Next Article in Journal
The Separability Problem in Two Qubits Revisited
Next Article in Special Issue
Investigation of Special Type-Π Smarandache Ruled Surfaces Due to Rotation Minimizing Darboux Frame in E3
Previous Article in Journal
Tricyclic Graph with Minimum Randić Index
Previous Article in Special Issue
A Surface Pencil with Bertrand Curves as Joint Curvature Lines in Euclidean Three-Space
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A Note on Shape Vector Fields on Hypersurfaces

by
Suha B. Al-Shaikh
Faculty of Computer Studies, Arab Open University, Riyadh 11681, Saudi Arabia
Symmetry 2023, 15(11), 2088; https://doi.org/10.3390/sym15112088
Submission received: 16 October 2023 / Revised: 11 November 2023 / Accepted: 16 November 2023 / Published: 20 November 2023
(This article belongs to the Special Issue Symmetry and Its Application in Differential Geometry and Topology II)

Abstract

:
In this paper, we initiate the study of shape vector fields on the hypersurfaces of a Riemannian manifold. We use a shape vector field on a compact hypersurface of a Euclidean space to obtain a characterization of round spheres. We also find a condition, under which a shape vector field that is on a compact hypersurface of a Euclidean space is a Killing vector field.

1. Introduction

In the geometry of the hypersurfaces of a Euclidean space, the main tools (as the functions), are defined on a hypersurface, and these are the mean curvature α , scalar curvature τ , the squared norms of the shape operator S, the Ricci operator Q and the curvature tensor field R. In [1], it was shown that a compact hypersurface M of the Euclidean space R n + 1 satisfying the inequality
S 2 τ 1 2 R 2 + Q 2 + 2 n ( n 1 ) α 2
is necessarily isometric to the n-sphere S n α 2 . In [2], the position vector field T of a compact hypersurface M in the Euclidean space R n + 1 was used to define a smooth vector field ξ on the hypersurface M, namely that ξ is the tangential component of the position vector field T, which is required to obtain the following integral formula
M R i c ( ξ , ξ ) + n ( n 1 ) ρ 2 τ = 0 ,
where ρ is the support function of the hypersurface. This integral formula leads to several important results on the geometry of the compact hypersurface M of the Euclidean space R n + 1 (cf. [2]). There is also another aspect in the study of hypersurfaces of a Euclidean space, namely the existence of special vector fields on a hypersurface, such as a conformal vector field or a Killing vector field. In [3], it was shown that a compact hypersurface M of the Euclidean space R n + 1 admits a Killing vector field unit ξ , such that the function g S ξ , ξ is a constant and is, necessarily, an odd dimensional sphere of a constant curvature.
There is yet another type of hypersurface in the Euclidean space R n + 1 , namely the translation hypersurface. In this type of hypersurface, the focus is on the analysis of the smooth function f : R n + 1 R , whose graph is the desired hypersurface under study. For instance, in [4], the author investigates translation hypersurfaces in the Euclidean space R n + 1 , i.e., those which have a Gauss–Kronecker curvature depending on either its first p variables or on the rest of the q variables. These hypersurfaces are graphs of a function, i.e., the sum of two functions in the p and q variables, whereby n = p + q . In [4], the authors found conditions on a translation hypersurface that had a Gauss–Kronecker curvature of zero. As generalizations of translation hypersurfaces, the notion of separable hypersurfaces in the Euclidean space R n + 1 has also been introduced (cf. [5,6,7,8]). Regarding separable hypersurfaces, the defining smooth function, whose graph is the translation hypersurface, is required to satisfy the additional conditions. In [8], the authors studied separable hypersurfaces in R n + 1 , as well as a vanishing Gauss–Kronecker curvature.
We should recall that specific types of vector fields, such as Killing and conformal vector fields, as well as differential equations, like Obata’s and Fischer–Marsden’s, have a great significance in characterizing important Riemannian manifolds (cf. [1,2,3,9,10,11,12,13,14,15]). On a Riemannian manifold ( M , g ) , a conformal vector field ξ —which is defined as a vector field for which the local flow corresponds to conformal transformations of ( M , g ) —is equivalently the following:
1 2 L ξ g = σ g ,
where the Lie derivative of the metric g with respect to the vector field ξ is represented as L ξ g , while σ denotes a smooth function on M (which is referred to as the conformal factor associated with ξ ). If σ = 0 , then ξ is said to be a Killing vector field. For M, an orientable hypersurface of the Riemannian manifold ( M ¯ , g ¯ ) is equipped with a shape operator S and a unit normal N. If we define a smooth vector field ξ on the hypersurface M, it is called a shape vector field if it meets the following conditions:
1 2 L ξ g = σ B ,
where B is defined on M by
B ( U , V ) = g S U , V , U , V X ( M ) ,
and the space of vector fields on the hypersurface M is denoted as X ( M ) . The potential function of the shape vector field ξ defined on the hypersurface M is denoted by the smooth function σ . In the case of the unit sphere S 2 n + 1 , as a hypersurface of Euclidean space R 2 ( n + 1 ) , the Reeb vector field ξ on S 2 n + 1 is a shape vector field with a potential function σ = 0 . We will see that there are shape vector fields on the sphere S n c , which have a constant curvature c as the hypersurfaces of the Euclidean space R n + 1 (which also have a nonzero potential functions, as shown in Section 3).
Consider an orientable hypersurface M in R n + 1 that is characterized by a unit normal vector N and a shape operator S. Let ξ be a shape vector field on M with a corresponding potential function σ . We will use η to denote the 1-form that is dual to the shape vector field ξ , that is,
η ( U ) = g U , ξ , U X ( M ) ,
where g is the induced metric on M. Then, the operator F is defined on the hypersurface M by
1 2 d η U , V = g F U , V , U , V X ( M ) ,
such that d η is the exterior derivative of η . We call F the operator associated with the shape vector field. For a local orthonormal frame v 1 , , v n , which is defined on the hypersurface, the squared length of F is represented as
F 2 = i = 1 n g F v i , F v i .
We denote the above by α as the mean curvature of the hypersurface M, and this is defined by
n α = T r · S ,
where T r · S is the trace of S. In this paper, we employed the shape vector field that was defined on the hypersurface, with a constant curvature c, to establish the subsequent characterization of the sphere S n ( c ) .
Theorem 1.
Let M be compact and simply connected to the hypersurface of R n + 1 , n 2 , with mean curvature α, and let ξ be a shape vector field with a nonzero potential function σ, as well as with the associated operator F on M. Then, the Ricci curvature R i c ( ξ , ξ ) satisfies
M R i c ξ , ξ M n ( n 1 ) σ 2 α 2 + F 2 ,
if and only if α is a constant and M is isometric to the sphere S n α 2 .
In the methods of the proof of Theorem 1, we found that, due to the restriction on the Ricci curvature, the shape vector field ξ turned out to be a conformal vector field. Thus, a natural question arises: can a shape vector field on the hypersurface of a Euclidean space be a Killing vector field? In the next result, we answer this question and determine the conditions that arise when a shape vector field on a compact hypersurface of a Euclidean space is a Killing vector field.
Theorem 2.
Let M be a compact and connected hypersurface with a nonzero scalar curvature ø of R n + 1 , and let ξ be a shape vector field with potential function σ on M. If the function σ ø does not change sign on M, then ξ will be a Killing vector field.
Now, let us discuss an example of a shape vector field on a hypersurface within the Euclidean space R n + 1 . The geometry of Einstein hypersurfaces in a Euclidean space is very rich (cf. [16,17]). Thus, let M be a non-total geodesic that is a completely umbilical Einstein hypersurface of the Euclidean space R n + 1 . Then, we will obtain the scalar curvature τ of the hypersurface, which is a constant and has a Ricci curvature that is given by
R i c = τ n g .
The Fischer–Marsden equation is considered one of the most significant differential equations on a Riemannian manifold ( M , g ) , and it is given by
Δ ρ g + ρ R i c = H e s s ( g )
We shall call the above equation the FM-equation for convenience (cf. [18]). It has been widely acknowledged that, in a Riemannian manifold ( M , g ) , the scalar curvature τ admits a nontrivial solution of FM-equation that is a constant (cf. [18]). Indeed, it has been conjectured that a compact ( M , g ) has a nontrivial solution ρ for the FM-Equation (4), which should imply that ( M , g ) is an Einstein manifold. However, this conjecture has turned out to be false. This is due to the fact that it is well known that the FM-equation on a Riemannian manifold facilitates the study of its geometry (cf. [19,20,21]). As such, let M be a non-total geodesic that is a completely umbilical Einstein hypersurface of the Euclidean space R n + 1 , wherein a nontrivial solution ρ of the FM-equation is admitted. Then, by undertaking a trace in Equation (4), we can obtain
Δ ρ = τ n 1 ρ .
By taking ξ = ρ on the hypersurface M, i.e., when using Equation (4), we can conclude
g U ξ , V = τ n 1 ρ g U , V + τ n ρ g U , V = τ n ( n 1 ) ρ g U , V .
Since M is non-total geodesic completely umbilical hypersurface, we can obtain S U = α U , the constant α 0 and—consequently—Equation (6), which implies that
£ ξ g U , V = 2 τ n ( n 1 ) α ρ g A U , V ,
that is,
1 2 £ ξ g = σ B , σ = τ n ( n 1 ) α ρ .
Hence, on the hypersurface M of the Euclidean space R n + 1 , ξ is a shape vector field.

2. Preliminaries

For an orientable hypersurface M of the Euclidean space R n + 1 (with a unit normal vector N and a shape operator S), we can denote this by , for the Euclidean metric. In addition, we can denote the induced metric on M by g, as well as denote ¯ and ∇ by the Euclidean connection and the Riemannian connection on R n + 1 and M, respectively. Then, for the hypersurface M, the fundamental equations are as follows:
¯ U V = U V + g S U , V N , U , V X ( M )
and
¯ U N = S U , U X ( M ) .
For the hypersurface M, the curvature tensor R is given by
R U , V W = g S V , W S U g S U , W S V , U , V , W X ( M ) ,
For the hypersurface M, the Ricci tensor is given by
R i c U , V = n α g S U , V g S U , S V , U , V X ( M ) ,
where the hypersurface M and the mean curvature α is denoted by n α = T r · S . In addition, the Ricci operator Q is defined by
R i c U , V = g Q U , V , U , V X ( M )
and the scalar curvature ø is introduced by
ø = T r · Q .
Using Equation (10), we can obtain
Q U = n α S U S 2 U , ø = n 2 α 2 S 2 ,
where
S 2 = i = 1 n g S v i , S v i
and v 1 , , v n is a local frame on the hypersurface M. As the Euclidean space R n + 1 is flat, the Codazzi equation for the hypersurface has the form
S U , V = S V , U , U , V X ( M ) ,
where
S U , V = U S V S U V .
Also, by considering a local frame v 1 , , v n on the hypersurface M, and by using n α = T r · S , i.e.,
n α = i = 1 n g S v i , v i ,
and then using the symmetry of the shape operator S, as well as Equation (14), we can compute
n U α = g U , i = 1 n S v i , v i , U X ( M ) .
Thus, the mean curvature α has the gradient
α = 1 n i = 1 n S v i , v i .
Regarding the hypersurface M, consider ξ as the shape vector field with the potential function σ and the associated operator F. Then, we can obtain
1 2 L ξ g = σ B , 1 2 d η U , V = g F U , V , U , V X ( M ) ,
where B ( U , V ) = g S U , V and η are the 1-form that is dual to ξ . Then, by applying Koszul’s formula (cf. [10]), we can obtain
U ξ = σ S U + F V , U X ( M ) .
When using the formula of the curvature tensor
R U , V W = U V W V U W [ U , V ] W , U , V , W X ( M ) ,
and Equations (14) and (16), we can confirm
R U , V ξ = U σ S V V σ S U + F ( U , V ) F V , U , U , V X ( M ) .
Regarding the hypersurface M, when using a local frame v 1 , , v n and the following formula of the Ricci tensor
R i c U , ξ = i = 1 n g R v i , U ξ , v i
as well as Equation (17), we arrive at
R i c U , ξ = g S σ , U n α U σ + i = 1 n g F ( v i , U ) , v i i = 1 n g F ( U , v i ) , v i ,
In this context, the symmetry of the shape operator S and the equation n α = Tr ( S ) have been utilized. Now, by applying the skew symmetry property of the operator F, its outcomes T r . F = 0 and
g F ( U , V ) , W = g V , F U , W ,
we can conclude
R i c U , ξ = g S σ , U n α U σ i = 1 n g U , F ( v i , v i ) , U X ( M ) ,
that is,
R i c ξ , ξ = S ξ σ n α ξ σ i = 1 n g ξ , F ( v i , v i ) .

3. Proof of Theorem 1

Regarding M, there is a hypersurface of R n + 1 , n 2 , which is compact and simply connected with a unit normal vector N and a shape operator S, where a shape vector field ξ that has a nonzero potential function σ and a related operator F is admitted. Suppose the Ricci curvature R i c ξ , ξ of hypersurface M satisfies
M R i c ξ , ξ M n ( n 1 ) σ 2 α 2 + F 2 .
Then, by regarding the hypersurface M and by considering a local frame v 1 , , v n , we can compute
d i v F ξ = i = 1 n g v i F ξ , v i = i = 1 n g F v i , ξ + F v i ξ , v i = i = 1 n g ξ , F v i , v i i = 1 n g v i ξ , F v i .
By applying Equation (16), we can obtain
d i v F ξ = i = 1 n g ξ , F v i , v i i = 1 n g σ S v i + F v i , F v i .
Note that, due to the symmetry property of S and the skew symmetry property of F, we ca obtain
i = 1 n g S v i , F v i = 0
Therefore, Equation (21) takes on the form
d i v F ξ = i = 1 n g ξ , F v i , v i F 2 ,
that is,
i = 1 n g ξ , F v i , v i = d i v F ξ + F 2 .
Next, we compute
d i v S ξ = i = 1 n g v i S ξ , v i = i = 1 n g S v i , ξ + S v i ξ , v i = i = 1 n g ξ , S v i , v i + i = 1 n g v i ξ , S v i .
By utilizing Equations (15) and (16), we can obtain
d i v S ξ = g ξ , n α + i = 1 n g σ S v i + F v i , S v i ,
which, because of Equation (21), implies
d i v S ξ = n ξ α + σ S 2 .
Note that d i v σ S ξ = S ξ σ + σ d i v S ξ and, consequently, Equation (23) imply
S ξ σ = d i v σ S ξ n σ ξ α σ 2 S 2 .
Moreover, we have n α ξ σ = n ξ α σ n σ ξ α , that is,
n α ξ σ = n σ ξ α n d i v σ α ξ σ α d i v ξ .
However, Equation (16) gives us d i v ξ = n σ α , and—by inserting this value in the above equation—we can thus obtain
n α ξ σ = n σ ξ α + n 2 σ 2 α 2 n d i v σ α ξ .
Inserting Equations (22), (24) and (25) into Equation (18) yields
R i c ξ , ξ = d i v σ S ξ σ 2 S 2 + n 2 σ 2 α 2 n d i v σ α ξ + d i v F ξ + F 2 ,
which, on integration, implies
M R i c ξ , ξ = M n 2 σ 2 α 2 σ 2 S 2 + F 2 .
Thus, we have
M σ 2 S 2 n α 2 = M n ( n 1 ) σ 2 α 2 + F 2 M R i c ξ , ξ .
In this equation, we use the inequality in (19), and we thus obtain the following:
M σ 2 S 2 n α 2 0 .
However, due to the Schwartz’s inequality, the integration on the left-hand side of the above inequality was found to be non-negative. Hence, we obtained
σ 2 S 2 n α 2 = 0 .
Note that by using the connectedness of M and σ 0 , the above equation gives
S 2 = n α 2 ,
which, with respect to the Schwartz’s inequality, is an equality that holds if and only if
S = α I .
Thus, we can obtain
S U , V = U α V , U , V X ( M ) .
On the hypersurface M, when choosing a local orthonormal frame v 1 , , v n and by considering the above equation, we can obtain
i = 1 n S v i , v i = i = 1 n v i α v i = α .
Using Equation (15) with the above equation, we can obtain
n α = α
Since n > 1 , we can infer that α is a constant. Consequently, when using Equation (27) in Equation (9), the curvature tensor of M is provided by
R U , V W = α 2 g V , W U g U , W V , U , V , W X ( M ) .
We can thus observe that, since M is a compact hypersurface of the Euclidean space R n + 1 , there exists a point on M where all sectional curvatures are of a positive nature. Consequently, we obtain the constant α 2 > 0 . As M is compact, it is thus complete. Therefore, M is complete and is a simply connected space of the constant positive curvature α 2 , thereby confirming that M is isometric to the sphere S n α 2 .
Conversely, suppose M is isometric to the sphere S n α 2 . By choosing a constant vector field u on the Euclidean space R n + 1 , and by defining a smooth function σ on S n α 2 by σ = u , N (where N is the unit normal to the sphere S n α 2 ), then let ξ be the tangential projection of u to S n α 2 . As such, we have
u = ξ + σ N
together with the shape operator of the sphere S n α 2 , which is given by A = α I . When differentiating Equation (28) and using the fundamental equation for a hypersurface, we obtain
0 = U ξ α g U , ξ N + U σ N + α σ U , U X ( S n α 2 ) .
By equating the tangential and normal components, we can obtain
U ξ = σ α U , σ = α ξ .
We can confirm
1 2 L ξ g U , V = σ α g U , V = σ g A U , V , U , V X ( S n α 2 ) ,
where ξ represents a shape vector field on S n α 2 with a potential function σ and the associated operator F = 0 . Consequently, the Ricci curvature R i c ξ , ξ of the sphere S n α 2 is given by the following:
R i c ξ , ξ = ( n 1 ) α 2 ξ 2 .
When using Equation (29), we obtain the Laplacian Δ σ = α d i v ξ = n α 2 σ , that is, σ Δ σ = n α 2 σ 2 . Integrating the last equation by parts yields
S n α 2 σ 2 = n S n α 2 α 2 σ 2
When using σ = α ξ from Equation (29) in Equation (30), we can obtain
S n α 2 R i c ξ , ξ = ( n 1 ) S n α 2 σ 2
In addition, by combining the above equation with Equation (31) and F = 0 , we can conclude
M R i c ξ , ξ = M n ( n 1 ) σ 2 α 2 + F 2 .
This completes the proof.

4. Proof of Theorem 2

Let M be a hypersurface that is compact and connected with a nonzero scalar curvature ø of R n + 1 , i.e., one that is associated with a unit normal N and the shape operator S, wherein the shape vector field ξ with a potential function σ is admitted. Suppose the function σ ø does not change the sign on M. From Equation (23), we can obtain
d i v S ξ = n ξ α + σ S 2 ,
which, in view of d i v ξ = n σ α and d i v α ξ = ξ α + n σ α 2 implies
d i v S ξ = n d i v α ξ n 2 σ α 2 + σ S 2 .
When using Equation (13), we can obtain
d i v S ξ = n d i v α ξ σ ø ,
which, on integration, yields
M σ ø = 0 .
Since the function σ ø does not change sign on M, the above equation implies σ ø = 0 . When connecting M with ø 0 , we can obtain σ = 0 . Hence, Equation (1) implies
L ξ g = 0 ,
which, in turn, implies that ξ is a Killing vector field.

5. Conclusions

In Theorem 1, we saw conditions under which the shape vector field ξ is a conformal vector field. In Theorem 2, we obtained conditions under which the same will be a Killing vector field. Moreover, there are some important vector fields, namely Jacobi-type vector fields (cf. [11]) and concurrent vector fields (cf. [10]), on a Riemannian manifold that have interesting impacts on the geometry of the host manifold. Finding the conditions under which a shape vector field on hypersurface of a Euclidean apace R n + 1 could be reduced to a Jacobi-type vector field or to a concurrent vector field could present an thought-provoking question.
The geodesic vector field (cf. [22]) is also considered another important special vector field on a Riemannian manifold. On a Riemannian manifold ( M , g ) , a vector field ζ is said to be a geodesic vector field if the integral curves of ζ are geodesics. Naturally, a Killing vector field has this property locally (i.e., that its integral curves are geodesics in a neighborhood), but this cannot be extended globally. There are numerous examples of geodesic vector fields that are not Killing vector fields, and these examples are provided by non-Sasakian structures such as Kenmotsu manifolds and nontrivial trans-Sasakian manifolds (cf. [23]). Therefore, it will be of interest to find the conditions on the shape vector field on a hypersurface M of the Euclidean space R n + 1 , such that it is reduced to a geodesic vector field. It is worth mentioning that we could analyze the applicational aspect of a shape vector field on the hypersurface of spacetime spaces, and it is also the case that the content in the books of (cf. [24,25]) would be helpful for this purpose.

Funding

This research received no external funding.

Data Availability Statement

Data are contained within the article.

Acknowledgments

The author would like to thank the Arab Open University for supporting this work.

Conflicts of Interest

The author declares no conflict of interest.

References

  1. Deshmukh, S. A note on spheres in a Euclidean space. Publ. Math. 2004, 64, 31–37. [Google Scholar] [CrossRef]
  2. Deshmukh, S. An integral formula for compact hypersurfaces in a Euclidean space and its applications. Glasgow Math. J. 1992, 34, 309–311. [Google Scholar] [CrossRef]
  3. Deshmukh, S. A note on compact hypersurfaces in a Euclidean space. C. R. Acad. Sci. Paris Ser. I 2012, 350, 971–974. [Google Scholar] [CrossRef]
  4. Ruiz-Hernández, G. Translation hypersurfaces whose curvature depends partially on its variables. J. Math. Anal. Appl. 2021, 479, 124913. [Google Scholar] [CrossRef]
  5. Hasanis, T.; López, R. Classification of separable surfaces with constant Gaussian curvature. Manuscripta Math. 2021, 166, 403–417. [Google Scholar] [CrossRef]
  6. Hasanis, T.; López, R. Translation surfaces in Euclidean space with constant Gaussian curvature. Commun. Anal. Geom. 2021, 29, 1415–1447. [Google Scholar] [CrossRef]
  7. Saglam, D.D.; Sunar, C. Translation hypersurfaces of semi-Euclidean spaces with constant scalar curvature. AIMS Math. 2022, 8, 5036–5048. [Google Scholar] [CrossRef]
  8. Chen, D.; Wang, C.X.; Wang, X.S. A Characterization of Separable Hypersurfaces in Euclidean Space. Math. Notes 2023, 113, 339–344. [Google Scholar] [CrossRef]
  9. Blair, D.E. On the characterization of complex projective space by differential equations. J. Math. Soc. Jpn. 1975, 27, 9–19. [Google Scholar] [CrossRef]
  10. Chen, B.Y. Total Mean Curvature and Submanifolds of Finite Type; World Scientific: Singapore, 1983. [Google Scholar]
  11. Deshmukh, S. Jacobi-type vector fields on Ricci solitons. Bull. Math. Soc. Sci. Math. Roum. 2012, 55, 41–50. [Google Scholar]
  12. Deshmukh, S. A Note on hypersurfaces of a Euclidean space. C. R. Acad. Sci. Paris Ser. I 2013, 351, 631–634. [Google Scholar] [CrossRef]
  13. Erkekoglu, F.; García-Río, E.; Kupeli, D.N.; Ünal, B. Characterizing specific Riemannian manifolds by differential equations. Acta Appl. Math. 2003, 76, 195–219. [Google Scholar] [CrossRef]
  14. García-Río, E.; Kupeli, D.N.; Ünal, B. Some conditions for Riemannian manifolds to be isometric with Euclidean spheres. J. Differ. Equ. 2003, 194, 287–299. [Google Scholar] [CrossRef]
  15. Tanno, S. Some differential equations on Riemannian manifolds. J. Math. Soc. Jpn. 1978, 30, 509–531. [Google Scholar] [CrossRef]
  16. Fialkow, A. Hypersurfaces of a space of constant curvature. Ann. Math. 1938, 39, 762–785. [Google Scholar] [CrossRef]
  17. Vlachos, T. Almost Einstein hypersurfaces in the Euclidean space. Ill. J. Math. 2009, 43, 1221–1235. [Google Scholar] [CrossRef]
  18. 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]
  19. Kobayashi, O. A dierential equation arising from scalar curvature function. J. Math. Soc. Jpn. 1982, 34, 665–675. [Google Scholar] [CrossRef]
  20. Geng, X.; Hou, S. Gradient estimates for the Fisher-KPP equation on Riemannian manifolds. Bound. Value Probl. 2018, 2018, 25. [Google Scholar] [CrossRef]
  21. Shen, Y. A note on Fischer-Marsden’s conjecture. Proc. Am. Math. Soc. 1997, 125, 901–905. [Google Scholar] [CrossRef]
  22. Deshmukh, S.; Khan, V.A. Geodesic vector fields and eikonal equation on a Riemannian manifold. Indag. Math. 2019, 30, 542–552. [Google Scholar] [CrossRef]
  23. Al-Dayel, I.; Deshmukh, S.; Vîlcu, G. Trans-Sasakian static spaces. Results Phys. 2021, 31, 105009. [Google Scholar] [CrossRef]
  24. Petrov, A.Z. New Methods in General Relativity Theory; NASA/ADS: Moscow, Russia, 1966. [Google Scholar]
  25. Petrov, A.Z. Einstein Spaces; Elsevier: Oxford, UK, 1969. [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

Al-Shaikh, S.B. A Note on Shape Vector Fields on Hypersurfaces. Symmetry 2023, 15, 2088. https://doi.org/10.3390/sym15112088

AMA Style

Al-Shaikh SB. A Note on Shape Vector Fields on Hypersurfaces. Symmetry. 2023; 15(11):2088. https://doi.org/10.3390/sym15112088

Chicago/Turabian Style

Al-Shaikh, Suha B. 2023. "A Note on Shape Vector Fields on Hypersurfaces" Symmetry 15, no. 11: 2088. https://doi.org/10.3390/sym15112088

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