ + F σ T E , E , F Γ ( N m ) .
Proof. 
Suppose that u is a σ -RVF on a T-manifold N m , g . Then, Equation (10) changes to
D E Ψ ( F ) D F Ψ ( E ) = R E , F u E σ T F + F σ T E .
Now, using the fact that the 2-form d β in Equation (8) is closed and the associated operator Ψ is skew symmetric, we have
g D E Ψ ( F ) D F Ψ ( E ) , G + g D G Ψ ( E ) , F = 0
and employing Equation (15) in the above equation yields
g R E , F u E σ T F + F σ T E , G + g D G Ψ ( E ) , F = 0 .
Thus, we have
g D G Ψ ( E ) , F = g R G , u E , F + E σ g T G , F R i c E , G g σ , F
and this proves the lemma. □
On a Riemannian manifold N m , g possessing a σ -RVF u , we have the second-order differential operator 2 u defined by
2 u E , F = D E D F u D D E F u , E , F Γ ( N m )
and its trace
Δ u = α = 1 m 2 u E α , E α
is the rough Laplacian of the σ -RVF u .
Lemma 2. 
On a connected T-manifold N m , g , the scalar curvature τ is a constant, and for a σ-RVF u on a connected T-manifold N m , g with associated operator Ψ, the rough Laplacian satisfies
Δ u = T σ + α = 1 m D E α Ψ E α ,
where E 1 , , E m is a local frame on N m .
Proof. 
First, note that for a T-manifold N m , g , using Equation (3), we have
E ( τ ) = E α = 1 m g T E α , E α = α = 1 m g D E T E α + T D E E α , E α + α = 1 m g T E α , D E E α = α = 1 m g D E α T E , E α + 2 α = 1 m g T E α , D E E α = α = 1 m g E , D E α T E α + 2 α = 1 m g T E α , D E E α .
Note that
D E E α = k α k E E k , T E α = j μ α j E j ,
where the connection forms α k are skew symmetric and coefficients μ α j are symmetric and, as such, we have
α = 1 m g T E α , D E E α = 0 .
Consequently, Equation (17) yields
τ = α = 1 m D E α T E α .
Combining it with Equation (11), we obtain τ = 0 , i.e., the scalar curvature τ of a T-manifold is a constant.
Employing Equation (9), we have
2 u E , F = E ( σ ) T ( F ) + σ D E T ( F ) + D E Ψ ( F )
and taking the trace in the above equation, while using Equation (11) with τ = 0 , we obtain
Δ u = T σ + α = 1 m D E α Ψ E α .
Next, the sphere S α m of constant curvature α possesses a σ -RVF induced by a coordinate unit vector field u on the Euclidean space R m + 1 . Indeed, on treating S α m as an embedded surface in R m + 1 with unit normal ζ and Weingarten operator— α I , we express u as
u = u + f ζ , f = u , ζ ,
where , is a Euclidean inner product and u Γ S α m . On taking g as the induced metric on S α m and D as the Riemannian connection with respect to g and differentiating the above equation with respect to the vector field E Γ S α m , we have
D E u = α f E , f = α u .
Using the first equation in (19), it follows that
£ u g = 2 α f g
and for the Ricci tensor of S α m , we have
R i c = ( m 1 ) α g , τ = m ( m 1 ) α .
Hence, the vector field u on S α m obeys
1 2 £ u g = σ R i c , σ = 1 ( m 1 ) α f ,
i.e., u is a σ -RVF on S α m .
Moreover, note that Equation (21) in view of Equation (19) confirms
H e s s ( σ ) E , F = g D E σ , F = 1 ( m 1 ) α g D E f , F = 1 m 1 g D E u , F = α f m = 1 g E , F ,
i.e.,
H e s s ( σ ) = α σ g , Δ σ = m α σ .
Combining Equations (20) and (22), we see that the function σ of the σ -RVF u on S α m satisfies the static fluid equation
σ R i c H e s s ( σ ) = 1 m τ σ Δ σ g .
We investigate now whether σ is a nontrivial solution. If σ was a constant, by virtue of Equation (21), it would mean that f was a constant, and, in turn, by (19), it would mean that u = 0 and, by the same equation, would imply f = 0 . Inserting this information in (18), we have u = 0 , a contradiction. Hence, σ is a nontrivial solution of the static fluid equation on S α m .

3. σ -Ricci Vector Fields on T -Manifolds

In this section, we consider an m-dimensional T-manifold ( N m , g ) that possesses a closed σ -RVF u . It is interesting to observe that, in this situation, the vector field u σ plays an interesting role while treating the Ricci operator T of ( N m , g ) . Note that, by Lemma 2, the scalar curvature τ of a T-manifold ( N m , g ) is a constant and we put τ = m ( m 1 ) α , for a constant α . Here, we prove the following result.
Theorem 1. 
An m-dimensional, m > 1 , complete, and simply connected T-manifold ( N m , g ) with positive scalar curvature τ admits a nonzero closed σ-RVF u , σ 0 satisfying
T u σ = τ m u σ ,
if and only if ( N m , g ) is isometric to S α m , where τ = m ( m 1 ) α .
Proof. 
Suppose that the complete and simply connected T-manifold ( N m , g ) , m > 1 , of scalar curvature τ > 0 , admits a nonzero closed σ -RVF u , σ 0 , which satisfies
T u σ = τ m u σ .
As the σ -RVF u is closed, its associated operator Ψ = 0 , and by Lemma 2, the scalar curvature τ is a constant, and Equation (14) becomes
T u = T σ τ σ .
Treating it with Equation (24) yields
τ m u σ = τ σ
and, as τ > 0 , it transforms into
u = ( m 1 ) σ .
Note that, by Equation (9), we have div u = σ τ , and taking the divergence in Equation (26) gives
σ τ = ( m 1 ) Δ σ .
Now, inserting the value of σ from Equation (26) into Equation (25), we arrive at
T u = m 1 m τ σ .
Note that as u is closed, Equation (9) has the form
D E u = σ T E , E Γ ( N m ) .
Next, we intend to compute the divergence div T u and we proceed by choosing a local frame E 1 , , E m and using Equation (29)
div T u = α = 1 m g E α T u , E α = α = 1 m g E α T ( u ) + T E α u , E α = α = 1 m g u , E α T E α + α = 1 m g E α u , T E α .
Note that on T-manifold ( N m , g ) , by Lemma 2, τ is a constant and, thus, employing Equations (12) and (29), we arrive at
div T u = σ T 2 .
Now, utilizing this equation in Equation (28) yields
σ T 2 = m 1 m τ Δ σ .
Inserting Equation (27) in the above equation gives
σ T 2 = 1 m σ τ 2 ,
i.e.,
σ T 2 1 m τ 2 = 0 .
As N k is connected (being simply connected) and σ 0 , in this situation, the above equation yields
T 2 = 1 m τ 2 .
However, Equation (31) is the equality in Schwartz’s inequality
T 2 1 m τ 2 .
Hence, equality (31) holds if and only if
T = τ m I
and Equation (29) changes to
D E u = τ m ρ E , E Γ ( N m ) .
Thus, on employing Equation (26) in the above equation, we confirm
D E σ = τ m ( m 1 ) σ E , E Γ N m .
Note that as u 0 by Equation (26), the function σ is a non-constant function and, also, τ being a positive constant, letting τ = m ( m 1 ) α , we obtain a positive constant α and Equation (32) is Obata’s equation
H e s s ( σ ) = α ρ g ,
proving that ( N m , g ) is isometric to the sphere S α m (cf. [18,21]).
Conversely, suppose that ( N m , g ) is isometric to the sphere S α m . Then, by Equations (19)–(21), there is a nonzero σ -RVF u on S α m and, as seen earlier, the function σ 0 and is a non-constant function. Moreover, the Ricci operator of S α m being
T = τ m I ,
the condition
T u σ = τ m u σ
holds, and this finishes the proof. □
In an earlier result, we considered a closed σ -RVF u on an m-dimensional T-manifold ( N m , g ) to find a characterization of the sphere S α m . Next, we consider a σ -RVF u on an m-dimensional T-manifold ( N m , g ) not necessarily closed and prove the following.
Theorem 2. 
An m-dimensional compact and connected T-manifold N m , g , m > 2 of positive scalar curvature τ admits a σ-RVF u with associated operator Ψ, σ 0 , Δ u = τ m ( m 1 ) u and the Ricci curvature R i c u , u satisfies
N m R i c u , u N m m 1 m τ 2 σ 2 + Ψ 2
if and only if N m , g is isometric to S α m , where τ = m ( m 1 ) α .
Proof. 
Let an m-dimensional T-manifold ( N m , g ) , m > 2 , with scalar curvature τ > 0 be equipped with a σ -RVF u with σ 0 and associated operator Ψ such that
Δ u = τ m ( m 1 ) u
and
N m R i c u , u N m m 1 m τ 2 σ 2 + Ψ 2 .
Using Lemma 1, we have
R E , u F = D E Ψ ( F ) + R i c E , F σ F σ T E , E , F Γ ( N m )
Employing a local frame E 1 , , E m in the above equation, we conclude
R i c u , F = R i c σ , F τ F σ α = 1 m g F , D E α Ψ ( E α ) , F Γ ( N m )
and the above equation implies
R i c u , u = R i c σ , u τ u σ α = 1 m g u , D E α Ψ ( E α ) .
Note that, by Equation (9), we have
div u = τ σ
and using
div σ u = u σ + τ σ 2 ,
in the above equation containing the expression of R i c u , u , we derive
R i c u , u = R i c σ , u + τ 2 σ 2 τ div σ u α = 1 m g u , D E α Ψ ( E α ) .
Next, using a local frame E 1 , , E m on ( N m , g ) , to compute the div Ψ u , we have, on using the skew symmetry of the associated operator Ψ and Equation (9),
div Ψ u = α = 1 m g D E α Ψ u , E α = α = 1 m g D E α Ψ ( u ) + Ψ σ T E α + Ψ E α , E α = α = 1 m g u , D E α Ψ E α σ α = 1 m g T E α , Ψ E α Ψ 2
Since T is symmetric and the associated operator Ψ is skew symmetric, it follows that
α = 1 m g T E α , Ψ E α = 0
and Equation (36) now becomes
div Ψ u = α = 1 m g u , D E α Ψ E α Ψ 2
and, inserting this equation into Equation (35), we arrive at
R i c u , u = R i c σ , u + τ 2 σ 2 τ div σ u + Ψ 2 + div Ψ u .
Note that on a T-manifold ( N m , g ) , τ is a constant and keeping this in mind and integrating the above equation brings us to
N m R i c u , u R i c σ , u τ 2 σ 2 Ψ 2 = 0 .
Observe that, by virtue of the symmetry of the operator T and Equations (9), (12) and (37), and the fact that τ is a constant, we have
div T u = α = 1 m g D E α T u , E α = α = 1 m g D E α T ( u ) + T σ T E α + Ψ E α , E α = σ T 2 .
Now, using the fact that
div σ T u = R i c σ , u + σ div T u
in Equation (39), we arrive at
R i c σ , u = div σ T u σ 2 T 2 .
Inserting the above equation in Equation (38), we confirm
N m R i c u , u + σ 2 T 2 τ 2 σ 2 Ψ 2 = 0
and the above integral could be rearranged as
N m σ 2 T 2 1 m τ 2 = N m m 1 m τ 2 σ 2 + Ψ 2 N m R i c u , u .
Treating the above equation with the inequality (34), we arrive at
N m σ 2 T 2 1 m τ 2 0 .
The integrand in the above inequality by virtue of Schwartz’s inequality is non-negative, and, therefore, we conclude
σ 2 T 2 1 m τ 2 = 0 .
As σ 0 and N m is connected, we conclude that
T 2 = 1 m τ 2 ,
which, being the equality in Schwartz’s inequality, it holds if and only if
T = τ m I .
Consequently, as τ is a constant, Equations (14) and (41) combine to arrive at
τ m u = τ m σ τ σ α = 1 m F α Ψ ( F α ) ,
for a local frame E 1 , , E m on ( N m , g ) , i.e., we have
τ m u = m 1 m τ σ α = 1 m F α Ψ ( F α ) .
Moreover, using Equations (33) and (41) with Lemma 2, we obtain the following:
τ m ( m 1 ) u = τ m σ + α = 1 m F α Ψ ( F α ) .
Adding Equations (42) and (43), we find
m 2 m m 1 τ u = m 2 m τ σ
and, as m > 2 , τ > 0 , it confirms
u = ( m 1 ) σ .
Differentiating the above equation and using Equations (9) and (41), we have
D E σ = 1 m 1 τ m σ E + Ψ E , E Γ ( N m ) ,
which, on taking the inner product with E and noticing that Ψ is a skew symmetric operator, leads to
H e s s ( σ ) E , E = α σ g E , E , E Γ ( N m ) ,
where τ = m ( m 1 ) α , i.e., α is a positive constant. Now, polarizing the above equation confirms
H e s s ( σ ) = α σ g .
Hence, N m , g is isometric to S α m (cf. [18,21]).
Conversely, suppose that N m , g is isometric to S α m . Then, by Equation (21), there is a nonzero σ -RVF u on S α m with σ 0 and, as u is is closed, the associated operator Ψ = 0 . Moreover, it is obvious that S α m is a T-manifold. Thus, using Equation (19), we have
2 u E , F = D E D F u D D E F u = α E f F
and, therefore, by treating the above equation with (16), we have
Δ u = α = 1 m 2 u E α , E α = α f ,
which, by virtue of Equation (19), implies
Δ u = α u = τ m ( m 1 ) u ,
where τ = m ( m 1 ) α . Finally, using Equations (19) and (21), we have
σ = 1 ( m 1 ) α f = 1 m 1 u ,
i.e.,
u 2 = ( m 1 ) 2 σ 2 .
Now, Equation (22) implies
σ Δ σ = m α σ 2 ,
which, on integrating by parts, confirms
S α m σ 2 = m α S α m σ 2 = 1 m ( m 1 ) 2 α S α m τ 2 σ 2 .
The Ricci curvature of S α m is
R i c u , u = ( m 1 ) α u 2
and, thus, using Ψ = 0 and Equations (44) and (45), we conclude
S α m R i c u , u = S α m ( m 1 ) α u 2 = S α m ( m 1 ) 3 α σ 2 = S α m m 1 m τ 2 σ 2 + Ψ 2
and this completes the proof. □

4. σ -Ricci Vector Fields on Static Spaces

Now, we are interested in a σ -RVF u , not necessarily closed, on a Riemannian manifold N m , g with function σ as a nontrivial solution of the static fluid Equation (5). Indeed, we prove the following.
Theorem 3. 
If an m-dimensional compact and connected Riemannian manifold N m , g admits a σ-RVF u with associated operator Ψ, such that σ is a nontrivial solution of the static perfect fluid equation, for a positive constant α and the Ricci curvature R i c u , u , it satisfies
N m R i c u , u N m m 1 m τ 2 σ 2 + 1 m Δ σ + m α σ 2 + Ψ 2 ,
and N m , g is isometric to S α m , and the converse also holds.
Proof. 
Assume that N m , g admits a σ -RVF u with associated operator Ψ , such that σ is a nontrivial solution of the static perfect fluid Equation (5) and the Ricci curvature R i c u , u satisfies
N m R i c u , u N m m 1 m τ 2 σ 2 + 1 m Δ σ + m α σ 2 + Ψ 2 .
Then, the Hessian operator H σ of the function σ defined by
g H σ E , F = H e s s ( σ ) E , F ,
by virtue of Equation (5) satisfies
H σ E = σ T E + 1 m Δ σ τ σ E , E Γ N m
Utilizing Equation (9) in the above equation, we arrive at
H σ E = D E u Ψ E + 1 m Δ σ τ σ E
and, for a positive constant α , the above equation could be rearranged as
H σ + α σ I ( E ) = D E u Ψ E + 1 m Δ σ + m α σ τ σ E , E Γ N m .
Choosing a local frame E 1 , , E m , and using the above equation, we compute
H σ + α σ I 2 = j = 1 m g H σ + α σ I ( E j ) , H σ + α σ I ( E j ) = j = 1 m g D E j u Ψ E j + 1 m Δ σ + m α σ τ σ E j , D E j u Ψ E j + 1 m Δ σ + m α σ τ σ E j = D u 2 + Ψ 2 + 1 m Δ σ + m α σ τ σ 2 2 j = 1 m g D E j u , Ψ E j + 2 m Δ σ + m α σ τ σ div u .
Now, using Equation (9) and div u = τ σ in the above equation, we arrive at
H σ + α σ I 2 = D u 2 Ψ 2 + 1 m Δ σ + m α σ τ σ 2 + 2 τ σ m Δ σ + m α σ τ σ ,
i.e.,
H σ + α σ I 2 = D u 2 Ψ 2 + 1 m Δ σ + m α σ τ σ Δ σ + m α σ + τ σ
or
H σ + α σ I 2 = D u 2 Ψ 2 + 1 m Δ σ + m α σ 2 1 m τ σ 2 .
We recall the integral formula (cf. [23])
N m D u 2 = N m R i c u , u + 1 2 £ u g 2 div u 2 .
Using div u = τ σ and the outcome of Equation (1) in the form
1 2 £ u g 2 = 2 σ 2 T 2
in the above integral equation, we have
N m D u 2 = N m R i c u , u + 2 σ 2 T 2 τ σ 2 .
Now, integrating Equation (48) and using the above equation, we arrive at
N m H σ + α σ I 2 = N m R i c u , u + 2 σ 2 T 2 m + 1 m τ σ 2 Ψ 2 + 1 m Δ σ + m α σ 2
Notice that
2 σ 2 T 2 m + 1 m τ σ 2 = 2 σ 2 T 2 1 m τ 2 m 1 m τ σ 2
and, by Equation (5), we have that
σ T E 1 m τ σ E = H σ E 1 m Δ σ E ,
which implies
σ 2 T τ m I 2 = H σ 1 m Δ σ I 2 .
Combining it with Equation (50), we arrive at
2 σ 2 T 2 m + 1 m τ σ 2 = 2 H σ 1 m Δ σ I 2 m 1 m τ σ 2 .
Moreover, we have
H σ 1 m Δ σ I 2 = H σ 2 + 1 m Δ σ 2 2 m Δ σ j = 1 m g H σ E j , E j = H σ 2 1 m Δ σ 2
Similarly, we have
H σ + α σ I 2 = H σ 2 + 2 α σ Δ σ + m α 2 σ 2
and utilizing it in Equation (52), we obtain
H σ 1 m Δ σ I 2 = H σ + α σ I 2 2 α σ Δ σ m α 2 σ 2 1 m Δ σ 2 ,
i.e.,
H σ 1 m Δ σ I 2 = H σ + α σ I 2 1 m Δ σ + m α σ 2 .
Thus, in view of the above equation, (51) assumes the form
2 σ 2 T 2 m + 1 m τ σ 2 = 2 H σ + α σ I 2 2 m Δ σ + m α σ 2 m 1 m τ σ 2 .
Now, inserting this value in Equation (49), we arrive at
N m H σ + α σ I 2 = N m R i c u , u + 2 H σ + α σ I 2 m 1 m τ σ 2 Ψ 2 1 m Δ σ + m α σ 2 ,
i.e.,
N m H σ + α σ I 2 = N m m 1 m τ σ 2 + 1 m Δ σ + m α σ 2 + Ψ 2 N m R i c u , u .
Using inequality (47) in the above equation, we conclude
N m H σ + α σ I 2 0 ,
which proves
H e s s ( σ ) = α σ g ,
where α > 0 is a constant and σ , being a nontrivial solution of a static perfect fluid, is a non-constant function. Hence, N m , g is isometric to S α m (cf. [18,21]).
The converse is trivial, because, by Equation (21), S α m admits a σ -RVF u , and by Equation (23) and the paragraph that follows (23), σ is a nontrivial solution of the static perfect fluid equation. Moreover, we have, by Equation (22), that
Δ σ + m α σ = 0
and, by Equation (46), we have
N m R i c u , u = N m m 1 m τ 2 σ 2 + 1 m Δ σ + m α σ 2 + Ψ 2
This finishes the proof. □

Author Contributions

Conceptualization, H.A., S.D. and G.-E.V.; methodology, H.A., S.D. and G.-E.V.; software, H.A., S.D. and G.-E.V.; validation, H.A., S.D. and G.-E.V.; formal analysis, H.A., S.D. and G.-E.V.; investigation, H.A., S.D. and G.-E.V.; resources, H.A., S.D. and G.-E.V.; data curation, H.A., S.D. and G.-E.V.; writing—original draft preparation, H.A., S.D. and G.-E.V.; writing—review and editing, H.A., S.D. and G.-E.V.; visualization, H.A., S.D. and G.-E.V.; supervision, H.A., S.D. and G.-E.V.; project administration, H.A., S.D. and G.-E.V.; funding acquisition, H.A. All authors have read and agreed to the published version of the manuscript.

Funding

The authors would like to extend their sincere appreciation to Supporting Project Number (RSPD2023R860), King Saud University, Riyadh, Saudi Arabia.

Data Availability Statement

Data are contained within the article.

Acknowledgments

The authors would like to extend their sincere appreciations to Supporting project number (RSPD2024R860) King Saud University, Riyadh, Saudi Arabia.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Alohali, H.; Deshmukh, S. Ricci vector fields. Mathematics 2023, 11, 4622. [Google Scholar] [CrossRef]
  2. Chen, B.-Y. Some results on concircular vector fields and their applications to Ricci solitons. Bull. Korean Math. Soc. 2015, 52, 1535–1547. [Google Scholar] [CrossRef]
  3. Deshmukh, S.; Turki, N.B.; Vîlcu, G.-E. A note on static spaces. Results Phys. 2021, 27, 104519. [Google Scholar] [CrossRef]
  4. Falcitelli, M.; Sarkar, A.; Halder, S. Conformal vector fields and conformal Ricci solitons on α-Kenmotsu manifolds. Mediterr. J. Math. 2023, 20, 127. [Google Scholar] [CrossRef]
  5. Hwang, S.; Yun, G. Conformal vector fields and their applications to Einstein-type manifolds. Results Math. 2024, 79, 45. [Google Scholar] [CrossRef]
  6. Khan, S.; Bukhari, M.; Alkhaldi, A.; Ali, A. Conformal vector fields of Bianchi type-I spacetimes. Mod. Phys. Lett. A 2021, 36, 2150254. [Google Scholar] [CrossRef]
  7. Narmanov, A.; Rajabov, E. On the geometry of orbits of conformal vector fields. J. Geom. Symmetry Phys. 2019, 51, 29–39. [Google Scholar] [CrossRef]
  8. Poddar, R.; Balasubramanian, S.; Sharma, R. Quasi-Einstein manifolds admitting conformal vector fields. Colloq. Math. 2023, 174, 81–87. [Google Scholar] [CrossRef]
  9. Sharma, R. Gradient Ricci solitons with a conformal vector field. J. Geom. 2018, 109, 33. [Google Scholar] [CrossRef]
  10. Sharma, R. Conformal flatness and conformal vector fields on umbilically synchronized space-times. Acta Phys. Pol. B 2023, 54, A3. [Google Scholar] [CrossRef]
  11. Filho, J.F.S. Quasi-Einstein manifolds admitting a closed conformal vector field. Differ. Geom. Appl. 2024, 92, 102083. [Google Scholar] [CrossRef]
  12. Nikonorov, Y. Spectral properties of Killing vector fields of constant length. J. Geom. Phys. 2019, 145, 103485. [Google Scholar] [CrossRef]
  13. Lynge, W.C. Sufficient conditions for periodicity of a Killing vector field. Proc. Am. Math. Soc. 1973, 38, 614–616. [Google Scholar] [CrossRef]
  14. Rong, X. Positive curvature, local and global symmetry, and fundamental groups. Am. J. Math. 1999, 121, 931–943. [Google Scholar] [CrossRef]
  15. Yorozu, S. Killing vector fields on complete Riemannian manifolds. Proc. Am. Math. Soc. 1982, 84, 115–120. [Google Scholar] [CrossRef]
  16. Blaga, A.M.; Chen, B.-Y. Harmonic forms and generalized solitons. Results Math. 2024, 79, 16. [Google Scholar] [CrossRef]
  17. Lee, K.-H. Stability and moduli space of generalized Ricci solitons. Nonlinear Anal. 2024, 240, 113458. [Google Scholar] [CrossRef]
  18. Obata, M. The conjectures about conformal transformations. J. Differ. Geom. 1971, 6, 247–258. [Google Scholar] [CrossRef]
  19. Chow, B.; Lu, P.; Ni, L. Hamilton’s Ricci Flow; Graduate Studies in Mathematics; American Mathematical Society: Providence, RI, USA, 2006; Volume 77. [Google Scholar]
  20. Fischer, A.E.; Marsden, J.E. Manifolds of Riemannian metrics with prescribed scalar curvature. Bull. Am. Math. Soc. 1974, 80, 479–484. [Google Scholar] [CrossRef]
  21. Obata, M. Conformal transformations of Riemannian manifolds. J. Differ. Geom. 1970, 4, 311–333. [Google Scholar] [CrossRef]
  22. Besse, A.L. Einstein Manifolds; Springer: Berlin/Heidelberg, Germany, 1987. [Google Scholar]
  23. Yano, K. Integral Formulas in Riemannian Geometry; Marcel Dekker Inc.: New York, NY, USA, 1970. [Google Scholar]
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

Alohali, H.; Deshmukh, S.; Vîlcu, G.-E. Ricci Vector Fields Revisited. Mathematics 2024, 12, 144. https://doi.org/10.3390/math12010144

AMA Style

Alohali H, Deshmukh S, Vîlcu G-E. Ricci Vector Fields Revisited. Mathematics. 2024; 12(1):144. https://doi.org/10.3390/math12010144

Chicago/Turabian Style

Alohali, Hanan, Sharief Deshmukh, and Gabriel-Eduard Vîlcu. 2024. "Ricci Vector Fields Revisited" Mathematics 12, no. 1: 144. https://doi.org/10.3390/math12010144

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop