Next Article in Journal
A Hybrid Deep Learning Model for Brain Tumour Classification
Previous Article in Journal
Medical Image Authentication Method Based on the Wavelet Packet and Energy Entropy
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Casorati Inequalities for Statistical Submanifolds in Kenmotsu Statistical Manifolds of Constant ϕ-Sectional Curvature with Semi-Symmetric Metric Connection

by
Simona Decu
1,2,*,† and
Gabriel-Eduard Vîlcu
3,4,5,†
1
Department of Applied Mathematics, Bucharest University of Economic Studies, 6 Piaţa Romană, 010374 Bucharest, Romania
2
Romanian Academy, Center of Mountain Economy (CE-MONT), “Costin C. Kiriţescu” National Institute of Economic Research, 13 Calea 13 Septembrie, 030508 Bucharest, Romania
3
Department of Mathematics and Informatics, Faculty of Applied Sciences, University Politehnica of Bucharest, Splaiul Independenţei 313, 060042 Bucharest, Romania
4
Romanian Academy, “Gheorghe Mihoc-Caius Iacob” Institute of Mathematical Statistics and Applied Mathematics, 050711 Bucharest, Romania
5
Department of Cybernetics, Economic Informatics, Finance and Accountancy, Petroleum-Gas University of Ploieşti, Bd. Bucureşti 39, 100680 Ploieşti, Romania
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Entropy 2022, 24(6), 800; https://doi.org/10.3390/e24060800
Submission received: 23 May 2022 / Revised: 5 June 2022 / Accepted: 6 June 2022 / Published: 8 June 2022

Abstract

:
In this paper, we prove some inequalities between intrinsic and extrinsic curvature invariants, namely the normalized δ -Casorati curvatures and the scalar curvature of statistical submanifolds in Kenmotsu statistical manifolds of constant ϕ -sectional curvature that are endowed with semi-symmetric metric connection. Furthermore, we investigate the equality cases of these inequalities. We also describe an illustrative example.

1. Introduction

The study of simple relationships between the main intrinsic and extrinsic invariants of submanifolds is a fundamental problem in submanifold theory [1]. Recent research shows a growing trend in approaching this fascinating problem through an approach that proves some types of geometric inequalities (see, e.g., [2,3,4,5,6,7,8,9,10]).
The interest in such inequalities goes back in 1993, when B.-Y. Chen introduced the intrinsic δ-invariants, now called Chen invariants, satisfying optimal inequalities for submanifolds in real space forms [11]. Later, the notion of normalized δ-Casorati curvatures (extrinsic invariants) was defined in [12,13], giving rise to new inequalities. Unlike the Gauss and mean curvature, F. Casorati in 1890 proposed to measure the curvature of a surface at a point according to common intuition of curvature [14]. Currently, this measure is named the Casorati curvature, defined by C = k 1 2 + k 2 2 2 , where k 1 and k 2 are the principal curvatures of the surface in E 3 . L. Verstraelen geometrically modeled the perception as the Casorati curvature of sensation in the context of early human vision [15]. The Casorati curvature is also assessed as a natural measure or a measure of the normal deviations from planarity in some models of computer vision [16,17]. In mechanics and modern computer science, the Casorati curvature has become known as bending energy [17].
The topic of δ -Casorati curvatures will appeal to more geometers focused on finding new solutions of the above problem. In this respect, some recent developments are devoted to inequalities on various submanifolds of a statistical manifold, notion defined by Amari [18] in 1985 in the realm of information geometry [3,4,5,6,7,8,9,10]. In this setting, the Fisher information metric is one of the most important metrics that can be considered on statistical models [19]. Actually, it is known that modulo rescaling is the only Riemannian metric invariant under sufficient statistics and it is seen as an infinitesimal form of the relative entropy [20]. In particular, Fisher information metrics play a key role in the multiple linear regressions by maximizing the likelihood [21]. Statistical manifolds are also applied in fields such as physics, machine learning, statistics, etc. There is a natural relationship between statistical manifolds and entropy. For example, P. Pessoa et al. studied the entropic dynamics on the statistical manifolds of Gibbs distributions in [22]. Since each point of the space is a probability distribution, a statistical manifold has a profound effect on the dynamics.
Initiated by K. Kenmotsu in 1972 [23] as a branch of contact geometry, Kenmotsu geometry has generated a wide range of applications in physics (thermodynamics, classical mechanics, geometrical optics, geometric quantization, classical mechanics) and control theory [24]. The Kenmotsu statistical manifold, defined by H. Furuhata in [25], is obtained locally as a warped product between a holomorphic statistical manifold and a real line. In [8], the authors established some Casorati inequalities for Kenmotsu statistical manifolds of constant ϕ -sectional curvature.
The concept of semi-symmetric metric connection on a Riemannian manifold was introduced by H.A. Hayden in [26]. Later, interesting properties of a Riemannian manifold with semi-symmetric metric connection were obtained by K. Yano in [27] and T. Imai in [28]. In addition, T. Imai investigated hypersurfaces of a Riemannian manifold with semi-symmetric metric connection [29]. Z. Nakao generalized Imai’s approach of hypersurfaces by studying submanifolds of a Riemannian manifold with semi-symmetric metric connection [30]. The geometric inequalities on submanifolds in various manifolds with semi-symmetric metric connection have been extensively proven (see, e.g., [31,32,33,34,35,36,37]). However, only a few results are dedicated to the ambient of statistical manifolds endowed with semi-symmetric metric connection. S. Kazan and A. Kazan obtained some geometric properties of Sasakian statistical manifolds with a semi-symmetric metric connection [38]. Furthermore, M.B.K. Balgeshir and S. Salahvarzi studied new curvature properties and equations of statistical manifolds with a semi-symmetric metric connection as well as their submanifolds [39].
In this article, we establish some basic inequalities between the normalized δ -Casorati curvatures (that are known to be extrinsic invariants) and the scalar curvature (a fundamental intrinsic invariant) of statistical submanifolds in Kenmotsu statistical manifolds having a constant ϕ -sectional curvature, which are endowed with semi-symmetric metric connection. Moreover, we investigated the equality cases of such inequalities. A nontrivial example is also constructed in the last part of the article.

2. Preliminaries

Let ( M ¯ , g) be a Riemannian manifold, with g a Riemannian metric on M ¯ and ¯ an affine connection on M ¯ . A triplet ( M ¯ , g, ¯ ) is called a statistical manifold if the torsion tensor field of ¯ vanishes and ¯ g is symmetric [40]. With other words, the pair ( ¯ , g) is a statistical structure on M ¯ . Let ¯ * be an affine connection of M ¯ defined by
X g ¯ ( Y , Z ) = g ¯ ( ¯ X Y , Z ) + g ¯ ( Y , ¯ X * Z ) ,
for any X, Y, Z Γ ( T M ¯ ) , where Γ ( T M ¯ ) is the set of smooth tangent vector fields on M ¯ .
Then ¯ * is named the dual connection of ¯ with respect to g. Clearly, ( ¯ * ) * = ¯ . Moreover, the Levi-Civita connection on M ¯ is given by ¯ 0 = ¯ + ¯ * 2 [41]. If ( M ¯ , g, ¯ ) is a statistical manifold, then it is known that ( M ¯ , g, ¯ * ) is too.
Let M be a submanifold of a statistical manifold ( M ¯ , g, ¯ ) with g the induced metric on M, and ∇ the induced connection on M. Then ( M , g , ) is a statistical manifold as well.
Denote by h and h * the imbedding curvature tensor of M in M ¯ with respect to ¯ and ¯ * , respectively. Then Gauss’s formulas [40] are expressed by:
¯ X Y = X Y + h ( X , Y ) , ¯ X * Y = X * Y + h * ( X , Y ) ,
for any X , Y Γ ( T M ) .
Furthermore, denote by R, R ¯ , R * and R ¯ * the ( 0 , 4 ) -curvature tensors for the connections ∇, ¯ , * and ¯ * , respectively. Thus, the Gauss equations for the connections ¯ and ¯ * , respectively, hold as follows [41]:
g ( R ¯ ( X , Y ) Z , W ) = g ( R ( X , Y ) Z , W ) + g ( h ( X , Z ) , h * ( Y , W ) ) g ( h * ( X , W ) , h ( Y , Z ) ) ,
and
g ( R ¯ * ( X , Y ) Z , W ) = g ( R * ( X , Y ) Z , W ) + g ( h * ( X , Z ) , h ( Y , W ) ) g ( h ( X , W ) , h * ( Y , Z ) ) ,
for any X , Y , Z , W Γ ( T M ) .
We can define now the statistical curvature tensor field [40] on M and M ¯ , denoted by S and S ¯ , respectively:
S ( X , Y ) Z = 1 2 { R ( X , Y ) Z + R * ( X , Y ) Z } ,
for any X , Y , Z , W Γ ( T M ) , and
S ¯ ( X , Y ) Z = 1 2 { R ¯ ( X , Y ) Z + R ¯ * ( X , Y ) Z } ,
for any X , Y , Z , W Γ ( T M ¯ ) .
Set a tensor field K ¯ Γ ( T M ¯ ( 1 , 2 ) ) by:
K ¯ X Y = K ¯ ( X , Y ) = ¯ X Y ¯ X 0 Y .
Furthermore, we have:
K ¯ ( X , Y ) = ¯ X 0 Y ¯ X * Y = 1 2 ( ¯ X Y ¯ X * Y ) .
Then K ¯ has the properties:
K ¯ ( X , Y ) = K ¯ ( Y , X ) ,
g ( K ¯ ( X , Y ) , Z ) = g ( Y , K ¯ ( X , Z ) ) .
Next, we consider ( M ¯ , g , ϕ , ξ ) a ( 2 n + 1 ) -dimensional Kenmotsu manifold defined as an almost contact metric manifold M ¯ which satisfies for any X , Y Γ ( T M ¯ ) the relations:
( ¯ X 0 ϕ ) ( Y ) = g ( ϕ X , Y ) ξ η ¯ ( Y ) ϕ X ,
¯ X 0 ξ = X η ¯ ( X ) ξ ,
where ϕ Γ ( T M ¯ ( 1 , 1 ) ) , ξ Γ ( T M ¯ ) , η ¯ is a 1-form on M ¯ with η ¯ ( X ) = g ( X , ξ ) .
A Kenmotsu manifold M ¯ with a statistical structure ( ¯ , g ) is called a Kenmotsu statistical manifold [25] if the following formula holds for any X , Y Γ ( T M ¯ ) :
K ¯ ( X , ϕ Y ) = ϕ K ¯ ( X , Y ) ,
where K ¯ is the tensor field defined in (5),
A Kenmotsu statistical manifold ( M ¯ , ¯ , g , ϕ , ξ ) is said to be of constant ϕ -sectional curvature c if and only if [25]:
S ¯ ( X , Y ) Z = c 3 4 { g ( Y , Z ) X g ( X , Z ) Y ) } + c + 1 4 { g ( ϕ Y , Z ) ϕ X g ( ϕ X , Z ) ϕ Y 2 g ( ϕ X , Y ) ϕ Z g ( Y , ξ ) g ( Z , ξ ) X + g ( X , ξ ) g ( Z , ξ ) Y + g ( Y , ξ ) g ( Z , X ) ξ g ( X , ξ ) g ( Z , Y ) ξ } ,
for any X , Y , Z Γ ( T M ¯ ) .
On the other hand, assume that ˜ is a linear connection on M ¯ . Then ˜ is called a semi-symmetric connection if the torsion tensor T ˜ of ˜ defined by
T ˜ ( X , Y ) = ˜ X Y ˜ Y X [ X , Y ]
satisfies for any X , Y Γ ( T M ¯ ) the relation:
T ˜ ( X , Y ) = η ( Y ) X η ( X ) Y ,
where η is a 1-form. Moreover, the connection ˜ is called a semi-symmetric metric connection on M ¯ if we have ˜ g = 0 (see [27]).
Next, we will denote by γ the ( 1 , 2 ) -tensor field defined by
γ ( X , Y ) = ( ¯ X 0 η ) Y ( X 0 η ) Y .
Let ( M ¯ , g , ¯ ) be a statistical manifold endowed with a semi-symmetric metric connection ˜ . Then ˜ satisfies for any X , Y Γ ( T M ¯ ) [39]:
˜ X Y = ¯ X Y + η ( Y ) X g ( X , Y ) U K ¯ X Y ,
˜ X Y = ¯ X * Y + η ( Y ) X g ( X , Y ) U K ¯ X Y ,
where U is a vector field such that g ( U , X ) = η ( X ) , K ¯ is the difference tensor field defined in (5).
Let M be an ( m + 1 ) -dimensional submanifold of a statistical manifold M ¯ endowed with a semi-symmetric metric connection ˜ . Denote the induced connection and h the second fundamental form on M with respect to ˜ . Then the Gauss formula with respect to ˜ is:
˜ X Y = X Y + h ( X , Y ) .
In addition, the Gauss equation with respect to ˜ is [39]:
g ( R ˜ ( X , Y ) Z , W ) = g ( R ( X , Y ) Z , W ) + g ( h ( X , Z ) , h ( Y , W ) ) g ( h ( Y , Z ) , h ( X , W ) ) ,
where R ˜ and R are the curvature tensor fields associated with the connections ˜ and , respectively.
We notice that h coincides with the second fundamental form of the Levi-Civita connection (see, e.g., [39]). Thus, h becomes:
h ( X , Y ) = 1 2 ( h ( X , Y ) + h * ( X , Y ) ) .
According to Kazan et al. [38], the relations between the curvature tensor R ˜ of ˜ and the curvature tensors R ¯ and R ¯ * of the connections ¯ and ¯ * are as follows:
R ˜ ( X , Y ) Z = R ¯ ( X , Y ) Z + { η ( X ) U η ( U ) X ¯ X U + K ¯ ( X , U ) } g ( Y , Z ) + { η ( Y ) U η ( U ) Y ¯ Y U + K ¯ ( Y , U ) } g ( X , Z ) g ( η ( X ) U ¯ X U + K ¯ ( X , U ) , Z ) Y + g ( η ( Y ) U ¯ Y U + K ¯ ( Y , U ) , Z ) X ( ¯ X K ¯ ) ( Y , Z ) + ( ¯ Y K ¯ ) ( X , Z ) + K ¯ ( X , K ¯ ( Y , Z ) ) K ¯ ( Y , K ¯ ( X , Z ) ) ,
and
R ˜ ( X , Y ) Z = R ¯ * ( X , Y ) Z + { η ( X ) U η ( U ) X ¯ X * U K ¯ ( X , U ) } g ( Y , Z ) { η ( Y ) U η ( U ) Y ¯ Y * U K ¯ ( Y , U ) } g ( X , Z ) g ( η ( X ) U ¯ X * U K ¯ ( X , U ) , Z ) Y + g ( η ( Y ) U ¯ Y * U K ¯ ( Y , U ) , Z ) X + ( ¯ X * K ¯ ) ( Y , Z ) ( ¯ Y K ¯ ) ( X , Z ) + K ¯ ( X , K ¯ ( Y , Z ) ) K ¯ ( Y , K ¯ ( X , Z ) ) ,
for any X , Y , Z Γ ( T M ¯ ) .
On the other hand, since the induced connection of the semi-symmetric metric connection ˜ is also semi-symmetric metric connection [39], then the Gauss formula (10) becomes:
g ( R ˜ ( X , Y ) Z , W ) = g ( R ( X , Y ) Z , W ) + { η ( X ) η ( W ) η ( U ) g ( X , W ) g ( X U , W ) + g ( K X U , W ) } g ( Y , Z ) + { η ( Y ) η ( W ) η ( U ) g ( Y , W ) g ( Y U , W ) + g ( K Y U , W ) } g ( X , Z ) g ( η ( X ) U X U + K ( X , U ) , Z ) g ( Y , W ) + g ( η ( Y ) U Y U + K ( Y , U ) , Z ) g ( X , W ) g ( ( X K ) ( Y , Z ) , W ) + g ( ( Y K ) ( X , Z ) , W ) + g ( K X K ( Y , Z ) , W ) g ( K Y K ( X , Z ) , W ) 1 4 g ( h ( X , W ) + h * ( X , W ) , h ( Y , Z ) + h * ( Y , Z ) ) + 1 4 g ( h ( X , Z ) + h * ( X , Z ) , h ( Y , W ) + h * ( Y , W ) ) ,
where K X Y = 1 2 ( * ) and R is the curvature tensor of the induced statistical connection ∇ on the submanifold M.
Similarly, we can obtain the Gauss formula involving the curvature tensor R * of the induced statistical connection * on M as follows:
g ( R ˜ ( X , Y ) Z , W ) = g ( R * ( X , Y ) Z , W ) + { η ( X ) η ( W ) η ( U ) g ( X , W ) g ( X * U , W ) + g ( K X U , W ) } g ( Y , Z ) + { η ( Y ) η ( W ) η ( U ) g ( Y , W ) g ( Y * U , W ) + g ( K Y U , W ) } g ( X , Z ) g ( η ( X ) U X * U + K ( X , U ) , Z ) g ( Y , W ) + g ( η ( Y ) U Y * U + K ( Y , U ) , Z ) g ( X , W ) g ( ( X * K ) ( Y , Z ) , W ) + g ( ( Y * K ) ( X , Z ) , W ) + g ( K X K ( Y , Z ) , W ) g ( K Y K ( X , Z ) , W ) 1 4 g ( h ( X , W ) + h * ( X , W ) , h ( Y , Z ) + h * ( Y , Z ) ) + 1 4 g ( h ( X , Z ) + h * ( X , Z ) , h ( Y , W ) + h * ( Y , W ) ) .
If x M and π T x M is a non-degenerate 2-plane, then the sectional curvature  σ is defined as [40]:
σ ( π ) = σ ( X Y ) = g ( S ( X , Y ) Y , X ) g ( X , X ) g ( Y , Y ) g 2 ( X , Y ) ,
where { X , Y } is a basis of π .
The scalar curvature  τ of ( M , , g ) at a point x M is defined by:
τ ( x ) = 1 i < j m + 1 σ ( e i e j ) = 1 i < j m + 1 g ( S ( e i , e j ) e j , e i ) ,
where { e 1 , . . . , e m + 1 } is an orthonormal basis at x. On the other hand, the normalized scalar curvature ρ of ( M , , g ) at a point x M is given by
ρ ( x ) = 2 τ ( x ) m ( m + 1 ) .
The mean curvature vector fields of M are defined by, respectively:
H = 1 m + 1 i = 1 m + 1 h ( e i , e i ) , H * = 1 m + 1 i = 1 m + 1 h * ( e i , e i ) .
It follows that we have 2 h 0 = h + h * and 2 H 0 = H + H * , where h 0 and H 0 are the second fundamental form and the mean curvature field of M, respectively, with respect to the Levi–Civita connection 0 on M.
Then, the squared mean curvatures of the submanifold M in M ¯ are given by:
H 2 = 1 ( m + 1 ) 2 α = m + 2 2 n + 1 i = 1 m + 1 h i i α 2 , H * 2 = 1 ( m + 1 ) 2 α = m + 2 2 n + 1 i = 1 m + 1 h i i * α 2 ,
where h i j α = g ( h ( e i , e j ) , e α ) and h i j * α = g ( h * ( e i , e j ) , e α ) , for i , j { 1 , . . . , m + 1 } , α { m + 2 , . . . , 2 n + 1 } .
The Casorati curvatures of the submanifold M in M ¯ are defined by the squared norms of h and h * over the dimension ( m + 1 ) , denoted by C and C * , respectively, as follows:
C = 1 m + 1 h 2 = 1 m + 1 α = m + 2 2 n + 1 i , j = 1 m + 1 h i j α 2 ,
C * = 1 m + 1 h * 2 = 1 m + 1 α = m + 2 2 n + 1 i , j = 1 m + 1 h i j * α 2 ,
where h i j α and h i j * α are defined above.
Let L be an s-dimensional subspace of T x M , s 2 and let { e 1 , , e s } be an orthonormal basis of L. Then the Casorati curvatures C ( L ) and C * ( L ) of L are given by:
C ( L ) = 1 s α = m + 2 2 n + 1 i , j = 1 s h i j α 2 , C * ( L ) = 1 s α = m + 2 2 n + 1 i , j = 1 s h i j * α 2 .
The normalized δ-Casorati curvatures δ C ( m ) and δ ^ C ( m ) of the submanifold M are given by:
δ C ( m ) | x = 1 2 C x + m + 2 2 ( m + 1 ) inf { C ( L ) | L a hyperplane of T x M }
and
δ ^ C ( m ) | x = 2 C x 2 m + 1 2 ( m + 1 ) sup { C ( L ) | L a hyperplane of T x M } .
Furthermore, the dual normalized δ * -Casorati curvatures  δ C * ( m ) and δ ^ C * ( m ) of the submanifold M in M ¯ are defined as follows:
δ C * ( m ) | x = 1 2 C * x + m + 2 2 ( m + 1 ) inf { C * ( L ) | L a hyperplane of T x M }
and
δ ^ C * ( m ) | x = 2 C * x 2 m + 1 2 ( m + 1 ) sup { C * ( L ) | L a hyperplane of T x M } .
The generalized normalized δ-Casorati curvatures  δ C ( r ; m ) and δ ^ C ( r ; m ) of M in M ¯ are defined in [13] by:
δ C ( r ; m ) | x = r C x + a ( r ) inf { C ( L ) L a hyperplane of T x M } ,
if 0 < r < m ( m + 1 ) , and
δ ^ C ( r ; m ) | x = r C x + a ( r ) sup { C ( L ) L a hyperplane of T x M } ,
if r > m ( m + 1 ) , where a ( r ) is set as
a ( r ) = m ( r + m + 1 ) ( m 2 + m r ) ( m + 1 ) r ,
for any positive real number r, different from m ( m + 1 ) .
Moreover, the dual generalized normalized δ * -Casorati curvatures  δ C * ( r ; m ) and δ * ^ C ( r ; m ) of the submanifold M in M ¯ are given by:
δ C * ( r ; m ) | x = r C * x + a ( r ) inf { C * ( L ) L a hyperplane of T x M } ,
if 0 < r < m ( m + 1 ) , and
δ ^ C * ( r ; m ) | x = r C * x + a ( r ) sup { C * ( L ) L a hyperplane of T x M } ,
if r > m ( m + 1 ) , where a ( r ) is expressed above.
Next, we consider the following constrained extremum problem
min x M f ( x ) ,
where M is a submanifold of a Riemannian manifold ( M ¯ , g ) , and f : M ¯ R is a function of differentiability class C 2 . In this setting, we recall the following result which we will use later.
Theorem 1 
([42]). If the Riemannian submanifold M is complete and connected, ( g r a d f ) ( x 0 ) T x 0 M for a point x 0 M , and the bilinear form V : T x 0 M × T x 0 M R defined by:
V ( X , Y ) = Hess ( f ) ( X , Y ) + g ( h ^ ( X , Y ) , grad f ) ,
is positive definite in x 0 , then x 0 is the optimal solution of the problem (20), where h ^ is the second fundamental form of M.
Remark 1.
If the bilinear form V defined by (21) is positive semi-definite on the submanifold M, then the critical points of f | M are global optimal solutions of the problem (20). For more details see ([43], Remark 3.2).

3. Main Inequalities

Theorem 2.
Let ( M ¯ , ¯ , g , ϕ , ξ ) be a ( 2 n + 1 ) -dimensional Kenmotsu statistical manifold of constant ϕ-sectional curvature c, endowed with a semi-symmetric metric connection ˜ . Suppose M is an ( m + 1 ) -dimensional statistical submanifold of ( M ¯ , ¯ , g , ϕ , ξ ) such that ξ is a tangent vector field on M. Then the generalized normalized δ-Casorati curvatures fulfill the following inequalities:
(i) 
δ C 0 ( r ; m ) 2 τ c 3 4 m ( m + 1 ) 3 ( c + 1 ) 4 P 2 + 1 2 m ( c + 1 ) + 2 m t r a c e ( γ )
for any r R with 0 < r < m ( m + 1 ) , where δ C 0 ( r , m ) is defined by 2 δ C 0 ( r , m ) = δ C ( r ; m ) + δ C * ( r ; m ) , and
(ii) 
δ ^ C 0 ( r ; m ) 2 τ c 3 4 m ( m + 1 ) 3 ( c + 1 ) 4 P 2 + 1 2 m ( c + 1 ) + 2 m t r a c e ( γ )
for any r R with r > m ( m + 1 ) , where δ ^ C 0 ( r , m ) is defined by 2 δ ^ C 0 ( r , m ) = δ ^ C ( r ; m ) + δ ^ C * ( r ; m ) .
Furthermore, the case of equality of any of the inequalities (22) and (23) holds at all points p M if and only if:
h = h * .
Proof. 
From Equations (13) and (14), we obtain:
R ˜ ( X , Y ) Z = S ¯ ( X , Y ) , Z + { η ( X ) U η ( U ) X ¯ X 0 U } g ( Y , Z ) { η ( Y ) U η ( U ) Y ¯ Y 0 U } g ( X , Z ) g ( η ( X ) U ¯ X 0 U , Z ) Y + g ( η ( Y ) U ¯ Y 0 U , Z ) X 1 2 [ ( ¯ X K ¯ ) ( Y , Z ) ( ¯ X * K ¯ ) ( Y , Z ) ] + 1 2 [ ( ¯ Y K ¯ ) ( X , Z ) ( ¯ Y * K ¯ ) ( X , Z ) ] + K ¯ ( X , K ¯ ( Y , Z ) ) K ¯ ( Y , K ¯ ( X , Z ) ) .
Moreover, using the definition (5), the formula (24) becomes:
R ˜ ( X , Y ) Z = S ¯ ( X , Y ) , Z + { η ( X ) U η ( U ) X ¯ X 0 U } g ( Y , Z ) { η ( Y ) U η ( U ) Y ¯ Y 0 U } g ( X , Z ) g ( η ( X ) U ¯ X 0 U , Z ) Y + g ( η ( Y ) U ¯ Y 0 U , Z ) X .
Next, the relation (25) implies:
g ( R ˜ ( X , Y ) Z , W ) = g ( S ¯ ( X , Y ) Z , W ) λ ( Y , Z ) g ( X , W ) + λ ( X , Z ) g ( Y , W ) λ ( X , W ) g ( Y , Z ) + λ ( Y , W ) g ( X , Z ) ,
for any X , Y , Z , W Γ ( T M ¯ ) , where λ is expressed by:
λ ( X , Y ) = ( ¯ X 0 η ) Y η ( X ) η ( Y ) + 1 2 η ( U ) g ( X , Y ) .
On the other hand, from the formula (6), we obtain:
g ( S ¯ ( X , Y ) Z , W ) = c 3 4 { g ( Y , Z ) g ( X , W ) g ( X , Z ) g ( Y , W ) } + c + 1 4 { g ( ϕ Y , Z ) g ( ϕ X , W ) g ( ϕ X , Z ) g ( ϕ Y , W ) 2 g ( ϕ X , Y ) g ( ϕ Z , W ) g ( Y , ξ ) g ( Z , ξ ) g ( X , W ) + g ( X , ξ ) g ( Z , ξ ) g ( Y , W ) + g ( Y , ξ ) g ( Z , X ) g ( ξ , W ) g ( X , ξ ) g ( Z , Y ) g ( ξ , W ) } ,
for any X , Y , Z , W Γ ( T M ¯ ) .
For x M , let { e 1 , . . . , e m + 1 = ξ } and { e m + 2 , . . . , e 2 n + 1 } be orthonormal bases of T x M and T x M , respectively. Suppose X = W = e i and Y = Z = e j ( i j , with i , j { 1 , . . . , m + 1 } ) in the relations (26) and (27), then we obtain:
g ( R ˜ ( e i , e j ) e j , e i ) = c 3 4 + c + 1 4 { 3 g 2 ( ϕ e i , e j ) g 2 ( e j , ξ ) g 2 ( e i , ξ ) } λ ( e i , e i ) λ ( e j , e j ) .
On the other hand, from the Gauss formulas (15) and (16) we obtain:
g ( R ˜ ( X , Y ) Z , W ) = g ( S ( X , Y ) Z , W ) μ ( Y , Z ) g ( X , W ) + μ ( X , Z ) g ( Y , W ) μ ( X , W ) g ( Y , Z ) + μ ( Y , W ) g ( X , Z ) 1 4 g ( h ( X , W ) + h * ( X , W ) , h ( Y , Z ) + h * ( Y , Z ) ) + 1 4 g ( h ( X , Z ) + h * ( X , Z ) , h ( Y , W ) + h * ( Y , W ) ) ,
for any X , Y , Z , W Γ ( T M ) , where μ has the following expression:
μ ( X , Y ) = ( X 0 η ) Y η ( X ) η ( Y ) + 1 2 η ( U ) g ( X , Y ) ,
with 0 = + * 2 . Now, we can easily see that we have
λ ( X , Y ) μ ( X , Y ) = ( ¯ X 0 η ) Y ( X 0 η ) Y = γ ( X , Y ) ,
for any X , Y Γ ( T M ) .
For X = W = e i and Y = Z = e j , from (29) we have:
g ( R ˜ ( e i , e j ) e j , e i ) = g ( S ( ( e i , e j ) e j , e i ) μ ( e i , e i ) μ ( e j , e j ) 1 4 g ( h ( e i , e i ) + h * ( e i , e i ) , h ( e j , e j ) + h * ( e j , e j ) + 1 4 g ( h ( e i , e j ) + h * ( e i , e j ) , h ( e j , e i ) + h * ( e j , e i ) ) .
Next, from (28) and (30) it follows that:
c 3 4 + c + 1 4 { 3 g 2 ( ϕ e i , e j ) g 2 ( e i , ξ ) g 2 ( e j , ξ ) } λ ( e i , e i ) λ ( e j , e j ) + μ ( e i , e i ) + μ ( e j , e j ) = g ( S ( ( e i , e j ) e j , e i ) ) 1 4 g ( h ( e i , e i ) + h * ( e i , e i ) , h ( e j , e j ) + h * ( e j , e j ) ) + 1 4 g ( h ( e i , e j ) + h * ( e i , e j ) , h ( e j , e i ) + h * ( e j , e i ) ) .
We remind that any vector field X Γ ( T M ) admits a unique decomposition into its tangent and normal components P X and P Y , respectively, as follows:
ϕ X = P X + F X .
Next, by summation over 1 i , j m + 1 , Equation (31) becomes:
c 3 4 m ( m + 1 ) + c + 1 4 ( 3 P 2 2 m ) 2 m t r a c e ( γ ) = 2 τ + 1 4 ( m + 1 ) C 0 ( m + 1 ) 2 H 0 2 ,
where P 2 is the squared norm of P expressed by
P 2 = 1 i , j m + 1 g 2 ( P e i , e j ) .
Let P be a quadratic polynomial in the components of the second fundamental form given by:
P = r C 0 + a ( r ) C 0 ( L ) + c 3 4 m ( m 1 ) + 3 ( c + 1 ) 4 P 2 1 2 m ( c + 1 ) 2 m t r a c e ( γ ) 2 τ .
We will prove that P 0 .
Consider, without loss of generality, that L is spanned by e 1 , e 2 , . . . , e m . Then, the expression of P in (33) becomes:
P = ( r + m + 1 ) C 0 + a ( r ) C 0 ( L ) ( m + 1 ) 2 H 0 2 .
Moreover, the above relation implies:
P = α = m + 2 2 n + 1 m + r + 1 m + 1 i , j = 1 m + 1 ( h i j 0 α ) 2 + a ( r ) m i , j = 1 m ( h i j 0 α ) 2 i = 1 m + 1 h i i 0 α 2 .
Furthermore, P given by (34) can be written as:
P = α = m + 2 2 n + 1 { 2 ( m + r + 1 ) m + 1 + 2 a ( r ) m 1 i < j m ( h i j 0 α ) 2 + 2 ( m + r + 1 ) m + 1 i = 1 m ( h i m + 1 0 α ) 2 + m + r + 1 m + 1 + a ( r ) m 1 i = 1 m ( h i i 0 α ) 2 2 1 i < j m + 1 h i i 0 α h j j 0 α + r m + 1 ( h m + 1 m + 1 0 α ) 2 } .
The latter equation implies:
P α = m + 2 2 n + 1 r m + a ( r ) ( m + 1 ) m ( m + 1 ) i = 1 m ( h i i 0 α ) 2 + r m + 1 ( h m + 1 m + 1 0 α ) 2 2 1 i < j m + 1 h i i 0 α h j j 0 α .
Now, suppose that f α is a quadratic form expressed by f α : R m + 1 R , for α { m + 2 , . . . , 2 n + 1 } :
f α ( h 11 0 α , h 22 0 α , . . . , h m + 1 m + 1 0 α ) = m r + ( m + 1 ) a ( r ) m ( m + 1 ) i = 1 m ( h i i 0 α ) 2 + r m + 1 ( h m + 1 m + 1 0 α ) 2 2 1 i < j m + 1 h i i 0 α h j j 0 α .
Our aim is to investigate the constrained extremum problem
min f α
under the constraint
Q : h 11 0 α + h 22 0 α + . . . + h m + 1 m + 1 0 α = k α ,
where k α is a real constant. In this respect, we establish the following first order partial derivatives system:
f α h i i 0 α = 2 [ m r + ( m + 1 ) ( a ( r ) + m ) ] m ( m + 1 ) h i i 0 α 2 k = 1 m + 1 h k k 0 α = 0 f α h m + 1 m + 1 0 α = 2 r m + 1 h m + 1 m + 1 0 α 2 k = 1 m h k k 0 α = 0 ,
for all i { 1 , . . . , m } , α { m + 2 , . . . , 2 n + 1 } .
By using the constraint Q defined by (35), the above system provides the critical point:
h i i 0 α = m ( m + 1 ) k α ( m + 1 ) a ( r ) + m r + m ( m + 1 ) ,
h m + 1 m + 1 0 α = ( m + 1 ) k α m + r + 1 ,
for all i { 1 , . . . , m } , α { m + 2 , . . . , 2 n + 1 } .
For x Q, we define the 2-form V : T x Q × T x Q R by:
V ( X , Y ) = Hess ( f α ) ( X , Y ) + h ^ ( X , Y ) , ( grad f α ) ( x ) ,
where h ^ denotes the second fundamental form of Q in R m + 1 and · , · stands for the standard inner product on R m + 1 .
We achieve also the Hessian matrix of f α with the expression:
Hess ( f α ) = β 2 2 2 2 β 2 2 2 2 β 2 2 2 2 2 r m + 1 ,
where β is a real constant set as β = 2 [ m r + ( m + 1 ) a ( r ) ] m ( m + 1 ) .
Assume that X = ( X 1 , . . . , X m + 1 ) is a tangent vector field to the hyperplane Q at x such that i = 1 m + 1 X i = 0 . Then we have:
V ( X , X ) = β i = 1 m X i 2 + 2 r m + 1 X m + 1 2 4 1 C i < j m + 1 X i X j .
By using i = 1 m + 1 X i = 0 in (36), it follows that:
V ( X , X ) = β i = 1 m X i 2 + 2 r m + 1 X m + 1 2 + 4 i = 1 m + 1 X i 2 0 .
By virtue of the Remark 1, the critical point ( h 11 0 α , . . . , h m + 1 m + 1 0 α ) is the global minimum point of the problem. In particular, we have f α ( h 11 0 α , . . . , h m + 1 m + 1 0 α ) = 0 . As a result, we obtain the inequality P 0 , namely represented by the inequalities (22) and (23), related to the infimum and supremum, respectively, over all tangent hyperplanes L of T x M .
Finally, we pursue the equality cases of the inequalities (22) and (23). For this purpose, we reveal the critical points of P , i.e., the solutions of following equations system:
P h i i 0 α = 2 m + r + 1 m + 1 + a ( r ) m 1 h i i 0 α 2 k i , k = 1 m + 1 h k k 0 α = 0 , P h m + 1 m + 1 0 α = 2 r m + 1 h m + 1 m + 1 0 α 2 k = 1 m h k k 0 α = 0 , P h i j 0 α = 4 m + r + 1 m + 1 + a ( r ) m h i j 0 α = 0 , i j , P h i m + 1 0 α = 4 ( m + r + 1 ) m + 1 h i m + 1 0 α = 0 .
Since M ¯ is a Kenmotsu statistical manifold, then we obtain the solution h c = h i j 0 α = 0 , for all i , j { 1 , . . . , m + 1 } and α { m + 2 , . . . , 2 n + 1 } . Moreover, due to P 0 and P ( h c ) = 0 , then P has a minimum point h c indicated above. In conclusion, the case of equality of any of the inequalities (22) and (23) holds if and only if h i j α = h i j * α , for i , j { 1 , . . . , m + 1 } , α { m + 2 , . . . , 2 n + 1 } . □
As a consequence of Theorem 2, we can derive the following inequalities involving the normalized δ -Casorati curvatures δ ^ C ( m ) and δ C ( m ) , the dual normalized δ -Casorati curvatures δ C * ( m ) and δ ^ C * ( m ) , as well as the normalized scalar curvature ρ of the submanifold.
Theorem 3.
Let ( M ¯ , ¯ , g , ϕ , ξ ) be a ( 2 n + 1 ) -dimensional Kenmotsu statistical manifold of constant ϕ-sectional curvature c, endowed with a semi-symmetric metric connection ˜ . Suppose M is an ( m + 1 ) -dimensional statistical submanifold of ( M ¯ , ¯ , g , ϕ , ξ ) such that ξ is a tangent vector field on M. Then the normalized δ-Casorati curvatures fulfill the following inequalities:
(i) 
δ C 0 ( m ) ρ 3 ( c + 1 ) 4 m ( m + 1 ) P 2 + c + 1 2 ( m + 1 ) c 3 4 + 2 m + 1 t r a c e ( γ ) ,
where δ C 0 ( m ) is defined by 2 δ C 0 ( m ) = δ C ( m ) + δ C * ( m ) , and
(ii) 
δ ^ C 0 ( m ) ρ 3 ( c + 1 ) 4 m ( m + 1 ) P 2 + c + 1 2 ( m + 1 ) c 3 4 + 2 m + 1 t r a c e ( γ ) ,
where δ ^ C 0 ( m ) is defined by 2 δ ^ C 0 ( m ) = δ ^ C ( m ) + δ ^ C * ( m ) .
Furthermore, the case of equality in any of the inequalities (38) and (39) holds at all points p M if and only if:
h = h * .
Proof. 
The inequality (38) follows replacing r = m ( m + 1 ) 2 in (22), by using (19) and remarking that we have the relation
δ C 0 ( m ( m + 1 ) / 2 ; m ) = m ( m + 1 ) δ C 0 ( m ) .
Similarly, we obtain inequality (39) replacing r = 2 m ( m + 1 ) in (23), by taking account of (19) and
δ ^ C 0 ( 2 m ( m + 1 ) ; m ) = m ( m + 1 ) δ ^ C 0 ( m ) .
Remark 2.
As proved in Theorems 2 and 3, the equality case of any of the inequalities (22), (23), (38) and (39) is attained for those statistical submanifolds for which the imbedding curvature tensors h and h * are related by h = h * . Note that, in view of (12), this condition implies the vanishing of the second fundamental form of the semi-symmetric metric connection. Hence, the equality case of any of the inequalities (22), (23), (38) and (39) holds at all points only for statistical submanifolds that are totally geodesic with respect to the semi-symmetric metric connection, or equivalently with respect to the Levi-Civita connection. This is a consequence of a result recently stated in [39] (see Corollary 4.4), where it was proved that for a statistical submanifold of a statistical manifold equipped with a semi-symmetric metric connection ˜ , the second fundamental form of the Levi-Civita connection coincides with the second fundamental form of ˜ .

4. Example

Let us consider the ( 2 n + 1 ) -dimensional Kenmotsu statistical manifold ( H 2 n + 1 , ¯ = ¯ 0 + K ¯ , g , ϕ , ξ ) constructed in [25] (for details see Examples 3.3 and 3.10 in the above referenced article). For the sake of simplicity, we will limit to the case of dimension 5, but the example we are going to build can be extended to any odd dimension. We remind that
H 5 = { ( x 1 , x 2 , y 1 , y 2 , z ) R 5 | z > 0 }
and the structure tensors ( g , ϕ , ξ ) are defined by
g = 1 z 2 { ( d x 1 ) 2 + ( d x 2 ) 2 + ( d y 1 ) 2 + ( d y 2 ) 2 + ( d z ) 2 } ,
ϕ x 1 = y 1 , ϕ x 2 = y 2 , ϕ y 1 = x 1 , ϕ y 2 = x 2 , ϕ z = 0
and
ξ = z z .
Denote by ¯ and ¯ * the dual connections on H 5 such that ¯ = ¯ 0 + K ¯ . We obtain:
¯ x 1 x 1 = ¯ x 2 x 2 = ¯ y 1 y 1 = ¯ y 2 y 2 = 0 ,
¯ x 1 x 2 = ¯ x 2 x 1 = ¯ y 1 y 2 = ¯ y 2 y 1 = 0 ,
¯ x 1 y 1 = ¯ y 1 x 1 = ¯ x 2 y 1 = ¯ y 1 x 2 = 0 ,
¯ x 1 y 2 = ¯ y 2 x 1 = ¯ x 2 y 2 = ¯ y 2 x 2 = 0 ,
¯ x 1 z = ¯ z x 1 = 2 z x 1 , ¯ x 2 z = ¯ z x 2 = 2 z x 2 ,
¯ y 1 z = ¯ z y 1 = 2 z y 1 , ¯ y 2 z = ¯ z y 2 = 2 z y 2 ,
¯ z z = 3 z z .
Moreover, we obtain:
¯ x 1 * x 1 = ¯ x 2 * x 2 = ¯ y 1 * y 1 = ¯ y 2 * y 2 = 2 z z ,
¯ x 1 * x 2 = ¯ x 2 * x 1 = ¯ y 1 * y 2 = ¯ y 2 * y 1 = 0 ,
¯ x 1 * y 1 = ¯ y 1 * x 1 = ¯ x 2 * y 1 = ¯ y 1 * x 2 = 0 ,
* ¯ x 1 y 2 = ¯ y 2 * x 1 = * ¯ x 2 y 2 = ¯ y 2 * x 2 = 0 ,
¯ x 1 * z = ¯ z * x 1 = ¯ x 2 * z = ¯ z * x 2 = 0 ,
¯ y 1 * z = ¯ z * y 1 = * ¯ y 2 z = ¯ z * y 2 = 0 ,
¯ z * z = 1 z z .
For any X , Y Γ ( T H 5 ) , we assume that the ( 1 , 2 ) -tensor field K ¯ is given by:
K ¯ ( X , Y ) = ν η ¯ ( X ) η ¯ ( Y ) ξ ,
where ν C ( H 5 ) and η ¯ is the 1-form on H 5 dual to ξ , that is η ¯ ( X ) = g ( X , ξ ) .
Thus, it is known that ( H 5 , ¯ = ¯ 0 + K ¯ , g , ϕ , ξ ) is a Kenmotsu statistical manifold with constant ϕ -sectional curvature c = 1 (see ([25]).
Next, we prove that H 5 admits a semi-symmetric metric connection. First, we assume that ˜ is an affine connection defined as follows:
˜ x 1 x 1 = ˜ x 2 x 2 = ˜ y 1 y 1 = ˜ y 2 y 2 = z 1 z 2 z ,
˜ x 1 x 2 = ˜ x 2 x 1 = ˜ y 1 y 2 = ˜ y 2 y 1 = 0 ,
˜ x 1 y 1 = ˜ y 1 x 1 = ˜ x 2 y 1 = ˜ y 1 x 2 = 0 ,
˜ x 1 y 2 = ˜ y 2 x 1 = ˜ x 2 y 2 = ˜ y 2 x 2 = 0 ,
˜ x 1 z = 1 z z 2 x 1 , ˜ x 2 z = 1 z z 2 x 2 ,
˜ z x 1 = 1 z x 1 , ˜ z x 2 = 1 z x 2 ,
˜ y 1 z = 1 z z 2 y 1 , ˜ y 2 z = 1 z z 2 y 2 ,
˜ z y 1 = 1 z y 1 , ˜ z y 2 = 1 z y 2 ,
˜ z z = 1 z z .
Then the torsion tensor T ˜ of ˜ satisfies the relations:
T ˜ ( x i , x j ) = T ˜ ( y i , y j ) = T ˜ ( x i , y j ) = 0 ,
T ˜ ( x i , z ) = 1 z 2 x i , T ˜ ( y i , z ) = 1 z 2 y i ,
for all i , j { 1 , 2 } .
It follows that ˜ is a semi-symmetric connection satisfying (7) with η = 1 z η ¯ . Furthermore, the relation ˜ g = 0 holds, which implies that ˜ is a semi-symmetric metric connection on the Kenmotsu statistical manifold ( H 5 , ¯ = ¯ 0 + K ¯ , g , ϕ , ξ ) of constant ϕ -sectional curvature 1 .
Let M be a 3-dimensional submanifold of the Kenmotsu statistical manifold H 5 with coordinates ( u 1 , u 2 , u 3 ) given by:
u : M H 5 ,
u ( u 1 , u 2 , u 3 ) = ( 1 2 u 1 , 1 2 u 2 , 1 2 u 2 , 1 2 u 1 , u 3 ) .
Consider the following bases in the tangent bundle T M and normal bundle T M , respectively:
{ T 1 = 1 2 ( x 1 + y 2 ) , T 2 = 1 2 ( x 2 y 1 ) , T 3 = z }
and
{ N 1 = 1 2 ( x 1 y 2 ) , N 2 = 1 2 ( x 2 + y 1 ) } .
Then we obtain:
˜ T 1 T 1 = z 1 2 z 2 T 3 , ˜ T 2 T 2 = 0 , ˜ T 3 T 3 = 1 z T 3 ,
˜ T 1 T 2 = ˜ T 2 T 1 = 0 ,
˜ T 1 T 3 = 1 z z 2 T 1 , ˜ T 3 T 1 = 1 z T 1 ,
˜ T 2 T 3 = 1 z z 2 T 2 , ˜ T 3 T 2 = 1 z T 2
and it follows immediately that the submanifold M is totally geodesic with respect to the semi-symmetric metric connection ˜ . Moreover, we conclude that the inequalities (22), (23), (38) and (39) are all satisfied with equality sign.

5. Conclusions

The purpose of this paper is to establish new inequalities between intrinsic and extrinsic curvature invariants, related to the normalized δ -Casorati curvatures and the scalar curvature of statistical submanifolds in Kenmotsu statistical manifolds of constant ϕ -sectional curvature, which are endowed with semi-symmetric metric connection. In addition, we pursued the equality cases of these inequalities and provided a nontrivial example to illustrate the results. Therefore, we believe that the topic of this survey may be developed in new challenging approaches on various classes of submanifolds in some statistical manifolds endowed with semi-symmetric metric connection.

Author Contributions

Conceptualization, S.D. and G.-E.V.; methodology, S.D. and G.-E.V.; software, S.D. and G.-E.V.; validation, S.D. and G.-E.V.; investigation, S.D. and G.-E.V.; visualization, S.D. and G.-E.V. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Informed Consent Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Chen, B.-Y. Relations between Ricci curvature and shape operator for submanifolds with arbitrary codimension. Glasgow Math. J. 1999, 41, 33–41. [Google Scholar] [CrossRef] [Green Version]
  2. Chen, B.-Y. Pseudo-Riemannian Geometry, δ-Invariants and Applications; World Scientific: Hackensack, NJ, USA, 2011. [Google Scholar]
  3. Chen, B.-Y. Recent developments in δ-Casorati curvature invariants. Turk. J. Math. 2021, 45, 1–46. [Google Scholar] [CrossRef]
  4. Chen, B.-Y. Recent developments in Wintgen inequality and Wintgen ideal submanifolds. Int. Electron. J. Geom. 2021, 20, 6–45. [Google Scholar] [CrossRef]
  5. Chen, B.-Y.; Decu, S.; Vîlcu, G.-E. Inequalities for the Casorati Curvature of Totally Real Spacelike Submanifolds in Statistical Manifolds of Type Para-Kähler Space Forms. Entropy 2021, 23, 1399. [Google Scholar] [CrossRef] [PubMed]
  6. Chen, B.-Y.; Blaga, A.M.; Vîlcu, G.-E. Differential Geometry of Submanifolds in Complex Space Forms Involving δ-Invariants. Mathematics 2022, 10, 591. [Google Scholar] [CrossRef]
  7. Mihai, I. Statistical manifolds and their submanifolds. Results on Chen-like invariants. In Geometry of Submanifolds: AMS Special Session in Honor of Bang-Yen Chen’s 75th Birthday, 20–21 October 2018, University of Michigan, Ann Arbor, Michigan; Van der Veken, J., Carriazo, A., Dimitrić, I., Oh, Y.M., Suceavă, B., Vrancken, L., Eds.; Contemporary Mathematics: Singapore, 2020; Volume 756. [Google Scholar]
  8. Decu, S.; Haesen, S.; Verstraelen, L.; Vîlcu, G.-E. Curvature Invariants of Statistical Submanifolds in Kenmotsu Statistical Manifolds of Constant ϕ-Sectional Curvature. Entropy 2018, 20, 529. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  9. Decu, S.; Haesen, S.; Verstraelen, L. Inequalities for the Casorati Curvature of Statistical Manifolds in Holomorphic Statistical Manifolds of Constant Holomorphic Curvature. Mathematics 2020, 8, 251. [Google Scholar] [CrossRef] [Green Version]
  10. Decu, S.; Haesen, S. Chen Inequalities for Spacelike Submanifolds in Statistical Manifolds of Type Para-Kähler Space Forms. Mathematics 2022, 10, 330. [Google Scholar] [CrossRef]
  11. Chen, B.-Y. Some pinching and classification theorems for minimal submanifolds. Arch. Math. 1993, 60, 568–578. [Google Scholar] [CrossRef]
  12. Decu, S.; Haesen, S.; Verstraelen, L. Optimal inequalities involving Casorati curvatures. Bull. Transilv. Univ. Braşov Ser. B Suppl. 2007, 14, 85–93. [Google Scholar]
  13. Decu, S.; Haesen, S.; Verstraelen, L. Optimal inequalities characterising quasi-umbilical submanifolds. J. Inequal. Pure Appl. Math. 2008, 9, 79. [Google Scholar]
  14. Casorati, F. Mesure de la courbure des surfaces suivant l’idée commune. Acta Math. 1890, 14, 95–110. [Google Scholar] [CrossRef]
  15. Verstraelen, L. Geometry of submanifolds I. The first Casorati curvature indicatrices. Kragujevac J. Math. 2013, 37, 5–23. [Google Scholar]
  16. Koenderink, J.J. Shadows of Shapes; De Clootcrans Press: Utrecht, The Netherlands, 2012. [Google Scholar]
  17. Koenderink, J.; van Doorn, A.; Wagemans, J. Local solid shape. i-Perception 2015, 6, 188–204. [Google Scholar] [CrossRef] [Green Version]
  18. Amari, S. Differential-Geometrical Methods in Statistics. In Lecture Notes in Statistics; Berger, J., Fienberg, S., Gani, J., Krickeberg, K., Olkin, I., Singer, B., Eds.; Springer: Berlin, Germany, 1985; Volume 28. [Google Scholar]
  19. Vîlcu, G.-E. Almost products structures on statistical manifolds and para-Kähler-like statistical submersions. Bull. Sci. Math. 2021, 171, 103018. [Google Scholar] [CrossRef]
  20. Gomez, I.S.; Portesi, M.; Borges, E.P. Universality classes for the Fisher metric derived from relative group entropy. Phys. A Stat. Mech. Appl. 2020, 547, 123827. [Google Scholar] [CrossRef]
  21. Jäntschi, L.; Balint, D.; Bolboacă, S.D. Multiple linear regressions by maximizing the likelihood under assumption of generalized Gauss-Laplace distribution of the error. Comput. Math. Methods Med. 2016, 2016, 8578156. [Google Scholar] [CrossRef]
  22. Pessoa, P.; Costa, F.X.; Caticha, A. Entropic Dynamics on Gibbs Statistical Manifolds. Entropy 2021, 23, 494. [Google Scholar] [CrossRef]
  23. Kenmotsu, K. A class of almost contact Riemannian manifolds. Tohoku Math. J. 1972, 2, 93–103. [Google Scholar] [CrossRef]
  24. Pitiş, G. Contact Forms in Geometry and Topology. In Topics in Modern Differential Geometry; Haesen, S., Verstraelen, L., Eds.; Atlantis Press: Paris, France, 2017. [Google Scholar]
  25. Furuhata, H.; Hasegawa, I.; Okuyama, Y.; Sato, K. Kenmotsu statistical manifolds and warped product. J. Geom. 2017, 108, 1175–1191. [Google Scholar] [CrossRef]
  26. Hayden, H.A. Subspaces of a space with torsion. Proc. London Math. Soc. 1932, 34, 27–50. [Google Scholar] [CrossRef]
  27. Yano, K. On semi-symmetric metric connection. Rev. Roumaine Math. Pures Appl. 1970, 15, 1579–1586. [Google Scholar]
  28. Imai, T. Notes on semi-symmetric metric connections. Tensor 1972, 24, 293–296. [Google Scholar]
  29. Imai, T. Hypersurfaces of a Riemannian manifold with semi-symmetric metric connection. Tensor 1972, 23, 300–306. [Google Scholar]
  30. Nakao, Z. Submanifolds of a Riemannian manifold with semisymmetric metric connections. Proc. Amer. Math. Soc. 1976, 54, 261–266. [Google Scholar] [CrossRef]
  31. Decu, S. Optimal inequalities for submanifolds in quaternion-space-forms with semi-symmetric metric connection. Bull. Transilv. Univ. Braşov 2009, 2, 175–184. [Google Scholar]
  32. He, G.; Liu, H.; Zhang, L. Optimal inequalities for the Casorati curvatures of submanifolds in generalized space forms endowed with semi-symmetric non-metric connections. Symmetry 2016, 8, 113. [Google Scholar] [CrossRef] [Green Version]
  33. Lee, C.W.; Lee, J.W.; Vîlcu, G.-E.; Yoon, D.W. Optimal inequalities for the Casorati curvatures of submanifolds of generalized space forms endowed with semi-symmetric metric connections. Bull. Korean Math. Soc. 2015, 52, 1631–1647. [Google Scholar] [CrossRef] [Green Version]
  34. Mihai, A.; Özgür, C. Chen inequalities for submanifolds of real space forms with a semi-symmetric metric connection. Taiwan. J. Math. 2010, 14, 1465–1477. [Google Scholar] [CrossRef]
  35. Poyraz, N.; Dogan, B.; Yaşar, E. Chen inequalities on lightlike hypersurface of a Lorentzian manifold with semi-symmetric metric connection. Int. Electron. J. Geom. 2017, 10, 1–14. [Google Scholar] [CrossRef]
  36. Zhang, P.; Zhang, L. Casorati Inequalities for Submanifolds in a Riemannian Manifold of Quasi-Constant Curvature with a Semi-Symmetric Metric Connection. Symmetry 2016, 8, 19. [Google Scholar] [CrossRef] [Green Version]
  37. Zhang, P.; Zhang, L.; Song, W. Chen’s inequalities for submanifolds of a Riemannian manifold of quasi-constant curvature with a semi-symmetric metric connection. Taiwan J. Math. 2014, 18, 1841–1862. [Google Scholar] [CrossRef]
  38. Kazan, S.; Kazan, A. Sasakian Statistical Manifolds with Semi-Symmetric Metric Connection. Univers. J. Math. Appl. 2018, 1, 226–232. [Google Scholar] [CrossRef]
  39. Balgeshir, M.B.K.; Salahvarzi, S. Curvatures of semi-symmetric metric connections of statistical manifolds. Commun. Korean Math. Soc. 2021, 36, 149–164. [Google Scholar]
  40. Furuhata, H.; Hasegawa, I. Submanifold theory in holomorphic statistical manifolds. In Geometry of Cauchy-Riemann Submanifolds; Dragomir, S., Shahid, M.H., Al-Solamy, F.R., Eds.; Springer Science+Business Media Singapore: Singapore, 2016; pp. 179–214. [Google Scholar]
  41. Vos, P. Fundamental equations for statistical submanifolds with applications to the Barlett correction. Ann. Inst. Statist. Math. 1989, 41, 429–450. [Google Scholar] [CrossRef]
  42. Oprea, T. Constrained Extremum Problems in Riemannian Geometry; University of Bucharest Publishing House: Bucharest, Romania, 2006. [Google Scholar]
  43. Vîlcu, G.-E. An optimal inequality for Lagrangian submanifolds in complex space forms involving Casorati curvature. J. Math. Anal. Appl. 2018, 465, 1209–1222. [Google Scholar] [CrossRef]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Decu, S.; Vîlcu, G.-E. Casorati Inequalities for Statistical Submanifolds in Kenmotsu Statistical Manifolds of Constant ϕ-Sectional Curvature with Semi-Symmetric Metric Connection. Entropy 2022, 24, 800. https://doi.org/10.3390/e24060800

AMA Style

Decu S, Vîlcu G-E. Casorati Inequalities for Statistical Submanifolds in Kenmotsu Statistical Manifolds of Constant ϕ-Sectional Curvature with Semi-Symmetric Metric Connection. Entropy. 2022; 24(6):800. https://doi.org/10.3390/e24060800

Chicago/Turabian Style

Decu, Simona, and Gabriel-Eduard Vîlcu. 2022. "Casorati Inequalities for Statistical Submanifolds in Kenmotsu Statistical Manifolds of Constant ϕ-Sectional Curvature with Semi-Symmetric Metric Connection" Entropy 24, no. 6: 800. https://doi.org/10.3390/e24060800

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