Keywords

2010 Mathematics Subject Classification

5.1 Introduction

K. Yano [21] introduced in 1963 the notion of f-structure on a \((2m+s)\)-dimensional manifold as a tensor field f of type (1, 1) and rank 2m satisfying \(f^3+f=0\). Almost complex (\(s=0\)) and almost contact (\(s=1\)) structures are well-known examples of f-structures. A Riemannian manifold endowed with an f-structure (\(s\ge 2\)) compatible with the Riemannian metric is called a metric f-manifold (for \(s=0\) we have almost Hermitian manifolds and for \(s=1\), metric almost contact manifolds). In this context, D.E. Blair [5] defined K-manifolds (and particular cases of S-manifolds and C-manifolds) as the analogue of Kaehlerian manifolds in the almost complex geometry and of quasi-Sasakian manifolds (and particular cases of Sasakian manifolds and cosymplectic manifolds) in the almost contact geometry.

He also showed that the curvature of S-manifolds is completely determined by their f-sectional curvatures. Later, M. Kobayashi and S. Tsuchiya [15] got expressions for the curvature tensor field of S-manifolds when their f-sectional curvature is constant depending on such a constant. Such spaces are called S-space-forms and they generalize complex and Sasakian space-forms. Nice examples of S-space-forms can be found in [5, 6, 8, 13]. In particular, it is proved in [5, 8] that certain principal toroidal bundles over complex-space-forms are S-space-forms and a generalization of the Hopf fibration denoted by \(\mathbb H^{2m+s}\) is introduced as a canonical example of such manifolds playing the role of complex projective space in Kaehler geometry and the odd-dimensional sphere in Sasakian geometry [5, 6].

When we want to study the submanifolds of a metric f-manifold, the natural first step is to consider such submanifolds depending on their behavior with respect to the f-structure. So, invariant and anti-invariant submanifolds (in the terminology of the complex geometry, holomorphic and totally real submanifolds) appear if all the tangent vector fields to the submanifold are transformed by f into tangent vector fields or into normal vector fields. But since an hypersurface of a metric f-manifold tangent to the structure vector fields is neither invariant nor anti-invariant, it is necessary to introduce a wider class of submanifolds: the CR-submanifolds. This work was made firstly by A. Bejancu and B.-Y. Chen [1, 10, 11] in the case \(s=0\) and by A. Bejancu and N. Papaghiuc, M. Kobayashi and K. Yano and M. Kon in the case \(s=1\) (we refer to the books [3, 22] for the background of these cases where a large list of fundamental references can be found). For \(s\ge 2\), I. Mihai [16] introduced the notion of CR-submanifold in a natural way.

Many authors have studied the geometry of submanifolds of locally conformal almost Kaehler (l.c.a.K.) manifolds [10, 11, 14, 20], which are almost Hermitian manifolds \((\widetilde{M},J,g)\) such that every \(x\in \widetilde{M}\) has an open neighborhood U such that for some differentiable function \(h:U\longrightarrow \mathbb R,\tilde{g}_U=e^{-h}g|_{U}\) is a (l.c.a.) Kaehler metric on U. If one can take \(U=\widetilde{M}\), the manifold is then called globally conformal almost Kaeler (g.c.a.K) manifold. Examples of l.c.K. manifolds are provided by the Hopf manifolds. So, it seems interesting to study CR-submanifolds of l.c.a.K. manifolds.

On the other hand, M. Okumura [17, 18] studied normal real hypersurfaces of Kaehlerian manifolds and obtained nice properties. For this reason, it also seems interesting to introduce and study normal CR-submanifolds. In the cases \(s=0\) and \(s=1\), the papers [2] and [4] can be consulted.

The aim of the present work is to briefly summarize our contributions to the study of CR-submanifolds of l.c.a.K. manifolds, normal CR-submanifolds of S-manifolds. To this end, we separate them into two different sections, which can be read independently.

5.2 CR-Submanifolds of (l.c.a.) Kaehler Manifolds

Let \((\widetilde{M},J,g)\) be an almost Hermitian manifold (\(dim(\widetilde{M})=2m\)) with almost complex structure J and Hermitian metric g and let M be a Riemannian submanifold isometrically immersed in \(\widetilde{M}.\)

A. Bejancu [1] introduced the notion of a CR-submanifold of \(\widetilde{M}.\) In fact, M is a CR-submanifold of the almost Hermitian manifold \(\widetilde{M}\) if there exists on M a differentiable holomorphic distribution \(\mathcal {D}\), i.e., \(J(\mathcal {D}_x)\subseteq \mathcal {D}_x\) for any \(x\in M\) such that its orthogonal complement \(\mathcal {D}^{\bot }\) in M is totally real in \(\widetilde{M}\), i.e., \(J(\mathcal {D}_x^{\bot })\subseteq T^{\perp }_x(M)\) for any \(x\in M,\) where \(T^{\perp }_x(M)\) is the normal space at x. If \(dim(\mathcal {D})=0,\) M is called a totally real submanifold, and if \(dim(\mathcal {D}^{\perp })=0\) M is a holomorphic submanifold.

We first discuss the Gauss–Weingarten equations of the submanifold with respect to the metric g and with respect to the local conformal Kaehler metrics and then we shall establish thereby the analytical conditions that characterize the important types of submanifolds.

5.2.1 Preliminaries

Let \((\widetilde{M},J,g)\) be an almost Hermitian manifold. It is easy to see [20] that \((\widetilde{M},J,g)\) is a l.c.(a).K. manifold if and only if there is a global closed 1-form \(\omega \) on \(\widetilde{M}\) (the Lee form) such that \(d\Omega =\omega \wedge \Omega \) (\(\Omega \) the fundamental form of the manifold) and \((\widetilde{M},J,g)\) is a g.c.(a).K. manifold if and only if \(\omega \) is also exact. In case \(\omega =0\), the manifold is an (almost) Kaehler manifold.

Let \((\widetilde{M},J,g)\) be a l.c.(a).K. manifold and consider the Lee vector field B [20] of \((\widetilde{M},J,g)\) defined by \(g(X,B)=\omega (X).\) Denote by \(\widetilde{\nabla }\) the Levi-Civita connection of g and define

$$\begin{aligned} \overline{\nabla }_XY=\widetilde{\nabla }_XY - \frac{1}{2} \omega (X)Y-\frac{1}{2}\omega (Y)X+\frac{1}{2}g(X,Y)B. \end{aligned}$$
(5.1)

Then \(\overline{\nabla }\) is a torsionless linear connection on \(\widetilde{M}\) which is called the Weyl connection of g. It is easy to see that \(\overline{\nabla }_Xg=\omega (X)g\). We have

Theorem 5.1

([20]) The almost Hermitian manifold \((\widetilde{M},J,g)\) is a l.cK. manifold if and only if there is a closed 1-form \(\omega \) on \(\widetilde{M}\) such that the Weyl connection is almost complex, That is, \(\overline{\nabla }J=0.\)

Let \((\widetilde{M},J,g)\) be a l.c.K. manifold and M a Riemannian manifold isometrically immersed in \(\widetilde{M}.\) We denote by g the metric tensor of \(\widetilde{M}\) as well as that induced on M,  and let \(\nabla ,\) \(\nabla ^M\) be the covariant derivations on M induced by \(\widetilde{\nabla }\) and \(\overline{\nabla }\), respectively. Then, the Gauss–Weingarten formulas for M with respect to \(\widetilde{\nabla }\) and \(\overline{\nabla }\) are given by

$$\begin{aligned} \widetilde{\nabla }_XY=\nabla _XY+\sigma (X,Y), \; \widetilde{\nabla }_XV = -A_{V}X+D_XV, \end{aligned}$$
(5.2)
$$\begin{aligned} \overline{\nabla }_XY=\nabla ^M_XY+\overline{\sigma } (X,Y), \; \overline{\nabla }_XV = -\overline{A}_{V}X+\overline{D}_XV, \end{aligned}$$
(5.3)

for any vector fields XY tangent to M and V normal to M,  where \(\sigma \) (respectively, \(\overline{\sigma }\)) is the second fundamental form of M with respect to \(\widetilde{\nabla } (\overline{\nabla })\) and D (respectively, \(\overline{D}\)) is the normal connection. The formulas (5.3) are the Gauss–Weingarten equations of \(M|_U\) in \((\widetilde{M}|_U,e^{-h}g|_U).\) The second fundamental tensors \(A_{V}, \overline{A}_{V}\) are related to \(\sigma , \overline{\sigma }\) respectively by

$$\begin{aligned} g(A_{V}X,Y)=g(\sigma (X,Y),V), \, g(\overline{A}_{V}X,Y)=g(\overline{\sigma }(X,Y),V). \end{aligned}$$
(5.4)

For any vector X tangent to M and V normal to M write

$$\begin{aligned} JX=TX+NX, \; JV =tV + nV, \end{aligned}$$
(5.5)

where TX and NX (respectively, tV and nV) are the tangential and normal component of J(X) (respectively JV). For the Lee field B, we have

$$\begin{aligned} B_x=(B_x)_1+(B_x)_2, \quad x\in M, \end{aligned}$$
(5.6)

where \((B_x)_1\) (resp.y \((B_x)_2)\) is the tangential (resp. normal) component of \(B_x.\)

If M is a CR-submanifold of an almost Hermitian manifold \((\widetilde{M},J,g)\) let us denote by \(\nu \) the complementary orthogonal subbundle of \(J\mathcal {D}^\perp \) in \(T^\perp (M).\) Hence we have, \(T^{\perp }(M)=J\mathcal {D}^{\perp } \oplus \nu .\)

5.2.2 Integrability Conditions of the Basic Distributions

First we give some general identities.

Lemma 1

Let M be a CR-submanifold of a l.c.K. manifold \((\widetilde{M},J,g).\) Then, we have

$$\begin{aligned} \nabla ^M_XY=\nabla _XY-\frac{1}{2}\omega (X)Y-\frac{1}{2}\omega (Y)X+\frac{1}{2}g(X,Y)B_1 \end{aligned}$$
(5.7)
$$\begin{aligned} \overline{\sigma }(X,Y)=\sigma (X,Y)+\frac{1}{2}g(X,Y)B_2 \end{aligned}$$
(5.8)
$$\begin{aligned} \overline{A}_{V}X=A_{V}X+\frac{1}{2}\omega (V)X \end{aligned}$$
(5.9)
$$\begin{aligned} \overline{D}_{V}X=D_{V}X-\frac{1}{2}\omega (X)V \end{aligned}$$
(5.10)

for any vector fields XY tangent to M and V normal to M.

Proof

The assertions follow immediately from (5.1)–(5.3).    \(\blacksquare \)

The following result is well known:

Theorem 5.2

([7]) The totally real distribution \(\mathcal {D}^{\perp }\) of any CR-submanifold of a l.c.K. manifold is integrable.

For the holomorphic distribution \(\mathcal {D}\), we have

Theorem 5.3

Let M be a submanifold of a l.c.K. manifold \(\widetilde{M}\) and let \(\mathcal {D}_x\) de maximal holomorphic subspace of \(T_x(M)\) and assume \(dim(\mathcal {D}_x)\) is a constant. Then, the holomorphic distribution \(\mathcal {D}\) is integrable if and only if the second fundamental form \(\overline{\sigma }\) satisfies \(\overline{\sigma }(X,JY)=\overline{\sigma }(JX,Y)\) or, equivalently, \(\sigma (X,JY)-\sigma (JX,Y)+\Omega (X,Y)B_2=0,\) for all vector fields \(X,Y \in \mathcal {D}.\)

If M is a CR-submanifold, the integrability condition on \(\mathcal {D}\) in Theorem 5.3 can be replaced by a weaker condition.

Theorem 5.4

Let M be a CR-submanifold of a l.c.K. manifold \(\widetilde{M}.\) The holomorphic distribution \(\mathcal {D}\) is integrable if and only if

$$g\left( \sigma (X,JY)-\sigma (JX,Y)+\Omega (X,Y)B,J\mathcal {D}^{\perp }\right) =0,$$

for all \(X,Y \in \mathcal {D}.\)

Theorems 5.3 and 5.4 follow easily from similar theorems in the Kaehlerian case ([7]), from (5.8) and the fact that, locally, \(\widetilde{M}\) is endowed with Kaehler metrics \(\widetilde{g}_U\) whose Levi-Civita connection is \(\overline{\nabla }.\)

With regard to integral submanifolds of \(\mathcal {D}^{\perp }\) and \(\mathcal {D}\) (provided \(\mathcal {D}\) is integrable), we have the following theorem.

Theorem 5.5

For a CR-submanifold M of a l.c.K. manifold \(\widetilde{M}\), the leaf \(M^{\perp }\) is totally geodesic in M if and only if

$$g\left( A_{JW}Z+\frac{1}{2}g(Z,W)JB,\mathcal {D}\right) =0,$$

that is,

$$g(\sigma (Z,X),JW)=\frac{1}{2}g(Z,W)\omega (JW),$$

for any \(X\in \mathcal {D},\) \(Z,W\in \mathcal {D}^{\perp }.\)

Proof

From (5.1), (5.2) and \(\overline{\nabla } J=0,\) for any \(X\in \mathcal {D},\) \(Z,W \in \mathcal {D}^{\perp }\), we obtain

$$\begin{aligned} g( J\nabla _ZW,X)+\frac{1}{2}g(Z,W)g(JB,X)=-g(A_{JW}Z,X). \end{aligned}$$
(5.11)

But \(M^{\perp }\) is totally geodesic in M if and only if \(\nabla _ZW\in \mathcal {D}^{\perp }\) for all \(Z,W\in \mathcal {D}^{\perp },\) and then (5.11) gives the theorem.    \(\blacksquare \)

Theorem 5.6

Let M be a CR-submanifold of a l.c.K manifold \(\widetilde{M}.\) If the holomorphic distribution \(\mathcal {D}\) is integrable and \(M^{T}\) is an integral submanifold of \(\mathcal {D},\) then \(M^{T}\) is totally geodesic if and only if

$$g\left( J\sigma (X,Y) +\frac{1}{2}g(X,Y)JB-\frac{1}{2}\Omega (X,Y)B,\mathcal {D}^{\perp } \right) =0,$$

for any \(X,Y \in \mathcal {D}.\)

Proof

From (5.1), (5.3), and \(\overline{\nabla } J=0,\) for any \(X,Y\in \mathcal {D}\) and \(Z\in \mathcal {D}^{\perp }\), we have

$$\begin{aligned} g(J\sigma (X,Y),Z)+\frac{1}{2}g(X,Y)g(JB,Z)=g(\nabla _X(JY),Z)+\frac{1}{2}\Omega (X,Y)g(B,Z).\quad \end{aligned}$$
(5.12)

But \(M^T\) is totally geodesic in M if and only if \(\nabla _XY \in \mathcal {D}\) for all \(X,Y \in \mathcal {D},\) and hence Eq. (5.12) gives the theorem.    \(\blacksquare \)

5.2.3 CR-Submanifolds of l.c.K. Manifolds

First of all, we shall give some identities for later use. Let T, N t, and n be the endomorphisms and vector-valued 1-forms defined in (5.5). The following lemma can be easily obtained from (5.3), (5.9), and \(\overline{\nabla }J=0.\)

Lemma 2

Let M be an isometrically immersed submanifold of a l.c.K. manifold \(\widetilde{M}.\) Then, we have

$$\begin{aligned} \nabla ^M_X(TY)-\overline{A}_{NY}X= T\nabla ^M_XY+t\overline{\sigma }(X,Y) \end{aligned}$$
(5.13)
$$\begin{aligned} \overline{\sigma }(X,TY)+D_X(NY)=N\nabla ^M_XY+n\overline{\sigma }(X,Y), \end{aligned}$$
(5.14)
$$\begin{aligned} \nabla ^M_X(tV)-\overline{A}_{nV}X=-T\overline{A}_{V}X+t\overline{D}_XV, \end{aligned}$$
(5.15)
$$\begin{aligned} \overline{\sigma }(X,tV)+\overline{D}_X(nV)=-N\overline{A}_{V}X+n\overline{D}_XV, \end{aligned}$$
(5.16)
$$\begin{aligned} \left[ \overline{A}_{V},\overline{A}_{\overline{V}} \right] = \left[ A_{V},A_{\overline{V}} \right] , \end{aligned}$$
(5.17)

for any vector fields XY tangent to M and \(V, \, \overline{V}\) normal to M.

Now, we shall study totally umbilical and totally geodesic CR-submanifolds.

Theorem 5.7

Let M be a totally umbilical CR-submanifold of a l.c.K. manifold \(\widetilde{M}.\) Then, we have

  1. (i)

    Either \(dim(\mathcal {D}^{\perp })=1\) or the component \(H_{J(TM)}\) of the mean curvature tensor H in J(TM) is given by \(H_{J(TM)}=-\frac{1}{2}B_2 \).

  2. (ii)

    If \(dim(\mathcal {D}^{\perp })>1\) and M is proper (neither holomorphic nor totally real) such that B is tangent to M,  then M is totally geodesic.

Proof

First, since M is totally umbilical, \(\sigma (X,Y)=g(X,Y)H\) for any XY tangent to M,  and hence

$$\begin{aligned} g(\sigma (X,X),JW)=g(X,X)g(H,JW). \end{aligned}$$
(5.18)

From (5.3) and (5.4) it is easy to see that

$$\begin{aligned} \overline{A}_{JZ}W=\overline{A}_{JW}Z \end{aligned}$$
(5.19)

and, then, if we take an unit vector field \(X=Z\in \mathcal {D}^{\perp }\) orthogonal to W,  (5.9), (5.18), and (5.19) give

$$\begin{aligned} \begin{aligned} g(H,JW)=&g(A_{JW}Z,Z)= g(A_{JZ}W+\frac{1}{2}\omega (JZ)W-\frac{1}{2}\omega ((JW)Z,Z)\\ =&-\frac{1}{2}\omega (JW)= g(-\frac{1}{2}B_2,JW), \end{aligned} \end{aligned}$$

so that (i) holds.

Now, since \(dim(\mathcal {D}^{\perp })>1,\) from (5.5) and assertion (i), we have \(tH=0.\) Thus, (5.15) gives \(t\overline{D}_YH=\overline{A}_{nH}Y-T\overline{A}_HY\), for any Y tangent to M. Therefore, for any Z tangent to M,  from (5.8) and (5.9) we get

$$\begin{aligned} g(t\overline{D}_YH,Z)=-g(\overline{A}_HY,TZ)-g(\overline{\sigma }(Y,Z),nH)=-g(Y,TZ)g(H,H) \end{aligned}$$
(5.20)

and, if we take \(Z=TY\), we have

$$\begin{aligned} -g(Y,T^2Y)g(H,H)=g(t\overline{D}_YH,TY)=g(Tt\overline{D}_YH,Y)=0. \end{aligned}$$
(5.21)

The last equation holds because \(Tt=0\) for any CR-submanifold of an almost Hermitian manifold [22]. Moreover, it is easy to see [22] that \(T^2=-I+tN\) and then (5.21) gives

$$\begin{aligned} g(Y,Y)g(H,H)-g(NY,NY)g(H,H)=0. \end{aligned}$$
(5.22)

Since M is proper, we can choose an unit vector field X in \(\mathcal {D}.\) Thus, \(NX=0\) and from (5.22) we have \(H=0.\)    \(\blacksquare \)

Theorem 5.8

Let M be a totally geodesic CR-submanifold of a l.c.K. manifold \(\widetilde{M}.\) We have

  1. (i)

    If \(B_x \in \mathcal {D}_x\), for all \(x\in M,\) then \(\mathcal {D}\) is integrable and any integral submanifold \(M^T\) of \(\mathcal {D}\) is totally geodesic in \(\widetilde{M}\).

  2. (ii)

    If B is normal to M,  any integral submanifold \(M^{\perp }\) of \(\mathcal {D}^{\perp }\) is totally geodesic in \(\widetilde{M}.\) Furthermore, \(\mathcal {D}\) is integrable if and only if \(B_x \in \nu _x\), for any \(x\in M,\) and in this case any integral submanifold \(M^T\) of \(\mathcal {D}\) is totally geodesic in \(\widetilde{M}.\)

Proof

Firstly, since B is tangent to M,  from Theorem 5.7 the distribution \(\mathcal {D}\) is integrable. Let \(M^T\) be an integral submanifold of \(\mathcal {D}.\) For any vector field X tangent to M\(Y\in \mathcal {D},\) \(Z\in \mathcal {D}^{\perp },\) from (5.3) and (5.4) we get \(g(\nabla ^{M}_XZ,Y)=-g(\overline{\sigma }(X,JY),JZ).\) But from (5.7) and (5.8) we find

$$\begin{aligned}&g(\nabla _XZ,Y)-\frac{1}{2}\omega (Z)g(X,Y)+\frac{1}{2}g(X,Z)g(B,Y) = -g(\sigma (X,JY),JZ)=0. \end{aligned}$$
(5.23)

If \(X\in \mathcal {D}\), (5.23) gives \(g(\nabla _XZ,Y)=0,\) or, equivalently, \(g(\nabla _XY,Z)=0\) and therefore, \(\nabla _XY \in \mathcal {D}.\) Thus \(M^T\) is totally geodesic in M and hence in \(\widetilde{M}.\)

Next, if B is normal to M,  from Theorem 5.5, any integral submanifold \(M^{\perp }\) of \(\mathcal {D}^{\perp }\) is totally geodesic in \(\widetilde{M}.\) The second statement follows immediately from Theorems 5.6 and 5.7.    \(\blacksquare \)

Corollary 1

Let M be a totally geodesic proper CR-submanifold of a l.c.K. manifold \(\widetilde{M}\) such that \(B_x\in \nu _x\), for any \(x\in M.\) Then, M is locally the Riemannian product of a Kaehler submanifold and a totally real submanifold of \(\widetilde{M}.\)

Proof

From Theorem 5.8, M is locally the product of a holomorphic submanifold \(M^{T}\) and a totally real submanifold \(M^{\perp }\) of \(\widetilde{M}.\) But \(\omega =0\) on M,  so that we have induced on \(M^{T}\) a Kaehlerian structure. Moreover, it can be easily seen that the projection map p (resp., q) onto \(\mathcal {D}\) (resp., \(\mathcal {D}^{\perp }\)) is parallel with respect to \(\nabla ,\) so that this local product is actually a local Riemannian product.    \(\blacksquare \)

Next, we consider the particular case when M is either holomorphic or totally real.

Lemma 3

Let M be a holomorphic submanifold of a l.c.K. manifold \(\widetilde{M}.\) Then the subbundles TM and \(T^{\perp }(M)\) are holomorphic. Moreover, we have

$$\begin{aligned} \overline{\sigma }(JX,Y)=\overline{\sigma }(X,JY)= J\overline{\sigma }(X,Y), \end{aligned}$$
(5.24)
$$\begin{aligned} \overline{A}_{JV}=J\overline{A}_{V}=-\overline{A}_{V}J, \end{aligned}$$
(5.25)
$$\begin{aligned} \overline{D}_X(JV)=J\overline{D}_XV, \end{aligned}$$
(5.26)
$$\begin{aligned} \nabla ^M_X(JY)=J\nabla ^M_XY, \end{aligned}$$
(5.27)

for any vector fields XY tangent to M and V normal to M.

Proof

As \(\widetilde{M}\) is locally endowed with Kaehler metrics \(\widetilde{g}_U\) whose Levi-Civita connection is \(\overline{\nabla },\) these formulas follow from similar formulas in the Kaehlerian case.    \(\blacksquare \)

Theorem 5.9

Let M be a holomorphic submanifold of a l.c.K. manifold \(\widetilde{M}.\) Then, we have

  1. (i)

    The mean curvature vector H of M is given by \(H=-\frac{1}{2}B_2\).

  2. (ii)

    M is totally umbilical if and only if the Weingarten endomorphisms are commutative.

Proof

Firstly, if \(dim(M)=2k>0,\) let \(\left\{ e_1,\ldots ,e_k,Je_1,\ldots ,Je_k \right\} \) be an orthonormal basis for \(T_x(M),\) \(x\in M.\) Then

$$\begin{aligned} 2kH_x=(tr(\sigma ))_x=\sum _{i=1}^k\sigma _x(e_i,e_i)+\sum _{i=1}^k\sigma _x(Je_i,Je_i). \end{aligned}$$
(5.28)

But from (5.8) and (5.24), (5.28) gives \(2kH_x=-k(B_2)_x.\)

Next, let V be a vector field normal to M. From (5.17) and (5.25), we have

$$0=[A_{V},A_{JV}]=[\overline{A}_{V},\overline{A}_{JV}]= -2J(\overline{A}_{V})^2,$$

Thus \(\overline{A}_{V}=0\) and from (5.9), we have \(A_{V}=-\frac{1}{2}\omega (V)I\)    \(\blacksquare \)

The endomorphism n of the normal bundle \(T^{\perp }M\) defined in (5.5) induces an f-structure in \(T^{\perp }M\) [22]. For any vector field X tangent to M and V normal to M, we write

$$\begin{aligned} (\widetilde{\nabla }^{\, \prime }_Xn)V = D_X(nV) - nD_XV, \end{aligned}$$
$$\begin{aligned} (\overline{\nabla }^{\, \prime }_Xn)V = \overline{D}_X(nV) - n\overline{D}_XV. \end{aligned}$$

When \(\widetilde{\nabla }^{\, \prime }n = 0,\) the f-structure n is said to be parallel [10].

Lemma 4

Let M be an r-dimensional totally real submanifold of a 2m-dimensional l.c.K. manifold \(\widetilde{M}\). Then we have

  1. (i)

    \(\overline{A}_{JX}Y=\overline{A}_{JY}X\), for any XY tangent to M.

  2. (ii)

    If \(r=m,\) then \(\overline{D}_X(JY)=J\nabla ^{M}_XY\), \(\nabla ^M_X(JV)=J\overline{D}_XV\), and \(\overline{\sigma }(X,JV)=-J\overline{A}_{V}X.\)

  3. (iii)

    \(\widetilde{\nabla }^{\,\prime }n=\overline{\nabla }^{\prime }n\).

  4. (iv)

    If the f-structure n is parallel, then

    $$\begin{aligned} A_{V}=-\frac{1}{2}\omega (V)I, \end{aligned}$$
    (5.29)

    for any \(V \in \nu \).

  5. (v)

    If the Weingarten endomorphisms are commutative, then there is an orthonormal local basis \(\left\{ e_1,\ldots ,e_r \right\} \) in M such that with respect to this basis \(\overline{A}_{Je_i}\) is a diagonal matrix

    $$\begin{aligned} \overline{A}_{Je_i}=\left( 0\ldots 0 \, \lambda _i \, 0\ldots 0 \right) , \quad i=1,\ldots ,r. \end{aligned}$$
    (5.30)

Proof

Assertions (i) and (ii) follow immediately from similar formulas in the Kaehlerian case. From Eq. (5.10), we easily obtain (iii).

In order to prove (iv), we take \(V \in \nu ,\) and \(X \in T(M).\) Then, (iii) gives \((\nabla ^{\prime }_Xn)V= \overline{D}_X(nV)-n\overline{D}_XV =0.\) By using (5.25) and (5.26) this yields \(J\overline{A}_{V}X=0.\) Therefore, \(\overline{A}_{V}=0\) and from (5.9), we obtain (iv).

Finally, from (5.17) we have \([\overline{A}_{V},\overline{A}_{\overline{V}}]=0\), for any \(V,\overline{V}\) normal to M. Then, we can find a local orthonormal basis \(\left\{ \widetilde{e}_1\ldots ,\widetilde{e}_r \right\} \) in M (with respect to the local Kaehlerian metrics \(\widetilde{g}_U=e^{-h}g|_U\)) such that \(\overline{A}_{Je_i} = \left( 0\ldots \mu _i\ldots 0 \right) \), \(i=1,\ldots ,r.\) If we start by using this basis, we can obtain an orthonormal (with respect to the metric g) local basis \(\left\{ e_1,\ldots ,e_n\right\} \) in M such that (v) holds.    \(\blacksquare \)

Theorem 5.10

Let M be an r-dimensional totally real and minimal submanifold of a l.c.K. manifold \(\widetilde{M}\) such that their Weingarten endomorphisms are commutative and the f-structure n is parallel. Then, we have

  1. (i)

    If \(r\ge 2\), M is totally geodesic if and only if the Lee vector field B is tangent to M.

  2. (ii)

    If \(r=1\) and B is orthogonal to \(\nu \), then M is a geodesic curve.

Proof

First, since the Weingarten endomorphisms are commutative, let \(\left\{ e_1,\ldots ,e_r \right\} \) be an orthonormal local basis as in Lemma 4 (v). From Eq. (5.8), we have

$$\begin{aligned} \begin{aligned} 0=g(H,Je_i)=&\frac{1}{n}\sum _{j=1}^ng\left( \sigma (e_j,e_j),Je_i \right) =\frac{1}{n}\sum _{j=1}^ng\left( \overline{A}_{Je_i}e_j,e_j \right) - \frac{1}{2} \omega (Je_i)\\ =&\frac{1}{n} \lambda _i - \frac{1}{2} \omega (Je_i), \quad i=1,\ldots ,r. \end{aligned} \end{aligned}$$

Therefore,

$$\begin{aligned} \overline{A}_{Je_i}e_j=\delta _{ij}\lambda _ie_j=\delta _{ij}\frac{n}{2}\omega (Je_i)e_j, \quad i=1,\ldots ,r. \end{aligned}$$
(5.31)

Now, from (5.9) and (5.31) we obtain,

$$\begin{aligned} A_{Je_i}e_j=\frac{1}{2}(n \delta _{ij}-1)\omega (Je_i)e_j, \quad i=1,\ldots ,r. \end{aligned}$$
(5.32)

Thus, if \(r\ge 2\) and B is tangent to M, Eq. (5.32) gives \(A_{Je_i}=0\), \(i=1,\ldots ,r.\) Moreover, from (iv) in Lemma 4, \(A_{V}=-\frac{1}{2}\omega (V)I=0\), for any \(V \in \nu .\) Then, \(A_{\overline{V}}=0\), for any vector field \(\overline{V}\) normal to M.

On the other hand, if there is \(x\in M\) such that \((B_2)_x\ne 0,\) from (5.32) and (iv) in Lemma 4, we can take a vector field V normal to M such that \(A_{V}\ne 0.\) This gives (i).

In order to prove (ii), let us take a unit vector field X tangent to M. We have \(0=g(H,JX)=g(\sigma (X,X),JX)=g(A_{JX}X,X),\) and, then, \(A_{JX}=0.\) But, if B is orthogonal to \(\nu ,\) from (iv) in Lemma 4, \(A_{V}=-\frac{1}{2}\omega (V)=0\), for any \(V \in \nu .\) This means that \(A_{\overline{V}}=0\), for any vector field \(\overline{V}\) normal to M.    \(\blacksquare \)

Theorem 5.11

Let M be an r-dimensional (\(r\ge 2\)) totally real and totally umbilical submanifold of a l.c.K. manifold \(\widetilde{M}\) such that the f-structure n is parallel. Then M is totally geodesic if and only if B is tangent to M.

Proof

Let \(\left\{ u_1,\ldots ,u_r\right\} \) be an orthonormal local basis in U. Since M is totally umbilical, for any vector field X tangent to M,  by using Eqs. (5.8) and (5.9), we find

$$\begin{aligned} g(\overline{A}_{JX}u_j,u_k)=\frac{1}{r}\delta _{jk}tr(\overline{A}_{JX}). \end{aligned}$$
(5.33)

But from Eq. (5.9) and (iv) in Lemma 4 we also have

$$\begin{aligned} \overline{A}_{V}=0, \end{aligned}$$
(5.34)

for any \(V\in \nu .\) On the other hand \([A_{\overline{V}},A_{\overline{\overline{V}}}]= [ \overline{A}_{\overline{V}},\overline{A}_{\overline{\overline{V}}}]=0,\) for any vector fields \(\overline{V},\overline{\overline{V}}\) normal to M. Therefore, from (v) in Lemma 4, there is an orthonormal local basis \(\{e_1,\ldots ,e_r\}\) in M such that, with respect to this basis, Eq. (5.30) holds. But, from Eq. (5.33), we also have

$$\begin{aligned} g(\overline{A}_{Je_i}e_j,e_j)=\frac{1}{r}\lambda _i, \quad i.j=1,\ldots ,n. \end{aligned}$$
(5.35)

Since \(r\ge 2\), we can take \(j\ne i\) and then, Eqs. (5.30) and (5.35) give \(\lambda _i=0\), \(i=1,\ldots ,r.\) Thus, we get \(\overline{A}_{Je_i}=0\), \(i=1,\ldots ,r\), which, together with (5.34) gives \(\overline{A}_{\overline{V}}=0\), for any vector field \(\overline{V}\) normal to M. Now, if B is tangent to M,  Eq. (5.9) proves that M is totally geodesic.

Conversely, if M is totally geodesic, from (5.29) we have \(0=A_{V}=-\frac{1}{2}\omega (V)I\), for any \(V \in \nu .\) This means that B is normal to \(\nu .\) Furthermore, from (v) in Lemma 4, we can find an orthonormal local basis \(\left\{ e_1,\ldots ,e_r \right\} \) in M such that \(\overline{A}_{Je_i}\) has a diagonal matrix \(\overline{A}_{Je_i} = \left( 0\ldots 0 \, \lambda _i \,0\ldots 0 \right) =\frac{1}{2}\omega (Je_i)I\), \(i=1,\ldots ,r.\) Since \(r\ge 2\), this means \(\omega (Je_i)=0\), \(i=1,\ldots ,r\) so that B is normal to J(T(M)). Thus, B is tangent to M.    \(\blacksquare \)

5.2.4 CR-products in l.c.K. Manifolds

Let TNtn be the endomorphisms and vector-valued 1-forms defined by (5.5). Let us write

$$\begin{aligned} \begin{aligned} (\widetilde{\nabla }^{\; \prime }_ZT)W&=\nabla _Z(TW)-T\nabla _ZW, \\ (\overline{\nabla }^{\, \prime }_ZT)W&=\nabla ^{M}_Z(TW)-T\nabla ^M_ZW, \end{aligned} \end{aligned}$$
(5.36)

for all ZW tangent to M. On the other hand, T is said to be parallel if \(\overline{\nabla }^{\,}T=0. \) From (5.1)–(5.3) it is easy to prove that

$$\begin{aligned} (\overline{\nabla }^{\, \prime }_ZT)W&=(\widetilde{\nabla }^{\, \prime }_ZT)W+\frac{1}{2}\omega (W)TZ-\frac{1}{2}\omega (TW)Z \nonumber \\&\quad +\frac{1}{2}g(Z,TW)B_1-\frac{1}{2}g(Z,W)TB_1. \end{aligned}$$
(5.37)

But, from (5.36) we see that

$$\begin{aligned} (\overline{\nabla }^{\, \prime }_Z T)W=t\overline{\sigma }(Z,W) + \overline{A}_{NW}Z. \end{aligned}$$
(5.38)

Definition 1

A CR-submanifold of a l.c.K. manifold \(\widetilde{M}\) is called a CR-product if it is locally a Riemannian product of a holomorphic submanifold \(M^{T}\) and a totally real submanifold \(M^{\perp }\) of \(\widetilde{M}.\)

Theorem 5.12

Let M be a CR-submanifold of a l.c.K. manifold \(\widetilde{M}\) such that the Lee field B is normal to M. Then M is a CR-product if and only if T is parallel.

Proof

Since B is normal to M,  from Eq. (5.37), we have \(\widetilde{\nabla }^{\, \prime }T=\overline{\nabla }^{\, \prime }T.\) If T is parallel, from (5.8), (5.9), and (5.38), we find

$$\begin{aligned} t\sigma (Z,W) + \frac{1}{2}g(Z,W)tB = - A_{NW}Z-\frac{1}{2}\omega (NW)Z. \end{aligned}$$
(5.39)

But for any \(X\in \mathcal {D}\), \(NX=0,\) and the last equation gives

$$0=g(A_{NW}Z,X)+\frac{1}{2}\omega (NW)g(Z,X),$$

or, equivalently, \(g(\sigma (Z,X),JW)+\frac{1}{2}g(JW,B)g(Z,X)=0\), for any W tangent to M. Therefore,

$$\begin{aligned} \sigma (Z,X)=-\frac{1}{2} g(Z,X)B. \end{aligned}$$
(5.40)

If we take \(Z\in \mathcal {D}\), the last equation gives \(\sigma (X,JY)-\sigma (JX,Y)=-\Omega (X,Y)B,\) and, from Theorem 5.3, \(\mathcal {D}\) is integrable. Let \(M^{T}\) be an integral submanifold of \(\mathcal {D}.\) For any \(Z\in \mathcal {D}^{\perp }\), Eq. (5.40) yields \(g(\sigma (Z,X),JZ)=-\frac{1}{2}g(Z,Z)g(B,JZ)\) and, from Theorem 5.6, the submanifold \(M^{T}\) is totally geodesic in M. Now, let \(M^{\perp }\) be an integral submanifold od \(\mathcal {D}^{\perp }.\) From (5.40), if \(Z\in \mathcal {D}^{\perp }\), then \(\sigma (Z,X)=0\) and, from Theorem 5.5, \(M^{\perp }\) is totally geodesic.

Conversely, assume that M is a CR-product. First, we prove that \(\nabla ^{M}_ZX \in \mathcal {D}\), for any \(X\in \mathcal {D}\) and Z tangent to M. As M is locally a Riemannian product of \(M^{T}\) (holomorphic submanifold) and \(M^{\perp }\) (totally real submanifold), it suffices to prove that \(\nabla ^M_ZX \in \mathcal {D}\), for any \(X\in \mathcal {D}\) and \(Z\in \mathcal {D}^{\perp }.\) In fact, from (5.3) we have

$$J\nabla ^{M}_ZX=\nabla ^M_Z(JX)+\overline{\sigma }(Z,JX)-J\overline{\sigma }(Z,X).$$

Thus, if \(W\in \mathcal {D}^{\perp }\), \(g(J\nabla ^M_ZX,JW)=g(\overline{\sigma }(Z,JX),JW).\) Since \(M^{\perp }\) is totally geodesic in M,  from (5.8) and Theorem 5.5 we have \(g(\nabla ^M_ZX,W)=0\), for any \(W\in \mathcal {D}^{\perp }.\) So, \(\nabla ^M_ZX \in \mathcal {D}\) and \(\nabla ^M_ZX \in \mathcal {D}\), for any Z tangent to M. From \(\overline{\nabla }J=0\), we find

$$J\nabla ^M_ZX+J\overline{\sigma }(Z,X)=\nabla ^M_Z(JX)+\overline{\sigma }(Z,JX),$$

and then, \(J\nabla ^M_ZX=\nabla ^M_Z(JX)\), \(J\overline{\sigma }(Z,X)=\overline{\sigma }(Z,JX).\) Now, from (5.36) we get

$$\begin{aligned} (\overline{\nabla }^{\, \prime }_ZT)X=\nabla ^M_Z(TX)-T\nabla ^M_ZX = \nabla ^M_Z(JX)-J\nabla ^M_ZX=0, \end{aligned}$$
(5.41)

for any \(X\in \mathcal {D}\) and Z tangent to M.

In a similar way, we prove that \(\nabla ^M_ZZ \in \mathcal {D}^{\perp }\) for any \(Z\in \mathcal {D}^{\perp }\) and Z tangent to M. Since M is a CR-product, it suffices to show this for \(Z=X\in \mathcal {D}.\) In fact, from (5.3), given any \(Y\in \mathcal {D}\) we find that

$$g(J\nabla ^M_XZ,Y)=-g(\overline{A}_{JZ}X,Y)-g(J\overline{\sigma }(X,Z),Y)= -g(\overline{\sigma }(X,Y),JZ)=0,$$

where the last equation holds from (5.8) and Theorem 5.6. Then, \(J\nabla ^M_XZ\) is orthogonal to \(\mathcal {D}.\) On the other hand, if \(W\in \mathcal {D}^{\perp }\), we have

$$g\left( \nabla ^M_XZ,W\right) =-g(\overline{\sigma }(X,W),JZ)+g(\overline{\sigma }(X,Z),JW).$$

But, from Theorem 5.5 we have \(g(J\nabla ^M_XZ,W)=0,\) That is, \(J\nabla ^M_XZ\) is normal to M,  so that \(\nabla ^M_XZ\in \mathcal {D}^{\perp }.\) Therefore, we have

$$\begin{aligned} (\overline{\nabla }^{\, \prime }_ZT)Z=\nabla ^M_Z(TZ)-T\nabla ^M_ZZ=0. \end{aligned}$$
(5.42)

Now, from (5.37), (5.41), and (5.42), we have \(\widetilde{\nabla }^{\, \prime }T=0.\)    \(\blacksquare \)

Theorem 5.13

Let M be a CR-submanifold of a l.c.K. manifold \(\widetilde{M}\) such that \(B_x\in \mathcal {D}_x\) for each \(x\in M.\) If T is parallel, then M is a CR-product. The converse does not holds unless \(dim(\mathcal {D})=2\) or \(B=0\) on M.

Proof

Since T is parallel, Eqs. (5.37) and (5.38) give

$$\begin{aligned} t\overline{\sigma }(Z,W)+\overline{A}_{NW}Z&=\frac{1}{2}\omega (W)TZ- \frac{1}{2}\omega (TW)Z \nonumber \\&\quad +\frac{1}{2}g(Z,TW)B-\frac{1}{2}g(Z,W)TB. \end{aligned}$$
(5.43)

If \(X\in \mathcal {D}\), then \(NX=0\) and (5.43) gives

$$\begin{aligned} -g(J\overline{\sigma }(X,Z),W)&=\frac{1}{2}g(B,W)g(JZ,X)+\frac{1}{2}g(W,JB)g(Z,X) \nonumber \\&\quad -\frac{1}{2}g(JZ,W)g(B,X)-\frac{1}{2}g(Z,W)g(JB,X), \end{aligned}$$
(5.44)

for any vector field W tangent to M. From (5.8), (5.44) yields

$$\begin{aligned} -J\sigma (X,Z)&=\frac{1}{2}g(JZ,X)B+\frac{1}{2}g(Z,X)JB \nonumber \\&\quad - \frac{1}{2}g(B,X)JZ-\frac{1}{2}g(JB,X)Z. \end{aligned}$$
(5.45)

For any \(Z\in \mathcal {D}^{\perp },\) Eq. (5.45) gives \(g(A_{JZ},Z)=\frac{1}{2}\omega (JX)g(Z,Z)\), for any Z tangent to M and, hence, we have

$$\begin{aligned} A_{JZ}X=\frac{1}{2}\omega (JX)Z. \end{aligned}$$
(5.46)

Next, for \(Y\in \mathcal {D},\) from (5.46), we have

$$\begin{aligned} g(\sigma (X,Y),JZ)=0, \quad for \; X \in \mathcal {D}, \, Z\in \mathcal {D}^{\perp }. \end{aligned}$$
(5.47)

Therefore, \(g(\sigma (X,JY)-\sigma (JX,Y,J\mathcal {D}^{\perp })=0\) and, from Theorem 5.4, the distribution \(\mathcal {D}\) is integrable. Moreover, any integral submanifold \(M^{\perp }\) od \(\mathcal {D}\) is totally geodesic in M because of (5.47) and Theorem 5.6. Now, let \(M^{\perp }\) be an integral submanifold of \(\mathcal {D}^{\perp }.\) For any \(W\in \mathcal {D}^{\perp },\) Eq. (5.46) gives

$$g\left( A_{JZ}+\frac{1}{2}g(Z,W)JB,X\right) =0$$

and this means that \(M^{\perp }\) is totally geodesic in M (Theorem 5.5). Thus M is a CR-product.    \(\blacksquare \)

In order to prove the converse, we first give the following Lemma.

Lemma 5

If M is a CR-product in a l.c.K. manifold \(\widetilde{M}\) such that \(B_x\in \mathcal {D}_x\) for any \(x\in M,\) then

$$\begin{aligned} \nabla _ZX\in \mathcal {D}, \end{aligned}$$
(5.48)
$$\begin{aligned} \nabla _XZ\in \mathcal {D}^{\perp }, \end{aligned}$$
(5.49)
$$\begin{aligned} J\, \nabla _ZX=\nabla _Z(JX), \end{aligned}$$
(5.50)

for any \(X\in \mathcal {D}\) and \(Z\in \mathcal {D}^{\perp }.\)

Proof

If \(X\in \mathcal {D}\) and \(Z\in \mathcal {D}^{\perp },\) then from (5.7) and (5.8), we obtain

$$\begin{aligned} J\,\nabla _ZX&=\frac{1}{2}\omega (X)JZ-J\sigma (Z,X)+\nabla _Z(JX)+\nabla _Z(JX) \nonumber \\&\quad -\frac{1}{2}\omega (JX)Z+\sigma (Z,JX). \end{aligned}$$
(5.51)

Now, for any \(W\in \mathcal {D}^{\perp }\), (5.51) yields

$$g(J\, \nabla _ZX,JW) = g(\nabla _ZX,W) = g\left( A_ {JW}Z+\frac{1}{2}g(Z,W)JB,JX\right) =0.$$

The last equation holds because any leaf \(M^{\perp }\) of \(\mathcal {D}^{\perp }\) is totally geodesic in M (Theorem 5.5). Thus \(\nabla _ZX \in \mathcal {D}\) and this is assertion (5.48). Now, take \(X,Y \in \mathcal {D}\) and \(Z\in \mathcal {D}^{\perp }.\) From (5.1) and (5.2), we find that

$$\begin{aligned} g(J\,\nabla _XZ,Y)=-g(A_{JZ},Y)=-g(\sigma (X,Y),JZ)=0 \end{aligned}$$
(5.52)

The last equation holds because of Theorem 5.6. If \(X\in \mathcal {D}\) and \(Z,W\in \mathcal {D}^{\perp },\) from (5.1) and (5.2) again we have

$$g(J\, \nabla _XZ,W)=g(A_{JW}Z,X)-g(A_{JZ}W,X).$$

But, from Theorem 5.5 we obtain

$$\begin{aligned} g(A_{JW}Z,X)-g(A_{JZ}W,X)= -\frac{1}{2}g(Z,W)g(JB,X)+\frac{1}{2}g(W,Z)g(JB,X) =0 \end{aligned}$$

and, hence

$$\begin{aligned} g(J\, \nabla _XZ,W)=0. \end{aligned}$$
(5.53)

Now, (5.49) follows from (5.52) and (5.53). Finally, (5.48) and (5.51) give (5.50).    \(\blacksquare \)

Now we prove the converse of Theorem 5.13. From (5.36) and (5.48), for any \(X\in \mathcal {D}\) and Z tangent to M we have

$$\begin{aligned} (\widetilde{\nabla }^{\; \, \prime }_Z,T)X = \nabla _Z(JX)-J(\nabla _ZX). \end{aligned}$$

On the other hand, we write \(Z=Y+Z\), where \(Y\in \mathcal {D}\) and \(Z\in \mathcal {D}^{\perp }.\) Then, from (5.50) we have

$$\begin{aligned} (\widetilde{\nabla }^{\; \, \prime }_ZT)X=\nabla _Y(JX)-J\, \nabla _YX. \end{aligned}$$
(5.54)

But (5.1)–(5.3) give

$$\begin{aligned} \nabla _Y(JX)-J\nabla _YX = \frac{1}{2}\omega (Y)JX-\frac{1}{2}(JY)X-\frac{1}{2}g(X,Y)JB+\frac{1}{2}g(X,JY)B. \end{aligned}$$
(5.55)

Now we have

(a) If \(dim(\mathcal {D})\ge 4\) and \(B_x \ne 0\) for some \(x\in M,\) there are \(X,Y \in \mathcal {D}\) such that the right-hand side of (5.55) does not vanish at x. Therefore, T is not parallel.

(b) If \(dim(\mathcal {D})=2\), then the right-hand side of (5.55) vanishes and, hence

$$\begin{aligned} (\widetilde{\nabla }^{\; \prime }_ZT)X=0, \end{aligned}$$
(5.56)

for any \(X\in \mathcal {D}\) and Z tangent to M. But (5.49) implies \(\nabla _ZZ \in \mathcal {D}^{\perp }\), for any \(Z\in \mathcal {D}^{\perp }\) and Z tangent to M,  so that

$$\begin{aligned} (\widetilde{\nabla }^{\; \prime }_ZT)Z=\nabla _Z(TZ)-T\nabla _ZZ=-T\nabla _ZZ=0. \end{aligned}$$
(5.57)

Then, (5.56) and (5.57) prove that T is parallel.   \(\blacksquare \)

5.3 Normal CR-Submanifolds of S-manifolds

We want to study here the normal CR-submanifolds for general S-manifolds. In fact, the normal CR-submanifolds become to be a very wide class of CR-submanifolds. Actually, either totally f-umbilical submanifolds (see [19] for more details) or CR-products (see [12]) of an S-manifold are normal CR-submanifolds. We also study normal CR-submanifolds of an S-space-form, specially in the concrete cases of \(\mathbb R^{2m+s}\) (with constant f-sectional curvature \(c=-3s\)) and \(\mathbb H^{2m+s}\) (with constant f-sectional curvature \(c=4-3s\)).

5.3.1 Preliminaries

A \((2m+s)\)-dimensional Riemannian manifold \((\wedge M,g)\) endowed with an f-structure f (that is, a tensor field of type (1, 1) and rank 2m satisfying \(f^3+f=0\) [21]) is said to be a metric f-manifold if, moreover, there exist s global vector fields \(\xi _1,\dots ,\xi _s\) on \(\wedge M\) (called structure vector fields) such that, if \(\eta _1,\dots ,\eta _s\) are the dual 1-forms of \(\xi _1,\dots ,\xi _s\), then

$$f\xi _{\alpha }=0; \eta _{\alpha }\circ f=0; f^2=-I+\sum _{{\alpha }=1}^s\eta _{\alpha }\otimes \xi _{\alpha };$$
$$\begin{aligned} g(X,Y)=g(fX,fY)+\sum _{{\alpha }=1}^s\eta _{\alpha }(X)\eta _{\alpha }(Y),\end{aligned}$$
(5.58)

for any \(X,Y\in \mathcal {X}(\wedge M)\) and \({\alpha }=1,\dots ,s\).

Let F be the 2-form on \(\wedge M\) defined by \(F(X,Y)=g(X,fY)\), for any \(X,Y\in \mathcal {X}(\wedge M)\). Since f is of rank 2m, then

$$\eta _1\wedge \cdots \wedge \eta _s\wedge F^m\ne 0$$

and, particularly, \(\wedge M\) is orientable.

The f-structure f is said to be normal if

$$[f,f]+2\sum _{{\alpha }=1}^s\xi _{\alpha }\otimes d\eta _{\alpha }=0,$$

where [ff] is the Nijenhuis torsion of f.

A metric f-manifold is said to be a K-manifold [5] if it is normal and \(\mathrm{d}F=0\). A K-manifold is called an S-manifold if \(F=\mathrm{d}\eta _{\alpha }\), for any \({\alpha }\). Note that, for \(s=0\), a K-manifold is a Kaehlerian manifold and, for \(s=1\), a K-manifold is a quasi-Sasakian manifold and an S-manifold is a Sasakian manifold. When \(s\ge 2\), nontrivial examples can be found in [5, 13]. Moreover, a K-manifold \(\wedge M\) is an S-manifold if and only if

$$\begin{aligned} \wedge \nabla _X\xi _{\alpha }=-fX, \end{aligned}$$
(5.59)

for any \(X\in \mathcal {X}(\wedge M)\) and any \({\alpha }=1,\dots ,s\), where \(\wedge \nabla \) denotes the Levi-Civita connection of g. It is easy to show that in any S-manifold

$$\begin{aligned} (\wedge \nabla _Xf)Y=\sum _{{\alpha }=1}^s\left\{ g(fX,fY)\xi _{\alpha }+\eta _{\alpha }(Y)f^2X\right\} , \end{aligned}$$
(5.60)

for any \(X,Y\in \mathcal {X}(\wedge M)\). A plane section \(\pi \) on a metric f-manifold \(\wedge M\) is said to be an f-section if it is determined by a unit vector X, normal to the structure vector fields and fX. The sectional curvature of \(\pi \) is called an f-sectional curvature. An S-manifold is said to be an S-space-form if it has a constant f-sectional curvature c and then, it is denoted by \(\wedge M(c)\). In such case, the curvature tensor field \(\wedge R\) of \(\wedge M(c)\) satisfies [15]

$$\begin{aligned}&\wedge R(X,Y,Z,W) \nonumber \\&\quad = \sum _{{\alpha },{\beta }}(g(fX,fW)\eta _{\alpha }(Y)\eta _{\beta }(Z)-g(fX,fZ)\eta _{\alpha }(Y)\eta _{\beta }(W) \nonumber \\&\qquad +g(fY,fZ)\eta _{\alpha }(X)\eta _{\beta }(W)-g(fY,fW)\eta _{\alpha }(X)\eta _{\beta }(Z))\nonumber \\&\qquad +\frac{c+3s}{4}(g(fX,fW)g(fY,fZ)-g(fX,fZ)g(fY,fW))\nonumber \\&\qquad +\frac{c-s}{4}(F(X,W)F(Y,Z)-F(X,Z)F(Y,W)\nonumber \\&\qquad \qquad \quad \, - 2F(X,Y)F(Z,W)), \end{aligned}$$
(5.61)

for any \(X,Y,Z,W\in \mathcal {X}(\wedge M)\). Next, let M be a isometrically immersed submanifold of a metric f-manifold \(\wedge M\) (for the general theory of submanifolds, we refer to [3, 22]). We denote by \(\mathcal {X}(M)\) the Lie algebra of tangent vector fields to M and by \(T(M)^\perp \) the set of tangent vector fields to \(\wedge M\) which are normal to M. For any vector field \(X\in \mathcal {X}(M)\), we write

$$\begin{aligned} fX=TX+NX, \end{aligned}$$
(5.62)

where TX and NX are the tangential and normal components of fX, respectively. Then, T is an endomorphism of the tangent bundle of M and N is a normal bundle valued 1-form on such tangent bundle. It is easy to show that if T does not vanish, it defines an f-structure in the tangent bundle of M. The submanifold M is said to be invariant if N is identically zero, that is, if fX is tangent to M, for any \(X\in \mathcal {X}(M)\). On the other hand, M is said to be an anti-invariant submanifold if T is identically zero, that is, if fX is normal to M, for any \(X\in \mathcal {X}(M)\). In the same way, for any \(V\in T(M)^\perp \), we write

$$\begin{aligned} fV=tV+nV, \end{aligned}$$
(5.63)

where tV and nV are the tangential and normal components of fV, respectively. Then, t is a tangent bundle valued 1-form on the normal bundle of M and n is an endomorphism of the normal bundle of M. It is easy to show that if n does nor vanish, it defines an f-structure in the normal bundle of M. From now on, we suppose that all the structure vector fields are tangent to the submanifold M and so, \(dim(M)\ge s\). Then, the distribution on M spanned by the structure vector fields is denoted by \(\mathcal {M}\) and its complementary orthogonal distribution is denoted by \(\mathcal {L}\). Consequently, if \(X\in \mathcal {L}\), then \(\eta _{\alpha }(X)=0\), for any \({\alpha }=1,\dots ,s\) and if \(X\in \mathcal {M}\), then \(fX=0\). In this context, M is said to be a CR-submanifold of \(\wedge M\) if there exist two differentiable distributions \(\mathcal {D}\) and \(\mathcal {D}^\perp \) on M satisfying

  1. (i)

    \(\mathcal {X}(M)=\mathcal {D}\oplus \mathcal {D}^\perp \oplus \mathcal {M}\), where \(\mathcal {D}\), \(\mathcal {D}^\perp \) and \(\mathcal {M}\) are mutually orthogonal to each other;

  2. (ii)

    The distribution \(\mathcal {D}\) is invariant by f, that is, \(f\mathcal {D}_x=\mathcal {D}_x\), for any \(x\in M\);

  3. (iii)

    The distribution \(\mathcal {D}^\perp \) is anti-invariant by f, that is, \(f\mathcal {D}^\perp _x\subseteq T_x(M)^\perp \), for any \(x\in M\).

This definition is motivated by the following theorem.

Theorem 5.14

([16]) Let \(\wedge M\) be an S-manifold which is the bundle space of a principal toroidal bundle over a Kaehler manifold \(\wedge M'\), \(\wedge \pi :\wedge M\longrightarrow \wedge M'\), M a submanifold immersed in \(\wedge M\), tangent to the structure vector fields and \(M'\) a submanifold immersed in \(\wedge M'\) such that there exists a fibration \(\pi :M\longrightarrow M'\), the diagram

commutes and the immersion i is a diffeomorphism on the fibers. Then, M is a CR-submanifold of \(\wedge M\) if and only if \(M'\) is a CR-submanifold of \(\wedge M'\).

We denote by 2p and q the real dimensions of \(\mathcal {D}\) and \(\mathcal {D}^\perp \), respectively. Then, we see that for \(p=0\) we obtain an anti-invariant submanifold tangent to the structure vector fields and for \(q=0\) an invariant submanifold. A CR-submanifold of an S-manifold is said to be a generic submanifold if given any \(V\in T(M)^\perp \), there exists \(Z\in \mathcal {D}^\perp \) such that \(V=fZ\), a \((\mathcal {D},\mathcal {D}^\perp )\) -geodesic submanifold if \(\sigma (X,Z)=0\), for any \(X\in \mathcal {D}\) and any \(Z\in \mathcal {D}^\perp \) and a \(\mathcal {D}^\perp \) -geodesic submanifold if \(\sigma (Y,Z)=0\), for any \(Y,Z\in \mathcal {D}^\perp \). As an example, it is easy to show that each hypersurface of \(\wedge M\) which is tangent to the structure vector fields is a CR-submanifold. Now, we write by P and Q the projections morphisms of \(\mathcal {X}(M)\) on \(\mathcal {D}\) and \(\mathcal {D}^\perp \), respectively. Thus, for any \(X\in \mathcal {X}(M)\), we have that

$$X=PX+QX+\sum _{{\alpha }=1}^s\eta _{\alpha }(X)\xi _{\alpha }.$$

We define the tensor field v of type (1, 1) by \(vX=fPX\) and the non-null, normal bundle valued 1-form u by \(uX=fQX\), for any \(X\in \mathcal {X}(M)\). Then, it is easy to show that \(u\circ v=0\) and \(\eta _{\alpha }\circ u=\eta _{\alpha }\circ v=0\), for any \({\alpha }=1,\dots ,s\). Moreover, a direct computation gives

$$\begin{aligned} g(X,Y)=g(uX,uY)+g(vX,vY)+\sum _{{\alpha }=1}^s\eta _{\alpha }(X)\eta _{\alpha }(Y), \end{aligned}$$
(5.64)
$$\begin{aligned} F(X,Y)=g(X,vY), \quad F(X,Y)=F(vX,vY), \end{aligned}$$
(5.65)

for any \(X,Y\in \mathcal {X}(M)\). From Gauss–Weingarten formulas and by using (5.59), for any \(X\in \mathcal {X}(M)\), \(V\in T(M)^\perp \), and \({\alpha }=1,\dots ,s\), we have

$$\begin{aligned} \nabla _X\xi _{\alpha }=-vX, \end{aligned}$$
(5.66)
$$\begin{aligned} \sigma (X,\xi _{\alpha })=-uX, \end{aligned}$$
(5.67)
$$\begin{aligned} A_V\xi _{\alpha }\in \mathcal {D}^\perp . \end{aligned}$$
(5.68)

Moreover, from (5.60) and the Gauss–Weingarten formulas, if \(X,Y\in \mathcal {X}(M)\), comparing the components in \(\mathcal {D}\), \(\mathcal {D}^\perp \) and \(T(M)^\perp \) respectively, we get

$$\begin{aligned} P\nabla _XvY-PA_{uY}X=v\nabla _XY-\sum _{{\alpha }=1}^s\eta _{\alpha }(Y)PX, \end{aligned}$$
(5.69)
$$\begin{aligned} Q\nabla _XvY-QA_{uY}X=t\sigma (X,Y)-\sum _{{\alpha }=1}^s\eta _{\alpha }(Y)QX, \end{aligned}$$
(5.70)
$$\begin{aligned} \sigma (X,vY)+D_XuY=u\nabla _XY+n\sigma (X,Y). \end{aligned}$$
(5.71)

From the above formulas and (5.60) we obtain

$$\begin{aligned} (\nabla _Xv)Y=A_{uY}X+t\sigma (X,Y)-\sum _{a=1}^s\{\eta _{\alpha }(Y)f^2X+g(fX,fY)\xi _{\alpha }\}, \end{aligned}$$
(5.72)
$$\begin{aligned} (\nabla _Xu)Y=n\sigma (X,Y)-\sigma (X,vY), \end{aligned}$$
(5.73)

for any \(X,Y\in \mathcal {X}(M)\). Also, from (5.60) and the Gauss–Weingarten formulas again, we have

$$\begin{aligned} \nabla _XZ=vA_{fZ}X-tD_XfZ, \end{aligned}$$
(5.74)
$$\begin{aligned} tD_XfZ=-Q\nabla _XZ, \end{aligned}$$
(5.75)

for any \(X\in \mathcal {X}(M)\) and any \(Z\in \mathcal {D}^\perp \). With regard to the integrability of the distributions involved in the definition of a CR-submanifold, I. Mihai [16] proved that the distributions \(\mathcal {D}^\perp \) and \(\mathcal {D}^\perp \oplus \mathcal {M}\) are always integrable. On the other hand, if \(p>0\), the distributions \(\mathcal {D}\) and \(\mathcal {D}\oplus \mathcal {D}^\perp \) are not integrable and the distribution \(\mathcal {D}\oplus \mathcal {M}\) is integrable if and only if

$$\begin{aligned} \sigma (X,fY)=\sigma (fX,Y), \end{aligned}$$
(5.76)

for any \(X,Y\in \mathcal {D}\). In [12], CR-products of S-manifolds are defined as CR-submanifolds such that the distribution \(\mathcal {D}\oplus \mathcal {M}\) is integrable and locally they are Riemannian products \(M_1\times M_2\), where \(M_1\) (resp., \(M_2\)) is a leaf of \(\mathcal {D}\oplus \mathcal {M}\) (resp., \(\mathcal {D}^\perp \)). From Theorem 3.1 and Proposition 3.2 in [12], we know that a CR-submanifold M of an S-manifold is a CR-product if and only if one of the following assertions is satisfied:

$$\begin{aligned} A_{f\mathcal {D}^\perp }f\mathcal {D}=0, \end{aligned}$$
(5.77)
$$\begin{aligned} g(\sigma (X,Y),fZ)=0, X\in \mathcal {D}, Y\in \mathcal {X}(M), Z\in \mathcal {D}^\perp , \end{aligned}$$
(5.78)
$$\begin{aligned} \nabla _XY\in \mathcal {D}\oplus \mathcal {M}, X\in \mathcal {D}, Y\in \mathcal {X}(M). \end{aligned}$$
(5.79)

5.3.2 Normal CR-Submanifolds of an S-manifold

Let M be a CR-submanifold of an S-manifold \(\wedge M\). We say that M is a normal CR-submanifold if

$$\begin{aligned} N_v(X,Y)=2t\mathrm{d}u(X,Y)-2\sum _{{\alpha }=1}^sF(X,Y)\xi _{\alpha }, \end{aligned}$$
(5.80)

for any \(X,Y\in \mathcal {X}(M)\), where \(N_v\) is denoting the Nijenhuis torsion of v, that is

$$N_v(X,Y)=(\nabla _{vX}v)Y-(\nabla _{vY}v)X+v((\nabla _Yv)X-(\nabla _Xv)Y).$$

We notice that (5.80) is equivalent to

$$\begin{aligned} S^*(X,Y)=N_v(X,Y)-t((\nabla _Xu)Y-(\nabla _Yu)X)+2\sum _{{\alpha }=1}^sF(X,Y)\xi _{\alpha }=0, \end{aligned}$$

for any \(X,Y\in \mathcal {X}(M)\). Now, we can prove the following characterization theorem in terms of the shape operator.

Theorem 5.15

A CR-submanifold M of an S-manifold \(\wedge M\) is normal if and only if

$$\begin{aligned} A_{uY}vX=vA_{uY}X, \end{aligned}$$
(5.81)

for any \(X\in \mathcal {D}\) and any \(Y\in \mathcal {D}^\perp \).

Proof

A direct expansion by using (5.72) and (5.73) gives that

$$\begin{aligned} S^*(X,Y)=A_{uY}vX-vA_{uY}X-A_{uX}vY+vA_{uX}Y, \end{aligned}$$
(5.82)

for any \(X,Y\in \mathcal {X}(M)\). Now, if M is a normal CR-submanifold of \(\wedge M\), (5.81) follows form (5.82) since \(uX=0\), for any \(X\in \mathcal {D}\). Conversely, if (5.81) holds, we use the decomposition \(\mathcal {X}(M)=\mathcal {D}\oplus \mathcal {D}^\perp \oplus \mathcal {M}\). First, since \(uX=0\) for any \(X\in \mathcal {D}\) and \(v\xi _{\alpha }=0=u\xi _{\alpha }\), for any \({\alpha }\), we deduce from (5.81) and (5.82) that \(S^*(X,Y)=0\), for any \(X\in \mathcal {D}\) and any \(Y\in \mathcal {X}(M)\). Moreover, if \(Y\in \mathcal {D}^\perp \), from (5.68) we have \(A_{uY}\xi _{\alpha }\in \mathcal {D}^\perp \) and so, \(vA_{uY}\xi _{\alpha }=0\) dfor any \({\alpha }\). Consequently, \(S^*(X,\xi _{\alpha })=0\), for any \(X\in \mathcal {X}(M)\). Finally, if \(X,Y\in \mathcal {D}^\perp \), (5.82) becomes

$$S^*(X,Y)=v(A_{fX}Y-A_{fY}X),$$

since \(vX=vY=0\) and \(uX=fX\), \(uY=fY\). But, from (5.60) we easily show that \(A_{fX}Y=A_{fY}X\).    \(\blacksquare \)

Corollary 2

A CR-submanifold M of an S-manifold is normal if and only if

$$\begin{aligned} g(\sigma (X,vY)+\sigma (Y,vX),fZ)=0, \end{aligned}$$
(5.83)
$$\begin{aligned} g(\sigma (X,Z)fW)=0, \end{aligned}$$
(5.84)

for any \(X,Y\in \mathcal {D}\) and any \(Z,W\in \mathcal {D}^\perp \).

Proof

Since v is skew-symmetric, from (5.81) we see that M is normal if and only if

$$\begin{aligned} g(\sigma (X,vY),uZ)=-g(\sigma (Y,vX),uZ)m \end{aligned}$$
(5.85)

for any \(X\in \mathcal {X}(M)\), \(Y\in \mathcal {D}\) and \(Z\in \mathcal {D}^\perp \). Now, if M is normal, from (5.85) we get (5.83) taking \(X\in \mathcal {D}\) and (5.84) taking \(X\in \mathcal {D}^\perp \). Conversely, if (5.83) and (5.84) are satisfied, we observe that (5.85) is satisfied too if \(X\in \mathcal {D}\) and \(X\in \mathcal {D}^\perp \), respectively. Finally, if \(X\in \mathcal {M}\), we have \(vX=0\) and, by using that \(u\circ v=0\) and (5.67), \(\sigma (X,vY)=0\), for any \(Y\in \mathcal {D}\). Thus, (5.85) holds for any \(X\in \mathcal {X}(M)\).    \(\blacksquare \)

Corollary 3

Any normal generic submanifold of an S-manifold is a \((\mathcal {D},\mathcal {D}^\perp )\)-geodesic submanifold.

From (5.60), (5.67), (5.83), and (5.84), we have

$$\begin{aligned} \sigma (fX,Z)=f\sigma (X,Z), \end{aligned}$$
(5.86)
$$\begin{aligned} t\sigma (fX,fX)=t\sigma (X,X), \end{aligned}$$
(5.87)
$$\begin{aligned} A_{fZ}X\in \mathcal {D}, \end{aligned}$$
(5.88)

for any \(X\in \mathcal {}\) and any \(Z\in \mathcal {D}^\perp \). On the other hand, from (5.78) and (5.83)–(5.84), we deduce

Proposition 1

Each CR-product in an S-manifold is a normal CR-submanifold.

For the converse we prove the following theorems.

Theorem 5.16

Let M be a normal CR-submanifold of an S-manifold. Then, M is a CR-product if and only if the distribution \(\mathcal {D}\oplus \mathcal {M}\) is integrable.

Proof

The necessary condition is obvious. Conversely, let \(X\in \mathcal {D}\). If \(Y\in \mathcal {D}^\perp \), then (5.78) is (5.84). Further, if \(Y\in \mathcal {M}\), from (5.67) we get \(\sigma (X,Y)=0\). Finally, if \(Y\in \mathcal {D}\), from (5.76) and (5.83) we obtain (5.78).    \(\blacksquare \)

Theorem 5.17

Let M be a normal CR-submanifold of an S-manifold such that \(\mathrm{d}u=0\). Then, M is a CR-product.

Proof

A straightforward computation gives, by using the hypothesis and (5.72),

$$\begin{aligned} g((\nabla _Xv)Y,Z)=\sum _{{\alpha }=1}^s\{\mathrm{d}\eta _{\alpha }(vX,Y)\eta _{\alpha }(Z)-\mathrm{d}\eta _{\alpha }(vZ,X)\eta _{\alpha }(Y)\}, \end{aligned}$$
(5.89)

for any \(X,Y,Z\in \mathcal {X}(M)\). Now, if \(Y\in \mathcal {D}\), from (5.64) and (5.65) we get \(\mathrm{d}\eta _{\alpha }(vX,Y)=F(vX,Y)=g(vX,vY)=g(X,Y)\). So, (5.89) becomes

$$(\nabla _Xv)Y=\sum _{{\alpha }=1}^sg(X,Y)\xi _{\alpha }$$

for any \(X\in \mathcal {X}(M)\) and any \(Y\in \mathcal {D}\). Comparing with (5.72) we have \(\sigma (X,Y)=0\) and so (5.78) holds.    \(\blacksquare \)

We say that v is \(\eta \) -parallel if

$$\begin{aligned} (\nabla _Xv)Y=\sum _{{\alpha }=1}^s\{g(PX,PY)\xi _{\alpha }-\eta _{\alpha }(Y)PX\}, \end{aligned}$$

for any \(X,Y\in \mathcal {X}(M)\). Then, from (5.64), (5.65), and (5.89), we prove

Proposition 2

Any normal CR-submanifold of an S-manifold such that \(\mathrm{d}u=0\) is \(\eta \)-parallel.

Given a CR-submanifold M of an S-manifold, a vector field \(X\in \mathcal {X}(M)\) is said to be \(\mathcal {D}\) -Killing if

$$\begin{aligned} g(P\nabla _ZX,PY)+g(P\nabla _YX,PZ)=0, \end{aligned}$$
(5.90)

for any \(Y,Z\in \mathcal {X}(M)\). We notice that it is possible to characterize normal CR-submanifolds in terms of \(\mathcal {D}\)-Killing vector fields.

Theorem 5.18

A CR-submanifold M of an S-manifold is a normal CR-submanifold if and only if any \(Z\in \mathcal {D}^\perp \) is a \(\mathcal {D}\)-Killing vector field

Proof

Given \(X,Y\in \mathcal {X}(M)\) and \(Z\in \mathcal {D}^\perp \), from (5.74) we get

$$\begin{aligned} g(\nabla _XZ,Y)+g(\nabla _YZ,X)&=g(vA_{fZ}X,Y)-g(tD_XfZ,Y) \nonumber \\&\quad + g(vA_{fZ}Y,X)-g(tD_YfZ,X). \end{aligned}$$
(5.91)

But \(g(vA_{fZ}Y,X)=-g(A_{fZ}vX,Y)\) and so, from (5.91)

$$\begin{aligned}&g(P\nabla _XZ,PY)+g(P\nabla _YZ,PX)+g(Q\nabla _XZ,QY)+g(Q\nabla _YZ,QX) \nonumber \\&\qquad +\sum _{{\alpha }=1}^s\{\eta _{\alpha }(\nabla _XZ)\eta _{\alpha }(Y)+\eta _{\alpha }(\nabla _YZ)\eta _{\alpha }(X)\} \nonumber \\&\quad = g((vA_{fZ}-A_{fZ}v)X,Y)-g(tD_XfZ,Y)-g(tD_YfZ,X). \end{aligned}$$
(5.92)

Now, since it is easy to show that \(\eta _{\alpha }(\nabla _XZ)=0\) for any \({\alpha }=1,\dots ,s\), by using (5.75), we deduce that (5.92) becomes

$$\begin{aligned} g(P\nabla _XZ,PY)+g(P\nabla _YZ,PX)=g((vA_{fZ}-A_{fZ}v)X,Y). \end{aligned}$$
(5.93)

Consequently, if Z is a \(\mathcal {D}\)-Killing vector field, from (5.81) we obtain that M is a normal CR-submanifold. Conversely, if \(X\in \mathcal {D}\), the right part of the equality (5.93) vanishes by using (5.81). If \(X\in \mathcal {D}^\perp \), then \(vX=0\) and from (5.84), \(A_{fZ}X\in \mathcal {D}^\perp \), that is, \(vA_{fZ}X=0\) and the right part of (5.93) vanishes again. Finally, if \(X\in \mathcal {M}\), \(vX=0\) and from (5.68), \(A_{fZ}X\in \mathcal {D}^\perp \). In any case, from (5.93) we have (5.90).    \(\blacksquare \)

To end this section, we recall that a submanifold M of an S-manifold is said to be totally f-umbilical [19] if there exists a normal vector field V such that

$$\begin{aligned} \sigma (X,Y)=g(fX,fY)V+\sum _{{\alpha }=1}^s\{\eta _a(Y)\sigma (X,\xi _{\alpha })+\eta _{\alpha }(X)\sigma (Y,\xi _{\alpha })\}, \end{aligned}$$
(5.94)

for any \(X,Y\in \mathcal {X}(M)\). These submanifolds have been studied and classified in [9]. Since from (5.94) we easily get (5.83) and (5.84), then we have the following theorem.

Theorem 5.19

Any totally f-umbilical CR-submanifold of an S-manifold is a normal CR-submanifold.

5.3.3 Normal CR-Submanifolds of an S-space-form

Let \(\wedge M(c)\) a \((2m+s)\)-dimensional S-space-form, where c is denoting the constant f-sectional curvature and let M be a CR-submanifold. Firstly, we can prove

Proposition 3

If M is a normal CR-submanifold, then

$$\begin{aligned} \Vert A_{fZ}X\Vert ^2+\Vert \sigma (X,Z)\Vert ^2-g(t\sigma (Z,Z),t\sigma (X,X))=\frac{c+3s}{4}, \end{aligned}$$
(5.95)

for any unit vector fields \(X\in \mathcal {D}\) and \(Z\in \mathcal {D}^\perp \).

Proof

From the Codazzi equation, we have

$$\begin{aligned} \wedge R(X,fX,Z,fZ)&=g(D_X\sigma (fX,Z)-D_{fX}\sigma (X,Z),fZ)\nonumber \\&\quad -g(\sigma ([X,fX],Z),fZ)\nonumber \\&\quad +g(\sigma (X,\nabla _{fX}Z)-\sigma (fX,\nabla _XZ),fZ). \end{aligned}$$
(5.96)

Now, from (5.60), (5.84), and (5.86), a direct expansion gives

$$\begin{aligned} g(D_X\sigma (fX,Z)-D_{fX}\sigma (X,Z),fZ)=-2\Vert \sigma (X,Z)\Vert ^2. \end{aligned}$$
(5.97)

On the other hand, since \(X\in \mathcal {D}\) is a unit vector field (and so, fX too), we see from (5.59) that \(\eta _{\alpha }([X,fX])=2\) for any \({\alpha }\) and from (5.70) that \(Q[X,fX]=t\sigma (X,X)+t\sigma (fX,fX)\). Thus, taking into account (5.67), (5.84), and (5.87), we get

$$\begin{aligned} g(\sigma ([X,fX],Z),fZ)=2g(\sigma (t\sigma (X,X),Z),fZ)-2s. \end{aligned}$$
(5.98)

However, since \(Z\in \mathcal {D}^\perp \), by using (5.70) it is easy to show that

$$g(\sigma (t\sigma (X,X),Z),fZ)=-g(t\sigma (X,X),t\sigma (Z,Z)).$$

Therefore, from (5.98) we have

$$\begin{aligned} g(\sigma ([X,fX],Z),fZ)=-2s-2g(t\sigma (X,X),t\sigma (Z,Z)). \end{aligned}$$
(5.99)

Next, since \(\eta _{\alpha }(\nabla _{fX}Z)=\eta _{\alpha }(\nabla _XZ)=0\) for any \({\alpha }\), from (5.69), (5.83), (5.84), and (5.88), we obtain

$$\begin{aligned} g(\sigma (X,\nabla _{fX}Z)-\sigma (fX,\nabla _XZ),fZ)=-2\Vert A_{fZ}X\Vert ^2. \end{aligned}$$
(5.100)

Finally, from (5.61) we deduce \(\wedge R(X,fX,Z,fZ)=-(c-s)/2\). Then, substituting (5.97), (5.99), and (5.100) into (5.96), we complete the proof.    \(\blacksquare \)

Corollary 4

If M is a normal \(\mathcal {D}^\perp \)-geodesic CR-submanifold of an S-space-form \(\wedge M(c)\), then \(c\ge -3s\).

Proposition 4

If M is a normal CR-submanifold of an S-space-form \(\wedge M(c)\) such the distribution \(\mathcal {D}\oplus \mathcal {M}\) is integrable, then \(c\ge -3s\) and M is a CR-product.

Proof

It is clear that M is a CR-product due to Theorem 5.16. Moreover, from (5.78) we have \(g(\sigma (X,Y),fZ)=0\). for any \(X,Y\in \mathcal {D}\). Then, if \(X\in \mathcal {D}\) is a unit vector field, \(t\sigma (X,X)=0\) and, by using (5.95), \(c\ge -3s\).    \(\blacksquare \)

Now, we are going to study the concrete case of the \((2m + s)\)-dimensional euclidean S-space-form \(\mathbb R^{2m+s}(-3s)\) (see [13] for the details of this structure). In this context, we can prove

Theorem 5.20

If M is a normal \((\mathcal {D},\mathcal {D}^\perp )\)-geodesic and \(\mathcal {D}^\perp \)-geodesic CR-submanifold of \(\mathbb R^{2m+s}(-3s)\), then it is a CR-product.

Proof

From (5.95), we have \(A_{fZ}X=0\) for any \(X\in \mathcal {D}\) and any \(Z\in \mathcal {D}^\perp \). So, from (5.77), M is a CR-product.    \(\blacksquare \)

Corollary 5

A normal \(\mathcal {D}^\perp \)-geodesic generic submanifold of \(\mathbb R^{2m+s}(-3s)\) is a CR-product.

Another interesting example of S-space-form is \(\mathbb H^{2m+s}(4-3s)\), a generalization of the Hopf fibration \(\pi :\S ^{2m+1}\longrightarrow \mathbb PC^{m}\), introduced by Blair in [5] as a canonical example of an S-manifold playing the role of the complex projective space in Kaehler geometry and the odd-dimensional sphere in Sasakian geometry. This space is given by (see [5, 6] for more details)

and its f-sectional curvature is constant equal to \(4-3s\). Let M be a CR-submanifold of \(\mathbb H^{2m+s}(4-3s)\) (we always suppose \(s\ge 2\)). Denote by \(\nu \) the orthogonal complementary distribution of \(f\mathcal {D}^\perp \) in \(T(M)^\perp \). Then, \(f\nu \subseteq \nu \). Let

$$\begin{aligned} \{E_1,\dots ,E_{2p}\}, \;\;\; \{F_1,\dots ,F_q\}, \;\;\; \{V_1,\dots ,V_r,fV_1,\dots ,fV_r\} , \end{aligned}$$

be local fields of orthonormal frames on \(\mathcal {D}\), \(\mathcal {D}^\perp \) and \(\nu \), respectively, where 2r is the real dimension of \(\nu \). First, we prove

Lemma 6

If M is a CR-product in \(\mathbb H^{2m+s}(4-3s)\), then

$$\begin{aligned} \Vert \sigma (X,Z)\Vert =1, \end{aligned}$$
(5.101)

for any unit vector fields \(X\in \mathcal {D}\) and \(Z\in \mathcal {D}^\perp \).

Proof

We know, from Proposition 1, that M is a normal CR-submanifold. Since, \(c=4-3s\), from (5.77), (5.78) and (5.95) we complete the proof.    \(\blacksquare \)

Lemma 7

If M is a CR-product in \(\mathbb H^{2m+s}(4-3s)\), the vector field \(\sigma (E_i,F_a)\), \(i=1,\dots ,2p\) and \(a=1,\dots ,q\), are 2pq orthonormal vector fields on \(\nu \).

Proof

From (5.101) and by the linearity, we get \(g(\sigma (E_i,Z),\sigma (E_j,Z))=0\), for any \(i,j=1,\dots ,2p\), \(i\ne j\) and any unit vector field \(Z\in \mathcal {D}^\perp \). Now, from (5.84), if \(q=1\), we complete the proof. If \(q\ge 2\), by linearity again, we have \(g(\sigma (E_i,F_a),\sigma (E_j,F_b))+g(\sigma (E_i,F_b),\sigma (E_j,F_a))=0\), for any \(i,j=1,\dots ,2p\), \(i\ne j\), \(a,b=1,\dots ,q\), \(a\ne b\). Next, by using (5.79) and the Bianchi identity, we obtain \(R(X,Y,Z,W)=0\), for any \(X,Y\in \mathcal {D}\), \(Z,W\in \mathcal {D}^\perp \), where R is denoting the curvature tensor field of M. But, if \(i\ne j\) and \(a\ne b\), (5.61) gives \(\wedge R(E_i,E_j,F_a,F_b)=0\). Then, from the Gauss equation we get

$$g(\sigma (E_i,F_a),\sigma (E_j,F_b))-g(\sigma (E_i,F_b),\sigma (E_j,F_a))=0,$$

for any \(i,j=1,\dots ,2p\), \(i\ne j\), \(a,b=1,\dots ,q\), \(a\ne b\) and this completes the proof.    \(\blacksquare \)

Now, we study the normal CR-submanifolds of \(\mathbb H^{2m+s}(4-3s)\).

Theorem 5.21

Let M be a normal CR-submanifold of \(\mathbb H^{2m+s}(4-3s)\), \(s\ge 2\), such that the distribution \(\mathcal {D}\oplus \mathcal {M}\) is integrable. Then

  1. (i)

    M is a CR-product \(M_1\times M_2\).

  2. (ii)

    \(m\ge pq+p+q\).

  3. (iii)

    If \(n=pq+p+q\), then \(M_1\) is an invariant totally geodesic submanifold immersed in \(\mathbb H^{2m+s}(4-3s)\).

  4. (iv)

    \(\Vert \sigma \Vert ^2\ge 2q(2p+s)\).

  5. (v)

    If \(\Vert \sigma \Vert ^2=2q(2p+s)\), then \(M_1\) is an S-space-form of constant f-sectional curvature \(4-3s\) and \(M_2\) has constant curvature 1.

  6. (vi)

    If M is a minimal submanifold, then \(\rho \le 4p(p+1)+2p(q+s)+q(q-1)\), where \(\rho \) denotes the scalar curvature and the equality holds if and only if \(\Vert \sigma \Vert ^2=2q(2p+s)\).

Proof

(i) follows directly from Proposition 4. Now, from Lemma 7, \(dim(\nu )=2(m-p)-2q\ge 2pq\). So (ii) holds. Next, suppose that \(m=pq+p+q\). If \(X,Y,Z\in \mathcal {D}\) and \(W\in \mathcal {D}^\perp \), from (5.61), \(\wedge R(X,Y,Z,W)=0\) and, by using a similar proof to that one of Lemma 7, \(R(X,Y,Z,W)=0\). So, the Gauss equation gives

$$\begin{aligned} g(\sigma (X,W),\sigma (Y,Z))-g(\sigma (X,Z),\sigma (Y,W))=0. \end{aligned}$$
(5.102)

Since from Proposition 3.2 of [12], \(\sigma (fX,Z)=f\sigma (X,Z)\), if we put \(Y=fX\), we have, by using (5.86), \(g(\sigma (fX,W),(\sigma (X,Z))=0\). Now, if we put \(Z=fY\), then \(g(\sigma (X,Y),\sigma (X,W))=0\). Thus, by linearity, we get \(g(\sigma (X,W),\sigma (Y,Z))+g(\sigma (X,Z),\sigma (Y,W))=0\). Consequently, from (5.102)

$$\begin{aligned} g(\sigma (X,W),\sigma (Y,Z)) =0, \end{aligned}$$
(5.103)

for any \(X,Y,Z\in \mathcal {D}\) and \(W\in \mathcal {D}^\perp \). Since now \(dim(\nu )=2pq\), (5.103) implies that \(\sigma (X,Y)=0\), for any \(X,Y\in \mathcal {D}\) and so, (iii) holds from Theorem 2.4(ii) of [12]. Assertions (iv) and (v) follow from Theorem 4.2 of [12]. Finally, if M is a minimal normal CR-submanifold of \(\mathbb H^{2m+s}(4-3s)\), a straightforward computation gives

$$\rho =4p(p+1)+2s(p+q)+q(q-1)+6pq-\Vert \sigma \Vert ^2.$$

Then, by using (iv), the proof is complete.    \(\blacksquare \)

Theorem 5.22

Let M be a normal, \((\mathcal {D},\mathcal {D}^\perp )\)-geodesic and \(\mathcal {D}^\perp \)-geodesic CR-submanifold of \(\mathbb H^{2m+s}(4-3s)\). Then,

  1. (i)

    \(\Vert A_{fZ}X\Vert =1\), for any unit vector fields \(X\in \mathcal {D}\) and \(Z\in \mathcal {D}^\perp \).

  2. (ii)

    \(\Vert \sigma \Vert ^2\ge 2q(p+s)\) and the equality hold if and only if \(\sigma (\mathcal {D},\mathcal {D})\in f\mathcal {D}^\perp \).

Proof

(i) follows immediately from (5.95). Now, considering the above-mentioned local fields of orthonormal frames for \(\mathcal {D}\), \(\mathcal {D}^\perp \), and \(\nu \), a straightforward computation using the hypothesis gives (ii).    \(\blacksquare \)

Finally, from (5.84) and (5.95), we can prove

Corollary 6

Let M be a normal \(\mathcal {D}^\perp \)-geodesic generic submanifold of \(\mathbb H^{2m+s}(4-3s)\). Then

  1. (i)

    \(\Vert A_{fZ}X\Vert =1\), for any unit vector fields \(X\in \mathcal {D}\) and \(Z\in \mathcal {D}^\perp \).

  2. (ii)

    \(\Vert \sigma \Vert ^2=2q(p+s)\).