Research article

Some new characterizations of spheres and Euclidean spaces using conformal vector fields

  • Received: 09 July 2024 Revised: 20 September 2024 Accepted: 30 September 2024 Published: 11 October 2024
  • MSC : 53C21, 53C24

  • Given a conformal vector field X defined on an n-dimensional Riemannian manifold (Nn,g), naturally associated to X are the conformal factor σ, a smooth function defined on Nn, and a skew symmetric (1,1) tensor field Ω, called the associated tensor, that is defined using the 1-form dual to X. In this article, we prove two results. In the first result, we show that if an n-dimensional compact and connected Riemannian manifold (Nn,g), n>1, of positive Ricci curvature admits a nontrivial (non-Killing) conformal vector field X with conformal factor σ such that its Ricci operator Rc and scalar curvature τ satisfy

    Rc(X)=(n1)σ and X(τ)=2σ(n(n1)cτ)

    for a constant c, necessarily c>0 and (Nn,g) is isometric to the sphere Snc of constant curvature c. The converse is also shown to be true. In the second result, it is shown that an n-dimensional complete and connected Riemannian manifold (Nn,g), n>1, admits a nontrivial conformal vector field X with conformal factor σ and associated tensor Ω satisfying

    Rc(X)=divΩ and Ω(X)=0,

    if and only if (Nn,g) is isometric to the Euclidean space (En,,).

    Citation: Sharief Deshmukh, Mohammed Guediri. Some new characterizations of spheres and Euclidean spaces using conformal vector fields[J]. AIMS Mathematics, 2024, 9(10): 28765-28777. doi: 10.3934/math.20241395

    Related Papers:

    [1] Hanan Alohali, Sharief Deshmukh . Some generic hypersurfaces in a Euclidean space. AIMS Mathematics, 2024, 9(6): 15008-15023. doi: 10.3934/math.2024727
    [2] Yanlin Li, Nasser Bin Turki, Sharief Deshmukh, Olga Belova . Euclidean hypersurfaces isometric to spheres. AIMS Mathematics, 2024, 9(10): 28306-28319. doi: 10.3934/math.20241373
    [3] Mohammed Guediri, Sharief Deshmukh . Hypersurfaces in a Euclidean space with a Killing vector field. AIMS Mathematics, 2024, 9(1): 1899-1910. doi: 10.3934/math.2024093
    [4] Nasser Bin Turki, Sharief Deshmukh, Olga Belova . A note on closed vector fields. AIMS Mathematics, 2024, 9(1): 1509-1522. doi: 10.3934/math.2024074
    [5] Sharief Deshmukh, Mohammed Guediri . Characterizations of Euclidean spheres. AIMS Mathematics, 2021, 6(7): 7733-7740. doi: 10.3934/math.2021449
    [6] Sharief Deshmukh, Amira Ishan, Olga Belova . On eigenfunctions corresponding to first non-zero eigenvalue of the sphere $ S^n(c) $ on a Riemannian manifold. AIMS Mathematics, 2024, 9(12): 34272-34288. doi: 10.3934/math.20241633
    [7] Ibrahim Al-Dayel, Meraj Ali Khan . Ricci curvature of contact CR-warped product submanifolds in generalized Sasakian space forms admitting nearly Sasakian structure. AIMS Mathematics, 2021, 6(3): 2132-2151. doi: 10.3934/math.2021130
    [8] Ali H. Alkhaldi, Meraj Ali Khan, Shyamal Kumar Hui, Pradip Mandal . Ricci curvature of semi-slant warped product submanifolds in generalized complex space forms. AIMS Mathematics, 2022, 7(4): 7069-7092. doi: 10.3934/math.2022394
    [9] Ghodratallah Fasihi-Ramandi, Hajar Ghahremani-Gol . Some geometric vector fields on 5-dimensional 2-step homogeneous nilmanifolds. AIMS Mathematics, 2020, 5(1): 546-556. doi: 10.3934/math.2020036
    [10] Najmeddine Attia, Bilel Selmi . Different types of multifractal measures in separable metric spaces and their applications. AIMS Mathematics, 2023, 8(6): 12889-12921. doi: 10.3934/math.2023650
  • Given a conformal vector field X defined on an n-dimensional Riemannian manifold (Nn,g), naturally associated to X are the conformal factor σ, a smooth function defined on Nn, and a skew symmetric (1,1) tensor field Ω, called the associated tensor, that is defined using the 1-form dual to X. In this article, we prove two results. In the first result, we show that if an n-dimensional compact and connected Riemannian manifold (Nn,g), n>1, of positive Ricci curvature admits a nontrivial (non-Killing) conformal vector field X with conformal factor σ such that its Ricci operator Rc and scalar curvature τ satisfy

    Rc(X)=(n1)σ and X(τ)=2σ(n(n1)cτ)

    for a constant c, necessarily c>0 and (Nn,g) is isometric to the sphere Snc of constant curvature c. The converse is also shown to be true. In the second result, it is shown that an n-dimensional complete and connected Riemannian manifold (Nn,g), n>1, admits a nontrivial conformal vector field X with conformal factor σ and associated tensor Ω satisfying

    Rc(X)=divΩ and Ω(X)=0,

    if and only if (Nn,g) is isometric to the Euclidean space (En,,).



    Conformal geometry is one of the oldest branches of differential geometry, as one can follow it through [1] as old as 1959. It is evolving with time and getting enriched constantly till one can find the most recent work in [2]. The main topic in the conformal geometry is about studying the influence of a conformal vector field X on an n-dimensional Riemannian manifold (Nn,g). We shall abbreviate a conformal vector field X by CONFVF X for the sake of convenience. There is a smooth function σ that is naturally associated to a CONFVFX on (Nn,g) called the conformal factor satisfying

    12£Xg=σg, (1.1)

    where £ is the Lie derivative operator. A CONFVF X is said to be Killing if the conformal factor σ=0, and consequently, a nontrivial CONFVF X must have a conformal factor σ0. There is naturally associated a (1,1) skew symmetric tensor field Ω to a CONFVF X on (Nn,g) called the associated tensor of CONFVF X, defined by

    12dη(E,F)=g(ΩE,F), (1.2)

    for smooth vector fields E,F on Nn, where η is the 1-form dual to X. This associated tensor Ω plays a crucial role in studying the impact of a CONFVF X on the geometry of (Nn,g) (see [2]).

    The sphere Snc of constant curvature as a hypersurface of the Euclidean space (En+1,,), where , is the Euclidean metric, has unit normal ζ, induced metric g, and the Weingarten map

    S=cI.

    Choosing a unit constant vector field Z on the Euclidean space (En+1,,), its tangential component X to the sphere Snc satisfies

    EX=cρE and ρ=cX, (1.3)

    where

    ρ=Z,ζ

    and ρ is the gradient of ρ on Snc with respect to the induced metric g. Thus, we see that X is a CONFVF on the sphere Snc with conformal factor

    σ=cρ.

    This CONFVF X is closed, and therefore the associated tensor Ω=0.

    Also, using complex structure J on (E2n,,), define a unit vector field ξ=Jζ, which has covariant derivative

    Eξ=(JE)T, (1.4)

    where (JE)T is the tangential component of JE to S2n1 and ξ is a Killing vector field, that is,

    £ξg=0.

    Now, define a vector field

    ¯X=X+ξ,

    we obtain a CONFVF¯X with conformal factor σ that is not closed and indeed has an associated tensor

    ΩE=(JE)T.

    Non-closed CONFVF are in abundance, for instance on the Euclidean space (E2n,,), if ξ is the position vector field on E2n, then

    X=ξ+Jξ,

    where J is the complex structure, is a CONFVF on (E2n,,) that is not closed and has an associated tensor

    ΩE=JE.

    Riemannian manifolds that admit closed CONFVF have been extensively studied [3,4,5]. For related work on conformal vector fields that are not necessarily closed, refer to [6,7,8,9]. We see that while studying the impact of a CONFVF X on the geometry of the Riemannian manifold (Nn,g), on which it is defined, the associated tensor Ω offers some resistance in analyzing the geometry. Therefore, this study becomes smooth once one assumes that the CONFVF X is closed, which forces the associated tensor Ω to vanish. The study of the impact of the existence of a non-closed CONFVF on Riemannian manifolds is relatively difficult, and therefore this difficulty is softened by imposing geometric restrictions on Riemannian manifolds, such as the scalar curvature τ being constant. Riemannian manifolds admitting non-closed CONFVF have been studied in [8,10,11]. Moreover, apart from the fact that the presence of a CONFVF on a Riemannian manifold influences its geometry, they are also used in the theory of relativity [12,13,14].

    Observe that the Ricci operator Rc on a Riemannian manifold (Nn,g) is related to the Ricci tensor Ric by

    Ric(E,F)=g(RcE,F),

    and for the sphere Snc the Ricci operator is given by

    Rc=(n1)cI.

    Moreover, the conformal factor σ of the CONFVF X on Snc described in Eq (1.3) satisfies

    Rc(X)=(n1)σ. (1.5)

    It naturally raises a question: Under what conditions is a compact and connected Riemannian manifold (Nn,g) admitting a nontrivial CONFVFX with conformal factor σ satisfying Eq (1.5) is isometric to the sphere Snc?

    The reader may refer to the following sources, as well as the references in [10,15,16,17], for additional information on this question. In this article, we answer this question and indeed find a new characterization of the sphere Snc (see Theorem 1). Finally, in this paper, we find a characterization of the Euclidean space (En,,) using a nontrivial CONFVF X with conformal factor σ on a complete and connected Riemannian manifold (Nn,g) (see Theorem 2).

    Let X be a nontrivial CONFVF on an n-dimensional Riemannian manifold (Nn,g) with conformal factor σ. Then employing Eqs (1.1) and (1.2) in Koszul's formula, we have

    2(EX,F)=(£Xg)(E,F)+dη(E,F),

    where E is the covariant derivative with respect to the Riemannian connection on (Nn,g) and E,F are smooth vector fields on Nn. Consequently, we have

    EX=σE+ΩE, (2.1)

    where Ω is the associated tensor associated with CONFVF X.

    Using Eq (2.1), we find the following expression for the curvature tensor field R

    R(E,F)X=E(σ)FF(σ)E+(EΩ)(F)(FΩ)(E), (2.2)

    for smooth vector fields E,F on Nn. Choosing a local frame {E1,,En} on (Nn,g) in the above equation in order to compute the Ricci tensor Ric, we obtain

    Ric(E,X)=(n1)E(σ)g(E,divΩ), (2.3)

    where

    divΩ=j(EjΩ)(Ej).

    Thus, for the CONFVF X on an n-dimensional Riemannian manifold (Nn,g) with conformal factor σ, on using Eq (2.3), for Ricci operator Rc, we have

    Rc(X)=(n1)σdivΩ. (2.4)

    On a Riemannian manifold (Nn,g), the scalar curvature τ=Tr.Rc satisfies the following [18,19]

    12τ=divRc, (2.5)

    where

    divRc=j(EjRc)(Ej).

    We see that for the CONFVF X on an n-dimensional Riemannian manifold (Nn,g) with conformal factor σ, on using Eq (2.1), we have

    divX=nσ. (2.6)

    Also, we compute the divergence of the vector field Rc(X) as follows:

    divRc(X)=jg(EjRc(X),Ei)=j((EjRc)(X)+Rc(EjX),Ej),

    which, on using the symmetry of Rc and Eqs (2.1) and (2.5), yields

    divRc(X)=12X(τ)+jg(Rc(Ej),σEj+ΩEj),

    and Ω being skew symmetric, we reach at

    divRc(X)=12X(τ)+στ. (2.7)

    Similarly, on using Eq (2.1) and the skew symmetry of the associated operator Ω for the CONFVF X on an n-dimensional Riemannian manifold (Nn,g) with conformal factor σ, we find

    div(ΩX)=Ω2g(X,divΩ), (2.8)

    where

    Ω2=jg(ΩEj,ΩEj).

    Now, for the conformal factor σ of the CONFVF X on an n -dimensional Riemannian manifold (Nn,g), the Hessian Hess(σ) of σ is the symmetric bilinear form defined by

    Hess(σ)(E,F)=E(Fσ)(EF)(σ),

    and the Hessian operator Hσ of σ is defined by

    Hess(σ)(E,F)=g(HσE,F)

    for smooth vector fields E,F on Nn. The Laplacian Δσ of σ is defined by

    Δσ=div(σ),

    which is also given by

    Δσ=Tr.Hσ.

    If (Nn,g) is a compact Riemannian manifold, then we have the following Bochner's formula

    NnRic(σ,σ)=Nn((Δσ)2Hσ2), (2.9)

    where for a local frame {E1,..,En} on Nn

    Hσ2=jg(HσEj,HσEj).

    Let X be a nontrivial CONFVF on an n-dimensional Riemannian manifold (Nn,g) with conformal factor σ. In this section, we shall answer the questions raised in the introduction. Indeed, first we prove the following:

    Theorem 1. An n-dimensional compact and connected Riemannian manifold (Nn,g), n>1, of positive Ricci curvature and scalar curvature τ admits a nontrivial CONFVF X with conformal factor σ satisfying

    Rc(X)=(n1)σandX(τ)=2σ(n(n1)cτ),

    for a constant c, if and only if c>0 and (Nn,g) is isometric to the sphere Snc.

    Proof. Suppose X is a nontrivial CONFVF on an n-dimensional compact and connected Riemannian manifold (Nn,g), n>1, of positive Ricci curvature with conformal factor σ, which satisfies

    Rc(X)=(n1)σ (3.1)

    and

    X(τ)=2σ(n(n1)cτ), (3.2)

    where c is a constant. Using Eqs (2.4) and (3.1), we obtain divΩ=0 and inserting it in Eq (2.8) yields

    div(ΩX)=Ω2. (3.3)

    Now, in Eq (3.1), taking the inner product with X, provides

    Ric(X,X)=(n1)X(σ),

    which, in light of Eq (2.6) used in the formula

    div(σX)=X(σ)+σdivX=X(σ)+nσ2,

    takes the form

    Ric(X,X)=n(n1)σ2(n1)div(σX).

    Integrating the above equation leads to

    NnRic(X,X)=n(n1)Nnσ2. (3.4)

    Again, using Eq (3.1), we immediately have

    NnRic(X,σ)=(n1)Nnσ2. (3.5)

    Next, taking divergence on both sides of Eq (3.1) and making use of Eq (2.7), we get

    (n1)Δσ=12X(τ)+στ,

    which, on treating with Eq (3.2), reduces to

    Δσ=ncσ. (3.6)

    Multiplying the above equation by σ and then integrating leads to

    Nnσ2=ncNnσ2. (3.7)

    Also, we have

    Ric(σ+cX,σ+cX)=Ric(σ,σ)+2cRic(σ,X)+c2Ric(X,X).

    Integrating the above equation and using Eqs (2.9), (3.4), and (3.5), we arrive at

    NnRic(σ+cX,σ+cX)=Nn{(Δσ)2Hσ22(n1)cσ2+n(n1)c2σ2},

    which is rearranged as

    NnRic(σ+cX,σ+cX)=Nn{(Hσ21n(Δσ)2)+n1n(Δσ)22(n1)cσ2+n(n1)c2σ2}.

    Now, inserting Eqs (3.6) and (3.7) in the above equation reveals

    NnRic(σ+cX,σ+cX)+Nn(Hσ21n(Δσ)2)=0. (3.8)

    Observe that the integrand in the second integral in Eq (3.8) is non-negative by Schwarz inequality, and the hypothesis requires that the Ricci curvatures of the Riemannian manifold (Nn,g) are positive. This traps Eq (3.8) to come forward with only the following solutions:

    σ+cX=0Hσ2=1n(Δσ)2. (3.9)

    Interestingly, both equations in Eq (3.9) reach to the same conclusions. First, take the equation

    σ+cX=0,

    which on differentiation with respect to a vector field E on Nn and using Eq (2.1), leads to the equation

    HσE+cσE=cΩE, (3.10)

    where the left-hand side is symmetric and the right-hand side is skew symmetric. Hence, we have both

    HσE+cσE=0

    and

    cΩE=0

    for arbitrary E. Thus, we have two choices: either c=0 or Ω=0. If c=0, Eq (3.6) will imply σ is a constant, and then the integral of Eq (2.6) will produce σ=0: and that is contrary to the fact that X is nontrivial. Hence, Ω=0 and Eq (3.10) reduces to

    Hσ=cσI, (3.11)

    where σ has to be a non-constant function, for σ a constant implies σ=0, which is forbidden by X being nontrivial. The second equation in Eq (3.9) also reaches the same conclusion as Eq (3.11). For it is the equality in Schwarz's inequality

    Hσ21n(Δσ)2,

    which holds if and only if

    Hσ=ΔσnI,

    and combining it with Eq (3.6), gives Eq (3.11). As seen in the above paragraph, the constant c0, indeed, as σ is a non-constant function, Eq (3.7) reveals that c>0. Thus, we have reached the conclusion that Eq (3.11) is Obata's differential equation [9,17]. Hence, (Nn,g) is isometric to the sphere Snc.

    Conversely, if (Nn,g) is isometric to the sphere Snc, then through Eq (1.3), we have a CONFVF X on Snc with conformal factor

    σ=cρ.

    First, we claim that X is a nontrivial CONFVF on Snc. For if σ=0, that is, ρ=0, which on using Eq (1.3) would imply X=0, and in turn it would give

    Z=X+ρζ=0

    a contradiction to the fact that Z is a unit vector. Hence, X is a nontrivial CONFVF on Snc. Also, for Snc, we have

    Rc=(n1)cI,

    which, on treating with Eq (1.3), implies

    Rc(X)=(n1)cX=(n1)cρ=(n1)σ,

    that is, Eq (3.1) holds. Also, the scalar curvature τ of Snc is

    τ=n(n1)c

    a constant, and therefore, Eq (3.2) holds. Finally, Snc is compact and has positive Ricci curvature. Hence, converse is true.

    It is worth noticing that the condition (3.2) is essential in the above characterization of Snc, as there are compact manifolds admitting a nontrivial CONFVF that are not isometric to Snc and on them the Eq (3.2) does not hold.

    For example, consider the compact Riemannian manifold (Nn,g), where

    Nn=S1×ρS(n1)c

    is the warped product, with ρ a smooth positive function on the unit circle S1 and the warped product metric

    g=dθ2+ρ2¯g,

    θ is a coordinate function on S1 and ¯g is the canonical metric on the sphere S(n1)c of constant curvature c. Then the vector field

    X=ρθ

    on (Nn,g) satisfies [19]

    EX=ρE,

    where E is any vector field on Nn. Thus, we obtain

    12£Xg=σg,

    that is, X is a CONFVF (Nn,g) with conformal factor σ=ρ. We have the following expression for the Ricci operator Rc on the warped product manifold (Nn,g) [19]

    Rc(E)=n1ρHρE (3.12)

    for a horizontal vector field E on S1 and

    Rc(V)=(n2)cV(ρρ+(n2)(ρρ)2)V

    for vertical vector field V on Snc. As the CONFVF X is horizontal, we see by Eq (3.12) that

    Rc(X)=(n1)Hρ(θ)=(n1)σ,

    that is, the condition (3.1) holds for the CONFVF X on the compact warped product manifold (Nn,g). The scalar curvature τ of (Nn,g) is given by

    τ=n1ρ2(2ρρ+(n2)ρ2(n2)c),

    which does not satisfy Eq (3.2).

    On the Euclidean space (En,,), there are finitely many nontrivial conformal vector fields. For instance, the position vector field ξ

    ξ=jujuj (4.1)

    satisfies

    12£ξg=g,

    that is, ξ is a CONFVF on (En,,) with conformal factor σ=1. However, ξ is closed, and therefore its associated tensor Ω=0. Next, we construct a non-closed nontrivial CONFVL on (En,,). Define a vector field X on En, n>2, by

    X=ξ+u2u1u1u2.

    Then, we see that

    EX=E+ΩE, (4.2)

    where

    ΩE=E(u2)u1E(u1)u2,

    and it follows that

    ΩE,F=E,ΩF,

    that is, Ω is a skew symmetric (1,1) tensor on the Euclidean space (En,,). Using Eq (4.2), one confirms that

    12£Xg=g,

    that is, X is a nontrivial CONFVF on (En,,) with conformal factor σ=1, and it is not a closed vector field. Moreover, we see that there are finitely many of these types of nontrivial conformal vector fields on the Euclidean space (En,,).

    In this section, we find the following characterization for a Euclidean space.

    Theorem 2. Let X be a nontrivial CONFVF on an n-dimensional complete and connected Riemannian manifold (Nn,g), n>1, with conformal factor σ and associated tensor Ω. Then the following conditions hold:

    Rc(X)=divΩ,Ω(X)=0,

    if and only if (Nn,g) is isometric to the Euclidean space (En,,).

    Proof. Suppose an n-dimensional complete and connected Riemannian manifold (Nn,g) admits a nontrivial CONFVL X with conformal factor σ and associated tensor Ω satisfying

    Rc(X)=divΩ (4.3)

    and

    Ω(X)=0. (4.4)

    Using Eqs (2.4), (4.3), and n>1, we reach the conclusion that σ is a constant. It is clear that the constant σ0 due to the fact that X is nontrivial. Next, define a smooth function α by

    2α=X2.

    Then, using Eqs (2.1) and (4.4), we find the gradient of the smooth function α is given by

    α=σX. (4.5)

    Since the CONFVF X is nontrivial and the constant σ0, the above equation confirms that the function α is not a constant. Now, differentiating Eq (4.5) with respect to a vector field E on Nn and using Eq (2.1), we arrive at

    HαE=σ2E+σΩE,

    and taking the inner product in the above equation by E, yields

    Hess(α)(E,E)=σ2g(E,E).

    On polarizing the above equation, we conclude

    Hess(α)=σ2g, (4.6)

    where α is a non-constant function and σ2 is a nonzero constant. Equation (4.6) guarantees that (Nn,g) is isometric to the Euclidean space (En,,).

    The converse is trivial, for the position vector field ξ given in (4.1) is a nontrivial CONFVF on the Euclidean space (En,,) with conformal factor σ=1, and it being closed, we have associated tensor Ω=0. Also, for the Euclidean space (En,,) being flat, we have Ricci operator Rc=0, and we see that the conditions (4.3) and (4.4) are automatically satisfied. Hence, the converse.

    It is needless to mention that one of the most interesting and most sought-out questions in differential geometry is to find different characterizations of the model spaces, namely, the sphere Snc of constant curvature c, the Euclidean space En, and the hyperbolic space Hnc of constant curvature c (c>0). In the present paper, we witnessed that an n-dimensional compact and connected Riemannian manifold (Nn,g), n>1, of positive Ricci curvature and scalar curvature admits a nontrivial CONFVF X with conformal factor σ satisfying the conditions

    Ric(X)=(n1)σ and X(τ)=2σ(n(n1)cτ), (5.1)

    for a constant c, if and only if c>0 and (Nn,g) is isometric to the sphere Snc. In the process of the proof, we have seen that the above conditions together with the assumption that (Nn,g) has positive Ricci curvature lead us to

    σ=cX. (5.2)

    Since X is a CONFVF, this indicates that σ is not constant, and furthermore, the compactness of Nn forces the constant c to be positive. Finally, combining Eq (5.2) with the definition of the CONFVF X leads us to Obata's differential equation, concluding that (Nn,g) is isometric to the sphere Snc.

    Note that the CONFVF X on the sphere Snc with conformal factor σ, by virtue of Eq (1.3), satisfies

    Ric(σ)=τnσ, (5.3)

    where

    τ=n(n1)c

    is the scalar curvature of the sphere Snc. This naturally leads to the question: Under what conditions a compact and connected Riemannian manifold (Nn,g) with Ricci operator Ric scalar curvature τ, admitting a nontrivial CONFVF X with conformal factor σ satisfying Eq (5.3), that is

    Ric(σ)=τnσ

    is isometric to the sphere Snc? We shall be interested in taking up this question in our future studies.

    S. Deshmukh: conceptualization, investigation, methodology, validation, writing-original draft, formal analysis, validation, inspection; M. Guediri: conceptualization, investigation, methodology, formal analysis, writing-review and editing, validation, inspection. All authors have read and agreed to the published version of the manuscript.

    This research was supported by Researchers Supporting Project number (RSPD2024R1053), King Saud University, Riyadh, Saudi Arabia.

    The authors declare that there are no conflicts of interest.



    [1] K. Yano, Integral formulas in Riemannian geometry, Marcel Dekker Inc., 1970.
    [2] S. Hwang, G. Yun, Conformal vector fields and their applications to Einstein-type manifolds, Results Math., 79 (2024), 45. https://doi.org/10.1007/s00025-023-02070-7 doi: 10.1007/s00025-023-02070-7
    [3] A. Caminha, The geometry of closed conformal vector fields on Riemannian spaces, Bull. Braz. Math. Soc. New Ser., 42 (2011), 277–300. https://doi.org/10.1007/s00574-011-0015-6 doi: 10.1007/s00574-011-0015-6
    [4] J. F. da S. Filho, Critical point equation and closed conformal vector fields, Math. Nach., 293 (2020), 2299–2305. https://doi.org/10.1002/mana.201900316 doi: 10.1002/mana.201900316
    [5] S. Tanno, W. Weber, Closed conformal vector fields, J. Differ. Geom., 3 (1969), 361–366. https://doi.org/10.4310/JDG/1214429058 doi: 10.4310/JDG/1214429058
    [6] W. Kuhnel, H. B. Rademacher, Conformal diffeomorphisms preserving the Ricci tensor, Proc. Amer. Math. Soc., 123 (1995), 2841–2848.
    [7] W. Kuhnel, H. B. Rademacher, Einstein spaces with a conformal group, Results Math., 56 (2009), 421. https://doi.org/10.1007/s00025-009-0440-7 doi: 10.1007/s00025-009-0440-7
    [8] W. Kühnel, H. B. Rademacher, Conformal vector fields on pseudo-Riemannian spaces, Differ. Geom. Appl., 7 (1997), 237–250. https://doi.org/10.1016/S0926-2245(96)00052-6 doi: 10.1016/S0926-2245(96)00052-6
    [9] M. Obata, The conjectures about conformal transformations, J. Differ. Geom., 6 (1971), 247–258. https://doi.org/10.4310/JDG/1214430407 doi: 10.4310/JDG/1214430407
    [10] S. Deshmukh, Characterizing spheres and Euclidean spaces by conformal vector field, Ann. Mat. Pura Appl., 196 (2017), 2135–2145. https://doi.org/10.1007/s10231-017-0657-0 doi: 10.1007/s10231-017-0657-0
    [11] K. Yano, T. Nagano, Einstein spaces admitting a one-parameter group of conformal transformations, North-holland Math. Stud., 70 (1982), 219–229. https://doi.org/10.1016/S0304-0208(08)72248-5 doi: 10.1016/S0304-0208(08)72248-5
    [12] B. Y. Chen, A simple characterization of generalized Robertson-Walker spacetimes, Gen. Relat. Gravit., 46 (2014), 1833. https://doi.org/10.1007/s10714-014-1833-9 doi: 10.1007/s10714-014-1833-9
    [13] G. S. Hall, Conformal vector fields and conformal-type collineations in space-times, Gen. Relat. Gravit., 32 (2000), 933–941. https://doi.org/10.1023/A:1001941209388 doi: 10.1023/A:1001941209388
    [14] G. S. Hall, J. D. Steele, Conformal vector fields in general relativity, J. Math. Phys., 32 (1991), 1847. https://doi.org/10.1063/1.529249 doi: 10.1063/1.529249
    [15] S. Deshmukh, M. Guediri, Characterization of Euclidean spheres, AIMS Math., 6 (2021), 7733–7740. https://doi.org/10.3934/math.2021449 doi: 10.3934/math.2021449
    [16] M. Guediri, S. Deshmukh, Hypersurfaces in a Euclidean space with a Killing vector field, AIMS Math., 9 (2024), 1899–1910. https://doi.org/10.3934/math.2024093 doi: 10.3934/math.2024093
    [17] M. Obata, Certain conditions for a Riemannian manifold to be isometric with a sphere, J. Math. Soc. Jpn., 14 (1962), 333–340. https://doi.org/10.2969/JMSJ/01430333 doi: 10.2969/JMSJ/01430333
    [18] A. L. Besse, Einstein manifolds, Springer-Verlag, 1987. https://doi.org/10.1007/978-3-540-74311-8
    [19] B. O'Neill, Semi-Riemannian geometry with applications to relativity, Academic Press, 1983.
  • Reader Comments
  • © 2024 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(642) PDF downloads(52) Cited by(0)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog