1 Introduction

For surfaces \(M^2\) of the Euclidean space \(\mathbb {E}^3\), the Euler inequality \(G\le \Vert H\Vert ^2\) is fulfilled, where G is the (intrinsic) Gauss curvature of \(M^2\) and \(\Vert H\Vert ^2\) is the (extrinsic) squared mean curvature of \(M^2\).

Furthermore, \(G=\Vert H\Vert ^2\) everywhere on \(M^2\) if and only if \(M^2\) is totally umbilical, or still, by a theorem of Meusnier, if and only if \(M^2\) is (a part of) a plane \(\mathbb {E}^2\) or, it is (a part of) a round sphere \( S^2\) in \(\mathbb {E}^3\).

In 1979, Wintgen [25] proved that the Gauss curvature G, the squared mean curvature \(\left\| H\right\| ^2\) and the normal curvature \(G^{\bot }\) of any surface \(M^{2}\) in \(\mathbb {E}^{4}\) always satisfy the inequality

$$\begin{aligned} G \le \left\| H\right\| ^2-|G^{\bot }|; \end{aligned}$$

the equality holds if and only if the ellipse of curvature of \(M^{2}\) in \(\mathbb {E}^{4}\) is a circle.

The Whitney 2-sphere satisfies the equality case of the Wintgen inequality identically.

A survey containing recent results on surfaces satisfying identically the equality case of Wintgen inequality can be read in [5].

Later, the Wintgen inequality was extended by Rouxel [20] and by Guadalupe and Rodriguez [10] independently, for surfaces \(M^{2}\) of arbitrary codimension m in real space forms \(\widetilde{M}^{2+m}(c)\); namely

$$\begin{aligned} G \le \left\| H\right\| ^2-|G^{\bot }|+c. \end{aligned}$$

The equality case was also investigated.

A corresponding inequality for totally real surfaces in n-dimensional complex space forms was obtained in [13]. The equality case was studied and a non-trivial example of a totally real surface satisfying the equality case identically was given.

In 1999, De Smet et al. [7] formulated the conjecture on Wintgen inequality for submanifolds of real space forms, which is also known as the DDVV conjecture.

This conjecture was proven by the authors for submanifolds \(M^n\) of arbitrary dimension \(n\ge 2\) and codimension 2 in real space forms \(\tilde{M}^{n+2}(c)\) of constant sectional curvature c.

Recently, the DDVV conjecture was finally settled for the general case by Lu [12] and independently by Ge and Tang [9].

One of the present authors obtained generalized Wintgen inequalities for Lagrangian submanifolds in complex space forms [14] and Legendrian submanifolds in Sasakian space forms [15], respectively. Moreover, two of the present authors established in [3] a version of the Euler inequality and the Wintgen inequality for statistical surfaces in statistical manifolds of constant curvature.

In this paper, using the sectional curvature defined in [19], we derive a generalized Wintgen inequality for statistical submanifolds in statistical manifolds of constant curvature.

2 Statistical manifolds and their submanifolds

A statistical manifold is a Riemannian manifold \((\tilde{M}^{n+k},\tilde{g})\) of dimension \((n+k),\) endowed with a pair of torsion-free affine connections \(\tilde{\nabla }\) and \(\tilde{\nabla }^{*}\) satisfying

$$\begin{aligned} Z\tilde{g}(X,Y)=\tilde{g}(\tilde{\nabla }_{Z}X,Y)+ \tilde{g}(X,\tilde{\nabla }_{Z}^{*}Y), \end{aligned}$$
(2.1)

for any \(X,Y,Z\in \Gamma (T\tilde{M}).\) The connections \(\tilde{\nabla }\) and \(\tilde{\nabla }^{*}\) are called dual connections (see [1, 17, 22]), and it is easily shown that \((\tilde{\nabla }^{*})^{*}=\tilde{\nabla }.\) The pair \((\tilde{\nabla },\tilde{g})\) is said to be a statistical structure. If \((\tilde{\nabla },\tilde{g})\) is a statistical structure on \(\tilde{M}^{n+k},\) so is \((\tilde{\nabla }^{*},\tilde{g})\) [1, 24].

On the other hand, any torsion-free affine connection \(\tilde{\nabla }\) always has a dual connection given by

$$\begin{aligned} \tilde{\nabla }+\tilde{\nabla }^{*}=2\tilde{\nabla }^{0}, \end{aligned}$$
(2.2)

where \(\tilde{\nabla }^{0}\) is Levi-Civita connection on \(\tilde{M}^{n+k}\).

Denote by \(\tilde{R}\) and \(\tilde{R}^{*}\) the curvature tensor fields of \(\tilde{\nabla }\) and \(\tilde{\nabla }^{*},\) respectively.

A statistical structure \((\tilde{\nabla },\tilde{g})\) is said to be of constant curvature \(c\in \mathbb {R}\) if

$$\begin{aligned} \tilde{R}(X,Y)Z=c\{\tilde{g}(Y,Z)X-\tilde{g}(X,Z)Y\}. \end{aligned}$$
(2.3)

A statistical structure \((\tilde{\nabla },\tilde{g})\) of constant curvature 0 is called a Hessian structure.

The curvature tensor fields \(\tilde{R}\) and \(\tilde{R}^{*}\) of dual connections satisfy

$$\begin{aligned} \tilde{g}(\tilde{R}^{*}(X,Y)Z,W) =-\tilde{g} (Z,\tilde{R}(X,Y)W). \end{aligned}$$
(2.4)

From (2.4) it follows immediately that if \((\tilde{\nabla },\tilde{g})\) is a statistical structure of constant curvature c,  then \((\tilde{\nabla }^{*},\tilde{g})\) is also a statistical structure of constant curvature c. In particular, if \((\tilde{\nabla },\tilde{g})\) is Hessian, so is \((\tilde{\nabla }^{*},\tilde{g})\) [8].

On a Hessian manifold (\(\tilde{M}^{n+k},\tilde{\nabla })\), let \(\gamma =\tilde{\nabla }^0-\tilde{\nabla }\). The tensor field Q of type (1,3) defined by the covariant differential \(Q=\tilde{\nabla }\gamma \) of \(\gamma \) is said to be the Hessian curvature tensor for \(\tilde{\nabla }\) (see [21]).

By using the Hessian curvature tensor Q, a Hessian sectional curvature can be defined on a Hessian manifold.

A Hessian manifold has constant Hessian sectional curvature \(\tilde{c}\) if and only if (see [21])

$$\begin{aligned} Q(X,Y,Z,W)=\frac{\tilde{c}}{2}[g(X,Y)g(Z,W)+g(X,W)g(Y,Z)], \end{aligned}$$

for all vector fields on \(\tilde{M}^{n+k}\).

If \((\tilde{M}^{n+k},\tilde{g})\) is a statistical manifold and \(M^n\) a submanifold of dimension n of \(\tilde{M}^{n+k},\) then \((M^{n},g)\) is also a statistical manifold with the induced connection by \(\tilde{\nabla }\) and induced metric g. In the case that \((\tilde{M}^{n+k},\tilde{g})\) is a semi-Riemannian manifold, the induced metric g has to be non-degenerate. For details, see [23, 24].

In the geometry of Riemannian submanifolds (see [4]), the fundamental equations are the Gauss and Weingarten formulas and the equations of Gauss, Codazzi and Ricci.

Let denote the set of the sections of the normal bundle to \(M^{n}\) by \(\Gamma (TM^{n\perp }).\)

In our case, for any \(X,Y\in \Gamma (TM^{n}),\) according to [24], the corresponding Gauss formulas are

$$\begin{aligned}&\tilde{\nabla }_{X}Y =\nabla _{X}Y+h(X,Y),\end{aligned}$$
(2.5)
$$\begin{aligned}&\tilde{\nabla }_{X}^{*}Y =\nabla _{X}^{*}Y+h^{*}(X,Y), \end{aligned}$$
(2.6)

where h, \(h^{*}:\Gamma (TM^n)\times \Gamma (TM^n)\rightarrow \Gamma ({TM^n}^\perp )\) are symmetric and bilinear, called the imbedding curvature tensor of \(M^{n}\) in \(\tilde{M}^{n+k}\) for \(\tilde{\nabla }\) and the imbedding curvature tensor of \(M^{n}\) in \(\tilde{M}^{n+k}\) for \(\tilde{\nabla }^{*}, \) respectively.

In [24], it is also proved that \((\nabla ,g)\) and \((\nabla ^{*},g)\) are dual statistical structures on \(M^{n}.\)

Since h and \(h^{*}\) are bilinear, we have the linear transformations \( A_{\xi }\) and \(A_{\xi }^{*}\) on \(TM^n\) defined by

$$\begin{aligned}&g(A_{\xi }X,Y)=\tilde{g}(h(X,Y),\xi ),\end{aligned}$$
(2.7)
$$\begin{aligned}&g(A_{\xi }^{*}X,Y)=\tilde{g}(h^{*}(X,Y),\xi ), \end{aligned}$$
(2.8)

for any \(\xi \in \Gamma (TM^{n\perp })\) and \(X,Y\in \Gamma (TM^{n}).\) Further, see [24], the corresponding Weingarten formulas are

$$\begin{aligned}&\tilde{\nabla }_{X}\xi =-A_{\xi }^{*}X+\nabla _{X}^{\perp }\xi ,\end{aligned}$$
(2.9)
$$\begin{aligned}&\tilde{\nabla }_{X}^{*}\xi =-A_{\xi }X+\nabla _{X}^{*\perp }\xi , \end{aligned}$$
(2.10)

for any \(\xi \in \Gamma (TM^{n\perp })\) and \(X\in \Gamma (TM^{n}).\) The connections \(\nabla _{X}^{\perp }\) and \(\nabla _{X}^{*\perp }\) given by (2.9) and (2.10) are Riemannian dual connections with respect to induced metric on \(\Gamma (TM^{n\perp }).\)

Let \(\{e_{1},\ldots ,e_{n}\}\) and \(\{\xi _{1},\ldots ,\xi _{k}\}\) be orthonormal tangent and normal frames, respectively, on \(M^{n}\). Then the mean curvature vector fields are defined by

$$\begin{aligned} H=\frac{1}{n}\sum _{i=1}^{n}h\left( e_{i},e_{i}\right) =\frac{1}{n} \sum _{\alpha =1}^{k}\left( \sum _{i=1}^{n}h_{ii}^{\alpha }\right) \xi _{\alpha },\quad h_{ij}^{\alpha }=\tilde{g}\left( h\left( e_{i},e_{j}\right) ,\xi _{\alpha }\right) , \end{aligned}$$
(2.11)

and

$$\begin{aligned} H^{*}=\frac{1}{n}\sum _{i=1}^{n}h^{*}\left( e_{i},e_{i}\right) =\frac{ 1}{n}\sum _{\alpha =1}^{k}\left( \sum _{i=1}^{n}h_{ii}^{*\alpha }\right) \xi _{\alpha },\quad h_{ij}^{*\alpha }=\tilde{g}\left( h^{*}\left( e_{i},e_{j}\right) ,\xi _{\alpha }\right) ,\qquad \end{aligned}$$
(2.12)

for \(1\le i,j\le n\) and \(1\le \alpha \le k\) (see also [6]).

The corresponding Gauss, Codazzi and Ricci equations are given by the following result.

Proposition 2.1

[24] Let \(\tilde{\nabla }\) and \(\tilde{\nabla }^*\) be dual connections on \(\tilde{M}^{n+k}\) and \(\nabla \) the induced connection by \(\tilde{\nabla }\) on \(M^{n}.\) Let \(\tilde{R}\) and R be the Riemannian curvature tensors for \(\tilde{\nabla }\) and \(\nabla ,\) respectively. Then,

$$\begin{aligned} \tilde{g}(\tilde{R}(X,Y)Z,W)= & {} g(R(X,Y)Z,W)+\tilde{g}(h(X,Z),h^{*}(Y,W))\nonumber \\&-\tilde{g}(h^{*}(X,W),h(Y,Z)),\\ (\tilde{R}(X,Y)Z)^{\perp }= & {} \nabla _{X}^{\perp }h(Y,Z)-h(\nabla _{X}Y,Z)-h(Y,\nabla _{X}Z)\nonumber \\&-\{ \nabla _{Y}^{\perp }h(X,Z)-h(\nabla _{Y}X,Z) -h(X,\nabla _{Y}Z)\},\nonumber \end{aligned}$$
(2.13)
$$\begin{aligned} \tilde{g}(R^{\perp }(X,Y)\xi ,\eta )= & {} \tilde{g} (\tilde{R}(X,Y)\xi ,\eta )+g([A_{\xi }^{*},A_{\eta }]X,Y), \end{aligned}$$
(2.14)

where \(R^{\perp }\) is the Riemannian curvature tensor of \(\nabla ^\perp \) on \( TM^{n\perp },\xi ,\eta \in \Gamma ( TM^{n\perp })\) and \([A_{\xi }^{*},A_{\eta }]=A_{\xi }^{*}A_{\eta }-A_{\eta }A_{\xi }^{*}.\)

For the equations of Gauss, Codazzi and Ricci with respect to the connection \(\tilde{\nabla }^{*}\) on \(M^{n}\), we have

Proposition 2.2

[24]  Let \(\tilde{\nabla }\) and \(\tilde{\nabla }^*\) be dual connections on \(\tilde{M}^{n+k}\) and \(\nabla ^{*}\) the induced connection by \(\tilde{\nabla }^*\) on \(M^{n}.\) Let \(\tilde{R}^{*}\) and \(R^{*}\) be the Riemannian curvature tensors for \(\tilde{\nabla }^{*}\) and \(\nabla ^{*},\) respectively. Then, 

$$\begin{aligned} \tilde{g}(\tilde{R}^{*}(X,Y) Z,W)= & {} g(R^{*}(X,Y) Z,W)+\tilde{g}(h^{*}(X,Z),h(Y,W))\nonumber \\&-\tilde{g}(h(X,W),h^{*}(Y,Z)),\\ (\tilde{R}^{*}(X,Y)Z)^{\perp }= & {} \nabla _{X}^{*\perp }h^{*} (Y,Z)-h^{*}(\nabla _{X}^{*}Y,Z)-h^{*} (Y,\nabla _{X}^{*}Z)\nonumber \\&-\{\nabla _{Y}^{*\perp }h^{*}(X,Z)-h^{*}(\nabla _{Y}^{*}X,Z)-h^{*}(X,\nabla _{Y}^{*}Z) \},\nonumber \end{aligned}$$
(2.15)
$$\begin{aligned} \tilde{g}(R^{*\perp }(X,Y)\xi ,\eta )= & {} \tilde{g} (\tilde{R}^{*}(X,Y)\xi ,\eta )+g([A_{\xi },A_{\eta }^{*}]X,Y), \end{aligned}$$
(2.16)

where \(R^{*\perp }\) is the Riemannian curvature tensor of \(\nabla ^{\perp *}\) on \(TM^{n\perp }, \xi ,\eta \in \Gamma ( TM^{n\perp })\) and \([A_{\xi },A_{\eta }^{*}]=A_{\xi }A_{\eta }^{*}-A_{\eta }^{*}A_{\xi }.\)

Geometric inequalities for statistical submanifolds in statistical manifolds with constant curvature were obtained in [2].

3 Statistical surfaces in statistical manifolds of constant curvature

Let \((\tilde{M}^{3},\tilde{g})\) be a 3-dimensional statistical manifold of constant curvature c and \(M^{2}\) a surface of \(\tilde{M}.\) Denote the Gauss curvature, the mean curvature and the dual mean curvature of M,  by GH and \(H^{*},\) respectively. In [3], a version of the Euler inequality for statistical surfaces was given.

Proposition 3.1

[3] Let \(M^{2}\) be a surface in a 3-dimensional statistical manifold of constant curvature c. Then its Gauss curvature satisfies : 

$$\begin{aligned} G\le 2\Vert H\Vert \cdot \Vert H^{*}\Vert -c. \end{aligned}$$
(3.1)

Some examples of statistical surfaces satisfying the equality case of the above Euler inequality can be provided by the following.

Example 1

(A trivial example) Recall Lemma 5.3 of Furuhata [8].

Let \((\mathbb {H},\tilde{\nabla },\tilde{g})\) be a Hessian manifold of constant Hessian sectional curvature \(\tilde{c}\ne 0,\) \(( M,\nabla ,g)\) a trivial Hessian manifold and \(f:M\longrightarrow \mathbb {H}\) a statistical immersion of codimension one. Then one has:

$$\begin{aligned} A^{*}=0,\quad h^{*}=0,\quad \Vert H^{*}\Vert =0. \end{aligned}$$

Thus, if dim \(M=2\), the immersion f of codimension one satisfies the equality case of the statistical version of the Euler inequality given by Proposition 3.1.

Example 2

Let \((\mathbb {H}^{3},\tilde{g})\) be the upper half space of constant sectional curvature \(-1\), i.e.,

$$\begin{aligned} \mathbb {H}^{3}=\{ y=(y^{1},y^{2},y^{3}) \in \mathbb {R}^{3}:y^{3}>0\},\quad \tilde{g}=(y^{3})^{-2}\sum _{k=1}^{3}dy^{k}dy^{k}. \end{aligned}$$

An affine connection \(\tilde{\nabla }\) on \(\mathbb {H}^3\) is given by

$$\begin{aligned} \tilde{\nabla }_{\frac{\partial }{\partial y^{3}}}\frac{\partial }{\partial y^{3}}=(y^{3})^{-1}\frac{\partial }{\partial y^{3}},\quad \tilde{\nabla }_{\frac{\partial }{\partial y^{i}}}\frac{\partial }{\partial y^{j}}=2\delta _{ij}(y^{3}) ^{-1}\frac{\partial }{\partial y^{3}} ,\quad \tilde{\nabla }_{\frac{\partial }{\partial y^{i}}}\frac{\partial }{ \partial y^{3}}=\tilde{\nabla }_{\frac{\partial }{\partial y^{3}}}\frac{ \partial }{\partial y^{j}}=0, \end{aligned}$$

where \(i,j=1,2.\) The curvature tensor field \(\tilde{R}\) of \(\tilde{\nabla }\) is identically zero, i.e., \(c=0\). Thus \(( \mathbb {H}^{3},\tilde{\nabla }, \tilde{g})\) is a Hessian manifold of constant Hessian sectional curvature 4 (see [21]).

Now let consider a horosphere \(M^{2}\) in \(\mathbb {H}^{3}\) having null Gauss curvature, i.e., \(G\equiv 0\) (for details, see [11]). If \( f:M^{2}\longrightarrow \mathbb {H}^{3}\) is a statistical immersion of codimension one, then, by using Lemma 4.1 of [16], we deduce \(A^{*}=0\), and then \(H^{*}=0.\) This implies that the horosphere \(M^{2}\) satisfies the equality case of the statistical version of the Euler inequality given by Proposition 3.1.

More generally, let consider a 4-dimensional statistical manifold of constant curvature c, i.e. \((\tilde{M}^{4},c),\) and a surface \(M^{2}\) of \(\tilde{M}^{4}.\) We respectively denote the Gauss curvature, the normal curvature and the Gauss curvature with respect to the Levi-Civita connection by G\(G^{\perp }\) and \(G^{0}.\) Similarly, we respectively denote the mean vector field, the dual mean curvature and the sectional curvature with respect to the Levi-Civita connection by H\(H^{*}\) and \(\tilde{K}^{0}.\) We have the following Wintgen inequalities.

Theorem 3.2

[3] Let \(M^{2}\) be a statistical surface in a 4-dimensional statistical manifold \((\tilde{M}^{4},c)\) of constant curvature c. Then

$$\begin{aligned} G+\vert G^{\perp }\vert +2G^{0}\le \frac{1}{2}(\Vert H\Vert ^{2}+\Vert H^{*}\Vert ^{2})-c+2\tilde{K}^{0}(e_{1}\wedge e_{2}). \end{aligned}$$

In particular,  for \(c=0\) we derive the following.

Corollary 3.3

[3] Let \(M^{2}\) be a statistical surface of a Hessian 4-dimensional statistical manifold \(\tilde{M}^{4}\) of Hessian curvature 0. Then : 

$$\begin{aligned} G+\vert G^{\perp }\vert +2G^{0}\le \frac{1}{2}(\Vert H\Vert ^{2}+\Vert H^{*}\Vert ^{2}). \end{aligned}$$

4 Wintgen inequality for statistical submanifolds

Let \(M^{n}\) be an n-dimensional statistical submanifold of a \((n+m)\)-dimensional statistical manifold \((\tilde{M}^{n+m},c) \) of constant curvature c.

The sectional curvature K on \(M^{n}\) is defined by [3] (see also [18, 19])

$$\begin{aligned} K(X\wedge Y)=\frac{1}{2}[g(R( X,Y)X,Y)+g(R^{*}(X,Y)X,Y)], \end{aligned}$$

for any orthonormal vectors \(X,Y\in T_{p}M^{n},\) \(p\in M^{n}.\)

In the case of the Levi-Civita connection, the above definition coincides (up to the sign) to the standard definition of the sectional curvature.

Let \(p\in M^{n}\) and \(\{e_{1},e_{2},\ldots ,e_{n}\}\) an orthonormal basis of \(T_{p}M^{n}.\) Then the normalized scalar curvature \(\rho \) is defined by (see [7]):

$$\begin{aligned} \rho= & {} \frac{2}{n\left( n-1\right) }\sum _{1\le i<j\le n}K\left( e_{i}\wedge e_{j}\right) \\= & {} \frac{1}{n\left( n-1\right) }\sum _{1\le i<j\le n}\left[ g\left( R\left( e_{i},e_{j}\right) e_{i},e_{j}\right) +g\left( R^{*}\left( e_{i},e_{j}\right) e_{i},e_{j}\right) \right] . \end{aligned}$$

By using the Gauss equations for the dual connections \(\tilde{\nabla }\) and \( \tilde{\nabla }^{*}\), respectively, we obtain

$$\begin{aligned} \rho= & {} \frac{1}{n\left( n-1\right) }\sum _{1\le i<j\le n}\left[ -c-g\left( h\left( e_{i},e_{i}\right) ,h^{*}\left( e_{j},e_{j}\right) \right) +g\left( h^{*}\left( e_{i},e_{j}\right) ,h\left( e_{i},e_{j}\right) \right) \right. \\&\left. -\,c-g\left( h^{*}\left( e_{i},e_{i}\right) ,h\left( e_{j},e_{j}\right) \right) +g\left( h\left( e_{i},e_{j}\right) ,h^{*}\left( e_{i},e_{j}\right) \right) \right] . \end{aligned}$$

Denoting as usual by

$$\begin{aligned} h_{ij}^{r}= & {} g\left( h\left( e_{i},e_{j}\right) ,\xi _{r}\right) ,\quad h_{ij}^{*r}=g\left( h^{*}\left( e_{i},e_{j}\right) ,\xi _{r}\right) , \qquad \\&\forall i,j=1,\ldots ,n\text { and }r=1,\ldots ,m, \end{aligned}$$

the above equation becomes

$$\begin{aligned} \rho =-c+\frac{1}{n\left( n-1\right) }\sum _{r=1}^{m}\sum _{1\le i<j\le n}\left( 2h_{ij}^{r}h_{ij}^{*r}-h_{ii}^{*r}h_{jj}^{r}-h_{ii}^{r}h_{jj}^{*r}\right) . \end{aligned}$$
(4.1)

On the other hand, the normalized normal scalar curvature \(\rho ^{\perp }\) is defined by (see also [3]):

$$\begin{aligned} \rho ^{\perp }=\frac{1}{n\left( n-1\right) }\left\{ \sum _{1\le r<s\le m}\sum _{1\le i<j\le n}\left[ g\left( R^{\perp }\left( e_{i},e_{j}\right) \xi _{r},\xi _{s}\right) +g\left( R^{*\perp }\left( e_{i},e_{j}\right) \xi _{r},\xi _{s}\right) \right] ^{2}\right\} ^{\frac{1}{2}}. \end{aligned}$$

The Ricci equations for the dual connections \(\tilde{\nabla },\) and \(\tilde{ \nabla }^{*}\), respectively, imply

$$\begin{aligned} \rho ^{\perp }=\frac{1}{n\left( n-1\right) }\left\{ \sum _{1\le r<s\le m}\sum _{1\le i<j\le n}\left[ g\left( \left[ A_{\xi _{r}}^{*},A_{\xi _{s}}\right] e_{i},e_{j}\right) +g\left( \left[ A_{\xi _{r}},A_{\xi _{s}}^{*}\right] e_{i},e_{j}\right) \right] ^{2}\right\} ^{\frac{1}{2}} \end{aligned}$$

or equivalently,

$$\begin{aligned} \rho ^{\perp }=\frac{1}{n\left( n-1\right) }\left\{ \sum _{1\le r<s\le m}\sum _{1\le i<j\le n}\left[ \sum _{k=1}^{n}\left( h_{ik}^{s}h_{jk}^{*r}-h_{ik}^{*r}h_{jk}^{s}+h_{ik}^{*s}h_{jk}^{r}-h_{ik}^{r}h_{jk}^{*s}\right) \right] ^{2}\right\} ^{\frac{1 }{2}}. \end{aligned}$$

It follows that

$$\begin{aligned} \rho ^{\perp }= & {} \frac{1}{n\left( n-1\right) }\left\{ \sum _{1\le r<s\le m}\sum _{1\le i<j\le n}\left[ \sum _{k=1}^{n}\left( \left( h_{ik}^{s}+h_{ik}^{*s}\right) \left( h_{jk}^{r}+h_{jk}^{*r}\right) -h_{ik}^{s}h_{jk}^{r}-h_{ik}^{*s}h_{jk}^{*r}\right. \right. \right. \\&\left. \left. \left. -\left( h_{ik}^{r}+h_{ik}^{*r}\right) \left( h_{jk}^{s}+h_{jk}^{*s}\right) +h_{ik}^{r}h_{jk}^{s}+h_{ik}^{*r}h_{jk}^{*s}\right) \phantom {\sum _{1\le r<s\le m}\sum _{1\le i<j\le n}}\right] ^{2}\right\} ^{\frac{1}{2}}. \end{aligned}$$

It is known that the components of the second fundamental form \(h^0\) of \(M^n\) with respect to the Levi-Civita connection \(\tilde{\nabla }^0\) are given by \(2h_{ik}^{0r}=h_{ik}^{r}+h_{ik}^{*r},\) \(\forall i,k=1,\ldots ,n,\) \(r=1,\ldots ,m.\) Then we can write

$$\begin{aligned} \rho ^{\perp }= & {} \frac{1}{n\left( n-1\right) }\left\{ \sum _{1\le r<s\le m}\sum _{1\le i<j\le n}\left[ \sum _{k=1}^{n}\left( 4\left( h_{ik}^{0s}h_{jk}^{0r}-h_{ik}^{0r}h_{jk}^{0s}\right) +\left( h_{ik}^{r}h_{jk}^{s}-h_{ik}^{s}h_{jk}^{r}\right) \right. \right. \right. \nonumber \\&\left. \left. \left. ~~~~~~~~~~~~~~~~~+\left( h_{ik}^{*r}h_{jk}^{*s}-h_{ik}^{*s}h_{jk}^{*r}\right) \right) \right] ^{2}\right\} ^{\frac{1}{2}}. \end{aligned}$$
(4.2)

We shall use the algebraic inequality

$$\begin{aligned} \left( a+b+c\right) ^{2}\le 3(a^{2}+b^{2}+c^{2}),\quad \forall a,b,c\in \mathbb {R}. \end{aligned}$$

Therefore

$$\begin{aligned} \rho ^{\perp }\le & {} \frac{3}{n\left( n-1\right) }\left\{ \sum _{1\le r<s\le m}\sum _{1\le i<j\le n}\left( 16\left[ \sum _{k=1}^{n}\left( h_{ik}^{0s}h_{jk}^{0r}-h_{ik}^{0r}h_{jk}^{0s}\right) \right] ^{2}\right. \right. \nonumber \\&\left. \left. +\left[ \sum _{k=1}^{n}\left( h_{ik}^{r}h_{jk}^{s}-h_{ik}^{s}h_{jk}^{r}\right) \right] ^{2}+\left[ \sum _{k=1}^{n}\left( h_{ik}^{*r}h_{jk}^{*s}-h_{ik}^{*s}h_{jk}^{*r}\right) \right] ^{2}\right) \right\} ^{\frac{1}{2}}.\qquad \end{aligned}$$
(4.3)

Recall an inequality from [12] (see also [14])

$$\begin{aligned}&\sum _{r=1}^{m}\sum _{1\le i<j\le n}\left( h_{ii}^{r}-h_{jj}^{r}\right) ^{2}+2n\sum _{r=1}^{m}\sum _{1\le i<j\le n}\left( h_{ij}^{r}\right) ^{2}\\&\quad \ge 2n\left[ \sum _{1\le i<j\le n}\sum _{1\le r<s\le m}\left( \sum _{k=1}^n(h_{jk}^{r}h_{ik}^{s}-h_{ik}^{r}h_{jk}^{s}) \right) ^{2}\right] ^{\frac{1}{2}}. \end{aligned}$$

Similarly, we have

$$\begin{aligned}&\sum _{r=1}^{m}\sum _{1\le i<j\le n}\left( h_{ii}^{*r}-h_{jj}^{*r}\right) ^{2}+2n\sum _{r=1}^{m}\sum _{1\le i<j\le n}\left( h_{ij}^{*r}\right) ^{2}\\&\quad \ge 2n\left[ \sum _{1\le i<j\le n}\sum _{1\le r<s\le m}\left( \sum _{k=1}^n(h_{jk}^{*r}h_{ik}^{*s}-h_{ik}^{*r}h_{jk}^{*s})\right) ^{2} \right] ^{\frac{1}{2}} \end{aligned}$$

and

$$\begin{aligned}&\sum _{r=1}^{m}\sum _{1\le i<j\le n}\left( h_{ii}^{0r}-h_{jj}^{0r}\right) ^{2}+2n\sum _{r=1}^{m}\sum _{1\le i<j\le n}\left( h_{ij}^{0r}\right) ^{2}\\&\quad \ge 2n\left[ \sum _{1\le i<j\le n}\sum _{1\le r<s\le m}\left( \sum _{k=1}^n(h_{jk}^{0r}h_{ik}^{0s}-h_{ik}^{0r}h_{jk}^{0s})\right) ^{2}\right] ^{\frac{1 }{2}}. \end{aligned}$$

Summing up the above three inequalities, from (4.3) we obtain

$$\begin{aligned} \rho ^{\perp }\le & {} \frac{3}{2n^{2}\left( n-1\right) }\sum _{r=1}^{m}\sum _{1 \le i<j\le n}\left[ \left( h_{ii}^{r}-h_{jj}^{r}\right) ^{2}+\left( h_{ii}^{*r}-h_{jj}^{*r}\right) ^{2}+16\left( h_{ii}^{0r}-h_{jj}^{0r}\right) ^{2}\right] \nonumber \\&+\frac{3}{n\left( n-1\right) }\sum _{r=1}^{m}\sum _{1\le i<j\le n}\left[ \left( h_{ij}^{r}\right) ^{2}+\left( h_{ij}^{*r}\right) ^{2}+16\left( h_{ij}^{0r}\right) ^{2}\right] . \end{aligned}$$
(4.4)

Also, we can write

$$\begin{aligned} n^{2}\left\| H\right\| ^{2}=\sum _{r=1}^{m}\left( \sum _{i=1}^{n}h_{ii}^{r}\right) ^{2}=\frac{1}{n-1}\sum _{r=1}^{m}\sum _{1\le i<j\le n}\left( h_{ii}^{r}-h_{jj}^{r}\right) ^{2}+\frac{2n}{n-1} \sum _{r=1}^{m}\sum _{1\le i<j\le n}h_{ii}^{r}h_{jj}^{r} \end{aligned}$$

and similarly,

$$\begin{aligned} n^{2}\left\| H^{*}\right\| ^{2}=\frac{1}{n-1}\sum _{r=1}^{m}\sum _{1\le i<j\le n}\left( h_{ii}^{*r}-h_{jj}^{*r}\right) ^{2}+\frac{2n}{n-1}\sum _{r=1}^{m}\sum _{1\le i<j\le n}h_{ii}^{*r}h_{jj}^{*r} \end{aligned}$$

and

$$\begin{aligned} n^{2}\left\| H^{0}\right\| ^{2}=\frac{1}{n-1}\sum _{r=1}^{m}\sum _{1\le i<j\le n}\left( h_{ii}^{0r}-h_{jj}^{0r}\right) ^{2}+\frac{2n}{n-1} \sum _{r=1}^{m}\sum _{1\le i<j\le n}h_{ii}^{0r}h_{jj}^{0r}. \end{aligned}$$

Substituting in (4.4), we get

$$\begin{aligned} \rho ^{\perp }\le & {} \frac{3}{2}\left\| H\right\| ^{2}+\frac{3}{2}\left\| H^{*}\right\| ^{2}+24\left\| H^{0}\right\| ^{2} \\&-\frac{3}{n(n-1)}\sum _{r=1}^{m}\sum _{1\le i<j\le n}\left( h_{ii}^{r}h_{jj}^{r}+h_{ii}^{*r}h_{jj}^{*r}+16h_{ii}^{0r}h_{jj}^{0r}\right) \\&+\frac{3}{n(n-1)}\sum _{r=1}^{m}\sum _{1\le i<j\le n}\left[ \left( h_{ij}^{r}\right) ^{2}+\left( h_{ij}^{*r}\right) ^{2}+16\left( h_{ij}^{0r}\right) ^{2}\right] \\= & {} \frac{3}{2}\left\| H\right\| ^{2}+\frac{3}{2}\left\| H^{*}\right\| ^{2}+24\left\| H^{0}\right\| ^{2} \\&-\frac{3}{n(n-1)}\sum _{r=1}^{m}\sum _{1\le i<j\le n}\left[ \left( h_{ii}^{r}+h_{ii}^{*r}\right) \left( h_{jj}^{r}+h_{jj}^{*r}\right) -h_{ii}^{*r}h_{jj}^{r}-h_{ii}^{r}h_{jj}^{*r}+16h_{ii}^{0r}h_{jj}^{0r} \right] \\&+\frac{3}{n(n-1)}\sum _{r=1}^{m}\sum _{1\le i<j\le n}\left[ \left( h_{ij}^{r}+h_{ij}^{*r}\right) ^{2}-2h_{ij}^{r}h_{ij}^{*r}+16\left( h_{ij}^{0r}\right) ^{2}\right] . \end{aligned}$$

Using again \(2h_{ij}^{0r}=h_{ij}^{r}+h_{ij}^{*r},\) \(\forall i,j=1,\ldots ,n,\) \(r=1,\ldots ,m,\) we obtain

$$\begin{aligned} \rho ^{\perp }\le & {} \frac{3}{2} \left\| H\right\| ^{2}+\frac{3}{2}\left\| H^{*}\right\| ^{2}+24\left\| H^{0}\right\| ^{2}\nonumber \\&-\frac{3}{n(n-1)}\sum _{r=1}^{m}\sum _{1\le i<j\le n}\left[ 20h_{ii}^{0r}h_{jj}^{0r}-h_{ii}^{*r}h_{jj}^{r}-h_{ii}^{r}h_{jj}^{*r}-20(h_{ij}^{0r})^{2}+2h_{ij}^{r}h_{ij}^{*r}\right] .\nonumber \\ \end{aligned}$$
(4.5)

Substituting (4.1) in (4.5), one leads to

$$\begin{aligned} \rho ^{\perp }\le & {} \frac{3}{2}\left\| H\right\| ^{2}+\frac{3}{2}\left\| H^{*}\right\| ^{2}+ 24\left\| H^{0}\right\| ^{2}-3\rho -3c\nonumber \\&-\,\frac{60}{n\left( n-1\right) }\sum _{r=1}^{m}\sum _{1\le i<j\le n}\left[ h_{ii}^{0r}h_{jj}^{0r}-\left( h_{ij}^{0r}\right) ^{2}\right] . \end{aligned}$$
(4.6)

If we denote by

$$\begin{aligned} \tilde{\rho }^{0}=\frac{2}{n\left( n-1\right) }\sum _{1\le i<j\le n}\tilde{R} ^{0}\left( e_{i},e_{j},e_{i},e_{j}\right) , \end{aligned}$$

the Gauss equation for the Levi-Civita connection \(\tilde{\nabla }^{0}\) gives

$$\begin{aligned} \tilde{\rho }^{0}=\rho ^0 -\frac{2}{n(n-1)}\sum _{r=1}^{m}\sum _{1\le i<j\le n}\left[ h_{ii}^{0r}h_{jj}^{0r}-\left( h_{ij}^{0r}\right) ^{2}\right] . \end{aligned}$$
(4.7)

From (4.6) and (4.7) we obtain

$$\begin{aligned} \rho ^{\perp }\le \frac{3}{2}\left\| H\right\| ^{2}+\frac{3}{2}\left\| H^{*}\right\| ^{2}+ 24\left\| H^{0}\right\| ^{2}-3\rho -3c+ 30\left( \tilde{\rho }^{0}-\rho ^{0}\right) . \end{aligned}$$

Summarizing, we proved the following generalized Wintgen inequality.

Theorem 4.1

Let \(M^{n}\) be a submanifold in a statistical manifold \((\tilde{M}^{n+m},c)\) of constant curvature c. Then

$$\begin{aligned} \rho ^{\perp }+3\rho \le \frac{15}{2}\Vert H\Vert ^{2}+\frac{15}{2}\Vert H^{*}\Vert ^{2}+ 12g(H,H^{*})-3c+30(\tilde{\rho }^{0}-\rho ^{0}). \end{aligned}$$