1 Introduction

A non-flat complex space form is a Kähler manifold of constant holomorphic sectional curvature c, which is analytically isometric either to complex projective space \({\mathbb {C}}P^{n}\), if \(c>0\) or to complex hyperbolic space \({\mathbb {C}}H^{n}\), if \(c<0\). We will suppose that in the case of \({\mathbb {C}}P^{n}\) the holomorphic sectional curvature is equal to 4 and in the case of \({\mathbb {C}}H^{n}\) the holomorphic sectional curvature is equal to \(-\) 4. The symbol \(M_{n}(c), n\ge 2,\) is used to denote them when it is not necessary to distinguish them.

Let \(M_{n}(c)\), \(n\ge 2\), be a non-flat complex space form endowed with the Kählerian structure (JG), where J is the complex structure and G is the metric of constant holomorphic sectional curvature c. Suppose M to be a connected real hypersurface of \(M_{n}(c)\) without boundary and consider a locally defined unit normal vector field N on M. The shape operatorA of the real hypersurface M in \(M_{n}(c)\) with respect to N is given by

$$\begin{aligned} {\overline{\nabla }}_{X}N=-AX, \end{aligned}$$

where \({\overline{\nabla }}\) is the Levi-Civita connection of \(M_{n}(c)\). Furthermore, the Kählerian structure of \(M_{n}(c)\) induces on M an almost contact metric structure \((\phi , \xi , \eta , g)\), where \(\phi X\) is the tangent component of JX, \(\eta \) is an one-form given by \(\eta (X)=g(X,\xi )\) for any X tangent to M, \(\xi =-JN\) is the structure vector field and g is the induced Riemannian metric. The maximal holomorphic distribution on M is defined by \({\mathbb {D}}=Ker(\eta )\).

The eigenvectors of the shape operator A are called principal vectors and the corresponding eigenvalues are called principal curvatures. A real hypersurface in \(M_{n}(c)\) is a Hopf hypersurface, when the structure vector field \(\xi \) is a principal vector of the shape operator, i.e. \(A\xi =g(A\xi ,\xi )\xi \). The corresponding eigenvalue \(g(A\xi ,\xi )=\alpha \) is called Hopf principal curvature and is constant.

Takagi initiated the study of real hypersurfaces in \({\mathbb {C}}P^n,n\ge 2\). He classified the homogeneous real hypersurfaces in \({\mathbb {C}}P^n,n\ge 2\) (see [24,25,26]). In [10], Kimura proved that these are the unique Hopf hypersurfaces with constant principal curvatures in \({\mathbb {C}}P^n\). Takagi’s list contains 6 types of real hypersurfaces:

  • Type \((A_1)\) are geodesic hyperspheres of radius r, \(0< r < \frac{\pi }{2}\). They have 2 distinct constant principal curvatures, \(2{\hbox {cot}}(2r)\) with eigenspace \({\mathbb {R}}[\xi ]\) and \({\hbox {cot}}(r)\) with eigenspace \({\mathbb {D}}\).

  • Type \((A_2)\) are tubes of radius r, \(0< r < \frac{\pi }{2}\), over totally geodesic complex projective spaces \({\mathbb {C}}P^m\), \(0< m < n-1\). They have 3 distinct constant principal curvatures, \(2{\hbox {cot}}(2r)\) with eigenspace \({\mathbb {R}}[\xi ]\), \({\hbox {cot}}(r)\) and \(-{\hbox {tan}}(r)\). The corresponding eigenspaces of \({\hbox {cot}}(r)\) and \(-{\hbox {tan}}(r)\) are complementary and \(\phi \)-invariant distributions in \({\mathbb {D}}\).

  • Type (B) are tubes of radius r, \(0< r < \frac{\pi }{4}\), over the complex quadric. They have 3 distinct constant principal curvatures, \(2{\hbox {cot}}(2r)\) with eigenspace \({\mathbb {R}}[\xi ]\), \({\hbox {cot}}(r-\frac{\pi }{4})\) and \(-{\hbox {tan}}(r-\frac{\pi }{4})\) whose corresponding eigenspaces are complementary and equal dimensional distributions in \({\mathbb {D}}\) such that \(\phi T_{{\hbox {cot}}(r-\frac{\pi }{4})}=T_{-{\hbox {tan}}(r-\frac{\pi }{4})}\).

  • Type (C) are tubes of radius r, \(0< r < \frac{\pi }{4}\), over the Segre embedding of \({\mathbb {C}}P^1 \times {\mathbb {C}}P^n\), where \(2n+1=m\) and \(m \ge 5\). They have 5 distinct constant principal curvatures, \(2{\hbox {cot}}(2r)\) with eigenspace \({\mathbb {R}}[\xi ]\), \({\hbox {cot}}(r-\frac{\pi }{4})\) with multiplicity 2, \({\hbox {cot}}(r-\frac{\pi }{2})=-{\hbox {tan}}(r)\) with multiplicity \(m-3\), \({\hbox {cot}}(r-\frac{3\pi }{4})\), with multiplicity 2 and \({\hbox {cot}}(r-\pi )={\hbox {cot}}(r)\) with multiplicity \(m-3\). Moreover \(\phi T_{{\hbox {cot}}(r-\frac{\pi }{4})}=T_{{\hbox {cot}}(r-\frac{3\pi }{4})}\) and \(T_{-{\hbox {tan}}(r)}\) and \(T_{{\hbox {cot}}(r)}\) are \(\phi \)-invariant.

  • Type (D) are tubes of radius r, \(0< r < \frac{\pi }{4}\), over the Plucker embedding of the complex Grassmannian manifold G(2, 5) in \({\mathbb {C}}P^9\). They have the same principal curvatures as type (C) real hypersurfaces, \(2{\hbox {cot}}(2r)\) with eigenspace \({\mathbb {R}}[\xi ]\), and the other 4 principal curvatures have the same multiplicity 4 and their eigenspaces have the same behaviour with respect to \(\phi \) as in type (C).

  • Type (E) are tubes of radius r, \(0< r < \frac{\pi }{4}\), over the canonical embedding of the Hermitian symmetric space SO(10) / U(5) in \({\mathbb {C}}P^{15}\). They also have the same principal curvatures as type (C) real hypersurfaces, \(2 {\hbox {cot}}(2r)\) with eigenspace \({\mathbb {R}}[\xi ]\), \({\hbox {cot}}(r-\frac{\pi }{4})\) and \({\hbox {cot}}(r-\frac{3\pi }{4})\) have multiplicities equal to 6 and \(-{\hbox {tan}}(r)\) and \({\hbox {cot}}(r)\) have multiplicities equal to 8. Their corresponding eigenspaces have the same behaviour with respect to \(\phi \) as in type (C).

From now on, the notion of real hypersurfaces of type (A) refers to both type \((A_1)\) and type \((A_2)\).

In the case of \({\mathbb {C}}H^{n},n\ge 2\), Montiel classified real hypersurfaces with at most two constant principal curvatures (see [15]). Later, in [1] Berndt classified Hopf hypersurfaces in \({\mathbb {C}}H^{n},n\ge 2\), with constant principal curvatures. Berndt’s list contains 4 types of real hypersurfaces:

  • Type (\(A_{0}\)) are horospheres.

  • Type (\(A_{1}\)) are geodesic hyperspheres or tubes over a totally geodesic complex hyperbolic hyperplane \({\mathbb {C}}H^{n - 1}\).

  • Type \((A_{2}) \) are tubes over a totally geodesic \({\mathbb {C}}H^{k}\), \((1\le k\le n-2)\).

  • Type (B) are tubes over totally geodesic real hyperbolic space \({\mathbb {R}}H^{n}\).

From now on, the notion of real hypersurfaces of type (A) refers to types (\(A_{0}\)), (\(A_{1}\)) and (\(A_{2}\)). All the above Hopf hypersurfaces are homogeneous but in contrast to \({\mathbb {C}}P^{n}\) there are homogeneous real hypersurfaces which are not Hopf (see [2]).

Apart from Hopf hypersurfaces, there are ruled real hypersurfaces in \(M_{n}(c)\). A ruled real hypersurface can be described as follows: consider a regular curve \(\gamma \) in \(M_{n}(c)\) with tangent vector field X. At each point of \(\gamma \), there is a unique hyperplane of \(M_{n}(c)\) cutting \(\gamma \) in a way to be orthogonal to both X and JX. The union of all these hyperplanes is called ruled hypersurface. It will be an embedded hypersurface locally, although globally it will in general have self-intersections and singularities. Equivalently, a ruled real hypersurface satisfies that the maximal holomorphic distribution \({\mathbb {D}}\) of M at any point p is integrable and it has an integrable manifold \(M_{n-1}(c)\), i.e. \(g(A{\mathbb {D}}, {\mathbb {D}})=0\). For examples of ruled real hypersurfaces see [11, 13].

The Tanaka-Webster connection, [27, 29], is the canonical affine connection defined on a non-degenerate, pseudo-Hermitian CR-manifold. As a generalization of this connection, Tanno in [28], defined the generalized Tanaka-Webster connection for contact metric manifolds by

$$\begin{aligned} {\hat{\nabla }}_XY=\nabla _XY+(\nabla _X \eta )(Y)\xi -\eta (Y)\nabla _X \xi -\eta (X)\phi Y. \end{aligned}$$
(1.1)

Using the naturally extended affine connection of Tanno’s generalized Tanaka-Webster connection, in [5, 6] Cho defined the k-th generalized Tanaka-Webster connection\({\hat{\nabla }}^{(k)}\) for a real hypersurface M in \(M_{n}(c)\) given by

$$\begin{aligned} {\hat{\nabla }}^{(k)}_XY=\nabla _XY+g(\phi AX,Y)\xi -\eta (Y)\phi AX-k\eta (X)\phi Y \end{aligned}$$
(1.2)

for any XY tangent to M where k is a nonnull real number and A is the shape operator of M. Then relations \({\hat{\nabla }}^{(k)}\eta =0\), \({\hat{\nabla }}^{(k)}\xi =0\), \({\hat{\nabla }}^{(k)}g=0\), \({\hat{\nabla }}^{(k)}\phi =0\) hold. In particular, if the shape operator of a real hypersurface satisfies \(\phi A+A \phi =2k\phi \), the kth generalized Tanaka-Webster connection coincides with the Tanaka-Webster connection.

The tensor field of type (1,2) given by the difference of both connections \(F^{(k)}(X,Y)=g(\phi AX,Y)\xi -\eta (Y)\phi AX-k\eta (X)\phi Y\), for any XY tangent to M is called k-th Cho tensor on M (see [12] Proposition 7.10, pp. 234–235). Associated with it, for any X tangent to M and any nonnull real number k the tensor field of type (1,1) \(F_X^{(k)}\), is called k-th Cho operator corresponding toX and is given by \(F_X^{(k)}Y=F^{(k)}(X,Y)\) for any \(Y \in TM\). If \(X \in {\mathbb {D}}\), the corresponding Cho operator does not depend on k and is simply written by \(F_X\). Finally, the torsion of the connection \({\hat{\nabla }}^{(k)}\) is given by \({\hat{T}}^{(k)}(X,Y)=F_X^{(k)}Y-F_Y^{(k)}X\) for any XY tangent to M.

Let K be a symmetric operator on M. Then relation \(\nabla _XK={\hat{\nabla }}_X^{(k)}K\) for a vector field X tangent to M due to (1.2) is equivalent to the fact that \(KF_X^{(k)}=F_X^{(k)}K\), which implies that every eigenspace of K is preserved by the kth Cho operator \(F_X^{(k)}\). Since, at each point \(P \, \in \, M\) the tangent space is decomposed as \(T_{P}M={\hbox {Span}} \{ \xi \} \oplus {\mathbb {D}}\) the above problem either for any \(X \in {\mathbb {D}}\) or for \(X=\xi \) will be studied.

In the case of \({\mathbb {C}}P^{n},n\ge 3,\) in [23] we consider the problem when \(K=A\), obtaining ruled real hypersurfaces for any \(X \in {\mathbb {D}}\) and type (A) real hypersurfaces when \(X=\xi \). As a consequence, we proved non-existence of real hypersurfaces in \({\mathbb {C}}P^n,n\ge 3,\) such that \(\nabla A={\hat{\nabla }}^{(k)}A\) for any nonnull k.

A similar study in the case of the structure Jacobi operator \(R_{\xi }\) of M was made in [20], also obtaining ruled real hypersurfaces for any \(X \in {\mathbb {D}}\) and either real hypersurfaces of type (A) or tubes of radius \(\frac{\pi }{4}\) over a complex submanifold of \({\mathbb {C}}P^n,n\ge 3,\) if \(X=\xi \). This second type of real hypersurfaces are those such that \(A\xi =0\) (see [4]). As above, there do not exist real hypersurfaces M in \({\mathbb {C}}P^n,n\ge 3,\) such that \(\nabla R_{\xi }={\hat{\nabla }}^{(k)}R_{\xi }\), for any nonnull k. In [19], the previous question was answered for real hypersurfaces in \({\mathbb {C}}H^{n},n\ge 2\), and for three-dimensional real hypersurfaces in \(M_{2}(c)\). More precisely, in both cases in the case of \(X \, \in \, {\mathbb {D}}\) the real hypersurface is a ruled one and in the case of \(X=\xi \) the real hypersurfaces is locally congruent to a real hypersurface of type A or to a real hypersurface with \(A\xi =0\) (for the construction of real hypersurfaces with \(A\xi =0\) in \({\mathbb {C}}H^{2}\) see [7]).

In this paper, the analogous problem when \(K=\phi A-A\phi \) will be considered. Relation \(\phi A-A\phi =0\) implies that the shape operator commutes with structure tensor \(\phi \). Real hypersurfaces satisfying the previous relation were studied by Okumura in the case of \({\mathbb {C}}P^{n}\), \(n\ge 2\), in [18] and by Montiel and Romero in the case of \({\mathbb {C}}H^{n}\), \(n\ge 2\) in [16]. The following Theorem provides the above classification of real hypersurfaces in \(M_{n}(c)\), \(n\ge 2\).

Theorem 1.1

Let M be a real hypersurface of \(M_{n}(c)\), \(n\ge 2\). Then \(\phi A=A\phi \), if and only if M is locally congruent to a homogeneous real hypersurface of type (A). More precisely:

In the case of \({\mathbb {C}}P^{n}\)

\((A_{1})\):

a geodesic hypersphere of radius r , where \(0<r<\frac{\pi }{2}\),

\((A_{2})\):

a tube of radius r over a totally geodesic \({\mathbb {C}}P^{k}\),\((1\le k\le n-2)\), where \(0<r<\frac{\pi }{2}\).

In the case of \({\mathbb {C}}H^{n}\)

\((A_{0})\):

a horosphere in \( {\mathbb {C}}H^{n}\),

\((A_{1})\):

a geodesic hypersphere or a tube over a totally geodesic complex hyperbolic hyperplane \({\mathbb {C}}H^{n - 1}\),

\((A_{2})\):

a tube over a totally geodesic \({\mathbb {C}}H^{k}\, (1\le k\le n-2)\).

The question if there are real hypersurfaces in \(M_{n}(c),n\ge 2,\) whose eigenspaces of the symmetric tensor \(\phi A-A\phi \) are preserved by the kth Cho operator will be answered. More precisely, the following Theorems are proved

Theorem 1.2

Let M be a real hypersurface in \(M_{n}(c)\), \(n\ge 2\). Let k be a nonnull constant. Then \(\nabla _X(\phi A-A\phi )={\hat{\nabla }}_X^{(k)}(\phi A-A\phi )\) for any \(X \in {\mathbb {D}}\) if and only if M is locally congruent either to a ruled real hypersurface or to a real hypersurface of type (A).

Theorem 1.3

Let M be a real hypersurface in \(M_{n}(c)\), \(n \ge 2\). Let k be a nonnull constant. Then \(\nabla _{\xi }(\phi A-A\phi )={\hat{\nabla }}_{\xi }^{(k)}(\phi A-A\phi )\) if and only if M is locally congruent to a real hypersurface of type (A).

As a consequence of both Theorems, the following is obtained

Corollary 1.1

Let M be a real hypersurface in \(M_{n}(c)\), \(n \ge 2\), and a nonnull constant k. Then \(\nabla (\phi A-A\phi )={\hat{\nabla }}^{(k)}(\phi A-A\phi )\) if and only if M is locally congruent to a real hypersurface of type (A).

Let \({\mathcal {L}}\) denote the Lie derivative on M. Therefore, \(\mathcal{L}_XY=\nabla _XY-\nabla _YX\) for any XY tangent to M. Associated with the kth generalized Tanaka-Webster connection a differential operator of first order can be defined and will be called the derivative of Lie type associated with such a connection and is given by

$$\begin{aligned} {{\mathcal {L}}}^{(k)}_XY={\hat{\nabla }}^{(k)}_XY-{\hat{\nabla }}^{(k)}_YX=\mathcal{L}_XY+{\hat{T}}^{(k)}_XY, \end{aligned}$$

for any XY tangent to the real hypersurface M. Given X tangent to M, the k-th torsion operator associated with X, \({\hat{T}}_X^{(k)}\), is considered as the operator such that \({\hat{T}}_X^{(k)}Y={\hat{T}}^{(k)}(X,Y)\) for any Y tangent to M. Then, given a symmetric operator K on M, \({{\mathcal {L}}}_XK=\mathcal{L}_X^{(k)}K\) for a tangent vector field X is equivalent to \(K{\hat{T}}_X^{(k)}={\hat{T}}_X^{(k)}K\), which implies that any eigenspace of K is preserved by \({\hat{T}}_X^{(k)}\).

In [21], this problem for \(K=A\) has been studied in the case of \({\mathbb {C}}P^{n},n\ge 3\), and it has been proved that ruled real hypersurfaces for any \(X \in {\mathbb {D}}\) satisfy the above condition and real hypersurfaces of type (A) when \(X=\xi \). Thus, the non-existence of real hypersurfaces in \({\mathbb {C}}P^m\) such that \(\mathcal{L}A = {{\mathcal {L}}}^{(k)}A\), for any nonnull k is assured. In [22] the same problem is solved when \(K=R_{\xi }\), obtaining a non-existence result when \(X \in {\mathbb {D}}\) and either real hypersurfaces of type (A) or tubes of radius \(\frac{\pi }{4}\) over a complex submanifold in \({\mathbb {C}}P^m\) when \(X=\xi \).

In this paper, the case \(K=\phi A-A\phi \) is examined and the following Theorems are proved

Theorem 1.4

Let M be a real hypersurface in \(M_{n}(c)\), \(n\ge 2\), and k a nonnull constant. Then \({{\mathcal {L}}}_{\xi }(\phi A-A\phi )=\mathcal{L}_{\xi }^{(k)}(\phi A-A\phi )\) if and only if M is locally congruent to either

  1. 1.

    a real hypersurface of type (A), or

  2. 2.

    in the case of \({\mathbb {C}}P^{n}\), a real hypersurface of type (B), (C), (D) or (E) whose radius r, \(0< r < \frac{\pi }{4}\), satisfies \(\mathrm{tan}(2r)=-k\) and in the case of \({\mathbb {C}}H^{n}\), a real hypersurface of type (B), whose radius r satisfies \(\mathrm{tanh}(2r)=\frac{1}{k}\).

Theorem 1.5

Let M be a real hypersurface in \(M_{n}(c),n\ge 2\). Let k be a nonnull constant. Then \({{\mathcal {L}}}_X(\phi A-A\phi )=\mathcal{L}_X^{(k)}(\phi A-A\phi )\) for any \(X \in {\mathbb {D}}\) if and only if M is locally congruent to a real hypersurface of type (A).

As a direct consequence of these Theorems we have

Corollary 2

Let M be a real hypersurface in \(M_{n}(c),n\ge 2\), and a nonnull constant k. Then \({{\mathcal {L}}}(\phi A-A\phi )={{\mathcal {L}}}^{(k)}(\phi A-A\phi )\) if and only if M is locally congruent to a real hypersurface of type (A)

2 Preliminaries

Throughout this paper, all manifolds, vector fields, etc., will be considered of class \(C^{\infty }\) unless otherwise stated. Let M be a connected real hypersurface in \(M_{n}(c),n\ge 2\), without boundary. Let N be a locally defined unit normal vector field on M. Let \(\nabla \) be the Levi-Civita connection on M and (JG) the Kälerian structure of \(M_{n}(c)\).

For any vector field X tangent to M we write \(JX=\phi X+\eta (X)N\), and \(-JN=\xi \). Then \((\phi ,\xi ,\eta ,g)\) is an almost contact metric structure on M (see [3]). That is, we have

$$\begin{aligned} \phi ^2X=-X+\eta (X)\xi , \quad \eta (\xi )=1, \quad g(\phi X,\phi Y)=g(X,Y)-\eta (X)\eta (Y) \end{aligned}$$
(2.1)

for any tangent vectors XY to M. From (2.1), we obtain

$$\begin{aligned} \phi \xi =0, \quad \eta (X)=g(X,\xi ). \end{aligned}$$
(2.2)

From the parallelism of J, we get

$$\begin{aligned} (\nabla _X\phi )Y=\eta (Y)AX-g(AX,Y)\xi \;\;\text{ and }\;\;\nabla _X\xi =\phi AX \end{aligned}$$
(2.3)

for any XY tangent to M, where A denotes the shape operator of the immersion. As the ambient space has holomorphic sectional curvature c, the equations of Gauss and Codazzi are given, respectively, by

$$\begin{aligned} \begin{aligned} R(X,Y)Z&= \displaystyle \frac{c}{4}[ g(Y,Z)X - g(X,Z)Y + g(\phi Y,Z)\phi X - g(\phi X,Z)\phi Y \\&\quad - 2g(\phi X,Y)\phi Z] + g(AY,Z)AX - g(AX,Z)AY, \end{aligned} \end{aligned}$$
(2.4)

and

$$\begin{aligned} (\nabla _XA)Y-(\nabla _YA)X=\frac{c}{4}[\eta (X)\phi Y-\eta (Y)\phi X-2g(\phi X,Y)\xi ] \end{aligned}$$
(2.5)

for any tangent vectors XYZ to M, where R is the curvature tensor of M.

In the sequel the following result is needed. In the case of complex projective space \({\mathbb {C}}P^{n}\) is owed to Maeda [14] and in the case of complex hyperbolic space \({\mathbb {C}}H^{n}\) is owed to Ki and Suh [9] (see also Corollary 2.3 in [17]).

Theorem 2.1

Let M be a Hopf hypersurface in \(M_{n}(c)\), \(n\ge 2\). Then

  1. (i)

    \(\alpha =g(A\xi ,\xi )\) is constant.

  2. (ii)

    If W is a vector field which belongs to \({\mathbb {D}}\) such that \(AW=\lambda W\), then

    $$\begin{aligned} \left( \lambda -\frac{\alpha }{2}\right) A\phi W=\left( \frac{\lambda \alpha }{2}+\frac{c}{4}\right) \phi W. \end{aligned}$$
  3. (iii)

    If the vector field W satisfies \(AW=\lambda W\) and \(A\phi W=\mu \phi W\) then

    $$\begin{aligned} \lambda \mu =\frac{\alpha }{2}(\lambda +\mu )+\frac{c}{4}. \end{aligned}$$
    (2.6)

3 Proof of Theorem 1.2

Let M a real hypersurface satisfying \((\nabla _X(\phi A-A\phi ))Y=({\hat{\nabla }}_X^{(k)}(\phi A-A\phi ))Y\) for any \(X \in {\mathbb {D}}\), Y tangent to M, which is equivalent to relation \(F_X(\phi A-A\phi )Y=(\phi A-A\phi )F_XY\). Therefore,

$$\begin{aligned}&g(\phi AX,(\phi A-A\phi )Y)\xi -\eta ((\phi A-A\phi )Y)\phi AX\nonumber \\&\quad =(\phi A-A\phi )(g(\phi AX,Y)\xi -\eta (Y)\phi AX) \end{aligned}$$
(3.1)

for any \(X \in {\mathbb {D}}\), Y tangent to M. Suppose that M is a non-Hopf real hypersurface. Thus, locally we can write \(A\xi =\alpha \xi +\beta U\), where U is a unit vector field in \({\mathbb {D}}\), \(\alpha \) and \(\beta \) are functions on M and \(\beta \ne 0\). Moreover, \({{\mathbb {D}}}_U\) is defined to be the orthogonal complementary distribution in \({\mathbb {D}}\) to the one spanned by U and \(\phi U\).

Relation (3.1) for \(Y=\xi \) implies \(g(\phi AX,\phi A\xi )\xi =-(\phi A-A\phi )\phi AX\), for any \(X \in {\mathbb {D}}\), that is, \(\beta g(AX,U)\xi =-\phi A\phi AX+A\phi ^2AX\), for any \(X \in {\mathbb {D}}\). The scalar product of the latter relation with \(\xi \) yields \(\beta g(AX,U)=g(\phi ^2AX,A\xi )=\beta g(\phi ^2AX,U)=-\beta g(AX,U)\), for any \(X \in {\mathbb {D}}\). Therefore, \(g(AX,U)=0\), for any \(X \in {\mathbb {D}}\),which implies

$$\begin{aligned} AU=\beta \xi . \end{aligned}$$
(3.2)

Take now \(Y=U\) in (3.1). We obtain \(-g(\phi AX,A\phi U)\xi =g(\phi AX,U)(\phi A-A\phi )\xi =\beta g(\phi AX,U)\phi U\), for any \(X \in {\mathbb {D}}\). Its scalar product with \(\phi U\) gives \(g(\phi AX,U)=0\) for any \(X \in {\mathbb {D}}\) and then

$$\begin{aligned} A\phi U=0 \end{aligned}$$
(3.3)

Take \(Y=\phi U\) in (3.1). Then \(\eta (AU)\phi AX=0\) for any \(X \in {\mathbb {D}}\). As \(\eta (AU)=\beta \ne 0\), this yields \(\phi AX=0\) for any \(X \in {{\mathbb {D}}}_U\). Therefore

$$\begin{aligned} AX=0 \end{aligned}$$
(3.4)

for any \(X \in {{\mathbb {D}}}_U\). From (3.2), (3.3) and (3.4) we conclude that M is ruled. Conversely, if M is ruled \(F_X=0\) for any \(X \in {\mathbb {D}}\) and (3.1) is satisfied.

Next the case of M being a Hopf hypersurface is examined. In this case, \(A\xi =\alpha \xi \). Let \(W \, \in \, {\mathbb {D}}\) be a unit principal vector field, i.e. \(AW=\lambda W\), where \(\lambda \) is the corresponding principal curvature. There are two cases:

CASE I\(\alpha ^{2}+c\ne 0\).

In this case, \(\lambda \ne \frac{\alpha }{2}\) and Theorem 2.1 holds. So, \(\phi W\) is also a principal vector of the shape operator with corresponding eigenvalue \(\mu \). Relation (3.1) for \(Y=\xi \) implies \((\phi A-A\phi )\phi AX=0\) for any \(X \in {\mathbb {D}}\). The last relation for \(X=W\) and \(X=\phi W\) yields

$$\begin{aligned} \lambda (\lambda -\mu )=0\;\;\text{ and }\;\;\mu (\lambda -\mu )=0. \end{aligned}$$

Combination of the last two relations results in \(\lambda =\mu \), which implies \(\phi A=A\phi \). Thus, because of Theorem 1.1, M is locally congruent to a real hypersurface of type (A).

CASE II\(\alpha ^{2}+c=0\).

In this case due to the above relation, \(c=-4\) and \(\alpha \ne 0\). First the case, \(\lambda \ne \frac{\alpha }{2}\) is examined. In this case Theorem 2.1 holds and let \(A\phi W=\mu \phi W\). Then relation (2.6) due to \(\alpha ^{2}+c=0\) implies \(\mu =\frac{\alpha }{2}\). Relation (3.1) for \(X=\phi W\) and \(Y=\xi \) implies \(\alpha =0\), which is impossible.

The remaining case is relation \(AW=\frac{\alpha }{2}W\) holds, for all \(W \, \in \, {\mathbb {D}}\). In this case \(\phi A=A\phi \) and M is locally congruent to a horosphere. Relation (3.1) is satisfied for any X\(\in \, {\mathbb {D}}\) and any \(Y \, \in \, TM\) and this completes the proof of Theorem 1.2.

4 Proof of Theorem 1.3

Let M be a real hypersurface in \(M_{n}(c),n\ge 2\) satisfying relation \(F_{\xi }^{(k)}(\phi A-A\phi )Y=(\phi A-A\phi )F_{\xi }^{(k)}Y\) for any Y tangent to M. This yields

$$\begin{aligned}&g(\phi A\xi ,(\phi A-A\phi )Y)\xi +\eta (A\phi Y)\phi A\xi -k\phi (\phi A-A\phi )Y\nonumber \\&\quad =g(\phi A\xi ,Y)\phi A\xi -\eta (Y)(\phi A-A\phi )\phi A\xi -k(\phi A-A\phi )\phi Y \end{aligned}$$
(4.1)

for any \(Y \in TM\).

Let M be a non-Hopf real hypersurface satisfying relation (4.1). Then, relation \(A\xi =\alpha \xi +\beta U\) for a unit \(U \in {\mathbb {D}}\) holds and \(\alpha \), \(\beta \) are functions on M, with \(\beta \ne 0\). So, (4.1) becomes

$$\begin{aligned}&\beta g(\phi U,(\phi A-A\phi )Y)\xi +\beta \eta (A\phi Y)\phi U-k\phi (\phi A-A\phi )Y \nonumber \\&\quad =\beta ^2g(\phi U,Y)\phi U-\beta \eta (Y)(\phi A-A\phi ) \phi U-k(\phi A-A\phi )\phi Y \end{aligned}$$
(4.2)

for any \(Y \in TM\). Taking \(Y=\xi \) in (4.2) implies \(\beta g(\phi U,\phi A\xi )\xi -k\phi ^2A\xi =-\beta (\phi A-A\phi )\phi U\) and its scalar product with \(\xi \) results in \(\beta =0\), which is impossible.

Thus, M is a Hopf hypersurface and \(A\xi =\alpha \xi \). Relation (4.1) becomes, bearing in mind that \(k \ne 0\)

$$\begin{aligned} \phi (\phi A-A\phi )Y=(\phi A-A\phi )\phi Y \end{aligned}$$
(4.3)

for any Y tangent to M.

Take a unit vector field \(W \in {\mathbb {D}}\) such that \(AW=\lambda W\). There are two cases:

CASE I\(\alpha ^{2}+c\ne 0\).

In this case, \(\lambda \ne \frac{\alpha }{2}\) and Theorem 2.1 holds. So, \(\phi W\) is also a principal vector of the shape operator with corresponding eigenvalue \(\mu \). Relation (4.3) for \(Y=W\), implies \(\lambda =\mu \). The latter results in \(\phi A=A\phi \) and because of Theorem 1.1, M is locally congruent to a real hypersurface of type (A).

CASE II\(\alpha ^{2}+c=0\).

First suppose that \(\lambda \ne \frac{\alpha }{2}\). So, Theorem 2.1 holds. Suppose that \(A\phi W=\mu \phi W\), then relation (2.6) because of \(\alpha ^{2}+c=0\) implies \(\mu =\frac{\alpha }{2}\). Relation (4.3) for \(Y=W\) implies \(\lambda =\frac{\alpha }{2}\), which is a contradiction.

So, relation \(AW=\frac{\alpha }{2}W\) holds, for all \(W \, \in \, {\mathbb {D}}\). In this case, \(\phi A=A\phi \) and M is locally congruent to a horosphere. Relation (4.1) is satisfied for any \(Y \, \in \, TM\) and this completes the proof of Theorem 1.3.

5 Proof of Theorem 1.4

Let M be a real hypersurface in \(M_{n}(c)\), \(n\ge 2\), whose symmetric operator \(\phi A-A\phi \) satisfies relation \((\mathcal{L}_{\xi }^{(k)}(\phi A-A\phi ))X=({{\mathcal {L}}}_{\xi }(\phi A-A\phi ))X\) for any vector field X tangent to M. This yields

$$\begin{aligned} F_{\xi }^{(k)}(\phi A-A\phi )X-(\phi A-A\phi )F_{\xi }^{(k)}X=\phi A^2\phi X-A\phi ^2AX \end{aligned}$$
(5.1)

for any vector field X tangent to M.

Suppose that M is a non-Hopf real hypersurface and, as in the previous section we consider \(A\xi =\alpha \xi +\beta U\), with \(\beta \ne 0\). Taking \(X=\xi \) in (5.1) implies \(\beta F_{\xi }^{(k)}\phi U-(\phi A-A\phi )(-\phi A\xi )=-A\phi ^2A\xi =-A\phi (\beta \phi U)=\beta AU\). Since, \(\beta \ne 0\), the last one gives \(g(\phi A\xi ,\phi U)\xi -k\phi ^2U+\phi A\phi U-A\phi ^2U=AU\), that is,

$$\begin{aligned} \beta \xi +kU+\phi A\phi U=0. \end{aligned}$$
(5.2)

The scalar product of (5.2) with \(\xi \) yields \(\beta =0\), which is impossible.

Therefore, M must be a Hopf hypersurface and \(A\xi =\alpha \xi \). Let \(W \, \in \, {\mathbb {D}}\) be a unit principal vector field such that \(AW=\lambda W\). There are two cases:

CASE I\(\alpha ^{2}+c\ne 0\). In this case \(\lambda \ne \frac{\alpha }{2}\) and Theorem 2.1 holds. So, \(\phi W\) is also a principal vector of the shape operator with \(A\phi W=\mu \phi W\). Relation (5.1) for \(X=W\) implies \((\lambda -\mu )F_{\xi }^{(k)}\phi X+k(\phi A-A\phi )\phi X=-\mu ^2X+\lambda ^2X\) and this results in

$$\begin{aligned} 2k(\lambda -\mu )=(\lambda +\mu )(\lambda -\mu ). \end{aligned}$$

Thus we have two possibilities

  1. 1.

    \(\lambda = \mu \).

  2. 2.

    \(\lambda \ne \mu \). In this case \(\lambda +\mu =2k\).

If for any eigenvalue \(\lambda \), we have \(\lambda =\mu \), \(\phi A=A\phi \) and because of Theorem 1.1M is locally congruent to a real hypersurface of type (A).

In \({\mathbb {C}}P^n\), if for any eigenvalue \(\lambda \), \(\lambda \ne \mu \), \(\lambda +\mu =2k\) yields \(\lambda +\frac{\alpha \lambda +2}{2\lambda -\alpha } =2k\). This implies \(\lambda ^{2}-2k\lambda +1+k\alpha =0\). Thus \(\lambda =k \pm \sqrt{k^2-k\alpha -1}\). This means that M has, at most, three distinct constant principal curvatures.

As \(\lambda \ne \mu \), M must be locally congruent to a real hypersurface of type (B). In such a case we may suppose \(\alpha =2{\hbox {cot}}(2r)\), \(\lambda ={\hbox {cot}}(r-\frac{\pi }{4})=k+\sqrt{k^2-k\alpha -1}\) and \(\mu =-{\hbox {tan}}(r-\frac{\pi }{4})=k-\sqrt{k^2-k\alpha -1}\). Then \(-1=(k+\sqrt{k^2-k\alpha -1})(k-\sqrt{k^2-k\alpha -1})=k^2-(k^2-k\alpha -1)=k\alpha +1\). That is, \(\alpha =-\frac{2}{k}\), that means \({\hbox {cot}}(2r)=-\frac{1}{k}\).

Suppose now there exist \(X \in {\mathbb {D}}\) such that \(AX=\lambda X\), \(A\phi X=\lambda \phi X\) and there exists \(Y \in {\mathbb {D}}\) such that \(AY=\omega Y\), \(A\phi Y=\epsilon \phi Y\), \(\epsilon \ne \omega \) which yields \(\epsilon +\omega =2k\). Bearing in mind the cases we have discussed above, M has five distinct constant principal curvatures. Therefore it is locally congruent to a real hypersurface of type either (C) or (D) or (E). From the introduction in any of these cases \(\alpha =2{\hbox {cot}}(2r)\), \(\omega ={\hbox {cot}}(r-\frac{\pi }{4})\), \(\epsilon ={\hbox {cot}}(r-\frac{3\pi }{4})\), \(\lambda =-{\hbox {tan}}(r)\), \(\mu ={\hbox {cot}}(r)\).

As \({\hbox {cot}}(r-\frac{\pi }{4})+{\hbox {cot}}(r-\frac{3\pi }{4})=2k\), we have \(k=\frac{2{\hbox {sin}}(r){\hbox {cos}}(r)}{{\hbox {sin}}^2(r)-{\hbox {cos}}^2(r)}=-{\hbox {tan}}(2r)\).

In \({\mathbb {C}}H^{n}\) as \(\lambda \ne \mu \) and \(\lambda +\mu =2k\), as above \(\lambda \) and \(\mu \) must be constant and M must be locally congruent to a real hypersurface of type (B). In this case \(\alpha =2{\hbox {tanh}}(2r)\), \(\lambda ={\hbox {tanh}}(r)\) and \(\mu ={\hbox {coth}}(r)\). Substitution in \(\lambda +\mu =2k\) yields \({\hbox {tanh}}(2r)=\frac{1}{k}\).

CASE II\(\alpha ^{2}+c=0\).

First suppose that \(\lambda \ne \frac{\alpha }{2}\). So, Theorem 2.1 holds. Suppose that \(A\phi W=\mu \phi W\), then relation (2.6) because of \(\alpha ^{2}+c=0\) implies \(\mu =\frac{\alpha }{2}\). Relation (5.1) for \(X=W\) implies that \(\lambda =2k-\frac{\alpha }{2}\). Therefore, M is a Hopf real hypersurface with three constant principal curvatures and is locally congruent to a real hypersurface of type (B). Substitution of the eigenvalues in relation \(\mu =\frac{\alpha }{2}\) leads to a contradiction.

Therefore, relation \(AW=\frac{\alpha }{2}W\) holds, for all \(W \, \in \, {\mathbb {D}}\). In this case \(\phi A=A\phi \) and M is locally congruent to a horosphere. Relation (5.1) is satisfied for any \(Y \, \in \, TM\) and this completes the proof of Theorem 1.4.

6 Proof of Theorem 1.5

Suppose M satisfies \(({{\mathcal {L}}}_X^{(k)}(\phi A-A\phi ))Y=(\mathcal{L}_X(\phi A-A\phi ))Y\) for any \(X \in {\mathbb {D}}\) and any Y tangent to M. If \(n=2\), the proof is obtained in [8]. Therefore we suppose \(n \ge 3\). Then we have

$$\begin{aligned} F_X(\phi A-A\phi )Y-F_{(\phi A-A\phi )Y}^{(k)}X -(\phi A-A\phi )F_XY+(\phi A-A\phi )F_Y^{(k)}X=0,\nonumber \\ \end{aligned}$$
(6.1)

for any \(X \in {\mathbb {D}}\), Y tangent to M.

Let M be a non-Hopf real hypersurface and write \(A\xi \) as in previous Theorems. Taking \(Y=\xi \) in (6.1) we get \(\beta F_X\phi U-\beta F_{\phi U}X-(\phi A-A\phi )F_X\xi +(\phi A-A\phi )F_{\xi }^{(k)}X=0\). This implies

$$\begin{aligned}&\beta (g(AX,U)+g(A\phi U,\phi X))\xi +\phi A\phi AX-A\phi ^2AX \nonumber \\&\quad +\beta ^2g(\phi U,X)\phi U-k\phi A\phi X+kA\phi ^2X=0 \end{aligned}$$
(6.2)

for any \(X \in {\mathbb {D}}\). The scalar product of (6.2) with \(\xi \) gives \(\beta (g(AU,X)+g(A\phi U,\phi X))+\beta g(AX,U)-k\beta g(U,X)=0\). As \(\beta \ne 0\) we obtain

$$\begin{aligned} 2g(AU,X)+g(A\phi U,\phi X)-kg(U,X)=0 \end{aligned}$$
(6.3)

for any \(X \in {\mathbb {D}}\).

If in (6.3) we take \(X=\phi U\) we get \(2g(AU,\phi U)-g(A\phi U,U)=0\). Therefore \(g(AU,\phi U)=0\) and we can write \(AU=\beta \xi +\gamma U+\omega Z\), \(A\phi U=\mu \phi U+\epsilon W\), for certain functions \(\gamma \), \(\omega \), \(\mu \) and \(\epsilon \), where Z and W are unit vector fields in \({{\mathbb {D}}}_U\). Taking \(X=Z\) in (6.3) we get \(2g(AU,Z)+g(A\phi U,Z)=0\). That is

$$\begin{aligned} 2\omega +\epsilon g(W,\phi Z)=0 \end{aligned}$$
(6.4)

and taking \(X=\phi W\) in (6.3) we have \(2g(AU,\phi W)-g(A\phi U,W)=0\), that becomes

$$\begin{aligned} 2\omega g(Z,\phi W)- \epsilon =0. \end{aligned}$$
(6.5)

If \(\epsilon =0\), (6.4) yields \(\omega =0\). If \(\epsilon \ne 0\), from (6.4) \(g(Z,\phi W)=\frac{2\omega }{\epsilon }\) and from (6.5) \(4\omega ^2-\epsilon ^2=0\). That is,

$$\begin{aligned} \epsilon ^2 =4\omega ^2. \end{aligned}$$
(6.6)

Taking \(Y=U\) in (6.1) we obtain \(F_X(\phi A-A\phi )U-F_{(\phi A-A\phi )U}^{(k)}X-(\phi A-A\phi )F_XU+(\phi A-A\phi )F_UX=0\) for any \(X \in {\mathbb {D}}\). As \(\phi AU-A\phi U=(\gamma -\mu )\phi U+\omega \phi Z-\epsilon W \in {\mathbb {D}}\), \(F_X(\phi A-A\phi )U-F_{(\phi A-A\phi )U}X\) is proportional to \(\xi \). Therefore \(-g(F_XU,(\phi A-A\phi )\phi U)+g(F_UX,(\phi A-A\phi )\phi U)=0\). Now \((\phi A-A\phi )\phi U=\beta \xi +(\gamma -\mu )U+\epsilon \phi W+\omega Z\). Therefore \(-g(\phi AX,U)g(\xi ,\beta \xi +(\gamma -\mu )U+\epsilon \phi W+\omega Z)+g(\phi AU,X)g(\xi ,\beta \xi +(\gamma -\mu )U+\epsilon \phi W+\omega Z)=0\). As \(\beta \ne 0\), this yields

$$\begin{aligned} -g(\phi AX,U)+g(\phi AU,X)=0 \end{aligned}$$
(6.7)

for any \(X \in {\mathbb {D}}\). Taking \(X=\phi Z\) in (6.7) we have \(g(A\phi U,\phi Z)+g(AU,Z)=0\). This implies

$$\begin{aligned} \omega +\epsilon g(W,\phi Z)=0. \end{aligned}$$
(6.8)

From (6.4) and (6.8), we conclude \(\omega =0\) and from (6.6), \(\epsilon =0\). Thus we have proved

$$\begin{aligned} AU=\beta \xi +\gamma U\;\;\text{ and }\;\;A\phi U=\mu \phi U. \end{aligned}$$

Moreover, from (6.3)

$$\begin{aligned} 2\gamma +\mu =k. \end{aligned}$$
(6.9)

Taking \(Y=\phi U\) in (6.1), as \((\phi A-A\phi )\phi U=\beta \xi +(\gamma -\mu )U\) we get \(\beta F_X\xi +(\gamma - \mu )F_XU-\beta F_{\xi }^{(k)}X-(\gamma -\mu )F_UX-(\phi A-A\phi )(g(AX,U)\xi +g(A\phi X,\phi U)\xi )=0\), for any \(X \in {\mathbb {D}}\). Developing it we have

$$\begin{aligned}&-\beta \phi AX-\mu (\gamma -\mu )g(\phi U,X)\xi -\beta ^2g(\phi U,X)\xi +\beta k\phi X \nonumber \\&\quad -(\gamma -\mu )g(\phi AU,X)\xi -\beta g(AU,X)\phi U-\beta g(A\phi U,\phi X)\phi U =0\nonumber \\ \end{aligned}$$
(6.10)

for any \(X \in {\mathbb {D}}\).

If \(X \in {{\mathbb {D}}}_U\), from (6.10) we get \(-\beta \phi AX+\beta k\phi X=0\). As \(\beta \ne 0\), if we apply \(\phi \) we obtain

$$\begin{aligned} AX=kX \end{aligned}$$
(6.11)

for any \(X \in {{\mathbb {D}}}_U\).

The scalar product of (6.10) with U yields \(\beta g(A\phi U,X)+\beta kg(\phi X,U)=0\) for any \(X \in {\mathbb {D}}\). If \(X=\phi U\), bearing in mind that \(\beta \ne 0\) we get

$$\begin{aligned} \mu =k \end{aligned}$$
(6.12)

and from (6.9) it follows

$$\begin{aligned} \gamma =0. \end{aligned}$$
(6.13)

The scalar product of (6.10) with \(\xi \) gives \(-\mu (\gamma -\mu )g(\phi U,X)-\beta ^2g(\phi U,X)-(\gamma -\mu )g(\phi AU,X)=0\) for any \(X \in {\mathbb {D}}\). Taking \(X=\phi U\) we have \(\mu ^2 -\beta ^2=0\). From (6.12)

$$\begin{aligned} \beta ^2=k^2 \end{aligned}$$
(6.14)

and therefore \(\beta \) is constant.

By Codazzi equation, for any \(X \in {{\mathbb {D}}}_U\) we get \((\nabla _XA)\phi X-(\nabla _{\phi X}A)X=-\frac{c}{2}\xi \). This yields

$$\begin{aligned} k\nabla _X\phi X-A\nabla _X\phi X-k\nabla _{\phi X}X+A\nabla _{\phi X}X=-\frac{c}{2}\xi . \end{aligned}$$
(6.15)

The scalar product of (6.15) with \(\xi \) and with U implies respectively

$$\begin{aligned} \beta g([ \phi X,X ] ,U)=-\frac{c}{2}-2\alpha k+2k^2\;\;\text{ and }\;\;g([ \phi X,X ] ,U)=2\beta . \end{aligned}$$
(6.16)

From (6.16), bearing in mind (6.14), we get

$$\begin{aligned} \alpha =-\frac{c}{4k} . \end{aligned}$$
(6.17)

Now \((\nabla _UA)\xi -(\nabla _{\xi }A)U=-\frac{c}{4}\phi U\). If we develop it and bear in mind that \(\phi AU=0\) we have \(\beta \nabla _UU-\beta \phi A\xi +A\nabla _{\xi }U=-\frac{c}{4}\phi U\). Its scalar product with \(\phi U\) yields

$$\begin{aligned} \beta g(\nabla _UU,\phi U)+kg(\nabla _{\xi }U,\phi U)=\beta ^2-\frac{c}{4}. \end{aligned}$$
(6.18)

Developing \((\nabla _{\phi U}A)U-(\nabla _UA)\phi U=\frac{c}{2}\xi \) we obtain

$$\begin{aligned} -\beta kU-A\nabla _{\phi U}U-k\nabla _U\phi U+A\nabla _U\phi U=\frac{c}{2}\xi . \end{aligned}$$
(6.19)

Its scalar product with \(\xi \) yields \(\alpha g(U,\phi A\phi U)+\beta g(\nabla _U\phi U,U)=\frac{c}{2}\). From (6.17), we have

$$\begin{aligned} \beta g(\nabla _UU, \phi U)=-\frac{c}{4}. \end{aligned}$$
(6.20)

The scalar product of (6.19) and U, bearing in mind (6.20), gives

$$\begin{aligned} g(\nabla _UU, \phi U)=2\beta . \end{aligned}$$
(6.21)

From (6.20) and (6.21) \(2\beta ^2=-\frac{c}{4}\), which is impossible if the ambient space is \({\mathbb {C}}P^{n}\). In the case of \({\mathbb {C}}H^{n}\) we have \(c=-4\). Therefore \(2\beta ^2=1\) and from (6.17) \(k\alpha =1\). From (6.18) and (6.21) we also have \(kg(\nabla _{\xi }U,\phi U)=1-\beta ^2\).

The scalar product of the Codazzi equation for \(X=Z \in {{\mathbb {D}}}_U\) and \(Y=\xi \) with \(\xi \) yields \(g(\nabla _{\xi }U,Z)=0\) for any \(Z \in {{\mathbb {D}}}_U\). Then

$$\begin{aligned} \nabla _{\xi }U=g(\nabla _{\xi }U,\phi U)\phi U. \end{aligned}$$
(6.22)

Scalar product of the Codazzi equation for \(X=Z \in {{\mathbb {D}}}_U\) and \(Y=U\) with U implies \(g(\nabla _UU,Z)=0\) for any \(Z \in {{\mathbb {D}}}_U\). Thus

$$\begin{aligned} \nabla _UU=2\beta \phi U. \end{aligned}$$
(6.23)

Taking \(X=\phi U\) and \(Y=\xi \) in Codazzi equation and its scalar product with \(Z \in {{\mathbb {D}}}_U\) gives \(g(\nabla _{\phi U}U,Z)=0\), for any \(Z \in {{\mathbb {D}}}_U\). A similar calculation taking the scalar product with \(\phi U\) yields \(g(\nabla _{\phi U}U,\phi U)=0\). Therefore

$$\begin{aligned} \nabla _{\phi U}U=k\xi . \end{aligned}$$
(6.24)

From Gauss equation, bearing in mind \(c=-4\) we have \(R(U,\phi U)U=4\phi U\). On the other hand \(R(U,\phi U)U=\nabla _U\nabla _{\phi U}\phi U-\nabla _{\phi U}\nabla _UU-\nabla _{\nabla _U\phi U-\nabla _{\phi U}U}U=k\phi AU-2\beta \nabla _{\phi U}\phi U-\nabla _{\nabla _U\phi U}U+k\nabla _{\xi }U\). Its scalar product with \(\phi U\) implies \(4=-g(\nabla _{\nabla _U\phi U}U,\phi U)+kg(\nabla _{\xi }U,\phi U)\).

From (6.19) \(g(\nabla _U\phi U,Z)=0\) for any \(Z \in {{\mathbb {D}}}_U\). Then we get \(\nabla _U\phi U=-2\beta U\). As \(kg(\nabla _{\xi }U,\phi U)=1-\beta ^2\), we finally obtain \(4=3\beta ^2+1\). Thus \(\beta ^2=1=\frac{1}{2}\), which is impossible.

Therefore, M must be a Hopf hypersurface and \(A\xi =\alpha \xi \). Let \(W\, \in \, {\mathbb {D}}\) be unit principal vector field such that \(AW=\lambda W\). There are two cases:

CASE I\(\alpha ^{2}+c\ne 0\).

In this case \(\lambda \ne \frac{\alpha }{2}\) and Theorem 2.1 holds. So, \(\phi W\) is also a principal vector of the shape operator with \(A\phi W=\mu \phi W\). Relation (6.1) for \(Y=\xi \) implies

$$\begin{aligned} \phi A\phi AX+A^2X-k\phi A\phi X-kAX=0, \end{aligned}$$
(6.25)

for any \(X \in {\mathbb {D}}\). Suppose \(X=W\) then

$$\begin{aligned} (\lambda -k)(\lambda -\mu )=0. \end{aligned}$$

If \(\lambda \ne \mu \) then \(\lambda =k\). Relation (6.25) for \(X=\phi W\) implies \(\mu =k\). So \(\lambda =\mu \), which is a contradiction. So, on M relation \(\lambda =\mu \) holds and, M is locally congruent to a real hypersurface of type (A).

CASE II\(\alpha ^{2}+c=0\). First suppose that \(\lambda \ne \frac{\alpha }{2}\). So, Theorem 2.1 holds. Suppose that \(A\phi W=\mu \phi W\), then relation (2.6) because of \(\alpha ^{2}+c=0\) implies \(\mu =\frac{\alpha }{2}\). Relation (6.1) for \(X=W\) and \(Y=\xi \) implies \(\lambda =k\) and for \(X=\phi W\) and \(Y=\xi \) implies \(\mu =k\), which is impossible.

Therefore, relation \(AW=\frac{\alpha }{2}W\) holds, for all \(W \, \in \, {\mathbb {D}}\). In this case \(\phi A=A\phi \) and M is locally congruent to a horosphere. Relation (6.1) is satisfied for any \(X \, \in \, {\mathbb {D}}\) and any \(Y \, \in \, TM\) and this completes the proof of Theorem 1.5.