1 Introduction

Let \(\mathbb {C}P^m\), \(m\ge 2\), be a complex projective space endowed with the metric \(g\) of constant holomorphic sectional curvature 4. Let \(M\) be a connected real hypersurface of \(\mathbb {C}P^m\) without boundary. Let \(\nabla \) be the Levi-Civita connection on \(M\) and \(J\) the complex structure of \(\mathbb {C}P^m\). Take a locally defined unit normal vector field \(N\) on \(M\) and denote by \(\xi =-JN\). This is a tangent vector field to \(M\) called the structure vector field on \(M\). On \(M\), there exists an almost contact metric structure \((\phi ,\xi ,\eta ,g)\) induced by the Kaehlerian structure of \(\mathbb {C}P^m\), where \(\phi \) is the tangent component of \(J\) and \(\eta \) is an one form given by \(\eta (X)=g(X,\xi )\) for any \(X\) tangent to \(M\). The classification of homogeneous real hypersurfaces in \(\mathbb {C}P^m\) was obtained by Takagi, see [5, 1214]. His classification contains 6 types of real hypersurfaces. Among them, we find type \((A_1)\) real hypersurfaces that are geodesic hyperspheres of radius \(r\), \(0<r< \frac{\pi }{2}\) and type \((A_2)\) real hypersurfaces that are tubes of radius \(r\), \(0 < r <\frac{\pi }{2}\), over totally geodesic complex projective spaces \(\mathbb {C}P^n\), \(0 < n < m-1\). We will call both types of real hypersurfaces type \((A)\) real hypersurfaces.

Ruled real hypersurfaces can be described as follows: take a regular curve \(\gamma \) in \(\mathbb {C}P^m\) with tangent vector field \(X\). At each point of \(\gamma \), there is a unique \(\mathbb {C}P^{m-1}\) cutting \(\gamma \) so as to be orthogonal not only to \(X\) but also to \(JX\). The union of these hyperplanes is called a ruled real 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 on \(M\), \(\mathbb {D}\), given at any point by the vectors orthogonal to \(\xi \), is integrable or \(g(A\mathbb {D}, \mathbb {D})=0\). For examples of ruled real hypersurfaces, see [6] or [8].

The Tanaka-Webster connection, [1517], is the canonical affine connection defined on a non-degenerate, pseudo-Hermitian CR manifold. As a generalization of this connection, Tanno, [16], 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, Cho defined the g-Tanaka-Webster connection \(\hat{\nabla }^{(k)}\) for a real hypersurface \(M\) in \(\mathbb {C}P^m\) given, see [3, 4], 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 \(X,Y\) tangent to \(M\) where \(k\) is a nonzero real number. Then, \(\hat{\nabla }^{(k)}\eta =0\), \(\hat{\nabla }^{(k)}\xi =0\), \(\hat{\nabla }^{(k)}g=0\), \(\hat{\nabla }^{(k)}\phi =0\). In particular, if the shape operator of a real hypersurface satisfies \(\phi A+A \phi =2k\phi \), the g-Tanaka-Webster connection coincides with the Tanaka-Webster connection.

Here, we can consider the tensor field of type (1,2) given by the difference in both connections \(F^{(k)}(X,Y)=g(\phi AX,Y)\xi -\eta (Y) \phi AX-k\eta (X)\phi Y\), for any \(X,Y\) tangent to \(M\), see [7] Proposition 7.10, pages 234–235. We will call this tensor the \(k\)th Cho tensor on \(M\). Associated to it, for any \(X\) tangent to \(M\) and any non-null real number \(k\), we can consider the tensor field of type (1,1) \(F_X^{(k)}\), given by \(F_X^{(k)}Y=F^{(k)}(X,Y)\) for any \(Y \in TM\). This operator will be called the \(k\)th Cho operator corresponding to \(X\). 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 \(X,Y\) tangent to \(M\).

The Jacobi operator \(R_X\) with respect to a unit vector field \(X\) is defined by \(R_X=R(.,X)X\), where \(R\) is the curvature tensor field on \(M\). Then, we see that \(R_X\) is a self-adjoint endomorphism of the tangent space. It is related to Jacobi vector fields, which are solutions of the second-order differential equation (the Jacobi equation) \(\nabla _{\dot{\gamma }}(\nabla _{\dot{\gamma }}Y)+R(Y,\dot{\gamma })\dot{\gamma } =0\) along a geodesic \(\gamma \) in \(M\). The Jacobi operator with respect to the structure vector field \(\xi \), \(R_{\xi }\), is called the structure Jacobi operator on \(M\).

The purpose of the present paper was to study real hypersurfaces \(M\) in \(\mathbb {C}P^m\) such that the covariant and g-Tanaka-Webster derivatives of the structure Jacobi operator coincide. \(\nabla R_{\xi }=\hat{\nabla }^{(k)}R_{\xi }\) is equivalent to the fact that, for any \(X\) tangent to \(M\), \(R_{\xi }F_X^{(k)}=F_X^{(k)}R_{\xi }\). The meaning of this condition is that every eigenspace of \(R_{\xi }\) is preserved by the \(k\)th Cho operator \(F_X^{(k)}\) for any \(X\) tangent to \(M\).

On the other hand, \(TM=Span\{ \xi \} \oplus \mathbb {D}\). Thus, we will obtain the following

Theorem 1

Let \(M\) be a real hypersurface in \(\mathbb {C}P^m\), \(m\ge 3\). Let \(k\) be a non-null constant. Then, \(F_X^{(k)}R_{\xi }=R_{\xi }F_X^{(k)}\) for any \(X \in \mathbb {D}\) if and only if \(M\) is locally congruent to a ruled real hypersurface.

Theorem 2

Let \(M\) be a real hypersurface in \(\mathbb {C}P^m\), \(m \ge 3\). Let \(k\) be a non-null constant. Then, \(F_{\xi }^{(k)}R_{\xi }=R_{\xi }F_{\xi }^{(k)}\) if and only if \(M\) is locally congruent to either a tube of radius \(\frac{\pi }{4}\) over a complex submanifold of \(\mathbb {C}P^m\) or to a type \((A)\) real hypersurface with radius \(r \ne \frac{\pi }{4}\).

As a direct consequence of these Theorems, we have

Corollary

There do not exist real hypersurfaces \(M\) in \(\mathbb {C}P^m\), \(m\ge 3\), such that for a non-null constant \(k\), \(F_X^{(k)}R_{\xi }=R_{\xi }F_X^{(k)}\) for any \(X\) tangent to \(M\).

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 \(\mathbb {C}P^m\), \(m\ge 2\), without boundary. Let \(N\) be a locally defined unit normal vector field on \(M\). Let \(\nabla \) be the Levi-Civita connection on \(M\) and \((J,g)\) the Kaehlerian structure of \(\mathbb {C}P^m\).

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 [1], 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 \(X,Y\) 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

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

and

$$\begin{aligned} \nabla _X\xi = \phi AX \end{aligned}$$
(2.4)

for any \(X,Y\) tangent to \(M\), where \(A\) denotes the shape operator of the immersion. As the ambient space has holomorphic sectional curvature 4, the equations of Gauss and Codazzi are given, respectively, by

$$\begin{aligned} R(X,Y)Z&= \displaystyle g(Y,Z)X - g(X,Z)Y + g(\phi Y,Z)\phi X - g(\phi X,Z) \phi Y \nonumber \\&- 2g (\phi X,Y) \phi Z + g(AY,Z)AX - g(AX,Z)AY, \end{aligned}$$
(2.5)

and

$$\begin{aligned} (\nabla _XA)Y- (\nabla _YA) X = \eta (X)\phi Y-\eta (Y)\phi X-2g(\phi X,Y)\xi \end{aligned}$$
(2.6)

for any tangent vectors \(X,Y,Z\) to \(M\), where \(R\) is the curvature tensor of \(M\). We will call the maximal holomorphic distribution \(\mathbb {D}\) on \(M\) to the following one: at any \(p \in M\), \(\mathbb {D}(p)=\{ X\in T_pM \vert g(X,\xi )=0\}\). We will say that \(M\) is Hopf if \(\xi \) is principal, that is, \(A\xi =\alpha \xi \) for a certain function \(\alpha \) on \(M\).

From the above formulas, we have that the structure Jacobi operator on \(M\) is given by

$$\begin{aligned} R_{\xi }(X)=X-\eta (X)\xi +g(A\xi ,\xi )AX-g(AX,\xi )A\xi \end{aligned}$$
(2.7)

for any \(X\) tangent to \(M\)

In the sequel, we need the following results:

Theorem 2.1

[10] Let \(M\) be a real hypersurface of \(\mathbb {C}P^m\), \(m\ge 2\). Then, the following are equivalent:

  1. 1.

    \(M\) is locally congruent to either a geodesic hypersphere or a tube of radius \(r\), \(0 < r < \frac{\pi }{2}\) over a totally geodesic \(\mathbb {C}P^n\), \(0 < n < m-1\).

  2. 2.

    \(\phi A=A\phi \).

Theorem 2.2

[9] If \(\xi \) is a principal curvature vector with corresponding principal curvature \(\alpha \) and \(X \in \mathbb {D}\) is principal with principal curvature \(\lambda \), then \(\phi X\) is principal with principal curvature \(\frac{\alpha \lambda +2}{2\lambda -\alpha }\).

3 Proof of Theorem 1

If we suppose that \(F_X^{(k)}R_{\xi }=R_{\xi }F_X^{(k)}\) for any \(X \in \mathbb {D}\), we get

$$\begin{aligned}&g(Y,\phi AX)\xi +\eta (A\xi )g(\phi AX,AY)\xi -\eta (AY)g(\phi AX,A\xi )\xi \nonumber \\&\quad +\eta (Y)\phi AX+\eta (Y)\eta (A\xi )A\phi AX-\eta (Y)\eta (A\phi AX)A\xi =0 \end{aligned}$$
(3.1)

for any \(X \in \mathbb {D}\), \(Y \in TM\). Let us suppose that \(M\) is non-Hopf. 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 \(\beta \ne 0\). We also call \({\mathbb {D}}_U\) to the orthogonal complementary distribution in \(\mathbb {D}\) to the one spanned by \({ U, \phi U }\).

If we take \(X=Y=\phi U\) in (3.1), we get

$$\begin{aligned} g(AU,\phi U)=0. \end{aligned}$$
(3.2)

And taking \(Y=\xi \) in (3.1), we obtain

$$\begin{aligned} \phi AX+\alpha A\phi AX-\alpha \beta g(\phi AX,U)\xi -\beta ^2g(\phi AX,U)U=0 \end{aligned}$$
(3.3)

for any \(X \in \mathbb {D}\). In particular, from (3.2) and (3.3), we have

$$\begin{aligned} \phi AU+\alpha A\phi AU=0. \end{aligned}$$
(3.4)

The scalar product of (3.3) and \(U\) yields

$$\begin{aligned} (\beta ^2-1)g(A\phi U,X)-\alpha g(A\phi AU,X)=0 \end{aligned}$$
(3.5)

for any \(X \in \mathbb {D}\). Thus, \((\beta ^2-1)A\phi U-\alpha A\phi AU\) has not a component in \(\mathbb {D}\), and taking its scalar product with \(\xi \), it follows

$$\begin{aligned} (\beta ^2-1) A \phi U-\alpha A\phi AU=0. \end{aligned}$$
(3.6)

From (3.4) and (3.6), we get

$$\begin{aligned} \phi AU=(1-\beta ^2)A\phi U. \end{aligned}$$
(3.7)

Therefore, we can write \(A\phi U=\delta \phi U+\omega Z_1\), where \(Z_1 \in {\mathbb {D}}_U\) is a unit vector field. The scalar product of (3.3) and \(Y \in {\mathbb {D}}_U\) yields \(A\phi Y+\alpha A\phi Y\) has not component in \(\mathbb {D}\). Then,

$$\begin{aligned} A\phi Y+\alpha A\phi AY=-\alpha \beta g(A\phi U,Y)\xi \end{aligned}$$
(3.8)

for any \(Y \in {\mathbb {D}}_U\). Taking \(Y=\phi Z_1\), we obtain \(-AZ_1+\alpha A\phi A\phi Z_1=0\). Its scalar product with \(\xi \) gives

$$\begin{aligned} \alpha \beta \omega (\beta ^2-1)=0. \end{aligned}$$
(3.9)

As \(\beta \ne 0\), the following cases appear

Case 1. \(\alpha =0\).

Case 2. \(\beta ^2=1\). In this case, from (3.7), \(AU=\beta \xi \).

Case 3. \(\omega =0\), thus \({\mathbb {D}}_U\) is \(A\)-invariant.

Case 1. \(\alpha =0\). From (3.4) \(\phi AU=0\), that is, \(AU=\beta \xi \) and \(A\xi =\beta U\) and from (3.6) \((\beta ^2-1)A\phi U=0\). So we have the following subcases

Subcase 1.1. Let us suppose that \(\beta ^2 \ne 1\). Then, \(A\phi U=0\). Moreover, from (3.8) for any \(Y \in {\mathbb {D}}_U\) \(A\phi Y=0\). That means that \(M\) is a minimal ruled hypersurface.

Subcase 1.2. \(\alpha =0\), \(\beta ^2 =1\). We can suppose \(\beta =1\), maybe after changing \(\xi \) by \(-\xi \). As above, \(A\phi Y=0\) for any \(Y \in {\mathbb {D}}_U\), \(AU=\xi \), \(A\xi =U\). Then, \(AZ_1=0\) and \(\omega =g(A\phi U,Z_1)=0\). Thus, \(A\phi U=\delta \phi U\).

By the Codazzi equation \(g((\nabla _{\xi }A)U-(\nabla _UA)\xi ,\phi U)=1\) yields

$$\begin{aligned} \delta g(\nabla _{\xi }U,\phi U)+g(\nabla _UU,\phi U)=0. \end{aligned}$$
(3.10)

From \(g((\nabla _{\xi }A)\phi U-(\nabla _{\phi U}A)\xi ,\xi )=0\), we obtain

$$\begin{aligned} g(\nabla _{\xi } \phi U,U)=-3\delta . \end{aligned}$$
(3.11)

From (3.10) and (3.11), we have

$$\begin{aligned} g(\nabla _UU,\phi U)=-3\delta ^2. \end{aligned}$$
(3.12)

As \(g((\nabla _UA)\phi U-(\nabla _{\phi U}A)U,\xi )=-2\), it follows

$$\begin{aligned} g(\nabla _UU,\phi U)=-2 \end{aligned}$$
(3.13)

and from \(g((\nabla _UA)\phi U-(\nabla _{\phi U}A)U,U)=0\), we get

$$\begin{aligned} \delta g(\nabla _U\phi U,U)+2 \delta =0. \end{aligned}$$
(3.14)

From (3.13) and (3.14), we have \(\delta =0\). Therefore, \(M\) is still a minimal ruled real hypersurface.

Case 2. \(\beta ^2=1\). As above, we suppose \(\beta =1\). As the case \(\alpha =0\) has been studied, we suppose \(\alpha \ne 0\). Then, from (3.6), \(A\phi AU=0\), and from (3.4), \(\phi AU=0\). Therefore, \(A\xi =\alpha \xi +U\), \(AU=\xi \). Moreover, we know that \(-AZ_1+\alpha A\phi AZ_1=0\). Taking its scalar product with \(\phi U\), we get \(\omega + \alpha \omega g(A\phi Z_1,\phi Z_1)=0\). Supposing \(\omega \ne 0\), we have \(g(A\phi Z_1,\phi Z_1)=-\frac{1}{\alpha }\).

Taking \(X=Y\in {\mathbb {D}}_U\) in (3.1), we obtain \(\phi AY+\alpha A\phi AY+\omega g(Y,Z_1)A\xi =0\). Its scalar product with \(\phi U\) gives \(\alpha g(A\phi AY,\phi U)=0=-\alpha \omega g(Y,A\phi Z_1)\). As \(\alpha \omega \ne 0\), \(g(Y,A\phi Z_1)=0\) for any \(Y \in {\mathbb {D}}_U\). This yields \(A\phi Z_1=0\) and \(0=-\frac{1}{\alpha }\). This is a contradiction, and we have \(\omega =0\), \(A\xi =\alpha \xi + U\), \(AU=\xi \) and \(A\phi U=\delta \phi U\).

This yields \({\mathbb {D}}_U\) is \(A\)-invariant and \(\phi \)-invariant, and we arrive to Case 3. As also \(\phi AU =(1-\beta ^2)A\phi U\), we have two possible subcases:

Subcase 3.1. \(\beta ^2 =1\). In this case, \(AU=\beta \xi \).

Subcase 3.2. \(\beta ^2 \ne 1\) and \(AU=\beta \xi + \sigma U\), where \(\sigma =(1-\beta ^2)\delta \).

If we take \(Y=\phi X \in {\mathbb {D}}_U\) in (3.1) for \(X \in {\mathbb {D}}_U\) such that \(AX=\lambda X\), we have \(\lambda +\alpha \lambda g(\phi X,A\phi X)=0\). This yields that either any eigenvalue in \({\mathbb {D}}_U\) is \(0\) or that if there exists a non-null eigenvalue \(\lambda \) in \({\mathbb {D}}_U\), \(\alpha \ne 0\) and \(\lambda =-\frac{1}{\alpha }\). In this case, the eigenspace corresponding to this eigenvalue is \(\phi \)-invariant.

Let us suppose that for any \(Y \in {\mathbb {D}}_U\) \(AY=0\). As \(g((\nabla _YA)\phi Y-(\nabla _{\phi Y}A)Y,\xi )=-2\), we obtain

$$\begin{aligned} g([ \phi Y,Y ],U) = -\frac{2}{\beta }. \end{aligned}$$
(3.15)

And as \(g((\nabla _YA)\phi Y-(\nabla _{\phi Y}A)Y,U)=0\), it follows

$$\begin{aligned} \sigma g([ \phi Y,Y ],U)=0. \end{aligned}$$
(3.16)

From (3.15) and (3.16), we get \(\sigma =0\). If also \(\beta ^2 \ne 1\), \(\delta =0\) and our real hypersurface should be ruled.

Let us then suppose that \(\beta ^2=1\). As above, we will take \(\beta =1\) and \(\sigma =0\). If we develop \(g((\nabla _YA)\phi U-(\nabla _{\phi U}A)Y,\xi )=0\), we get

$$\begin{aligned} g(\nabla _Y\phi U,U)=g(\nabla _{\phi U}Y,U) \end{aligned}$$
(3.17)

and from \(g((\nabla _YA)\phi U-(\nabla _{\phi U}A)Y,U)=0\), it follows

$$\begin{aligned} \delta g(\nabla _Y\phi U,U)=0. \end{aligned}$$
(3.18)

From (3.17) and (3.18), suppose \(g(\nabla _Y\phi U,U)=g(\nabla _{\phi U}Y,U)=0\). As \(g((\nabla _{\phi U}A)U-(\nabla _UA)\phi U,\xi )=2\), we obtain

$$\begin{aligned} g(\nabla _U\phi U,U)=2+\alpha \delta \end{aligned}$$
(3.19)

and from \(g((\nabla _{\phi U}A)U-(\nabla _UA)\phi U,U)=0\), we have

$$\begin{aligned} 2\delta +\delta g(\nabla _U\phi U,U)=0. \end{aligned}$$
(3.20)

If \(\delta \ne 0\), from (3.20) \(g(\nabla _U\phi U,U)=-2\) and from (3.19)

$$\begin{aligned} \alpha \delta =-4. \end{aligned}$$
(3.21)

Now, \(g((\nabla _{\phi U}A)\xi -(\nabla _{\xi }A)\phi U,U)=1\) gives \(\delta g(\nabla _{\xi }\phi U,U)=2-\alpha \delta \). From (3.21)

$$\begin{aligned} \delta g(\nabla _{\xi }\phi U,U)=6. \end{aligned}$$
(3.22)

But from \(g((\nabla _{\xi }A)U-(\nabla _UA)\xi ,\phi U)=1\), we obtain

$$\begin{aligned} -\delta g(\nabla _{\xi }U,\phi U)=0. \end{aligned}$$
(3.23)

From (3.22) and (3.23), we arrive to a contradiction. Thus, \(\delta =0\) and \(M\) is also a ruled real hypersurface.

Therefore, we have only to study the following case: \(A\xi =\alpha \xi +\beta U\), \(AU=\beta \xi +\sigma U\), \(A\phi U=\delta \phi U\), \({\mathbb {D}}_U\) is \(A\)-invariant, and there exists \(Z \in {\mathbb {D}}_U\) such that \(AZ=-\frac{1}{\alpha } Z\), \(A\phi Z=-\frac{1}{\alpha } \phi Z\). As \((1-\beta ^2)A\phi U=\phi AU\), two subcases appear

Subcase 1. \(\beta ^2 =1\), and then \(\sigma =0\).

Subcase 2. \(\beta ^2 \ne 1\), \(\sigma =(1-\beta ^2)\delta \).

From \(g((\nabla _ZA)\phi Z-(\nabla _{\phi Z}A)Z,\xi )=-2\), we obtain

$$\begin{aligned} \beta g([ \phi Z,Z ],U)=\frac{2}{\alpha ^2} \end{aligned}$$
(3.24)

and from \(g((\nabla _ZA)\phi Z-(\nabla _{\phi Z}A)Z,U)=0\), we get

$$\begin{aligned} \left( \frac{1}{\alpha }+\sigma \right) g([ \phi Z,Z ],U)=\frac{2\beta }{\alpha }. \end{aligned}$$
(3.25)

From (3.24) and (3.25), we obtain

$$\begin{aligned} 1+\alpha \sigma =\alpha ^2\beta ^2. \end{aligned}$$
(3.26)

In Subcase 1, as \(\beta ^2=1\) and \(\sigma =0\), we should obtain \(\alpha ^2=1\). Changing, if necessary, \(\xi \) by \(-\xi \), we can take \(\alpha =1\). This case cannot occur by Proposition 3.2, page 1607 in [11]. Therefore, we have \(\beta ^2 \ne 1\) and from (3.26) \(1+\alpha \delta (1-\beta ^2)=\alpha ^2\beta ^2\). Thus,

$$\begin{aligned} \delta =\frac{\alpha ^2\beta ^2-1}{\alpha (1-\beta ^2)}. \end{aligned}$$
(3.27)

Now, \(g((\nabla _ZA)\phi Z-(\nabla _{\phi Z}A)Z,\phi U)=0\) yields

$$\begin{aligned} \left( \frac{1}{\alpha } +\delta \right) g([ Z,\phi Z ],\phi U)=0. \end{aligned}$$
(3.28)

Let us suppose that \(\delta =-\frac{1}{\alpha }\). Then, \(\sigma =\frac{\beta ^2-1}{\alpha }\) and from (3.26) \(\alpha ^2=1\). As above, we suppose \(\alpha =1\). Thus, \(A\xi =\xi +\beta U\), \(AU=\beta \xi +(\beta ^2-1)U\), \(A\phi U=-\phi U\), and there exists a unit \(Z \in {\mathbb {D}}_U\) such that \(AZ=-Z\), \(A\phi Z=-\phi Z\).

Suppose that there exists a unit \(W \in {\mathbb {D}}_U\) such that \(AW=A\phi W=0\). From \(g((\nabla _WA)\xi -(\nabla _{\xi }A)W,\xi )=0\), we obtain \(g(\nabla _{\xi }W,U)=0\), and from \(g((\nabla _WA)\xi -(\nabla _{\xi }A)W,U)=0\), we get \(W(\beta ) + (\beta ^2-1)g(\nabla _{\xi }W,U)=0\). Thus, \(W(\beta )=0\). This fact and the proof of Proposition 3.3, page 1608 in [11], yield \(grad(\beta )=-(2\beta ^2+1)\phi U\). The same proof yields this case cannot occur. Therefore, \(\delta \ne -\frac{1}{\alpha }\) and \(g([ Z,\phi Z ],\phi U)=0\).

Then, from \(g((\nabla _ZA)\phi Z-(\nabla _{\phi Z}A)Z,Z)=g((\nabla _ZA)\phi Z-(\nabla _{\phi Z}A)Z,\phi Z)=0\), we get

$$\begin{aligned} Z(\alpha )=(\phi Z)(\alpha )=0. \end{aligned}$$
(3.29)

From \(g((\nabla _ZA)\xi -(\nabla _{\xi }A)Z,\xi )=0\), it follows

$$\begin{aligned} Z(\alpha )+\beta g(\nabla _{\xi }Z,U)=0. \end{aligned}$$
(3.30)

From (3.29) and (3.30), we obtain

$$\begin{aligned} g(\nabla _{\xi }Z,U)=0. \end{aligned}$$
(3.31)

As \(g((\nabla _ZA)\xi -(\nabla _{\xi }A)Z,U)=0\), we have, bearing in mind (3.31),

$$\begin{aligned} Z(\beta )=0. \end{aligned}$$
(3.32)

From \(g((\nabla _{\xi }A)U-(\nabla _UA)\xi ,\xi )=0\), we get

$$\begin{aligned} \xi (\beta )=U(\alpha ) \end{aligned}$$
(3.33)

and as \(g((\nabla _{\xi }A)U- (\nabla _UA)\xi ,U)=0\), it follows

$$\begin{aligned} \xi (\sigma )=U(\beta ). \end{aligned}$$
(3.34)

Now, \(g((\nabla _ZA)U-(\nabla _UA)Z,\xi )=0\) yields

$$\begin{aligned} Z(\beta )+\sigma g(\nabla _UZ,U)=0 \end{aligned}$$
(3.35)

and from (3.26) and \(g((\nabla _ZA)U-(\nabla _UA)Z,U)=0\), we obtain

$$\begin{aligned} Z(\sigma )+\alpha \beta ^2g(\nabla _UZ,U)=0. \end{aligned}$$
(3.36)

From (3.32) and (3.36), we have \(\sigma g(\nabla _UZ,U)=0\). This and (3.36) yield \(Z(\sigma )+\frac{1}{\alpha } g(\nabla _UZ,U)=0\). As \(Z(\alpha )=Z(\beta )=0\), from (3.26) \(Z(\sigma )=0\). Therefore,

$$\begin{aligned} g(\nabla _UZ,U)=0. \end{aligned}$$
(3.37)

As \(g((\nabla _ZA)U-(\nabla _UA)Z,\phi U)=0\), this gives

$$\begin{aligned} (\sigma -\delta )g(\nabla _ZU,\phi U)+\left( \delta +\frac{1}{\alpha }\right) g(\nabla _UZ,\phi U)=0 \end{aligned}$$
(3.38)

and \(g((\nabla _ZA)U-(\nabla _UA)Z,Z)=0\) yields

$$\begin{aligned} U\left( \frac{1}{\alpha }\right) =\left( \sigma +\frac{1}{\alpha }\right) g(\nabla _ZZ,U). \end{aligned}$$
(3.39)

From \(g((\nabla _ZA)\xi -(\nabla _{\xi }A)Z,Z)=0\), we obtain

$$\begin{aligned} \xi \left( \frac{1}{\alpha }\right) =\beta g(\nabla _ZZ,U) \end{aligned}$$
(3.40)

and from \(g((\nabla _ZA)\phi U-(\nabla _ {\phi U}A)Z,U)=0\), we get

$$\begin{aligned} (\delta -\sigma )g(\nabla _Z\phi U,U)+\left( \sigma +\frac{1}{\alpha }\right) g(\nabla _{\phi U}Z,U)=0. \end{aligned}$$
(3.41)

We also have from \(g((\nabla _ZA)\phi U)-(\nabla _{\phi U}A)Z,\xi )=0\)

$$\begin{aligned} g([ \phi U,Z],U)=0. \end{aligned}$$
(3.42)

Thus, from (3.41) and (3.42), we have a homogeneous system of linear equations where \(g(\nabla _Z\phi U,U)\) and \(g(\nabla _{\phi U}Z,U)\) are unknown. The determinant of its matrix of coefficients is \(\delta + \frac{1}{\alpha }\). As \(\delta \ne -\frac{1}{\alpha }\), we obtain

$$\begin{aligned} g(\nabla _Z\phi U,U)= g(\nabla _{\phi U}Z,U)=0. \end{aligned}$$
(3.43)

As \(g((\nabla _ZA)\phi U-(\nabla _{\phi U}A)Z,\phi Z)=0\), we have \((\delta +\frac{1}{\alpha })g(\nabla _Z\phi U,\phi Z)=0\). As \(\delta \ne -\frac{1}{\alpha }\), \(g(\nabla _Z\phi U,\phi Z)=0\). By (2.3), this gives \(g(\nabla _ZU,Z)=0\). From (3.39) and (3.40), it follows

$$\begin{aligned} \xi (\alpha )=U(\alpha )=0. \end{aligned}$$
(3.44)

From \(g((\nabla _ZA)\phi U-(\nabla _{\phi U}A)Z,Z)=0\), we have \((\delta +\frac{1}{\alpha })g(\nabla _Z\phi U,Z)+(\phi U)(\frac{1}{\alpha })=0\). Now, from (2.3)

$$\begin{aligned} (\phi U)\left( \frac{1}{\alpha }\right) =\left( \frac{1}{\alpha } +\delta \right) g(\nabla _ZU,\phi Z). \end{aligned}$$
(3.45)

Developing \(g((\nabla _ZA)U-(\nabla _UA)Z,\phi Z)=0\) and bearing in mind (3.26), we get

$$\begin{aligned} \alpha ^2\beta ^2g(\nabla _ZU,\phi Z)=\beta . \end{aligned}$$
(3.46)

Now, from (3.45) and (3.46), we obtain

$$\begin{aligned} (\phi U) (\alpha )=-\frac{1+\alpha \delta }{\alpha \beta } =\frac{\beta (1-\alpha ^2)}{\alpha (1-\beta ^2)} . \end{aligned}$$
(3.47)

From (3.33) and (3.44), we have

$$\begin{aligned} \xi (\beta )=U(\beta )=0. \end{aligned}$$
(3.48)

The equality \(g((\nabla _{\xi }A)U-(\nabla _UA)\xi ,\phi U)=1\) yields

$$\begin{aligned} \beta g(\nabla _UU,\phi U)=\beta ^2+\sigma ^2-\alpha \sigma -1. \end{aligned}$$
(3.49)

From \(g((\nabla _UA)\phi U-(\nabla _{\phi U}A)U,\xi )=-2\), we arrive to \(-2\delta \sigma +\alpha \sigma +\alpha \delta -\beta g(\nabla _U\phi U,U)-(\phi U)(\beta )=2\). This and (3.49) yield

$$\begin{aligned} (\phi U)(\beta )=-2\delta \sigma +\alpha \delta +\beta ^2+\sigma ^2+1. \end{aligned}$$
(3.50)

Bearing in mind all these facts, we arrive to

$$\begin{aligned} grad(\alpha )&= \rho \phi U \nonumber \\ grad(\beta )&= \theta \phi U \end{aligned}$$
(3.51)

where \(\rho =-(\frac{1+\alpha \delta }{\alpha \beta })\) and \(\theta =-2\delta \sigma +\alpha \delta +\beta ^2+\sigma ^2+1\). As \(g(\nabla _Xgrad(\alpha ),Y)=g(\nabla _Ygrad(\alpha ),X)\) for any \(X,Y\) tangent to \(M\), we have, taking \(X=\xi \), \(\xi (\rho )g(\phi U,Y)+\rho g(\nabla _{\xi }\phi U,Y)=-\rho g(U,AY)\). If \(Y=\phi U\), this yields \(\xi (\rho )=0\). Thus, \(\rho g(\nabla _{\xi }\phi U,Y)=-\rho g(U,AY)\), for any \(Y\) tangent to \(M\). As \(\rho \ne 0\), taking \(Y=U\), we get

$$\begin{aligned} g(\nabla _{\xi }\phi U,U)=-\sigma . \end{aligned}$$
(3.52)

From \(g((\nabla _{\xi }A)\phi U-(\nabla _{\phi U}A)\xi ,\xi )=0\) and bearing in mind (3.52), we have

$$\begin{aligned} (\phi U)(\alpha )=-3\beta \delta +\alpha \beta -\beta \sigma . \end{aligned}$$
(3.53)

From (3.47) and (3.53) \(2+\alpha \delta +\alpha \sigma -3\alpha \beta ^2\delta +\alpha \sigma \beta ^2=0\), or equivalently

$$\begin{aligned} \alpha ^2(2\beta ^2-3\beta ^4-\beta ^6)+\beta ^2(2+\beta ^2)-1=0. \end{aligned}$$
(3.54)

If \(2-3\beta ^2-\beta ^4=0\), we should have \(\beta ^4+2\beta ^2-1=0\). Both equalities yield \(\beta ^2=1\), that is impossible. From (3.54), we have

$$\begin{aligned} \alpha ^2=\frac{1-\beta ^2(2+\beta ^2)}{2\beta ^2-3\beta ^4-\beta ^6}. \end{aligned}$$
(3.55)

If we take the derivative of (3.54) in the direction of \(\phi U\) and bear in mind (3.47), (3.48), and (3.54), we find that \(\beta \) is a root of a polynomial with constant coefficients. Therefore, \(\beta \) is constant. From (3.55), \(\alpha \) is also constant, which is impossible.

Thus, we have proved that if \(M\) is not Hopf, it is locally congruent to a ruled real hypersurface. It is easy to see that these real hypersurfaces satisfy (3.1).

Let us now suppose that \(M\) is a Hopf real hypersurface with \(A\xi = \alpha \xi \) and that \(M\) satisfies (3.1). Then, we have for any \(X \in \mathbb {D}\) that \(\phi AX+\alpha A\phi AX=0\). If \(\alpha =0\), we get \(\phi AX=0\). Thus, \(AX=0\) for any \(X \in \mathbb {D}\) and \(M\) should be totally geodesic, which is impossible.

Suppose now that \(\alpha \ne 0\) and take a unit \(X \in \mathbb {D}\) such that \(AX=\lambda X\). From (3.1), we get either \(\lambda =0\) or \(A\phi X=-\frac{1}{\alpha } \phi X\). Applying the same reasoning to \(\phi X\), we obtain that the eigenspaces in \(\mathbb {D}\) are \(\phi \)-invariant and correspond to the eigenvalues \(0\) and \(-\frac{1}{\alpha }\). This is impossible by Theorem 2.2, and we finish the proof.

4 Proof of Theorem 2

If we suppose that \(R_{\xi } F_{\xi }^{(k)}=F_{\xi }^{(k)}R_{\xi }\), we get

$$\begin{aligned}&g(Y,\phi A\xi )\xi +g(A\xi ,\xi )g(\phi A\xi ,AY)\xi +\eta (Y)\phi A\xi +\eta (Y)\eta (A\xi )A\phi A\xi \nonumber \\&\quad -\,\eta (Y)\eta (A\phi A\xi )A\xi -k\phi R_{\xi }(Y)+kR_{\xi }(\phi Y)=0 \end{aligned}$$
(4.1)

for any \(Y\in TM\). Let us suppose that \(M\) is non-Hopf. Thus, we write \(A\xi =\alpha \xi +\beta U\) for a unit \(U \in \mathbb {D}\) and functions \(\alpha \) and \(\beta \) on \(M\), \(\beta \) being nonvanishing. From (4.1), we have

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

for any \(Y\in TM\). Taking \(Y=\xi \) in (4.2), we obtain

$$\begin{aligned} \beta \phi U+\alpha \beta A\phi U=0. \end{aligned}$$
(4.3)

As \(\beta \ne 0\), this yields

$$\begin{aligned} \alpha&\ne 0,\nonumber \\ A\phi U&= -\frac{1}{\alpha }\phi U. \end{aligned}$$
(4.4)

If now we take \(Y=U\) in (4.2), as \(k \ne 0\), we get \(\phi R_{\xi }(U)=R_{\xi }(\phi U)\). This yields \(\alpha \phi AU=(\beta ^2-1)\phi U\), that is, \(\phi AU=\frac{\beta ^2-1}{\alpha } \phi U\). By applying \(\phi \) to such an equality, it follows

$$\begin{aligned} AU=\beta \xi +\frac{\beta ^2-1}{\alpha } U. \end{aligned}$$
(4.5)

From (4.4) and (4.5), we obtain that \({\mathbb {D}}_U\) is \(\phi \)-invariant and \(A\)-invariant. Take a unit \(Y \in {\mathbb {D}}_U\) such that \(AY=\lambda Y\). Introducing this \(Y\) in (4.2), we get \(\phi R_{\xi }(Y)=R_{\xi }(\phi Y)\). This yields \(\alpha \lambda \phi Y=\alpha A\phi Y\). As \(\alpha \ne 0\), \(A\phi Y=\lambda \phi Y\). Therefore, the eigenspaces in \({\mathbb {D}}_U\) are \(\phi \)-invariant.

The Codazzi equation gives \((\nabla _{\xi }A)\phi Y-(\nabla _{\phi Y}A)Y=-2\xi \). Taking its scalar product with \(\phi Y\), respectively, with \(Y\), we have

$$\begin{aligned} Y(\lambda )=(\phi Y)(\lambda )=0. \end{aligned}$$
(4.6)

Its scalar product with \(\xi \) implies

$$\begin{aligned} \beta g([\phi Y,Y ],U)=2\lambda ^2-2\alpha \lambda -2 \end{aligned}$$
(4.7)

and its scalar product with \(U\) gives

$$\begin{aligned} \left( \lambda -\frac{\beta ^2-1}{\alpha }\right) g([\phi Y,Y ],U)=2\beta \lambda . \end{aligned}$$
(4.8)

From (4.7) and (4.8), we get

$$\begin{aligned} (\alpha \lambda +1) (\lambda ^2-\alpha \lambda -1)=\beta ^2(\lambda ^2-1). \end{aligned}$$
(4.9)

From \(g((\nabla _{\phi U}A)Y-(\nabla _YA)\phi U,\phi Y)=0\), we have \((\lambda +\frac{1}{\alpha })g(\nabla _Y\phi U,\phi Y)=0\). Then, either \(g(\nabla _Y\phi U,\phi Y)=g(\nabla _YU,Y)=0\), where we have applied (2.3) or \(\lambda =-\frac{1}{\alpha }\). In this second case from (4.9), we have \(0=\beta ^2(\lambda ^2-1)\). As \(\beta \ne 0\), this yields \(\lambda ^2=1\) and \(\alpha ^2=1\). Changing, if necessary, \(\xi \) by \(-\xi \), we can suppose \(\alpha =1\) and then \(\lambda =-1\).

The scalar product of \((\nabla _{\xi }A)Y-(\nabla _YA)\xi =\phi Y\) and \(Y\) gives

$$\begin{aligned} \xi (\lambda )-\beta g(\nabla _YU,Y)=0. \end{aligned}$$
(4.10)

As either \(\lambda =-1\) or \(g(\nabla _YU,Y)=0\), we always have

$$\begin{aligned} \xi (\lambda )=0. \end{aligned}$$
(4.11)

Developing \((\nabla _UA)\phi U-(\nabla _{\phi U}A)U=-2\xi \) and taking its scalar product with \(\phi U\), we get

$$\begin{aligned} \alpha U\left( \frac{1}{\alpha }\right) =\beta ^2g(\nabla _{\phi U}\phi U,U) \end{aligned}$$
(4.12)

The same procedure applied to \(g((\nabla _{\xi }A)\phi U-(\nabla _{\phi U}A)\xi ,\phi U)=0\) yields

$$\begin{aligned} \xi \left( \frac{1}{\alpha }\right) =\beta g(\nabla _{\phi U}\phi U,U). \end{aligned}$$
(4.13)

From (4.12) and (4.13), we obtain

$$\begin{aligned} \alpha U(\alpha )=\beta \xi (\alpha ). \end{aligned}$$
(4.14)

From \(g((\nabla _{\xi }A)U-(\nabla _UA)\xi ,\xi )=0\), \(\xi (\beta )=U(\alpha )\) and from (4.14), we have

$$\begin{aligned} \beta \xi (\alpha ) = \alpha \xi (\beta ) . \end{aligned}$$
(4.15)

By derivating (4.9) in the direction of \(\xi \) and bearing in mind (4.11) and (4.15), it follows

$$\begin{aligned} (\lambda (\lambda ^2-\alpha \lambda -1)-\lambda (\alpha \lambda +1))\xi (\alpha )=\frac{2\beta ^2}{\alpha }(\lambda ^2-1)\xi (\alpha ) . \end{aligned}$$
(4.16)

If we suppose \(\xi (\alpha ) \ne 0\) and bear in mind (4.9), from (4.16), we have

$$\begin{aligned} \alpha \lambda ^3-2\alpha \lambda +2\lambda ^2-2=0. \end{aligned}$$
(4.17)

Derivating (4.17) in the direction of \(\xi \), we obtain \((\lambda ^3-2\lambda )\xi (\alpha )=0\). As we suppose \(\xi (\alpha ) \ne 0\), we have \(\lambda (\lambda ^2-2)=0\). If \(\lambda =0\) from (4.9), it follows \(\beta ^2=1\) and \(\beta \) should be constant. From (4.15), \(\xi (\alpha )=0\), and we arrive to a contradiction. Therefore, \(\lambda ^2=2\). From (4.9), we obtain \(1-2\alpha ^2=\beta ^2\). By derivating this equality in the direction of \(\xi \) and bearing in mind (4.15) and that we suppose \(\xi (\alpha ) \ne 0\), we get \(-2\alpha ^2=\beta ^2\) and we have a new contradiction. This proves

$$\begin{aligned} \xi (\alpha )=\xi (\beta )=U(\alpha )=0. \end{aligned}$$
(4.18)

The equality \(g((\nabla _{\xi }A)Y-(\nabla _YA)\xi ,\xi )=0\) yields

$$\begin{aligned} Y(\alpha )=-\beta g(\nabla _{\xi }Y,U). \end{aligned}$$
(4.19)

Analogously, from \(g((\nabla _{\xi }A)Y-g(\nabla _YA)\xi ,U)=0\), we obtain

$$\begin{aligned} Y(\beta )=\left( \lambda -\frac{\beta ^2-1}{\alpha }\right) g (\nabla _{\xi }Y,U). \end{aligned}$$
(4.20)

From (4.19) and (4.20), we get

$$\begin{aligned} \beta Y(\beta )=\left( \frac{\beta ^2-1}{\alpha } -\lambda \right) Y(\alpha ). \end{aligned}$$
(4.21)

As \(Y(\lambda )=0\), from (4.9), it follows

$$\begin{aligned} (\lambda (\lambda ^2-\alpha \lambda -1)-\lambda (\alpha \lambda +1)) Y (\alpha )=(\lambda ^2-1)2\beta Y(\beta ) \end{aligned}$$
(4.22)

and from (4.21) and (4.22), if we suppose \(Y(\alpha ) \ne 0\), we have

$$\begin{aligned} \alpha (3\lambda ^2-2\alpha \lambda ^2-4\lambda )=2(\alpha \lambda +1)(\lambda ^2-\alpha \lambda -1)-2(\lambda ^2-1). \end{aligned}$$
(4.23)

Derivating once again in the direction of \(Y\) and bearing in mind that we suppose \(Y(\alpha ) \ne 0\), we obtain \(\lambda ^3=0\), that is, \(\lambda =0\) and \(\beta ^2=1\). Therefore, \(\beta \) is constant and \(Y(\beta )=0\).

From \(g((\nabla _{\xi }A)\phi U-(\nabla _{\phi U}A)\xi ,\xi )=0\), we have

$$\begin{aligned} (\phi U)(\alpha )=\frac{3\beta }{\alpha } +\alpha \beta +\beta g(\nabla _{\xi }U, \phi U). \end{aligned}$$
(4.24)

And \(\beta ^2=1\) and \(g((\nabla _{\xi }A)\phi U-(\nabla _{\phi U}A)\xi ,U)=-1\) yield

$$\begin{aligned} g(\nabla _{\xi }U, \phi U)=-\alpha . \end{aligned}$$
(4.25)

From (4.24) and (4.25), we conclude

$$\begin{aligned} (\phi U) (\alpha )=\frac{3\beta }{\alpha }. \end{aligned}$$
(4.26)

As \(g((\nabla _{\xi }A)U-(\nabla _UA)\xi ,\phi U)=1\), we get

$$\begin{aligned} \frac{1}{\alpha }g(\nabla _{\xi }U,\phi U)-\beta g(\nabla _UU,\phi U)=0 \end{aligned}$$
(4.27)

and from the Codazzi equation \(g((\nabla _UA)\phi U-(\nabla _{\phi U}A)U,U) = 0\), it follows

$$\begin{aligned} g(\nabla _UU,\phi U)=2\beta . \end{aligned}$$
(4.28)

From (4.27) and (4.28), bearing in mind that \(\beta ^2=1\), we have

$$\begin{aligned} g(\nabla _{\xi }U,\phi U)=2\alpha . \end{aligned}$$
(4.29)

Now, from (4.27) and (4.29), \(\alpha \) should vanish. As this is a contradiction, we arrive to

$$\begin{aligned} Y(\alpha )=Y(\beta )=0 \end{aligned}$$
(4.30)

By linearity, we have \(X(\alpha )=X(\beta )=0\) for any \(X \in {\mathbb {D}}_U\).

The Codazzi equation \(g((\nabla _{\xi }A)\phi U-(\nabla _{\phi U}A)\xi ,U)=-1\) yields

$$\begin{aligned} (\phi U)(\beta )= \frac{\beta ^2-1}{\alpha ^2} +\beta ^2+\frac{\beta ^2}{\alpha }g(\nabla _{\xi }U,\phi U). \end{aligned}$$
(4.31)

As \(g((\nabla _{\xi }A)U-(\nabla _UA)\xi ,\phi U)=1\), we have

$$\begin{aligned} \frac{\beta ^2}{\alpha } g(\nabla _{\xi }U,\phi U)-\beta g(\nabla _UU,\phi U)=\frac{\beta ^2-1}{\alpha ^2} \end{aligned}$$
(4.32)

and \(g((\nabla _UA)\phi U-(\nabla _{\phi U}A)U,U)=0\) shows

$$\begin{aligned} \beta g(\nabla _UU,\phi U)+\beta ^2-3-2(\phi U)(\beta )-\frac{\beta ^2-1}{\alpha \beta } (\phi U)(\alpha )=0. \end{aligned}$$
(4.33)

From (4.31), (4.32), and (4.33), we have

$$\begin{aligned} \beta g(\nabla _UU\phi U)-\frac{\beta ^2-1}{\alpha } g(\nabla _{\xi }U,\phi U)+\frac{\beta ^2-1}{\alpha ^2} -4=0. \end{aligned}$$
(4.34)

Now, from (4.33) and (4.34), it follows \(g(\nabla _{\xi }U,\phi U)=-4\alpha \) and \(g(\nabla _UU,\phi U)=\frac{1-\beta ^2}{\alpha ^2\beta }-4\beta \). Then, from (4.24) and (4.32), we get

$$\begin{aligned} (\phi U) (\alpha )=3\beta \left( \frac{1-\alpha ^2}{\alpha }\right) \end{aligned}$$
(4.35)

and

$$\begin{aligned} (\phi U) (\beta )=-3\beta ^2+\frac{\beta ^2-1}{\alpha ^2}. \end{aligned}$$
(4.36)

From all the facts we have until now, we obtain \(grad(\alpha ) = \omega \phi U\), where \(\omega =3\beta (\frac{1-\alpha ^2}{\alpha })\). As \(g(\nabla _X grad(\alpha ),Y)=g(\nabla _Ygrad(\alpha ),X)\) for any \(X,Y \in TM\), we get \(X(\omega )g(\phi U,Y)-Y(\omega )g(\phi U,X)+\omega (g(\nabla _X\phi U,Y)-g(\nabla _Y\phi U,X))=0\). Taking \(Y=\xi \), this yields \(\omega (g(\nabla _X\phi U,\xi )-g (\nabla _{\xi }\phi U,X))=0\) for any \(X \in TM\). Thus, either \(\omega =0\) or \(g(\nabla _X\phi U,\xi )=g(\nabla _{\xi }\phi U,X)\) for any \(X \in TM\). If we take \(X=U\), we have \(-g(U,AU)=g(\nabla _{\xi }\phi U,U)\). Then, \(4\alpha ^2+\beta ^2=1\). This yields \(4\alpha (\phi U)(\alpha )+\beta (\phi U)(\beta )=0\). From (4.35) and (4.36), we have \(9\alpha ^2+\beta ^2=1\). Therefore, \(\alpha =0\), which is impossible. So we have \(\omega =0\) and \(\alpha ^2=1\). From (4.35) \((\phi U)(\alpha )=0\) and from (4.36) \((\phi U)(\beta )=-(2\beta ^2+1)\). Then, from (4.18) and the fact that \(g((\nabla _{\xi }A)U-(\nabla _UA)\xi , U)=0\), we have \(U(\beta )=0\) and then \(grad(\beta )=-(2\beta ^2+1)\phi U\).

Applying the same reasoning to \(grad(\beta )\), \(-(1+2\beta ^2)(g(\nabla _X\phi U,\xi )-g(\nabla _{\xi }\phi U,X))=0\) for any \(X \in TM\). This yields \(g(\nabla _X\phi U,\xi )=g(\nabla _{\xi }\phi U,X)\) for any \(X \in TM\). Taking \(X=U\), it follows \(4\alpha ^2+\beta ^2=1\) and being \(\alpha ^2=1\), \(\beta ^2=-3\), which is impossible and proves that \(M\) must be Hopf.

If \(M\) is Hopf with \(A\xi =\alpha \xi \), from (4.1), we get \(\phi R_{\xi }=R_{\xi }\phi \). Let \(Y\in \mathbb {D}\) a unit vector field such that \(AY=\lambda Y\). Therefore, \(\alpha \lambda Y=\alpha A\phi Y\). Then, either \(\alpha =0\) and \(M\) is locally congruent to a tube of radius \(\frac{\pi }{4}\) around a complex submanifold of \(\mathbb {C}P^m\), see [2], or \(A\phi =\phi A\) and from Theorem 2.1, \(M\) is locally congruent to a type \((A)\) real hypersurface.

It is very easy to see that these real hypersurfaces satisfy (4.1), and we finish the proof. \(\square \)