Next Article in Journal
Utilizing Cross-Ratios for the Detection and Correction of Missing Digits in Instrument Digit Recognition
Next Article in Special Issue
Alternative View of Inextensible Flows of Curves and Ruled Surfaces via Alternative Frame
Previous Article in Journal
SSRES: A Student Academic Paper Social Recommendation Model Based on a Heterogeneous Graph Approach
Previous Article in Special Issue
On Curvature Pinching for Submanifolds with Parallel Normalized Mean Curvature Vector
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Shape Operator of Real Hypersurfaces in S6(1)

by
Djordje Kocić
*,† and
Miroslava Antić
Faculty of Mathematics, University of Belgrade, 11000 Belgrade, Serbia
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Mathematics 2024, 12(11), 1668; https://doi.org/10.3390/math12111668
Submission received: 18 April 2024 / Revised: 23 May 2024 / Accepted: 24 May 2024 / Published: 27 May 2024

Abstract

:
The aim of the paper is to present two results concerning real hypersurfaces in the six-dimensional sphere S 6 ( 1 ) . More precisely, we prove that real hypersurfaces with the Lie-parallel shape operator A must be totally geodesic hyperspheres. Additionally, we classify real hypersurfaces in a nearly Kähler sphere S 6 ( 1 ) whose Lie derivative of the shape operator coincides with its covariant derivative.

1. Introduction

An almost Hermitian manifold with an almost complex structure J and the Levi-Civita connection ˜ is, respectively, Kähler or nearly Kähler if the tensor field G ( X , Y ) = ( ˜ X J ) Y is vanishing or skew-symmetric. In [1], it was shown that an arbitrary nearly Kähler manifold can be locally decomposed into manifolds of three particular types, with six-dimensional nearly Kähler manifolds being one of those types. We recall that it has been shown by Butruille, see [2], that only four homogeneous, six-dimensional, strictly nearly Kähler manifolds exist: the six-dimensional sphere S 6 ( 1 ) , the manifold S 3 × S 3 , the projective space C P 3 , and the flag manifold S U ( 3 ) / U ( 1 ) × U ( 1 ) . We note that, out of these four, only the sphere S 6 ( 1 ) is endowed with standard metrics.
If we denote by N the unit normal vector field on the hypersurface M of an almost Hermitian manifold, the tangent vector field ξ = J N is called the Reeb vector field or characteristic vector field. We denote by g the metric on S 6 ( 1 ) induced by the standard Euclidean metric , in the space R 7 . The shape operator of real hypersurfaces M of S 6 ( 1 ) is denoted by A and satisfies g ( A X , Y ) = g ( h ( X , Y ) , N ) for all X, Y tangent to M, where h is the second fundamental form of M.
We say that a hypersurface M is Hopf if the vector field ξ satisfies A ξ = α ξ for a certain differentiable function α , and then ξ is a principal vector field. We also note that the function α is locally constant, see [3]. It was shown in [3] that a connected Hopf hypersurface of a nearly Kähler S 6 ( 1 ) is an open part of either a geodesic hypersphere or a tube around an almost complex curve in S 6 ( 1 ) . Therefore, at each point of a Hopf hypersurface in S 6 ( 1 ) , there exist either one, two, or three different principal curvatures.
The parts of totally geodesic hyperspheres are, of course, umbilical and have only one principal curvature α . All other Hopf hypersurfaces in S 6 ( 1 ) are parts of tubes around almost complex curves in S 6 ( 1 ) , and the principal curvature α has a multiplicity of 3. If the almost complex curve is totally geodesic, there is only one other principal curvature μ that has a multiplicity of 2. If the almost complex curve is not totally geodesic, then it is one of types (I), (II), or (III), and α has two other principal curvatures, μ and λ , that have a multiplicity of 1. For details, we refer the reader to [4].
In recent years, the study of Riemannian hypersurfaces in different ambient spaces endowed with an almost complex structure has been an active field of research. In particular, many studies deal with the question of the existence and classification of hypersurfaces that satisfy various conditions related to their parallelism, particularly related to its shape operator, such as η -parallelness or pseudo-parallelness; we refer the reader to [5,6]. The non-existence of real hypersurfaces in complex space forms with a parallel shape operator was observed in [7]. Additionally, Kimura and Maeda [8] classified real hypersurfaces in complex projective space whose shape operator is ξ -parallel, i.e., ξ A = 0 .
Recall that Lie derivative L of a vector field on M is given by L X Y = X Y Y X , while the Lie derivative of the shape operator is given by ( L X A ) Y = X ( A Y ) + A Y X A Y X A X Y for the X, Y tangent of M. The Lie derivative, with respect to the vector field, has many applications in physics, in particular in mechanics, hydrodynamics, theory of relativity, and cosmology. Hence, hypersurfaces whose shape operator is invariant or has at least a certain regularity in its behavior with respect to the Lie derivative are of particular interest. Ki, Kim, and Lee, see [9], classified real hypersurfaces in complex space forms whose shape operator is Lie ξ -parallel, i.e., L ξ A = 0 . Suh, in [10], provides a characterization of real hypersurfaces of type A in a complex two-plane Grassmannian G 2 ( C m + 2 ) , which are tubes over totally geodesic G 2 ( C m + 1 ) in G 2 ( C m + 2 ) in terms of the vanishing Lie derivative of the shape operator A along the direction of the Reeb vector field ξ . In [11], the authors prove the nonexistence of real hypersurfaces in C P 2 and C H 2 , whose shape operator satisfies the relation L X A = X A , from the X orthogonal to ξ .
Motivated by these results, we consider a similar line of research regarding the nearly Kähler sphere S 6 and prove the following theorems.
Theorem 1.
Let M be a real hypersurface in S 6 ( 1 ) . The shape operator A on M is Lie-parallel, i.e., L X A = 0 , X T M , if and only if M is a totally geodesic hypersphere in S 6 ( 1 ) .
Theorem 2.
Let M be a real hypersurface in S 6 ( 1 ) . The shape operator A on M satisfies L ξ A = ξ A , if and only if M is a totally geodesic hypersphere in S 6 ( 1 ) .
Additionally, given that totally geodesic hyperspheres also satisfy the stronger condition L X A = X A , for all X T M , we obtain the following:
Corollary 1.
Let M be a real hypersurface in S 6 ( 1 ) . The shape operator A on M satisfies L X A = X A , for all X T M , if and only if M is a totally geodesic hypersphere in S 6 ( 1 ) .
In particular, we note that most of the known results dealing with this type of problem regard hypersurfaces of Kähler manifolds. In this case, the almost complex structure is parallel, making it easier to conduct calculations. For hypersurfaces of the nearly Kähler sphere S 6 , the covariant derivative of the almost complex structure, tensor G, is skew-symmetric, which, in a technical sense, imposes the need to approach the problem in a different manner, see [12,13]. As a consequence, we need to use a suitable moving frame along the hypersurface, which is nicely suited to the given structure, in order to obtain its properties and analyze them.

2. Preliminaries

First, we provide a brief exposition of how the standard nearly Kähler structure J on S 6 ( 1 ) arises in a natural manner from the multiplication of Cayley numbers O . A vector cross product × given as follows:
u × v = 1 2 ( u v v u ) ,
and is well defined on the space of purely imaginary Cayley numbers O , which we may identify with R 7 .
Then, vectors e 1 , , e 7 of the standard orthonormal basis of the space R 7 satisfy the relations provided in the following table of multiplication.
× e 1 e 2 e 3 e 4 e 5 e 6 e 7 e 1 0 e 3 e 2 e 5 e 4 e 7 e 6 e 2 e 3 0 e 1 e 6 e 7 e 4 e 5 e 3 e 2 e 1 0 e 7 e 6 e 5 e 4 e 4 e 5 e 6 e 7 0 e 1 e 2 e 3 e 5 e 4 e 7 e 6 e 1 0 e 3 e 2 e 6 e 7 e 4 e 5 e 2 e 3 0 e 1 e 7 e 6 e 5 e 4 e 3 e 2 e 1 0
Any orthonormal basis or frame for which the relations of this table hold is called a G 2 basis or frame. Then, for an arbitrary point p S 6 ( 1 ) and X T p S 6 ( 1 ) , the (1,1)-tensor field J is defined by
J p X = p × X
and is an almost complex structure.
We will denote by , the standard metric in the space R 7 and, by g, the induced metric on S 6 ( 1 ) . Further, we denote D and ¯ by the corresponding Levi–Civita connections.
Let M be a Riemannian submanifold of the sphere S 6 ( 1 ) with a Hermitian structure ( J , g ) . The tensor field G of type ( 2 , 1 ) , defined by G ( X , Y ) = ( ¯ X J ) Y , is skew symmetric, which makes the almost complex structure a nearly Kähler one. The tensor field G has the following properties:
G ( X , J Y ) + J G ( X , Y ) = 0 ,       g ( G ( X , Y ) , Z ) + g ( G ( X , Z ) , Y ) = 0 .
Additionally, for tangent vector fields X, Y, and Z, see [14], it holds that
( ¯ G ) ( X , Y , Z ) = g ( X , Z ) J Y g ( X , Y ) J Z g ( J Y , Z ) X .
Let and be, respectively, the Levi–Civita connection of the submanifold M and the normal connection in the normal bundle T M of M in S 6 ( 1 ) . The formulae of Gauss and Weingarten for the hypersurface are as follows:
¯ X Y = X Y + h ( X , Y ) ,       ¯ X N = A X + X N ,
where X and Y are tangent vector fields and N is a normal vector field on the hypersurface. Here, A and h denote the shape operator and the second fundamental form. The relationship between them is given by g ( h ( X , Y ) , N ) = g ( A X , Y ) . The Gauss equation for the hypersurface is as follows:
R ( X , Y , Z , W ) = g ( X , W ) g ( Y , Z ) g ( X , Z ) g ( Y , W ) + g ( h ( X , W ) , h ( Y , Z ) ) g ( h ( X , Z ) , h ( Y , W ) ) , X , Y , Z , W T M ,
where we denote by R the Riemannian curvature tensor of M.
Now, let M be a hypersurface in S 6 ( 1 ) . Using the almost complex structure J on S 6 ( 1 ) , the normal vector field N on M, we define the corresponding Reeb vector field ξ = J N with dual 1-form η ( X ) = g ( X , ξ ) . Then, D = Ker η = { X T M | η ( X ) = 0 } is a four-dimensional almost complex distribution on M.

3. The Moving Frame for Hypersurfaces in S 6 ( 1 )

Now, we will present the construction of one of the local moving frames that is, roughly speaking, compatible with the structure on the hypersurface, and hence, easier to work with. Also, we will present the relationship between the connection coefficients in this particular frame, for more details see [12,15].
For any unit vector field E 1 D , let E 2 = J E 1 , E 3 = G ( E 1 , ξ ) , and E 4 = J E 3 . Then, as shown in [15], the set { E 1 , E 2 , E 3 , E 4 , E 5 = ξ } is a local orthonormal moving frame. One such frame is uniquely determined by the choice of the vector field E 1 D .
We note that, additionally, the following relations are also valid for each frame.
Lemma 1 ([15]).
The orthonormal frame { E 1 , E 2 , E 3 , E 4 , E 5 = ξ } satisfies the following relations:
G ( E 1 , E 2 ) = 0 , G ( E 1 , E 3 ) = ξ , G ( E 1 , E 4 ) = N , G ( E 1 , ξ ) = E 3 , G ( E 1 , N ) = E 4 , G ( E 2 , E 3 ) = N , G ( E 2 , E 4 ) = ξ , G ( E 2 , ξ ) = E 4 , G ( E 2 , N ) = E 3 , G ( E 3 , E 4 ) = 0 , G ( E 3 , ξ ) = E 1 , G ( E 3 , N ) = E 2 , G ( E 4 , ξ ) = E 2 , G ( E 4 , N ) = E 1 .
We further denote the coefficients of the covariant derivatives in the given frame as follows:
g i j k = g ( D E i E j , E k ) , h i j = g ( D E i E j , N ) , 1 i , j , k 5 .
Since D is a metric connection and the second fundamental form is symmetric, straightforwardly, we obtain g i j k = g i k j , and h i j = h j i .
Using the definition of G ( X , Y ) = ( ¯ X J ) Y and the expression for its covariant derivative, we obtain, the following two lemmas, see [12].
Lemma 2 ([12]).
For the previously defined coefficients, we obtain the following:
g 12 3 = g 11 4 , g 12 4 = g 11 3 , h 11 = g 12 5 , h 12 = g 11 5 , g 22 3 = g 21 4 , g 22 4 = g 21 3 , g 22 5 = g 11 5 , h 22 = g 21 5 , g 32 3 = g 31 4 , g 32 4 = g 31 3 , h 13 = 1 g 32 5 , h 23 = g 31 5 , g 42 3 = g 41 4 , g 42 4 = g 41 3 , h 14 = g 42 5 , h 24 = 1 + g 41 5 , g 52 3 = 1 g 51 4 , g 52 4 = g 51 3 , h 15 = g 52 5 , h 25 = g 51 5 , g 32 5 = 2 + g 14 5 , g 42 5 = g 13 5 , g 31 5 = g 24 5 , g 41 5 = 2 + g 23 5 , h 33 = g 43 5 , h 34 = g 33 5 , g 44 5 = g 33 5 , h 44 = g 43 5 , h 35 = g 54 5 , h 45 = g 53 5 .
Proof. 
By taking X { E 1 , , ξ } and Y { E 1 , , ξ , N } in the relation
¯ X ( J Y ) = G ( X , Y ) + J ( ¯ X Y ) ,
and using (3), we obtain the proof of the lemma.
For X = E 1 and Y = E 1 , we obtain the following:
g 12 3 = g 11 4 , g 12 4 = g 11 3 , h 11 = g 12 5 , h 12 = g 11 5 .
For X = E 2 and Y = E 1 , we obtain the following:
g 22 3 = g 21 4 , g 22 4 = g 21 3 , g 22 5 = h 12 = g 11 5 , h 22 = g 21 5 .
In a similar manner, we obtain other relations. □
Lemma 3 ([12]).
For the coefficient (4), the following holds:
g 52 5 = g 11 2 + g 13 4 , g 51 5 = g 21 2 g 23 4 , g 54 5 = g 31 2 + g 33 4 , g 53 5 = g 41 2 g 43 4 , h 55 = g 51 2 g 53 4 .
Proof. 
By taking X = E 1 , Y = E 1 , and Z = E 3 in (1), we obtain the following: g 52 5 = g 11 2 + g 13 4 . Similarly, by taking X = E 2 , Y = E 1 , and Z = E 3 , we obtain g 51 5 = g 21 2 g 23 4 ; for X = E 3 , Y = E 1 , and Z = E 3 , we obtain g 54 5 = g 31 2 + g 33 4 . Finally, for X = E 4 , Y = E 1 , and Z = E 3 and X = ξ , Y = E 1 , and Z = E 3 , respectively, we get g 53 5 = g 41 2 g 43 4 and h 55 = g 51 2 g 53 4 . Straightforward computation shows that (1) is then satisfied for an arbitrary choice of vector fields which completes the proof. □
Recall that we still have the possibility of choosing E 1 D . We can decompose the vector field A ξ into an orthogonal sum of two vector fields, parallel to D and ξ , respectively. Let E 1 be the unit vector field collinear with the first component. Then, we can write
A ξ = β E 1 + α ξ ,
for differentiable functions α and β . Moreover, by choosing the direction of E 1 , we may assume that β 0 . Of course, in the case of the Hopf hypersurface, when β = 0 , the vector field E 1 is still not uniquely determined by this.
Since A ξ has no components in the direction of vector fields E 2 , E 3 , E 4 , causing it to vanish, we obtain:
g 13 4 = g 11 2 β ,     g 23 4 = g 21 2 ,     g 33 4 = g 31 2 ,     g 43 4 = g 41 2 ,     g 53 4 = g 51 2 α .
From the Gauss equation we obtain the next two relations between the coefficients. By implementing X = E 2 , Y = E 3 , Z = E 1 , and W = E 2 into the Gauss Equation (2) we obtain the following:
E 2 ( g 31 2 ) = g 21 2 g 21 4 3 g 21 5 2 g 14 5 g 21 5 + 2 g 11 5 g 24 5 + g 11 2 ( g 21 3 + g 31 2 ) 2 g 21 4 g 31 3 + 2 g 21 3 g 31 4 + g 31 2 g 31 4 g 21 2 g 41 2 g 31 3 g 41 2 2 g 51 2 g 14 5 g 51 2 + g 23 5 g 51 2 + E 3 ( g 21 2 ) ,
and then by implementing ( X , Y , Z , W ) = ( E 2 , E 3 , E 3 , E 4 ) in (2), we also obtain the following:
( 3 + 2 g 14 5 ) g 21 5 2 g 24 5 ( g 11 5 + g 33 5 ) + ( 1 + 2 g 23 5 ) g 34 5 + ( 2 g 14 5 + g 23 5 ) α + ( g 21 3 + g 31 2 ) β = 0 .

4. Proof of the Theorem 1

Let us denote L i j k = g ( ( L E i A ) E j , E k ) . The condition that the shape operator A is Lie-parallel is equivalent to L i j k = 0 , 1 i , j , k 5 .
Lemma 4.
Let M be a hypersurface in S 6 ( 1 ) with Lie ξ-parallel shape operator A. Then, M is a Hopf hypersurface.
Proof. 
Suppose that M is a non-Hopf hypersurface, i.e., β 0 . From L 55 i = 0 , 1 i 4 , we have, respectively,
ξ ( β ) = g 11 5 β ,     g 51 2 = g 12 5 ,     g 51 3 = g 13 5 ,     g 51 4 = g 14 5 .
Using this, from
0 = L 51 5 = 3 g 11 5 β , 0 = L 53 5 = ( g 13 5 + 2 g 24 5 ) β , 0 = L 54 5 = ( 3 g 14 5 2 g 23 5 ) β ,
we get, respectively g 11 5 = 0 , g 13 5 = 2 g 24 5 and g 14 5 = 3 2 g 23 5 . From 0 = L 51 2 = 6 ( 1 + g 23 5 ) , we obtain g 12 5 = 1 , and
0 = L 52 1 = ( g 12 5 + g 21 5 ) 2 + ( g 24 5 ) 2 + β 2
gives us a contradiction. □
Therefore, now we assume M to be a Hopf hypersurface, implying that β = 0 and α are constant.
Recall that, for Hopf hypersurfaces, we still have the option of choosing E 1 D , and we can assume that E 1 is an eigenvector field for the shape operator A. As
A E 1 = g 12 5 E 1 + g 11 5 E 2 ( 1 + g 14 5 ) E 3 + g 13 5 E 4 ,
we obtain g 11 5 = 0 , g 14 5 = 1 , g 13 5 = 0 .
Now, from 0 = L 22 5 = 2 g 24 5 , we obtain g 24 5 = 0 , and then from 0 = L 24 5 + L 13 5 = 4 g 33 5 , we obtain g 33 5 = 0 . Now let us see what the shape operator looks like:
A E 1 = g 12 5 E 1 , A E 2 = g 21 5 E 2 + ( 1 + g 23 5 ) E 4 , A E 3 = g 34 5 E 3 , A E 4 = ( 1 + g 23 5 ) E 2 + g 43 5 E 4 , A E 5 = α E 5 .
Lemma 5.
Let M be a Hopf hypersurface in S 6 ( 1 ) with Lie-parallel shape operator A. Then, the moving frame can be chosen so that both E 1 and E 3 are the eigenvectors of the shape operator A for the eigenvalues α.
Proof. 
From (8), we want to prove that g 12 5 = α = g 34 5 .
Let us first assume that g 34 5 α . From
0 = L 23 5 = ( 1 g 23 5 ) ( g 34 5 + α ) , 0 = L 45 3 = ( g 34 5 + α ) ( g 43 5 + g 51 2 + α ) , 0 = L 43 5 = ( g 34 5 g 43 5 ) ( g 34 5 + α ) ,
we obtain, respectively, g 23 5 = 1 , g 51 2 = g 43 5 α , and g 43 5 = g 34 5 . Next, from 0 = L 54 3 = 2 g 51 4 we have g 51 4 = 0 , and then, from 0 = L 53 4 = 4 ( 1 + ( g 34 5 ) 2 ) , we get a contradiction. Hence, g 34 5 = α .
Now, suppose that g 12 5 α . From
0 = L 41 5 = ( 3 + g 23 5 ) ( g 12 5 + α ) ,     0 = L 25 1 = ( g 21 5 g 51 2 ) ( g 12 5 + α ) ,
we have g 23 5 = 3 and g 51 2 = g 21 5 . Then, from 0 = L 45 1 = ( 1 g 51 4 ) ( g 12 5 + α ) , we obtain g 51 4 = 1 and now, from 0 = L 51 2 = 4 ( g 12 5 + g 21 5 ) 2 , we obtain a contradiction, which completes the proof of this lemma. □
Now, (7) becomes g 21 5 ( 2 + g 23 5 ) α = 0 , so g 21 5 = ( 2 + g 23 5 ) α , and we obtain the following:
0 = L 12 5 = ( 1 + g 23 5 ) ( 3 + g 23 5 ) ( 1 + α 2 ) .
If we assume that g 23 5 = 3 , then from 0 = L 32 5 = 2 g 43 5 6 α , we obtain g 43 5 = 3 α , and from 0 = L 34 5 = 8 ( 1 + α 2 ) , we obtain a contradiction. Therefore, it must be g 23 5 = 1 .
From (8), we now see that the shape operator A of the hypersurface M has a quadruple eigenvalue α and a single g 43 5 . Since M is a Hopf surface, it is possible that M is a totally geodesic hypersphere, i.e., g 34 5 = α = 0 .
Straightforward computation shows that all L i j k , 1 i , j , k 5 are equal to zero, which completes the proof.

5. Proof of the Theorem 2

Let us denote a i j = g ( A E i ξ A E i ξ , E j ) . The condition L ξ A = ξ A is equivalent to a i j = 0 , 1 i , j 5 .
Lemma 6.
Let M be a hypersurface in S 6 ( 1 ) . If the shape operator A on M satisfies L ξ A = ξ A , then M is a Hopf hypersurface.
Proof. 
Suppose that M is a non-Hopf hypersurface, i.e., β 0 . From a i 5 = 0 , 1 i 4 , we obtain, respectively,
g 11 5 β = 0 ,     g 21 5 β     = 0 ,     g 24 5 β     = 0 ,     ( 2 + g 23 5 ) β     = 0 ,
so g 11 5 = g 21 5 = g 24 5 = 0 and g 23 5 = 2 . Now, from 0 = a 54 = ( 1 + g 14 5 ) β , we obtain g 14 5 = 1 , and then, a 21 = 4 0 , which is a contradiction. □
Suppose now that M is a Hopf hypersurface, i.e., β = 0 .
To determine the moving frame, we can choose E 1 to be an eigenvector field for the shape operator A. As
A E 1 = g 12 5 E 1 + g 11 5 E 2 ( 1 + g 14 5 ) E 3 + g 13 5 E 4 ,
we obtain g 11 5 = 0 , g 14 5 = 1 , g 13 5 = 0 .
Now, from 0 = a 22 = 2 g 24 5 , we obtain g 24 5 = 0 ; then, we obtain 0 = a 12 = g 33 5 . From a 14 = 0 and a 32 = 0 , we obtain, respectively,
g 12 5 g 23 5 + g 43 5     = 0 ,       g 21 5 ( 2 + g 23 5 ) g 34 5     = 0 ,
so g 43 5 = g 12 5 g 23 5 and g 21 5 = ( 2 + g 23 5 ) g 34 5 . The only non-zero a i j are now the following:
a 12 = 1 ( g 12 5 ) 2 + g 23 5 + g 12 5 ( 2 + g 23 5 ) g 34 5 , a 21 = [ 2 + g 23 5 ] [ 1 + g 23 5 g 12 5 g 34 5 + ( 2 + g 23 5 ) ( g 34 5 ) 2 ] , a 23 = g 23 5 ( 1 + g 23 5 ) ( g 12 5 g 34 5 ) , a 34 = 1 g 23 5 g 34 5 ( g 12 5 g 23 5 + g 34 5 ) , a 41 = ( 1 + g 23 5 ) ( 2 + g 23 5 ) ( g 12 5 g 34 5 ) , a 43 = g 23 5 ( 1 + g 23 5 + ( g 12 5 ) 2 g 23 5 + g 12 5 g 34 5 ) .
We begin further discussion from a 23 = g 23 5 ( 1 + g 23 5 ) ( g 12 5 g 34 5 ) = 0 .
If we assume that g 23 5 = 0 , then a 34 = 0 becomes 1 ( g 34 5 ) 2 = 0 , which is a contradiction. If we assume that g 12 5 = g 34 5 , from 0 = a 12 = ( 1 + ( g 12 5 ) 2 ) ( 1 + g 23 5 ) , we obtain 1 + g 23 5 = 0 , so it is enough to consider only the case g 23 5 = 1 .
Now, 0 = a 12 = g 12 5 ( g 34 5 g 12 5 ) gives us g 12 5 = 0 or g 34 5 = g 12 5 .
Assuming g 12 5 = 0 , a 21 = 0 becomes ( g 34 5 ) 2 = 0 , so g 34 5 = 0 and now all a i j are zero. In this case, the shape operator A has a quadruple eigenvalue 0 and one is α . As M is a Hopf hypersurface, this is possible if and only if α = 0 , i.e., M is a totally geodesic hypersphere.
Assuming g 34 5 = g 12 5 , we assume that all a i j are zero. In this case, the shape operator A has a quadruple eigenvalue g 12 5 and one is α . As M is a Hopf hypersurface, this is possible if and only if g 12 5 = α = 0 , i.e., M is a totally geodesic hypersphere.
Hence, if the shape operator A on M satisfies L ξ A = ξ A , then M is a totally geodesic hypersphere.
If M is a totally geodesic hypersphere then the shape operator A on M obviously satisfies L ξ A = ξ A , and that completes the proof.

Author Contributions

Conceptualization, M.A.; methodology, M.A. and D.K.; software, D.K.; validation, M.A. and D.K.; formal analysis, M.A. and D.K.; writing—original draft preparation, D.K.; supervision, M.A. All authors have read and agreed to the published version of the manuscript.

Funding

The research of the authors was partially funded by the Faculty of Mathematics at the University of Belgrade (the contracts 451-03-47/2023-01/200104 and 451-03-66/2024-03/200104) through a grant by the Ministry of Science, Technological Development, and Innovation of the Republic of Serbia.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within the article.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Nagy, P.-A. Nearly Kähler geometry and Riemannian foliations. Asian J. Math. 2002, 6, 481–504. [Google Scholar] [CrossRef]
  2. Butruille, J.-B. Homogeneous nearly Kähler manifolds. In Handbook of Pseudo-Riemannian Geometry and Supersymmetry; EMS Press: Berlin, Germany, 2010; pp. 399–423. [Google Scholar]
  3. Berndt, J.; Bolton, J.; Woodward, L.M. Almost complex curves and Hopf hypersurfaces in the nearly Kähler 6-sphere. Geom. Dedicata 1995, 56, 237–247. [Google Scholar] [CrossRef]
  4. He, L.; Jiao, X.X.; Zhou, X.C. On almost complex curves and Hopf hypersurfaces in the nearly Kähler 6-sphere. Sci. China Math. 2014, 57, 1045–1056. [Google Scholar] [CrossRef]
  5. Kon, S.H.; Loo, T.H. Real hypersurfaces in a complex space form with η-parallel shape operator. Math. Z. 2011, 269, 47–58. [Google Scholar] [CrossRef]
  6. Lobos, G.A.; Ortega, M. Pseudo-parallel real hypersurfaces in complex space forms. Bull. Korean Math. Soc. 2004, 41, 609–618. [Google Scholar] [CrossRef]
  7. Niebergall, R.; Ryan, P.J. Real hypersurfaces in complex space forms. In Tight and Taut Submanifolds; Cambridge University Press: Cambridge, UK, 1997; pp. 233–305. [Google Scholar]
  8. Kimura, M.; Maeda, S. On real hypersrufaces of a complex projective space II. Tsukuba J. Math. 1991, 15, 299–311. [Google Scholar] [CrossRef]
  9. Ki, U.H.; Kim, S.J.; Lee, S.B. Some characterizations of a real hypersurface of type A. Kyngpook Math. J. 1991, 31, 73–82. [Google Scholar]
  10. Suh, Y.J. Real hypersurfaces in complex two-plane grassmannians with vanishing Lie derivative. Canad. Math. Bull. 2006, 49, 134–143. [Google Scholar] [CrossRef]
  11. Panagiotidou, K. The structure Jacobi operator and the shape operator of real hypersurfaces in CP2 and CH2. Beitr. Algebra Geom. 2014, 55, 545–556. [Google Scholar] [CrossRef]
  12. Antić, M.; Kocić, Dj. Non-Existence of Real Hypersurfaces with Parallel Structure Jacobi Operator in S6(1). Mathematics 2022, 10, 2271. [Google Scholar] [CrossRef]
  13. Kocić, Dj. Real hypersurfaces in S6(1) equipped with structure Jacobi operator satisfying LXl=Xl. Filomat 2023, 37, 8435–8440. [Google Scholar] [CrossRef]
  14. Gray, A. The structure of nearly Kähler manifolds. Math. Ann. 1976, 22, 233–248. [Google Scholar] [CrossRef]
  15. Deshmukh, S.; Al-Solamy, F.R. Hopf hypersurfaces in nearly Kaehler 6-sphere. Balk. J. Geom. Appl. 2008, 13, 38–46. [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

Kocić, D.; Antić, M. The Shape Operator of Real Hypersurfaces in S6(1). Mathematics 2024, 12, 1668. https://doi.org/10.3390/math12111668

AMA Style

Kocić D, Antić M. The Shape Operator of Real Hypersurfaces in S6(1). Mathematics. 2024; 12(11):1668. https://doi.org/10.3390/math12111668

Chicago/Turabian Style

Kocić, Djordje, and Miroslava Antić. 2024. "The Shape Operator of Real Hypersurfaces in S6(1)" Mathematics 12, no. 11: 1668. https://doi.org/10.3390/math12111668

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