1 Introduction

The study of real hypersurfaces in complex two-plane Grassmannians \(G_2({\mathbb C}^{m+2})\) was initiated by Berndt and Suh [1]. Let us denote by \(G_{2}(\mathbb C ^{m+2})\) the set of all complex two-dimensional linear subspaces in \(\mathbb C^{m+2}\). This set can be identified with the homogeneous space \(SU(m+2)/S(U(2) \times U(m))\). From this, we know that \(G_2({\mathbb C}^{m+2})\) becomes the unique compact, irreducible, Riemannian manifold being equipped with both a Kaehler structure \(J\) and a quaternionic Kaehler structure \(\mathfrak {J}\) not containing \(J\). In other words, \(G_{2}(\mathbb C ^{m+2})\) is the unique compact, irreducible, Kaehler, quaternionic Kaehler manifold which is not a hyper-Kaehler manifold [1, 2].

For real hypersurfaces \(M\) in \(G_{2}(\mathbb C^{m+2})\), we have the following two natural geometric conditions: the \(1\)-dimensional distribution \([\xi ]= \text {Span} \{\xi \}\) and the \(3\)-dimensional distribution \(\mathfrak {D}^{\bot } = \text {Span}\{\xi _{1},\xi _{2}, \xi _{3}\}\) are invariant under the shape operator \(A\) of \(M\). Here the almost contact structure vector field \(\xi \) defined by \(\xi = -JN\) is said to be a Reeb vector field, where \(N\) denotes a local unit normal vector field of \(M\) in \(G_{2}(\mathbb C ^{m+2})\). The almost contact 3-structure vector fields \(\xi _{1},\xi _{2},\xi _{3}\) spanning the 3-dimensional distribution \(\mathfrak {D}^{\bot }\) of \(M\) in \(G_{2}(\mathbb C ^{m+2})\) are defined by \(\xi _{\nu } = - J_{\nu } N\) \((\nu =1, 2, 3\)), where \(J_{\nu }\) denotes a canonical local basis of the quaternionic Kaehler structure \(\mathfrak J\) such that \(T_{x}M = \mathfrak D \oplus \mathfrak D^{\bot }\), \(x \in M\).

By using these two invariant conditions and the result in Alekseevskii [3], Berndt and Suh [1] proved the following:

Theorem A

Let \(M\) be a connected real hypersurface in \(G_{2}(\mathbb C ^{m+2})\), \(m \ge 3\). Then both \([\xi ]\) and \(\mathfrak D^{\bot }\) are invariant under the shape operator of \(M\) if and only if

  1. (A)

    \(M\) is an open part of a tube around a totally geodesic \(G_2({\mathbb C}^{m+1})\) in \(G_2({\mathbb C}^{m+2})\), or

  2. (B)

    \(m\) is even, say \(m = 2n\), and \(M\) is an open part of a tube around a totally geodesic \({\mathbb H}P^n\) in \(G_2({\mathbb C}^{m+2})\).

The Reeb vector field \(\xi \) is said to be Hopf if it is invariant under the shape operator \(A\). The 1-dimensional foliation of \(M\) by the integral curves of the Reeb vector field \(\xi \) is said to be a Hopf foliation of \(M\). We say that \(M\) is a Hopf hypersurface in \(G_2({\mathbb C}^{m+2})\) if and only if the Hopf foliation of \(M\) is totally geodesic. By the almost contact metric structure \((\phi , \xi , \eta , g)\) and the formula \(\nabla _{X}\xi =\phi AX\) for any \(X \in TM\) in Sect. 2, it can be easily checked that \(M\) is Hopf if and only if the Reeb vector field \(\xi \) is Hopf. We will give a brief review of \((\phi , \xi , \eta , g)\) on \(M\) in Sect. 2.

On the other hand, when the distribution \(\mathfrak D^{\bot }\) of a hypersurface \(M\) in \(G_2({\mathbb C}^{m+2})\) is invariant under the shape operator, we say that \(M\) is a \(\mathfrak D^{\bot }\)-invariant hypersurface. Moreover, we say that the Reeb flow on \(M\) in \(G_2({\mathbb C}^{m+2})\) is isometric, when the Reeb vector field \(\xi \) on \(M\) is Killing. This means that the metric tensor \(g\) is invariant under the Reeb flow of \(\xi \) on \(M\).

In [4], Berndt and Suh gave some equivalent conditions of isometric Reeb flow. They gave a characterization of real hypersurfaces of Type \((A)\) in Theorem A in terms of the Reeb flow on \(M\) as follows:

Theorem B

Let \(M\) be a connected orientable real hypersurface in \(G_2({\mathbb C}^{m+2})\), \(m \ge 3\). Then the Reeb flow on \(M\) is isometric if and only if \(M\) is an open part of a tube around a totally geodesic \(G_2({\mathbb C}^{m+1})\) in \(G_2({\mathbb C}^{m+2})\).

In the proof of our Main Theorems, we will use that the Reeb flow on \(M\) is isometric if and only if the shape operator \(A\) commutes with the structure tensor field \(\phi \), that is, \(A\phi = \phi A\). Related to this commuting property, recently, the authors gave many characterizations of model spaces of Type \((A)\) in \(G_2({\mathbb C}^{m+2})\) mentioned in Theorems A and B (see [5, 6]).

On the other hand, Suh [7] gave a characterization of real hypersurfaces of Type \((B)\) in \(G_2({\mathbb C}^{m+2})\) in terms of the contact hypersurface. Moreover, as another characterization of one of Type \((B)\) in \(G_2({\mathbb C}^{m+2})\) related to the Reeb vector field \(\xi \) Lee and Suh [8] obtained the following:

Theorem C

Let \(M\) be a connected orientable Hopf hypersurface in \(G_2({\mathbb C}^{m+2})\), \(m \ge 3\). Then the Reeb vector field \(\xi \) belongs to the distribution \(\mathfrak D\) if and only if \(M\) is locally congruent to an open part of a tube around a totally geodesic \(\mathbb H P^{n}\) in \(G_2({\mathbb C}^{m+2})\), where \(m=2n\).

Usually, any submanifold in Kaehler manifolds has many kinds of connections. Among them, we consider two connections, namely, Levi-Civita and Tanaka–Webster connections for real hypersurfaces \(M\) in \(G_2({\mathbb C}^{m+2})\). In fact, \(G_2({\mathbb C}^{m+2})\) is a Riemannian symmetric space with Riemannian metric and Levi-Civita connection. Using the induced connection from the Levi-Civita connection, many geometers gave some characterizations for real hypersurfaces in \(G_2({\mathbb C}^{m+2})\) related to the covariant derivative \(\nabla \) of the shape operator on \(M\) ([9, 10], etc). For real hypersurfaces in a Kaehler manifold, we consider a new affine connection \(\widehat{\nabla }^{(k)}\) different from the Levi-Civita connection \(\nabla \), namely, the generalized Tanaka–Webster connection (in short, the g-Tanaka–Webster connection). It becomes a generalization of the well-known connection defined by Tanno [11]. Besides, it coincides with Tanaka–Webster connection if a real hypersurface in Kaehler manifolds satisfies \(\phi A + A \phi = 2k \phi \) for a nonzero real number \(k\). The Tanaka–Webster connection is defined as the canonical affine connection on a non-degenerate, pseudo-Hermitian CR-manifold [1214]. Using the generalized Tanaka–Webster connection, \(\widehat{\nabla }^{(k)}\) defined in such a way that

$$\begin{aligned} \widehat{\nabla }^{(k)}_{X}Y = \nabla _{X}Y + g(\phi AX, Y) \xi - \eta (Y) \phi AX - k \eta (X) \phi Y \end{aligned}$$
(*)

for any \(X\), \(Y\) tangent to \(M\), where \(\nabla \) denotes the Levi-Civita connection on \(M\), \(A\) is the shape operator on \(M\) and \(k\) is a nonzero real number, the authors studied some characterizations of real hypersurfaces in \(G_2({\mathbb C}^{m+2})\)  ([15, 16], etc). The latter part of the generalized Tanaka–Webster connection \(g(\phi AX, Y) \xi - \eta (Y) \phi AX - k \eta (X) \phi Y\) is denoted by \(F_{X}Y\). Here the operator \(F_{X}\) is a kind of (1,1)-type tensor and said to be the Tanaka–Webster operator.

On the other hand, there are many results for the classification problem of real hypersurfaces in \(G_2({\mathbb C}^{m+2})\) related to the structure Jacobi operator and Ricci tensor, for example, [1724] and so on. Recently, Pérez and Suh [25] investigated the Levi-Civita and g-Tanaka–Webster covariant derivatives for the shape operator or the structure Jacobi operator of real hypersurfaces in complex projective space \(\mathbb C P^{m}\). In particular, the authors  [25] gave the result about the shape operator as follows:

Theorem D

There exist no real hypersurfaces \(M\) in \(\mathbb C P^{m}\), \(m \ge 2\) such that \(\nabla A = \widehat{\nabla }^{(k)}A\).

Motivated by Theorem D, in this paper, we study a real hypersurface \(M\) in \(G_2({\mathbb C}^{m+2})\) whose Levi-Civita covariant derivative coincides with generalized Tanaka–Webster derivative for the shape operator of \(M\), that is,

$$\begin{aligned} \left( \nabla _{X}A\right) Y = \left( \widehat{\nabla }^{(k)}_{X}A\right) Y \end{aligned}$$
(C-1)

for arbitrary tangent vector fields \(X\) and \(Y\) on \(M\).

The condition (C-1) has a geometric meaning such that the shape operator \(A\) commutes with the Tanaka–Webster operator \(F_{X}\), that is, \(A \cdot F_{X} = F_{X} \cdot A\). This meaning gives any eigenspaces of the shape operator \(A\) are invariant by the Tanaka–Webster operator \(F_{X}\).

From such a point of view, in Sect. 3, we prove that a real hypersurface in Kaehler manifolds satisfying (C-1) must be Hopf. Then from this result, we assert the following:

Theorem 1

There does not exist any real hypersurface in complex two-plane Grassmannians \(G_2({\mathbb C}^{m+2})\), \(m \ge 3\), satisfying (C-1).

First, if we restrict \(X=\xi \) in (C-1), then the following condition (C-2) along the Reeb vector field \(\xi \) becomes a generalized condition weaker than the condition (C-1). This also has a geometric meaning that any eigenspaces of the shape operator \(A\) are invariant by the restricted Tanaka–Webster operator \(F_{\xi }\) in the direction of the Reeb vector field \(\xi \). Thus, we assert the following:

Theorem 2

Let \(M\) be a Hopf hypersurface in complex two-plane Grassmannians \(G_2({\mathbb C}^{m+2})\), \(m \ge 3\). If \(M\) satisfies

$$\begin{aligned} \left( \nabla _{\xi }A\right) Y = \left( \widehat{\nabla }^{(k)}_{\xi }A\right) Y \end{aligned}$$
(C-2)

for any tangent vector field \(Y\) on \(M\), then \(M\) is locally congruent to a tube of radius \(r\) over a totally geodesic \(G_2({\mathbb C}^{m+1})\) in \(G_2({\mathbb C}^{m+2})\).

As a second, let us consider a distribution \(\mathfrak D^{\bot }\) spanned by \(\{{\xi }_1, \xi _{2}, {\xi }_3\}\). Accordingly, if we consider the condition (C-1) to the distribution \(\mathfrak D^{\bot }\), the derivatives of the shape operator \(A\) of \(M\) along the distribution \(\mathfrak D^{\bot }\) becomes a condition more weaker than (C-1). Obviously, this has a geometric meaning that any eigenspaces of the shape operator \(A\) are invariant by the restricted Tanaka–Webster operator \(F_{\xi _{\nu }}\), \(\nu =1,2,3\), along the distribution \(\mathfrak D^{\bot }\). Then we have the following:

Theorem 3

There does not exist a Hopf hypersurface in \(G_2({\mathbb C}^{m+2})\), \(m \ge 3\), satisfying

$$\begin{aligned} \left( \nabla _{\xi _{\nu }}A\right) Y = \left( \widehat{\nabla }^{(k)}_{\xi _{\nu }}A\right) Y, \quad \nu =1,2,3 \end{aligned}$$
(C-3)

for any tangent vector field \(Y\) on \(M\).

Finally, we consider a distribution \(\mathfrak D\) which is an orthogonal complement of \(\mathfrak D^{\bot }\) in \(TM\). Then by restricting the condition (C-1) to the distribution \(\mathfrak D\), we get the following condition (C-4), which becomes another condition more weaker than (C-1). Using this geometric notion, we get:

Theorem 4

There does not exist a Hopf hypersurface in \(G_2({\mathbb C}^{m+2})\), \(m \ge 3\), with

$$\begin{aligned} \left( \nabla _{X}A\right) Y = \left( \widehat{\nabla }^{(k)}_{X}A\right) Y \end{aligned}$$
(C-4)

for all vector fields \(X \in \mathfrak D\) and \(Y\) on \(M\).

In this paper, we refer to [1, 2, 4, 26] for Riemannian geometric structures of \(G_2({\mathbb C}^{m+2})\), and [11, 1316, 27] for generalized Tanaka–Webster connection of real hypersurfaces in Kaehler manifolds.

2 Key Lemmas

Let \(M\) be a real hypersurface in Kaehler manifolds \((\tilde{M}, \tilde{g})\). The induced Riemannian metric on \(M\) is denoted by \(g\). In addition, \(\tilde{\nabla }\) and \(\nabla \) denote the Levi-Civita connections of \(\tilde{M}\) and \(M\), respectively. Let \(N\) be a local unit normal vector field of \(M\) and \(A\) the shape operator of \(M\) with respect to \(N\).

From the Kaehler structure \(J\) of \(\tilde{M}\), we have a tensor field \(\phi \) of type (1,1) on \(M\), given by

$$\begin{aligned} g(\phi X, Y) = \tilde{g}(JX, Y) \end{aligned}$$

for all tangent vector fields \(X\) of \(M\). Moreover, we obtain the unit tangent vector field \(\xi \) and the 1-form \(\eta \) of \(M\) defined by

$$\begin{aligned} \xi =-JN \ \ \text{ and }\ \ \eta (X)=g(X, \xi )=\tilde{g}(JX, N), \end{aligned}$$

respectively. It implies that \(\phi ^{2}X=-X+\eta (X)\xi \), \(\eta (\xi )=1\), \(\phi \xi =0\) and

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

together with Gauss and Weingarten formulas. Thus, the Kaehler structure \(J\) of \(\tilde{M}\) induces an almost contact metric structure \((\phi , \xi , \eta , g)\) on \(M\).

Now let us assume that a real hypersurface \(M\) in \(\tilde{M}\) satisfies

$$\begin{aligned} \left( \nabla _{X}A\right) Y = \left( \widehat{\nabla }^{(k)}_{X}A\right) Y \end{aligned}$$
(C-1)

for all tangent vector fields \(X\) and \(Y\) on \(M\).

From the definition of the g-Tanaka–Webster connection (*), we have

$$\begin{aligned} (\widehat{\nabla }^{(k)}_{X} A)Y&= \widehat{\nabla }^{(k)}_{X}(AY) - A (\widehat{\nabla }^{(k)}_{X} Y) \\&= (\nabla _{X}A)Y + g(\phi AX, AY) \xi - \eta (AY) \phi AX - k \eta (X) \phi AY \\&\quad - g(\phi AX, Y) A \xi + \eta (Y) A \phi AX + k \eta (X) A \phi Y. \end{aligned}$$

Therefore, the condition (C-1) can be written as

$$\begin{aligned}&g(\phi AX, AY) \xi - \eta (AY) \phi AX - k \eta (X) \phi AY \nonumber \\&\quad - g(\phi AX, Y) A \xi + \eta (Y) A \phi AX + k \eta (X) A \phi Y=0 \end{aligned}$$
(2.1)

for all tangent vector fields \(X\) and \(Y\) on \(M\).

In a situation like this, we prove

Lemma 2.1

Let \(M\) be a real hypersurface in a Kaehler manifold \(\tilde{M}\) with the condition (C-1). Then \(M\) becomes a Hopf hypersurface.

Proof

The purpose of this lemma is to show that the structure vector field \(\xi \) is principal. In order to prove this, let us suppose that there is a point where the Reeb vector field \(\xi \) is not principal. Then there exists a neighborhood \(\mathfrak U\) of this point, on which we can define a unit vector field \(U\) orthogonal to \(\xi \) in such a way that

$$\begin{aligned} \beta U = A\xi - g(A\xi , \xi ) \xi = A\xi - \alpha \xi \end{aligned}$$

where \(\beta \) denotes the length of vector filed \(A\xi - \alpha \xi \) and \(\beta (x) \ne 0\) for any point \(x\) in \(\mathfrak U\). Hereafter, unless otherwise stated, let us continue our discussion on this neighborhood \(\mathfrak U\).

Taking \(X=Y=\xi \) in (2.1), we get \(\beta (\alpha + k) \phi U = \beta A \phi U\). Since \(\beta \ne 0\), it follows that

$$\begin{aligned} A \phi U = (\alpha + k) \phi U. \end{aligned}$$
(2.2)

Moreover, putting \(X=Y=U\) in (2.1), we have \(-\beta \phi AU =0\). It implies that

$$\begin{aligned} AU=\beta \xi , \end{aligned}$$
(2.3)

together with \(\beta \ne 0\) and \(\phi ^{2}AU = -AU + \eta (AU) \xi = -AU + \beta \xi \).

Replacing \(Y\) by \(U\) in (2.1), we have

$$\begin{aligned} -\beta \phi AX-g(\phi AX, U)A\xi + k \eta (X) A\phi U =0 \end{aligned}$$
(2.4)

for any tangent vector field \(X\) on \(M\). Substituting \(X=\xi \) in the above equation, we get

$$\begin{aligned} \big (-\beta ^{2} + k (\alpha + k)\big ) \phi U =0 \end{aligned}$$

together with \(\phi A \xi = \beta \phi U\) and (2.2). Taking the inner product with \(\phi U\), it turns to

$$\begin{aligned} \alpha + k = \frac{\beta ^{2}}{k} \end{aligned}$$
(2.5)

because \(k\) is nonzero real number from the definition of g-Tanaka–Webster connection on real hypersurfaces in Kaehler manifolds.

On the other hand, putting \(X=\phi U\) in (2.4), we get

$$\begin{aligned} 2\beta (\alpha + k) U + \alpha (\alpha +k)\xi =0 \end{aligned}$$
(2.6)

from (2.2) and \(\phi ^{2}U=-U\). Taking the inner product with \(\xi \), we obtain \(\alpha (\alpha +k)=0\). By (2.5), this equation is written as \(\frac{\alpha \beta ^{2}}{k}=0\). Since \(k \ne 0\) and \(\beta \ne 0\), we have \(\alpha =0\). Moreover, taking the inner product of (2.6) with \(U\), we have \(\beta (\alpha + k)=0\). It follows that \(\beta =0\), together with \(\alpha =0\) and \(k \ne 0\), which gives a contradiction. This is, the set \(\mathfrak U\) should be empty. Thus, there does not exist such an open neighborhood \(\mathfrak U\) in \(M\), which means that the structure vector field \(\xi \) is principle. Hence, \(M\) must be Hopf under our assumption.

By means of Lemma 2.1, the condition (C-1) implies

$$\begin{aligned}&g(\phi AX, AY) \xi - \alpha \eta (Y) \phi AX - k \eta (X) \phi AY \nonumber \\&\quad - \alpha g(\phi AX, Y) \xi + \eta (Y) A \phi AX + k \eta (X) A \phi Y=0 \end{aligned}$$
(2.7)

for all tangent vector fields \(X\) and \(Y\) on \(M\). Moreover, putting \(Y=\xi \) in the above equation, we obtain \(A \phi AX = \alpha \phi AX\) for any tangent vector field \(X\) on \(M\). From this, the Eq. (2.7) is reduced to

$$\begin{aligned} k \eta (X)(A \phi - \phi A)Y=0 \end{aligned}$$

for all tangent vector fields \(X\) and \(Y\) on \(M\). By the definition of generalized Tanaka–Webster connection for real hypersurfaces in a Kaehler manifold, it follows that

$$\begin{aligned} \eta (X)(A \phi - \phi A)Y=0 \end{aligned}$$

for all tangent vector fields \(X\) and \(Y\) on \(M\).

Summing up above discussions, we assert the following

Lemma 2.2

Let \(M\) be a real hypersurface in a Kaehler manifold \(\tilde{M}\) with the condition (C-1). Then we have

$$\begin{aligned}&A\phi AX = \alpha \phi AX, \end{aligned}$$
(2.8)
$$\begin{aligned}&\eta (X)(A \phi - \phi A)Y=0 \end{aligned}$$
(2.9)

for all tangent vector fields \(X, Y\) on \(M\).

3 Proof of Theorem 1

From now on, we will prove Theorem 1 in the introduction by using the above two Lemmas which are induced from our condition (C-1).

In fact, since \(M\) is a real hypersurface in \(G_2({\mathbb C}^{m+2})\) with the property (C-1), \(M\) becomes a Hopf hypersurface (Lemma 2.1). From this, we have

$$\begin{aligned} \eta (X) (A\phi -\phi A)Y =0, \end{aligned}$$
(3.1)

because \(k\) is a nonzero constant (Lemma 2.2).

Putting \(X=\xi \) in (3.1), it follows that \(A \phi - \phi A =0\). On the other hand, Berndt and Suh [4] gave a characterization of real hypersurfaces of Type \((A)\) in \(G_2({\mathbb C}^{m+2})\) when the shape operator \(A\) of \(M\) commutes with the structure tensor \(\phi \) of \(M\). By virtue of this result, we assert that if \(M\) is a real hypersurface in \(G_2({\mathbb C}^{m+2})\) satisfying (C-1), then \(M\) is locally congruent to an open part of a tube around a totally geodesic \(G_2({\mathbb C}^{m+1})\) in \(G_2({\mathbb C}^{m+2})\).

Let us check that whether the model space \(M_{A}\) of Type \((A)\) satisfies the condition (C-1). In order to do this, let us assume that the shape operator \(A\) of \(M_{A}\) satisfies the condition (C-1). According to Proposition \(3\) given in  [1], the Eq. (2.8) implies

$$\begin{aligned} \beta (\beta - \alpha )=0 \end{aligned}$$
(3.2)

if \(X=\xi _{2}\). But it does not hold, because \(\beta (\beta - \alpha )=2\) where \(\alpha = \sqrt{8} \cot (2\sqrt{2}r)\) and \(\beta = \sqrt{2} \cot (\sqrt{2}r)\), \(r \in (0, \pi /{2\sqrt{2}}\,)\). It completes the proof of Theorem 1. \(\square \)

4 Proofs of Theorems 2 and 3

In this section, we investigate Hopf hypersurfaces in \(G_2({\mathbb C}^{m+2})\) satisfying the property (C-2) and (C-3) which are weaker than (C-1), respectively. On the other hand, \(G_2({\mathbb C}^{m+2})\) is equipped with both a Kaehler and a quaternionic Kaehler structure. By applying these two structures to the normal vector field \(N\) of \(M\) in \(G_2({\mathbb C}^{m+2})\), we get 1- and 3-dimensional distributions on \(M\). For the sake of convenience, we denote \([\xi ]=\text {Span}\{\xi \}\) or \(\mathfrak D^{\bot }=\text {Span}\{{\xi }_1, \xi _{2}, {\xi }_3\}\), respectively. For these two distributions, we define a new distribution \(\mathfrak F\) given by \(\mathfrak F= [\xi ] \cup \mathfrak D^{\bot }\). If we restrict \(X \in \mathfrak F\) in (C-1), then it becomes a new weaker condition for (C-1). Accordingly, we also consider this case.

First, we assume that \(M\) is a Hopf hypersurface in \(G_2({\mathbb C}^{m+2})\) satisfying

$$\begin{aligned} (\nabla _{\xi }A) Y = (\widehat{\nabla }^{(k)}_{\xi }A)Y \end{aligned}$$
(C-2)

for any vector field \(Y \in TM\).

Under our assumptions, this condition means that the structure tensor field \(\phi \) commutes with the shape operator \(A\) of \(M\). In fact, putting \(X=\xi \) in (2.1), it follows that for any tangent vector field \(Y\) on \(M\)

$$\begin{aligned} \phi AY - A \phi Y =0, \end{aligned}$$

because \(M\) is Hopf and \(k\) is a nonzero real number. By Theorem B, we assert our Theorem \(2\) in the introduction. \(\square \)

Next, we observe the following condition of covariant derivatives with respect to the Levi-Civita and g-Tanaka–Webster connections for shape operator \(A\) on Hopf hypersurfaces \(M\) in \(G_2({\mathbb C}^{m+2})\) given by

$$\begin{aligned} (\nabla _{\xi _{\nu }}A) Y = (\widehat{\nabla }^{(k)}_{\xi _{\nu }}A)Y, \quad \nu =1,2,3 \end{aligned}$$
(C-3)

for any tangent vector field \(Y\) on \(M\).

According to (2.1), the condition (C-3) is equal to

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

where \(Y\) is any tangent vector field on \(M\) and \(\nu =1,2,3\).

Putting \(Y=\xi \) in (4.1), we have that

$$\begin{aligned} A \phi A \xi _{\nu }= \alpha \phi A \xi _{\nu }, \quad \nu =1,2,3. \end{aligned}$$
(4.2)

From this, (4.1) can be written as

$$\begin{aligned} \eta (\xi _{\nu }) (A \phi - \phi A ) Y =0 \end{aligned}$$

for any vector field \(Y \in TM\) and \(\nu =1,2,3\).

By virtue of this equation, we have the following two cases:

  • Case 1 \(\eta (\xi _{\nu })=0\), \(\nu =1,2,3\)   and

  • Case 2 \(A \phi = \phi A\).

First, we consider the case \(\eta (\xi _{\nu }) =0\) for any \(\nu =1,2,3\). It means that the Reeb vector field \(\xi \) belongs to the distribution \(\mathfrak D\). By Theorem C, it implies that \(M\) is of Type \((B)\) in Theorem A given in the introduction.

On the other hand, due to Berndt and Suh’s classification [1], all the principal curvatures on a model space of Type \((B)\) are given as follows: \(\alpha =-2\tan (2r)\), \(\beta =2\cot (2r)\), \(\gamma =0\), \(\lambda =\cot (r)\) and \(\mu =-\tan (r)\) for some \(r \in (0, \pi /4)\). Since \(\gamma =0\), we get

$$\begin{aligned} \left( \widehat{\nabla }^{(k)}_{\xi _{\nu }}A\right) \xi - (\nabla _{\xi _{\nu }}A)\xi&= A \phi A \xi _{\nu }- \alpha \phi A \xi _{\nu }\\&= - \alpha \beta \phi \xi _{\nu }\end{aligned}$$

for \(\nu =1,2,3\). In fact, since \(\alpha =-2\tan (2r)\), \(\beta =2\cot (2r)\) for some \(r \in (0, \pi /4)\), the constant \(\alpha \beta \) must be nonzero. It means that the model space of Type \((B)\) does not satisfy our condition (C-3).

Next we consider the remain case that the structure tensor \(\phi \) commutes with the shape operator \(A\) of \(M\). By virtue of Theorem B, we see that \(M\) must be a real hypersurface of Type \((A)\) in \(G_2({\mathbb C}^{m+2})\).

From now on, let us check the converse problem, that is, whether a model space \(M_{A}\) of Type \((A)\) satisfies the condition (C-3) or not. In fact, we suppose that \(M_{A}\) has the condition (C-3), that is, \(M_{A}\) satisfies (4.2). For \(\nu =2\), it becomes \(\beta (\beta -\alpha )=0\). In the proof of Theorem 1, we get \(\beta (\beta -\alpha )=2\), because \(\alpha =\sqrt{8}\cot (2\sqrt{2}r)\) and \(\beta =\sqrt{2}\cot {\sqrt{2}r}\) where \(r \in (0, \pi / 2 \sqrt{2})\). Hence, we assert that \(M_{A}\) does not satisfy the condition (C-3).

Summing up these subcases, we give a complete proof of Theorem 3. \(\square \)

As mentioned above, the distribution \(\mathfrak F\) is defined by \(\mathfrak F= [\xi ] \cup \mathfrak D^{\bot }\). From the structure of \(\mathfrak F\) and the proofs of Theorems 2 and 3, we naturally obtain

Corollary 4.1

There does not exist a Hopf hypersurface in complex two-plane Grassmannians \(G_2({\mathbb C}^{m+2})\), \(m \ge 3\), with

$$\begin{aligned} \left( \nabla _{X}A\right) Y = \left( \widehat{\nabla }^{(k)}_{X}A\right) Y \end{aligned}$$

for any \(X\in \mathfrak F\) and \(Y \in TM\).

5 Proof of Theorem 4

In this section, we observe the condition

$$\begin{aligned} \left( \nabla _{X}A\right) Y = \left( \widehat{\nabla }^{(k)}_{X}A\right) Y \end{aligned}$$
(C-4)

for all tangent vector fields \(X \in \mathfrak D\) and \(Y \in TM\). Putting \(Y=\xi \) in (2.1) and using the assumption that \(M\) is Hopf, we obtain

$$\begin{aligned} A \phi AX = \alpha \phi AX \end{aligned}$$
(5.1)

for any tangent vector field \(X \in \mathfrak D\). Thus, the condition (C-4) is equal to

$$\begin{aligned} \eta (X) ( A\phi - \phi A)Y =0 \end{aligned}$$
(5.2)

for any \(X \in \mathfrak D\) and \(Y \in TM\). From this, we have the following two cases:

  • Case 1    \(A \phi = \phi A\) and

  • Case 2    \(\eta (X) =0\) for any \(X \in \mathfrak D\).

For the first case \(A \phi = \phi A\), we know that \(M\) becomes a model space of Type \((A)\) by Theorem B in the introduction.

Now let us consider the remaining case \(\eta (X) =0\) for any \(X \in \mathfrak D\). It means that the Reeb vector field \(\xi \) belongs to the distribution \(\mathfrak D^{\bot }\). Thus, without loss of generality we may put \(\xi ={\xi }_1\). Under these assumptions, we now prove that \(M\) becomes to be a \(\mathfrak D^{\bot }\)-invariant hypersurface, that is, \(g(A\mathfrak D, \mathfrak D^{\bot })=0\).

Since \(M\) is Hopf, we have the following formula given by Berndt and Suh [4]:

$$\begin{aligned} 2 A\phi AX&=\alpha A\phi X + \alpha \phi A X + 2 \phi X + 2\displaystyle \sum _{\nu =1}^3 \Big \{ \eta _\nu (X) \phi \xi _{\nu } + \eta _\nu (\phi X) \xi _{\nu } \nonumber \\&\quad \ \ + \eta _\nu (\xi ) \phi _{\nu }X - 2\eta (X) \eta _\nu (\xi ) \phi \xi _{\nu } - 2 \eta _\nu (\phi X)\eta _\nu (\xi ) \xi \Big \} \end{aligned}$$

for any tangent vector field \(X\) on \(M\). It can be written as

$$\begin{aligned} 2 A\phi AX =\alpha A\phi X + \alpha \phi A X + 2 \phi X + 2 \phi _{1}X \end{aligned}$$
(5.3)

for any \(X \in \mathfrak D\) and \(\xi ={\xi }_1\). Substituting (5.1) into (5.3), we get

$$\begin{aligned} \alpha ( A\phi - \phi A) X = -2 (\phi X + \phi _{1}X) \end{aligned}$$
(5.4)

for any \(X \in \mathfrak D\).

Let \(\{e_{1}, e_{2}, \cdots , e_{4m-4}, e_{4m-3}=\xi , e_{4m-2}=\xi _{2}, e_{4m-1}={\xi }_3\}\) be an orthonormal basis for \(T_{x}M\), \(x\in M\). Then for any tangent vector field \(Y\) on \(M\) it follows that

$$\begin{aligned} \alpha (A \phi - \phi A)Y&= \sum _{i=1}^{4m-1} g(\alpha (A \phi -\phi A) Y, e_{i})e_{i}\\&= \sum _{i=1}^{4m-4} g(\alpha (A \phi -\phi A) Y, e_{i})e_{i} + \sum _{\nu =1}^{3} g(\alpha (A \phi -\phi A) Y, \xi _{\nu })\xi _{\nu }\\&= \sum _{i=1}^{4m-4} g(\alpha (A \phi -\phi A) e_{i}, Y )e_{i} + \sum _{\nu =1}^{3} g(\alpha (A \phi -\phi A) Y, \xi _{\nu })\xi _{\nu }. \end{aligned}$$

Putting \(Y=e_{k} \in \mathfrak D\,(k=1,2, \cdots , 4m-4)\), this equation can be changed

$$\begin{aligned} \alpha (A \phi - \phi A)e_{k} = \sum _{i=1}^{4m-4} g(\alpha (A \phi -\phi A) e_{i}, e_{k} )e_{i} + \sum _{\nu =1}^{3} g(\alpha (A \phi -\phi A) e_{k}, \xi _{\nu })\xi _{\nu }. \end{aligned}$$

From (5.4), it follows that

$$\begin{aligned} -2 (\phi e_{k} + \phi _{1}e_{k})&= \alpha (A \phi - \phi A)e_{k} \\&= \sum _{i=1}^{4m-4} g(\alpha (A \phi -\phi A) e_{i}, e_{k} )e_{i} + \sum _{\nu =1}^{3} g(\alpha (A \phi -\phi A) e_{k}, \xi _{\nu })\xi _{\nu }\\&= \sum _{i=1}^{4m-4} g( -2 ( \phi e_{i} + \phi _{1}e_{i}), e_{k} )e_{i} + \sum _{\nu =1}^{3} g( -2(\phi e_{k} + \phi _{1}e_{k}), \xi _{\nu })\xi _{\nu }\\&= -2 \sum _{i=1}^{4m-4} g( \phi e_{i}, e_{k}) e_{i} -2 \sum _{i=1}^{4m-4} g(\phi _{1}e_{i}, e_{k} )e_{i}\\&= 2 \sum _{i=1}^{4m-4} g( \phi e_{k}, e_{i}) e_{i} + 2 \sum _{i=1}^{4m-4} g(\phi _{1}e_{k}, e_{i} )e_{i}\\&= 2 \sum _{i=1}^{4m-4} g( \phi e_{k}, e_{i}) e_{i} + 2 \sum _{\nu =1}^{3} g(\phi e_{k}, \xi _{\nu })\xi _{\nu }\\&+ 2 \sum _{i=1}^{4m-4} g(\phi _{1}e_{k}, e_{i} )e_{i} + 2 \sum _{\nu =1}^{3} g(\phi _{1} e_{k}, \xi _{\nu })\xi _{\nu } \\&= 2 \sum _{i=1}^{4m-1} g( \phi e_{k}, e_{i}) e_{i} + 2 \sum _{i=1}^{4m-1} g(\phi _{1}e_{k}, e_{i} )e_{i}\\&= 2 \phi e_{k} + 2 \phi _{1}e_{k} \end{aligned}$$

where in the fourth and sixth equalities, we have used \(g(\phi e_{k}, \xi _{\nu }) = g(\phi _{1} e_{k}, \xi _{\nu })=0\) for any \(\nu \pmod 3\) and nonzero real number \(k\). Thus, we get

$$\begin{aligned} \phi X = -\phi _{1}X \end{aligned}$$
(5.5)

for any tangent vector field \(X\in \mathfrak D\). Differentiating this equation covariantly in the direction of \(Y\), we have

$$\begin{aligned} g(AX, Y)=0 \end{aligned}$$

for all tangent vector fields \(X \in \mathfrak D\) and \(Y \in TM\), where we have used the formulas about the covariant derivative of structure tensors \(\phi \) and \(\phi _{\nu }\) \((\nu =1,2,3)\). It implies that \(M\) must be a \(\mathfrak D^{\bot }\)-invariant hypersurface, if we restrict to \(Y \in \mathfrak D^{\bot }\). Accordingly, for this case we can assert that \(M\) is locally congruent to model spaces of Type \((A)\) by virtue of Theorem A in the introduction.

Summing up these cases, we consequently know that any Hopf hypersurface \(M\) in \(G_2({\mathbb C}^{m+2})\) satisfying the condition (C-4) is of Type \((A)\).

Now it remains only to show that whether a real hypersurface \(M_{A}\) of Type \((A)\) satisfies the condition (C-4) or not. To check this, let us assume that \(M_{A}\) has the condition \( (\nabla _{X}A)Y = (\widehat{\nabla }^{(k)}_{X}A)Y \) for any \(X \in \mathfrak D\) and \(Y \in TM_{A}\). It is equivalent that

$$\begin{aligned} A \phi AX = \alpha \phi AX, \end{aligned}$$
(5.6)

for \(X \in \mathfrak D\) as observed in this section.

From the structure of the tangent vector space \(T_{x}M_{A}\) for a model space of Type \((A)\) at any point \(x\) on \(M_{A}\), we see that the distribution \(\mathfrak D\) is composed with two eigenspaces \(T_{\lambda }\) and \(T_{\mu }\). In addition, since the eigenspace \(T_{\lambda }\) is given by \(T_{\lambda }=\{ X |\ X \bot \mathbb H \xi , JX = J_{1}X\}\) where \(\mathbb H \xi \) denotes quaternionic span of \(\xi \), we see that \(\phi X \in T_{\lambda }\) for any \(X \in T_{\lambda }\). Using these facts, the Eq. (5.6) is reformed as

$$\begin{aligned} (\lambda ^{2} - \alpha \lambda ) \phi X = 0 \end{aligned}$$

for any \(X \in T_{\lambda }\subset \mathfrak D\). From this, we get \(\lambda ^{2} - \alpha \lambda =0\).

On the other hand, from Proposition \(3\) in [1], we know that

$$\begin{aligned} \lambda ^{2} - \alpha \lambda = 2 \end{aligned}$$

where \(\lambda =-\sqrt{2}\tan (\sqrt{2}r)\) and \(\alpha = 2 \sqrt{2} \cot (2\sqrt{2}r)\) for some \(r \in (0, \pi /2\sqrt{2})\). This makes a contradiction, and therefore, we have Theorem 4 in the introduction. \(\square \)