1 Introduction

We denote by \(G_{2}(\mathbb C ^{m+2})\) the set of all complex two-dimensional linear subspaces in \(\mathbb C^{m+2}\). This Riemannian symmetric space \(G_{2}(\mathbb C ^{m+2})\) has a remarkable geometric structure. It is the unique compact irreducible Riemannian manifold with both a Kähler structure \(J\) and a quaternionic Kähler structure \(\mathfrak {J}\) not containing \(J\). Namely, \(G_{2}(\mathbb C ^{m+2})\) is a unique compact, irreducible, Kähler, quaternionic Kähler manifold which is not a hyper-Kähler manifold. Accordingly, in \(G_{2}(\mathbb C ^{m+2})\) we have two natural geometric conditions for real hypersurfaces \(M\): that 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 \(M\) (see [2, 3] and [4]).

The almost contact structure vector field \(\xi \) is defined by \(\xi = -JN\) and 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 \(\{\xi _{1},\xi _{2},\xi _{3}\}\) for 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 a quaternionic Kähler structure \(\mathfrak J\) and \(T_{x}M = \mathfrak D \oplus \mathfrak D^{\bot }\), \(x \in M\).

Using two invariant conditions mentioned above and the result in Alekseevskii [1], Berndt and Suh [3] proved the following:

Theorem A

Let \(M\) be a connected orientable 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 \(G_2({\mathbb C}^{m+2})\).

Furthermore, the Reeb vector field \(\xi \) is said to be Hopf if it is invariant under the shape operator \(A\). The one dimensional foliation of \(M\) by the integral manifolds 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 formulas in Sect. 3, it can be easily checked that \(M\) is Hopf if and only if the Reeb vector field \(\xi \) is Hopf. In particular, \(M\) is said to be a real hypersurface with non-vanishing geodesic Reeb flow in \(G_2({\mathbb C}^{m+2})\) if it has a nonzero principal curvature for the Reeb vector field \(\xi \), that is, \(A\xi =\alpha \xi \) where \(\alpha =g(A\xi , \xi ) \ne 0\).

Using Theorem \(\mathrm A\), many geometers have given characterizations for Hopf hypersurfaces in \(G_2({\mathbb C}^{m+2})\) under certain assumption for various geometry quantities, for instance, shape operator, normal (or structure) Jacobi operator, structure tensor, and so on.

In [4], Berndt and Suh considered some equivalent conditions of isometric Reeb flow. Here, the Reeb flow on \(M\) in \(G_2({\mathbb C}^{m+2})\) is isometric means the Reeb vector field \(\xi \) on \(M\) is Killing. Using this notion, they gave a characterization of real hypersurfaces of Type \((A)\) in Theorem \(\mathrm A\) 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 \(G_2({\mathbb C}^{m+2})\).

Among the equivalent conditions of isometric Reeb flow in [4], it is very useful to our proof in Sect. 5 that the Reeb flow on \(M\) is isometric if and only if the shape operator \(A\) and the structure tensor field \(\phi \) commute with each other, that is, \(A\phi = \phi A\).

Moreover, Lee and Suh [9] gave a characterization of real hypersurfaces of Type \((B)\) in \(G_2({\mathbb C}^{m+2})\) in terms of the Reeb vector field \(\xi \) as follows:

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 \(G_2({\mathbb C}^{m+2})\), \(m=2n\), where the distribution \(\mathfrak D\) denotes the orthogonal complement of \(\mathfrak D^{\bot }=\mathrm{Span}\{{\xi }_1, {\xi }_2, {\xi }_3\}\).

In [11], Suh proved the non-existence of real hypersurfaces in \(G_2({\mathbb C}^{m+2})\) with parallel shape operator, that is, \((\nabla _{X} A)Y =0\), where \(X\) and \(Y\) are any tangent vector field on \(M\). Moreover, he [12] also considered a new condition which is to restrict \(X\) to a distribution \(\mathfrak F=[\xi ] \cup \mathfrak D^{\bot }\), namely \(\mathfrak {F}\) -parallel shape operator, and gave two non-existence theorems to the following two cases of real hypersurfaces \(M\) in \(G_2({\mathbb C}^{m+2})\) with \(\mathfrak F\)-parallel shape operator: One is when \(M\) is a Hopf hypersurface. Another is when \(M\) is a real hypersurface in \(G_2({\mathbb C}^{m+2})\) satisfying \(\mathfrak D^{\bot }\)-invariance under the shape operator, that is, \(A \mathfrak D^{\bot } \subset \mathfrak D^{\bot }\). As regards a weaker condition for a real hypersurface in \(G_2({\mathbb C}^{m+2})\) with parallel shape operator, in [6] and [8] Kim, Yang and the first author considered recurrent and \(\eta \) -parallel shape operator and gave non-existence theorems of Hopf hypersurfaces in \(G_2({\mathbb C}^{m+2})\) satisfying such weaker parallelism conditions, respectively.

Motivated by these notions, it is natural to consider a condition weaker than parallel shape operator for real hypersurfaces \(M\) in \(G_2({\mathbb C}^{m+2})\). From such a point of view, the authors in  [5] studied a generalized parallelness for the shape operator of \(M\) in \(G_2({\mathbb C}^{m+2})\), namely \(\eta \)-parallel shape operator of \(M\). They defined the \(\eta \) -parallel shape operator of \(M\) in \(G_2({\mathbb C}^{m+2})\) if the shape operator \(A\) of \(M\) satisfies \(g((\nabla _{X}A)Y, Z) =0\) for any tangent vectors \(X,Y,Z \in \mathfrak {h}\), where \(\mathfrak h\) denotes the set of all tangent vectors being orthogonal to the Reeb vector \(\xi \) in \(T_{x}M\), \(x \in M\). From this definition, we see that it becomes a weaker condition than parallel shape operator.

Accordingly, we consider a new notion weaker than parallel shape operator, that is, Reeb parallel shape operator which is defined by

$$\begin{aligned} (\nabla _{\xi }A)Y=0 \end{aligned}$$
(*)

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

In this paper, we give a classification of real hypersurfaces in \(G_2({\mathbb C}^{m+2})\) with Reeb parallel shape operator as follows:

Theorem 1

Let \(M\) be a connected orientable real hypersurface in \(G_2({\mathbb C}^{m+2})\), \(m \ge 3\), with non-vanishing geodesic Reeb flow. Then the shape operator of \(M\) is Reeb parallel if and only if \(M\) is an open part of a tube of some radius \(r \in (0, \frac{\pi }{4\sqrt{2}})\cup (\frac{\pi }{4\sqrt{2}}, \frac{\pi }{2\sqrt{2}})\) around a totally geodesic \(G_{2}(\mathbb C^{m+1})\) in \(G_2({\mathbb C}^{m+2})\).

Actually, when the Reeb vector field \(\xi \) belongs to the distribution \(\mathfrak D^{\bot }\), the shape operator \(A\) for a Hopf hypersurface \(M\) in \(G_2({\mathbb C}^{m+2})\) with vanishing geodesic Reeb flow satisfies automatically the Reeb parallelness (see Sect. 6). Using this fact, we give:

Theorem 2

Let \(M\) be a connected orientable real hypersurface in \(G_2({\mathbb C}^{m+2})\), \(m \ge 3\), with Reeb parallel shape operator and vanishing geodesic Reeb flow. If the squared norm of the shape operator satisfies \(\text {Tr}A^{2}=||A||^{2} \le 4m\), then \(M\) is locally congruent to an open part of a tube of radius \(r= \frac{\pi }{4\sqrt{2}}\) around a totally geodesic \(G_{2}(\mathbb C^{m+1})\) in \(G_2({\mathbb C}^{m+2})\).

In order to give the proof of our theorems, in Sect. 2 we recall Riemannian geometry of complex two-plane Grassmannians \(G_2({\mathbb C}^{m+2})\). In Sect. 3, some fundamental formulas including the Codazzi and Gauss equations for real hypersurfaces in \(G_2({\mathbb C}^{m+2})\) will be also recalled. In Sect. 4, we will prove that the Reeb vector field \(\xi \) of a Hopf hypersurface in \(G_2({\mathbb C}^{m+2})\) with Reeb parallel shape operator belongs to either the distribution \(\mathfrak D\) or the distribution \(\mathfrak D^{\bot }\). And in the same section, we will check whether real hypersurfaces of Type \((A)\) or Type \((B)\) in Theorem \(A\) satisfy the condition (*) or not. In Sect. 5, we will give a complete proof of our Theorem 1 according to the non-vanishing geodesic Reeb flow. Finally, we will give the proof of Theorem 2 in Sect. 6.

2 Riemannian Geometry of \(G_2({\mathbb C}^{m+2})\)

In this section, we summarize basic material about \(G_2({\mathbb C}^{m+2})\), for details we refer to [2, 3], and [4]. By \(G_2({\mathbb C}^{m+2})\) we denote the set of all complex two-dimensional linear subspaces in \({\mathbb C}^{m+2}\). The special unitary group \(G = SU(m+2)\) acts transitively on \(G_2({\mathbb C}^{m+2})\) with stabilizer isomorphic to \(K = S(U(2) \times U(m))\subset G\). Then \(G_2({\mathbb C}^{m+2})\) can be identified with the homogeneous space \(G/K\), which we equip with the unique analytic structure for which the natural action of \(G\) on \(G_2({\mathbb C}^{m+2})\) becomes analytic. Denote by \({\mathfrak g}\) and \({\mathfrak k}\) the Lie algebra of \(G\) and \(K\), respectively, and by \({\mathfrak m}\) the orthogonal complement of \({\mathfrak k}\) in \({\mathfrak g}\) with respect to the Cartan-Killing form \(B\) of \({\mathfrak g}\). Then \({\mathfrak g} = {\mathfrak k} \oplus {\mathfrak m}\) is an \(Ad(K)\)-invariant reductive decomposition of \({\mathfrak g}\). We put \(o = eK\) and identify \(T_oG_2({\mathbb C}^{m+2})\) with \({\mathfrak m}\) in the usual manner. Since \(B\) is negative definite on \({\mathfrak g}\), its negative restricted to \({\mathfrak m} \times {\mathfrak m}\) yields a positive definite inner product on \({\mathfrak m}\). By \(Ad(K)\)-invariance of \(B\) this inner product can be extended to a \(G\)-invariant Riemannian metric \(g\) on \(G_2({\mathbb C}^{m+2})\). In this way, \(G_2({\mathbb C}^{m+2})\) becomes a Riemannian homogeneous space, even a Riemannian symmetric space. For computational reasons, we normalize \(g\) such that the maximal sectional curvature of \((G_2({\mathbb C}^{m+2}),g)\) is eight.

When \(m=1\), \(G_2({\mathbb C}^3)\) is isometric to the two-dimensional complex projective space \({\mathbb C}P^2\) with constant holomorphic sectional curvature eight.

When \(m=2\), we note that the isomorphism \(Spin(6) \simeq SU(4)\) yields an isometry between \(G_2({\mathbb C}^4)\) and the real Grassmann manifold \(G_2^+({\mathbb R}^6)\) of oriented two-dimensional linear subspaces in \({\mathbb R}^6\). In this paper, we will assume \(m{\ge }3\).

The Lie algebra \({\mathfrak k}\) of \(K\) has the direct sum decomposition \({\mathfrak k} = {\mathfrak su}(m) \oplus {\mathfrak su}(2) \oplus \mathfrak R\), where \(\mathfrak R\) denotes the center of \({\mathfrak k}\). Viewing \({\mathfrak k}\) as the holonomy algebra of \(G_2({\mathbb C}^{m+2})\), the center \(\mathfrak R\) induces a Kähler structure \(J\) and the \({\mathfrak su}(2)\)-part a quaternionic Kähler structure \({\mathfrak J}\) on \(G_2({\mathbb C}^{m+2})\). If \(J_\nu \) is any almost Hermitian structure in \({\mathfrak J}\), then \(JJ_\nu = J_\nu J\), and \(JJ_\nu \) is a symmetric endomorphism with \((JJ_\nu )^2 = I\) and \(\text{ tr }(JJ_\nu ) =0\) for \(\nu =1,2,3\).

A canonical local basis \(\{J_1,J_2,J_3\}\) of \({\mathfrak J}\) consists of three local almost Hermitian structures \(J_\nu \) in \({\mathfrak J}\) such that \(J_\nu J_{\nu +1} = J_{\nu +2}=-J_{\nu +1}J_\nu \), where the index \(\nu \) is taken modulo three. Since \({\mathfrak J}\) is parallel with respect to the Riemannian connection \(\widetilde{\nabla }\) of \((G_2({\mathbb C}^{m+2}),g)\), there exist for any canonical local basis \(\{J_1,J_2,J_3\}\) of \({\mathfrak J}\) three local one-forms \(q_1,q_2,q_3\) such that

$$\begin{aligned} \widetilde{\nabla }_XJ_\nu = q_{\nu +2}(X)J_{\nu +1} - q_{\nu +1}(X)J_{\nu +2} \end{aligned}$$
(2.1)

for all vector fields \(X\) on \(G_2({\mathbb C}^{m+2})\).

The Riemannian curvature tensor \(\widetilde{R}\) of \(G_2({\mathbb C}^{m+2})\) is locally given by

$$\begin{aligned} \widetilde{R}(X,Y)Z&= g(Y,Z)X - g(X,Z)Y + \ g(JY,Z)JX\nonumber \\&-\, g(J X,Z)JY- 2g(JX,Y)JZ\nonumber \\&+\, \ \sum _{\nu =1}^3\Big \{g(J_{\nu } Y,Z)J_\nu X - g(J_\nu X,Z)J_{\nu }Y - 2g(J_\nu X,Y)J_\nu Z\Big \}\nonumber \\&+\, \ \sum _{\nu =1}^3\Big \{g(J_\nu JY,Z)J_\nu JX- g(J_\nu JX,Z)J_{\nu }JY\Big \}, \end{aligned}$$
(2.2)

where \(\{J_1,J_2,J_3\}\) denotes a canonical local basis of \({\mathfrak J}\).

3 Some Fundamental Formulas

In this section, we derive some basic formulas and the Codazzi equation for a real hypersurface in \(G_2({\mathbb C}^{m+2})\) (see [912] and [7]).

Let \(M\) be a real hypersurface of \(G_2({\mathbb C}^{m+2})\), that is, a submanifold of \(G_2({\mathbb C}^{m+2})\) with real codimension one. The induced Riemannian metric on \(M\) will also be denoted by \(g\), and \(\nabla \) denotes the Riemannian connection of \((M,g)\). Let \(N\) be a local unit normal vector field of \(M\) and \(A\) the shape operator of \(M\) with respect to \(N\).

Now let us put

$$\begin{aligned} JX={\phi }X+{\eta }(X)N,\quad J_{\nu }X={\phi }_{\nu }X+{\eta }_{\nu }(X)N \end{aligned}$$
(3.1)

for any tangent vector field \(X\) of a real hypersurface \(M\) in \(G_2({\mathbb C}^{m+2})\). From the Kähler structure \(J\) of \(G_2({\mathbb C}^{m+2})\) there exists an almost contact metric structure \((\phi ,\xi ,\eta ,g)\) induced on \(M\) in such a way that

$$\begin{aligned} \phi ^{2}X=-X+\eta (X)\xi ,\quad \eta (\xi )=1,\quad \phi \xi =0, \quad \eta (X)=g(X,\xi ) \end{aligned}$$
(3.2)

for any vector field \(X\) on \(M\). Furthermore, let \(\{J_1,J_2,J_3\}\) be a canonical local basis of \({\mathfrak J}\). Then the quaternionic Kähler structure \(J_\nu \) of \(G_2({\mathbb C}^{m+2})\), together with the condition \(J_{\nu }J_{\nu +1} = J_{\nu +2} = -J_{\nu +1}J_{\nu }\) in Sect. 1, induces an almost contact metric 3-structure \((\phi _{\nu }, \xi _{\nu }, \eta _{\nu }, g)\) on \(M\) as follows:

$$\begin{aligned}&\phi _{\nu }^{2}X=-X+\eta _{\nu }(X)\xi _{\nu },\quad \eta _{\nu }(\xi _{\nu })=1, \quad \phi _{\nu } \xi _{\nu } =0,\nonumber \\&{\phi }_{\nu +1}{\xi }_{\nu }=-{\xi }_{{\nu }+2},\quad {\phi }_{\nu }{\xi }_{{\nu }+1}={\xi }_{{\nu }+2},\nonumber \\&{\phi }_{\nu }{\phi }_{{\nu }+1}X = {\phi }_{{\nu }+2}X+{\eta }_{{\nu }+1}(X){\xi }_{\nu },\nonumber \\&{\phi }_{{\nu }+1}{\phi }_{\nu }X=-{\phi }_{{\nu }+2}X+{\eta }_{\nu }(X) {\xi }_{{\nu }+1} \end{aligned}$$
(3.3)

for any vector field \(X\) tangent to \(M\). Moreover, from the commuting property of \(J_{\nu }J=JJ_{\nu }\), \({\nu }=1,2,3\) in Sect. 2 and (3.1), the relation between these two contact metric structures \((\phi , \xi , \eta , g)\) and \((\phi _{\nu }, \xi _{\nu }, \eta _{\nu }, g)\), \(\nu =1,2,3\), can be given by

$$\begin{aligned} {\phi }{\phi }_{\nu }X&= {\phi }_{\nu }{\phi }X+{\eta }_{\nu }(X){\xi }-{\eta }(X) {\xi }_{\nu },\nonumber \\ \eta _{\nu }(\phi X)&= \eta (\phi _{\nu }X), \quad {\phi }{\xi }_{\nu }={\phi }_{\nu }{\xi }. \end{aligned}$$
(3.4)

On the other hand, from the parallelism of Kähler structure \(J\), that is, \(\widetilde{\nabla }J=0\) and the quaternionic Kähler structure \(\mathfrak J\) (see (2.1)), together with Gauss and Weingarten formulas it follows that

$$\begin{aligned} ({\nabla }_X{\phi })Y&= {\eta }(Y)AX-g(AX,Y){\xi },\quad {\nabla }_X{\xi }={\phi }AX,\end{aligned}$$
(3.5)
$$\begin{aligned} {\nabla }_X{\xi }_{\nu }&= q_{{\nu }+2}(X){\xi }_{{\nu }+1}-q_{{\nu }+1}(X) {\xi }_{{\nu }+2}+{\phi }_{\nu }AX,\end{aligned}$$
(3.6)
$$\begin{aligned} ({\nabla }_X{\phi }_{\nu })Y&= -q_{{\nu }+1}(X){\phi }_{{\nu }+2}Y+q_{{\nu }+2}(X) {\phi }_{{\nu }+1}Y +{\eta }_{\nu }(Y)AX\nonumber \\&\ \ -g(AX,Y){\xi }_{\nu }. \end{aligned}$$
(3.7)

Combining these formulas, we find the following:

$$\begin{aligned} {\nabla }_X({\phi }_{\nu }{\xi })&= {\nabla }_X({\phi }{\xi }_{\nu })\nonumber \\&= ({\nabla }_X{\phi }){\xi }_{\nu }+{\phi }({\nabla }_X{\xi }_{\nu })\nonumber \\&= q_{{\nu }+2}(X){\phi }_{{\nu }+1}{\xi }-q_{{\nu }+1}(X){\phi }_{{\nu }+2} {\xi }+{\phi }_{\nu }{\phi }AX\nonumber \\&-g(AX,{\xi }){\xi }_{\nu }+{\eta }({\xi }_{\nu })AX. \end{aligned}$$
(3.8)

Using the above expression (2.2) for the curvature tensor \(\widetilde{R}\) of \(G_2({\mathbb C}^{m+2})\), the equations of Codazzi and Gauss are respectively given by

$$\begin{aligned} (\nabla _XA)Y - (\nabla _YA)X&= \eta (X)\phi Y - \eta (Y)\phi X - 2g(\phi X,Y)\xi \nonumber \\&+\, \sum _{\nu =1}^3 \Big \{\eta _\nu (X)\phi _\nu Y - \eta _\nu (Y)\phi _\nu X - 2g(\phi _\nu X,Y)\xi _\nu \Big \}\nonumber \\&+\, \sum _{\nu =1}^3 \Big \{\eta _\nu (\phi X)\phi _\nu \phi Y - \eta _\nu (\phi Y)\phi _\nu \phi X\Big \}\nonumber \\&+\, \sum _{\nu =1}^3 \Big \{\eta (X)\eta _\nu (\phi Y) - \eta (Y)\eta _\nu (\phi X)\Big \}\xi _\nu \end{aligned}$$
(3.9)

and

$$\begin{aligned} R (X,Y)Z&= g(Y,Z)X - g(X,Z)Y\nonumber \\&+\, g(\phi Y,Z)\phi X - g(\phi X, Z)\phi Y -2g(\phi X, Y)\phi Z\nonumber \\&+\, \ \sum _{\nu =1}^{3} \Big \{g(\phi _{\nu }Y, Z) \phi _{\nu }X - g(\phi _{\nu }X, Z) \phi _{\nu }Y -2g(\phi _{\nu }X, Y) \phi _{\nu } Z \Big \}\nonumber \\&+\, \ \sum _{\nu =1}^3\Big \{g(\phi _{\nu }\phi Y, Z) \phi _{\nu }\phi X - g(\phi _{\nu }\phi X, Z) \phi _{\nu }\phi Y \Big \}\nonumber \\&-\, \ \sum _{\nu =1}^3\Big \{\eta (Y) \eta _{\nu }(Z) \phi _{\nu }\phi X -\eta (X) \eta _{\nu }(Z) \phi _{\nu }\phi Y \Big \}\nonumber \\&-\,\sum _{\nu =1}^{3} \Big \{\eta (X) g(\phi _{\nu }\phi Y, Z) - \eta (Y) g(\phi _{\nu }\phi X, Z)\Big \}\xi _{\nu }\nonumber \\&+\,g(AY,Z)AX -g(AX, Z)AY, \end{aligned}$$
(3.10)

where \(R\) denotes the curvature tensor of a real hypersurface \(M\) in \(G_2({\mathbb C}^{m+2})\).

4 Hopf Hypersurfaces in \(G_2({\mathbb C}^{m+2})\) with Reeb Parallel Shape Operator

From now on, we assume that \(M\) is a Hopf hypersurface in \(G_2({\mathbb C}^{m+2})\) with Reeb parallel shape operator, that is, the shape operator \(A\) of \(M\) satisfies:

$$\begin{aligned} (\nabla _{\xi }A)Y=0 \end{aligned}$$
(*)

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

Then from the equation of Codazzi (3.9), we have

$$\begin{aligned} (\nabla _{\xi } A) Y =(\nabla _{Y}A)\xi + \phi Y + \sum _{\nu =1}^{3} \big \{ \eta _{\nu }(\xi ) \phi _{\nu } Y - \eta _{\nu }(Y) \phi _{\nu }\xi +3\eta _{\nu }(\phi Y) \xi _{\nu } \big \} \end{aligned}$$

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

Since \((\nabla _{Y}A)\xi = (Y\alpha ) \xi + \alpha \phi AY -A\phi A Y\), the condition (*) can be written as

$$\begin{aligned}&(Y\alpha ) \xi + \alpha \phi AY -A\phi A Y + \phi Y\nonumber \\&\quad \quad + \sum _{\nu =1}^{3} \big \{ \eta _{\nu }(\xi ) \phi _{\nu } Y - \eta _{\nu }(Y) \phi _{\nu }\xi +3\eta _{\nu }(\phi Y) \xi _{\nu } \big \}=0 \end{aligned}$$
(4.1)

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

Substituting \(Y=\xi \) in above equation, we have \((\xi \alpha )\xi =0\). From this, we obtain the following result:

Lemma 4.1

Let \(M\) be a Hopf hypersurface in \(G_2({\mathbb C}^{m+2})\), \(m \ge 3\), with Reeb parallel shape operator. Then the principal curvature \(\alpha \) is constant along the direction of \(\xi \), that is, \(\xi \alpha =0\).

In this section, our main purpose is to show that the Reeb vector field \(\xi \) belongs to either the distribution \(\mathfrak D\) or the orthogonal complement \(\mathfrak D^{\bot }\) such that \(T_{x}M=\mathfrak D \oplus \mathfrak D^{\bot }\) for any point \(x \in M\).

To show this fact, unless otherwise stated in this section, we consider that the Reeb vector field \(\xi \) satisfies

$$\begin{aligned} \xi = \eta (X_{0})X_{0}+\eta (\xi _{1})\xi _{1} \end{aligned}$$
(**)

for some unit vectors \(X_{0} \in \mathfrak D\) and \(\xi _{1} \in \mathfrak D^{\bot }\) and \(\eta (X_{0}) \eta (\xi _{1})\ne 0\).

Remark 4.2

Under this situation, in [7] the authors proved that \(\mathfrak D\) and \(\mathfrak D^{\bot }\)-components of the Reeb vector field \(\xi \) are invariant under the shape operator when a Hopf hypersurface in \(G_2({\mathbb C}^{m+2})\) satisfies the condition \(\xi \alpha =0\).

On the other hand, using the notion of the geodesic Reeb flow, Berndt and Suh ([3, 4]) proved the following:

Lemma A

If \(M\) is a connected orientable real hypersurface in \(G_2({\mathbb C}^{m+2})\) with geodesic Reeb flow, then we have the following two equations:

$$\begin{aligned} Y\alpha = (\xi \alpha ) \eta (Y) - 4 \sum _{\nu =1}^{3}\eta _{\nu }(\xi )\eta _{\nu }(\phi Y), \end{aligned}$$
(4.2)

and

$$\begin{aligned} \alpha A\phi Y \!+\! \alpha \phi A Y \!-\!2 A\phi AY \!+\! 2 \phi Y&= 2\displaystyle \sum _{\nu =1}^3 \Big \{\!-\!\eta _\nu (Y) \phi \xi _{\nu } \!-\! \eta _\nu (\phi Y) \xi _{\nu }\nonumber \\&\!-\! \eta _\nu (\xi ) \phi _{\nu }Y \!+\! 2\eta (Y) \eta _\nu (\xi ) \phi \xi _{\nu } \!+\! 2 \eta _\nu (\phi Y)\eta _\nu (\xi ) \xi \Big \}\nonumber \\ \end{aligned}$$
(4.3)

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

Remark 4.3

Assume that the \(\mathfrak D\)-component of \(\xi \) is invariant under the shape operator \(A\), that is, \(AX_{0}=\alpha X_{0}\). By putting \(Y=X_{0}\) in (4.3) and using the fact \(\phi X_{0}=-\eta (\xi _{1})\phi _{1}X_{0}\) which is induced by \(\phi \xi =0\), we see that

$$\begin{aligned} \alpha A\phi X_0 =\big (\alpha ^2 +4\eta ^2(X_0)\big )\,\phi X_0. \end{aligned}$$
(4.4)

Now, using these facts, we prove the following proposition:

Proposition 4.4

Let \(M\) be a real hypersurface in complex two-plane Grassmannians \(G_2({\mathbb C}^{m+2})\), \(m \ge 3\) with Reeb parallel shape operator. Then the Reeb vector field \(\xi \) belongs to either the distribution \(\mathfrak D\) or the distribution \(\mathfrak D^{\bot }\).

Proof

Under our assumption, \(M\) is a Hopf hypersurface with Reeb parallel shape operator, we see that \(\xi \alpha =0\) (Lemma 4.1). Moreover, we know that \(\mathfrak D\) and \(\mathfrak D^{\bot }\)-components of the Reeb vector field \(\xi \) are invariant under the shape operator \(A\) of \(M\), that is, \(A\xi _{1}=\alpha \xi _{1}\) and \(AX_{0}=\alpha X_{0}\) (see Remark 4.2).

Actually, when the smooth function \(\alpha =g(A\xi , \xi )\) identically vanishes, this proposition can be verified directly from (4.2).

Thus, in this proof, we consider only the case that the function \(\alpha \) is non-vanishing. In order to prove our proposition, we put \(Y=X_{0}\) in (4.1). It follows

$$\begin{aligned} (X_{0}\alpha ) \xi + \alpha \phi AX_{0}-A\phi A X_{0}+ \phi X_{0}+ \sum _{\nu =1}^{3} \big \{ \eta _{\nu }(\xi ) \phi _{\nu } X_{0}+3\eta _{\nu }(\phi X_{0}) \xi _{\nu } \big \}=0. \end{aligned}$$

Since \(\phi X_{0}\in \mathfrak {D}\) and \(AX_{0}=\alpha X_{0}\), it becomes

$$\begin{aligned} (X_{0}\alpha ) \xi + \alpha ^{2} \phi X_{0}- \alpha A\phi X_{0}+ \phi X_{0}+ \sum _{\nu =1}^{3}\eta _{\nu }(\xi ) \phi _{\nu } X_{0}=0. \end{aligned}$$

Moreover, using (**), we have

$$\begin{aligned} \eta (X_{0})(X_{0}\alpha ) X_{0}+ \eta ({\xi }_1) (X_{0}\alpha ) {\xi }_1+ \alpha ^{2} \phi X_{0}- \alpha A\phi X_{0}+ \phi X_{0}+ \eta ({\xi }_1) \phi _{1} X_{0}=0. \end{aligned}$$

From (4.4), we obtain

$$\begin{aligned} \eta (X_{0})(X_{0}\alpha ) X_{0}+ \eta ({\xi }_1) (X_{0}\alpha ) {\xi }_1- 4 \eta ^{2}(X_{0})\phi X_{0}+ \phi X_{0}+ \eta ({\xi }_1) \phi _{1} X_{0}=0. \end{aligned}$$

Using \(\phi \xi =0\) and (**), we see that \(\phi X_{0}= -\eta ({\xi }_1)\phi _{1}X_{0}\). It implies that

$$\begin{aligned} \eta (X_{0})(X_{0}\alpha ) X_{0}+ \eta ({\xi }_1) (X_{0}\alpha ) {\xi }_1+ 4 \eta ^{2}(X_{0})\eta ({\xi }_1)\phi _{1} X_{0}=0. \end{aligned}$$
(4.5)

Taking the inner product with \(\phi _{1}X_{0}\) in (4.5), we get \(4\eta ^{2}(X_{0})\eta ({\xi }_1)=0\). It contradicts our assumption \(\eta (X_{0})\eta ({\xi }_1)\ne 0\). Accordingly, we get a complete proof of our Proposition 4.4. \(\square \)

Before giving the proofs of our Theorems in the introduction, let us check whether the shape operator \(A\) of real hypersurfaces of Type \((A)\) and of Type \((B)\) in Theorem \(\mathrm A\) satisfies the condition (*) or not for any tangent vector field \(Y \in TM\).

First, let us check our problem for the case that \(M\) is locally congruent to a real hypersurface of Type \((A)\), that is, an open part of a tube around a totally geodesic \(G_2({\mathbb C}^{m+1})\) in \(G_2({\mathbb C}^{m+2})\) with some radius \(r \in (0, \frac{\pi }{2\sqrt{2}})\). In order to do this, we recall a proposition due to Berndt and Suh [3] as follows :

Proposition A

Let \(M\) be a connected real hypersurface of \(G_2({\mathbb C}^{m+2})\). Suppose that \(A{\mathfrak D} \subset {\mathfrak D}\), \(A\xi = \alpha \xi \), and \(\xi \) is tangent to \({\mathfrak D}^{\bot }\). Let \(J_1 \in \mathfrak {J}\) be the almost Hermitian structure such that \(JN=J_1 N\). Then \(M\) has three (if \(r=\pi /2\sqrt{8})\) or four (otherwise) distinct constant principal curvatures

$$\begin{aligned} \alpha = \sqrt{8}\cot (\sqrt{8}r),\quad \beta = \sqrt{2}\cot (\sqrt{2}r),\quad \lambda =-\sqrt{2}\tan (\sqrt{2}r),\quad \mu = 0 \end{aligned}$$

with some \(r \in (0,\pi /\sqrt{8})\). The corresponding multiplicities are

$$\begin{aligned} m(\alpha ) = 1,\quad m(\beta ) = 2, \quad m(\lambda )=2m-2=m(\mu ), \end{aligned}$$

and the corresponding eigenspaces are

$$\begin{aligned} T_\alpha&= {\mathbb R}\xi ={\mathbb R}JN={\mathbb R}\xi _1=\mathrm{Span}\big \{ \xi \big \}=\mathrm{Span}\big \{ \xi _{1} \big \},\\ T_\beta&= {\mathbb C}^{\bot }\xi ={\mathbb C}^{\bot }N={\mathbb R} \xi _2 \oplus {\mathbb R} \xi _3=\mathrm{Span}\big \{ \xi _{2},\ \xi _{3} \big \},\\ T_\lambda&= \{X| \, X \perp {\mathbb H}\xi ,\ JX=J_1 X\,\},\\ T_\mu&= \{X | \, X\perp {\mathbb H}\xi ,\ JX=-J_1 X\,\}, \end{aligned}$$

where \({\mathbb R}\xi \), \({\mathbb C}\xi \), and \({\mathbb H}\xi \), respectively, denote real, complex, and quaternionic span of the structure vector field \(\xi \), and \({\mathbb C}^{\bot }\xi \) denotes the orthogonal complement of \({\mathbb C}\xi \) in \({\mathbb H}\xi \).

For our convenience, let \(M_{A}\) be a real hypersurface of Type \((A)\) in \(G_2({\mathbb C}^{m+2})\). Using the equation of Codazzi (3.9) and the fact that the principal curvature \(\alpha \) of \(\xi \) is a constant, we have the following equation.

$$\begin{aligned} (\nabla _{\xi } A) Y = \alpha \phi AY - A\phi AY + \phi Y + \phi _{1}Y + 2\eta _{3}(Y) \xi _{2}- 2\eta _{2}(Y) \xi _{3}\end{aligned}$$
(4.6)

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

From now on, using (4.6), let us check whether each eigenspace, \(T_{\alpha }\), \(T_{\beta }\), \(T_{\lambda }\), and \(T_{\mu }\) of \(M_{A}\) in \(G_2({\mathbb C}^{m+2})\), has Reeb parallel shape operator or not.

Case A-1 :   \(Y=\xi (=\xi _1) \in T_\alpha \).

By putting \(Y=\xi \) into (4.6), we know that the shape operator \(A\) becomes Reeb parallel, that is, \((\nabla _{\xi }A) \xi =0\).

Case A-2 :   \(Y \in T_{\beta }\) where \(T_{\beta }=\text {Span}\{ \xi _{2}, \xi _{3}\}\).

Since \(T_{\beta }\) is spanned by \(\xi _{2}\) and \(\xi _{3}\), we put \(Y =\xi _{2}\) and \(Y=\xi _{3}\) in (4.6). Then we have

$$\begin{aligned} (\nabla _{\xi }A)\xi _{2}=(\beta ^2 -\alpha \beta -2)\xi _3 \end{aligned}$$

and

$$\begin{aligned} (\nabla _{\xi }A) \xi _{3}=-(\beta ^2 -\alpha \beta -2)\xi _2, \end{aligned}$$

respectively. On the other hand, we know that

$$\begin{aligned} \beta ^2 -\alpha \beta -2=0, \end{aligned}$$

because \(\alpha = \sqrt{8} \cot (\sqrt{8}r)\) and \(\beta = \sqrt{2} \cot (\sqrt{2}r)\) in Proposition \(\mathrm A\). So we conclude that the shape operator of \(M_{A}\) also satisfies \((\nabla _{\xi }A)Y=0\) for any eigenvector \(Y \in T_{\beta }\).

Case A-3 :   \(Y \in T_\lambda =\{\,Y\,| \ Y \in \mathfrak D,\ JY=J_1 Y\,\}\).

We naturally see that if \(Y \in T_{\lambda }\) then \(\phi Y = \phi _{1} Y\). Moreover, the vector \(\phi Y\) also belong to the eigenspace \(T_{\lambda }\) for any \(Y\in T_{\lambda }\), that is, \(\phi T_{\lambda } \subset T_{\lambda }\). Putting \(Y \in T_{\lambda }\) in (4.6) and together with these facts, we obtain

$$\begin{aligned} (\nabla _{\xi } A)Y=-(\lambda ^2 -\alpha \lambda -2)\phi Y. \end{aligned}$$

But in Proposition \(\mathrm A\), since the principal curvatures \(\alpha \) and \(\lambda \) are given by \(\alpha =\sqrt{8} \cot (\sqrt{8}r)\) and \(\lambda = -\sqrt{2} \tan (\sqrt{2}r)\) for \(r\in (0,\pi /\sqrt{8})\), respectively, we get

$$\begin{aligned} \lambda ^2 -\alpha \lambda -2=0. \end{aligned}$$

It implies that \((\nabla _{\xi } A)Y=0\) for any tangent vector field \(Y \in T_{\lambda }\).

Case A-4 :   \(Y \in T_\mu =\{\,Y\,| \ Y \in {\mathfrak D},\ JY=-J_1 Y\,\}\).

In this case, if \(Y\in T_{\mu }\) then \(\phi Y = -\phi _{1}Y\). Moreover, we see \(\phi T_{\mu } \subset T_{\mu }\). So, (4.6) is reduced to \((\nabla _{\xi } A)X=0\), because \(\mu =0\).

Summing up all cases mentioned above, we can assert that:

Remark 4.5

The shape operator \(A\) of real hypersurfaces of Type \((A)\) in \(G_2({\mathbb C}^{m+2})\) is Reeb parallel.

Next, let us check whether the shape operator \(A\) of real hypersurfaces of Type \((B)\) satisfies the condition (*) for any tangent vector field \(Y \in TM\). From now on, we will denote such real hypersurfaces by \(M_{B}\) for the sake of convenience. As is well known, \(M_{B}\) has five distinct constant principal curvatures as follows [3]:

Proposition B

Let \(M\) be a connected real hypersurface of \(G_2({\mathbb C}^{m+2})\). Suppose that \(A \mathfrak {D} \subset \mathfrak {D}\), \(A\xi =\alpha \xi \), and \(\xi \) is tangent to \(\mathfrak {D}\). Then the quaternionic dimension \(m\) of \(G_2({\mathbb C}^{m+2})\) is even, say \(m=2n\), and \(M\) has five distinct constant principal curvatures

$$\begin{aligned} \alpha = -2 \tan (2r), \quad \beta =2 \cot (2r), \quad \gamma =0, \quad \lambda =\cot (r), \quad \mu =-\tan (r) \end{aligned}$$

with some \(r \in ( 0, \pi /4)\). The corresponding multiplicities are

$$\begin{aligned} m(\alpha )=1, \quad m(\beta )=3=m(\gamma ), \quad m(\lambda )=4n-4=m(\mu ) \end{aligned}$$

and the corresponding eigenspaces are

$$\begin{aligned} T_{\alpha }&= \mathbb {R}\xi =\mathrm{Span}\big \{ \xi \big \},\nonumber \\ T_{\beta }&= \mathfrak {J}J\xi = \mathrm{Span}\big \{ \xi _{\nu }\,|\ \nu =1,2,3 \, \big \},\nonumber \\ T_{\gamma }&= \mathfrak {J}\xi = \mathrm{Span}\big \{ \phi _{\nu }\xi \,|\ \nu =1,2,3 \,\big \},\nonumber \\&T_{\lambda },\quad T_{\mu }, \end{aligned}$$

where

$$\begin{aligned} T_{\lambda }\oplus T_{\mu } = ( \mathbb {H}\mathbb {C}\xi )^{\bot }, \quad \mathfrak {J}T_{\lambda }=T_{\lambda },\quad \mathfrak {J}T_{\mu }= T_{\mu },\quad JT_{\lambda }=T_{\mu }. \end{aligned}$$

The distribution \((\mathbb {H}\mathbb {C}\xi )^{\bot }\) is the orthogonal complement of \(\mathbb {H}\mathbb {C}\xi \) where

$$\begin{aligned} \mathbb {H}\mathbb {C}\xi =\mathbb {R}\xi \oplus \mathbb {R}J\xi \oplus \mathfrak {J}\xi \oplus \mathfrak {J}J \xi . \end{aligned}$$

Now, to prove the claim, we suppose that \(M_{B}\) has Reeb parallel shape operator. Then, since \(\xi \in \mathfrak D\), \(M_{B}\) satisfies the following equation:

$$\begin{aligned} \alpha \phi AY - A \phi AY + \phi Y - \sum _{\nu =1}^{3} \big \{\eta _{\nu }(Y) \phi _{\nu }\xi - 3 \eta _{\nu } (\phi Y) \xi _{\nu }\big \}=0, \quad \forall Y \in TM , \end{aligned}$$

from (4.1).

If we put \(Y=\xi _{2}\in T_{\beta }\) in above equation, it becomes

$$\begin{aligned} \alpha \beta \phi \xi _{2}=0 \end{aligned}$$

because \(A\phi _{2}\xi = \gamma \phi _{2}\xi \) and \(\gamma =0\). From this, it follows that

$$\begin{aligned} \alpha \beta =0. \end{aligned}$$

But, from Proposition \(\mathrm B\), we see that \(\alpha \beta =-4\) for some \(r \in (0, \pi / 4)\). This is a contradiction. So this case can not occur.

Therefore we also have the following:

Remark 4.6

The shape operator \(A\) of real hypersurfaces of Type \((B)\) in \(G_2({\mathbb C}^{m+2})\) does not satisfy the Reeb parallel condition (*).

5 The Proof of Theorem 1

In this section, let \(M\) be a real hypersurface in \(G_2({\mathbb C}^{m+2})\) with Reeb parallel shape operator and non-vanishing geodesic Reeb flow.

In [4], Berndt and Suh proved that

Lemma B

Let \(M\) be a connected orientable real hypersurface in Kähler manifolds. Then the following statements are equivalent:

  1. (a)

    The Reeb flow on \(M\) is geodesic;

  2. (b)

    The Reeb vector field \(\xi \) is a principal curvature vector of \(M\) everywhere;

  3. (c)

    The maximal complex subbundle \(\mathfrak {B}\) of \(TM\) is invariant under the shape operator \(A\) of \(M\).

From this, we see that a real hypersurface \(M\) satisfying our condition becomes Hopf, since our real hypersurface \(M\) has non-vanishing geodesic Reeb flow. Thus by Proposition 4.4, we consider the following two cases:

  • Case I :   the Reeb vector field \(\xi \) belongs to the distribution \(\mathfrak D^{\bot }\),

  • Case II :   the Reeb vector field \(\xi \) belongs to the distribution \(\mathfrak D\).

First of all, let us consider the Case I, that is, \(\xi \in \mathfrak D^{\bot }\). Accordingly, we may put \(\xi =\xi _1\). Under these assumptions, we prove the following:

Proposition 5.1

Let \(M\) be a real hypersurface in \(G_2({\mathbb C}^{m+2})\), \(m \ge 3\), with Reeb parallel shape operator and non-vanishing geodesic Reeb flow. If the Reeb vector field \(\xi \) belongs to the distribution \(\mathfrak D^{\bot }\), then the shape operator \(A\) commutes with the structure tensor field \(\phi \).

Proof

From our assumptions, (4.1) can be written as

$$\begin{aligned} (Y \alpha ) \xi + \alpha \phi AY - A\phi AY + \phi Y + \phi _{1}Y + 2\eta _{3}(Y) \xi _{2}-2\eta _{2}(Y)\xi _{3}=0 \end{aligned}$$

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

$$\begin{aligned} 2 A\phi AY = 2(Y \alpha ) \xi + 2 \alpha \phi AY + 2 \phi Y + 2 \phi _{1}Y + 4\eta _{3}(Y) \xi _{2}- 4\eta _{2}(Y)\xi _{3}. \end{aligned}$$

On the other hand, from (4.3) we also obtain

$$\begin{aligned} 2 A \phi AY = \alpha A\phi Y + \alpha \phi AY + 2 \phi Y + 2 \phi _{1}Y + 4\eta _{3}(Y) \xi _{2}- 4\eta _{2}(Y) \xi _{3}. \end{aligned}$$

Thus from the preceding two equations, we have finally

$$\begin{aligned} 2 (Y\alpha )\xi + \alpha \phi AY - \alpha A \phi Y =0 \end{aligned}$$
(5.1)

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

But, under our assumptions, we have already seen that \(\xi \alpha =0\) (see Lemma 4.1). From this fact, (4.2) can be written as

$$\begin{aligned} Y \alpha = - 4 \sum _{\nu =1}^{3} \eta _{\nu }(\xi )\eta _{\nu }(\phi Y) \end{aligned}$$

for any \(Y \in TM\). Therefore since \(\xi = {\xi }_1\), it follows that \(Y \alpha =0\). Substituting this result into (5.1), it follows that

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

for any tangent vector field \(Y\) on \(M\). It means that the shape operator \(A\) commutes with the structure tensor field \(\phi \) on \(M\) in \(G_2({\mathbb C}^{m+2})\), since \(M\) has non-vanishing geodesic Reeb flow. It completes the proof of our Proposition 5.1.

Remark 5.2

As mentioned in the introduction, the structure tensor field \(\phi \) and the shape operator \(A\) of \(M\) commute with each other if and only if the Reeb flow on \(M\) is isometric (see [4]).

Therefore from Theorem \(\mathrm B\) and Remark 4.5, we have the following:

Theorem 5.3

Let \(M\) be a connected real hypersurface in \(G_2({\mathbb C}^{m+2})\), \(m \ge 3\) with non-vanishing geodesic Reeb flow. The shape operator \(A\) of \(M\) is Reeb parallel and the Reeb vector field \(\xi \) belongs to the distribution \(\mathfrak D^{\bot }\) if and only if \(M\) is locally congruent to an open part of a tube around radius \(r\) on a totally geodesic \(G_{2}(\mathbb C^{m+1})\) in \(G_2({\mathbb C}^{m+2})\) where \(r \in (0, \frac{\pi }{4\sqrt{2}}) \cup (\frac{\pi }{4\sqrt{2}}, \frac{\pi }{2\sqrt{2}})\).

Next, we consider the case \(\xi \in \mathfrak D\). By Theorem \(\mathrm C\), we see that \(M\) is locally congruent to a real hypersurface of Type \((B)\) under our assumptions. But as mentioned in Sect. 4, a real hypersurface of Type \((B)\) does not have Reeb parallel shape operator (see Remark 4.6). From these facts, we obtain the following theorem:

Theorem 5.4

There does not exist any real hypersurface in \(G_2({\mathbb C}^{m+2})\), \(m \ge 3\), with Reeb parallel shape operator and non-vanishing geodesic Reeb flow when the Reeb vector field \(\xi \) belongs to the distribution \(\mathfrak D\).

Combining Proposition 4.4, and Theorems 5.3 and 5.4, this completes the proof of our Theorem 1 in the introduction.

6 The Proof of Theorem 2

From now on, let \(M\) be a real hypersurface in \(G_2({\mathbb C}^{m+2})\), \(m \ge 3\), with Reeb parallel shape operator and vanishing geodesic Reeb flow. By virtue of Lemma \(\mathrm B\) given in the previous section, \(M\) must be Hopf, that is, \(A\xi =\alpha \xi \) where \(\alpha =g(A\xi , \xi )=0\). Then by Proposition 4.4, we consider the following two cases:

  • Case I :   the Reeb vector field \(\xi \) belongs to the distribution \(\mathfrak D^{\bot }\),

  • Case II :   the Reeb vector field \(\xi \) belongs to the distribution \(\mathfrak D\).

By virtue of Theorem \(\mathrm C\) and Proposition \(\mathrm B\), we assert that when the Reeb vector field \(\xi \) belongs to the distribution \(\mathfrak D\), there does not exist any real hypersurface in \(G_2({\mathbb C}^{m+2})\), \(m \ge 3\), with Reeb parallel shape operator and vanishing geodesic Reeb flow. In fact, a real hypersurface of Type \((B)\) in Theorem \(A\) due to Berndt and Suh [3] does not have vanishing geodesic Reeb flow (see Proposition \(\mathrm B\) in Sect. 4).

From such a point of view, from now on we only consider the Case I, that is, \(\xi \in \mathfrak D^{\bot }\). Accordingly, we may put \(\xi =\xi _1\). Under these assumptions, we prove the following:

Proposition 6.1

Let \(M\) be a real hypersurface in \(G_2({\mathbb C}^{m+2})\), \(m \ge 3\), with Reeb parallel shape operator and vanishing geodesic Reeb flow. If the Reeb vector field \(\xi \) belongs to the distribution \(\mathfrak D^{\bot }\) and the squared norm of the shape operator satisfies \(||A||^{2} \le 4m\), then the Reeb flow on \(M\) is isometric.

Proof

Since \(M\) has vanishing geodesic Reeb flow, that is, \(A\xi =0\), we obtain

$$\begin{aligned} (\nabla _{X}A)\xi = - A \phi AX \end{aligned}$$

for any tangent vector field \(X\) on \(M\). Using the equation of Codazzi (3.9), we get

$$\begin{aligned} (\nabla _{X}A)\xi = -\phi X - \phi _{1}X - 2 \eta _{3}(X) \xi _{2}+ 2 \eta _{2}(X) \xi _{3}\end{aligned}$$

together with our assumptions that \(M\) has Reeb parallel shape operator and \(\xi ={\xi }_1\) (since we now consider the case \(\xi \in \mathfrak D^{\bot }\), we may put \(\xi ={\xi }_1\)). Hence the above two equations give us

$$\begin{aligned} A \phi AX = \phi X + \phi _{1}X + 2 \eta _{3}(X) \xi _{2}- 2 \eta _{2}(X) \xi _{3}\end{aligned}$$
(6.1)

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

Moreover, applying the structure tensor \(\phi \) to (6.1), it can be written as

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

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

Let \(\{e_{1}, e_{2}, \cdots , e_{4m-1} \}\) be an orthonormal basis for \(T_{x}M\) where \(x\) is any point of \(M\). Then we get

$$\begin{aligned} \phi A \phi A e_{i}= -e_{i}+ \eta (e_{i})\xi + \phi \phi _{1}e_{i}- 2 \eta _{3}(e_{i}) \xi _{3}- 2 \eta _{2}(e_{i}) \xi _{2}\end{aligned}$$
(6.2)

for \(i=1,2, \cdots , 4m-1\). From this, we calculate the trace of the matrix \(\phi A \phi A\), that is,

$$\begin{aligned} \text {Tr}(\phi A \phi A)&= \sum _{i=1}^{4m-1} g(\phi A \phi A e_{i}, e_{i})\nonumber \\&= -\sum _{i=1}^{4m-1} g(e_{i}, e_{i}) + \sum _{i=1}^{4m-1} \eta (e_{i})g(\xi , e_{i}) + \sum _{i=1}^{4m-1} g(\phi \phi _{1}e_{i}, e_{i})\nonumber \\&- 2 \sum _{i=1}^{4m-1}\eta _{3}(e_{i}) g(\xi _{3}, e_{i}) - 2 \sum _{i=1}^{4m-1}\eta _{2}(e_{i}) g(\xi _{2}, e_{i})\nonumber \\&= -(4m-1) + g(\xi , \xi ) + \text {Tr}(\phi \phi _{1}) -2g(\xi _{3}, \xi _{3}) -2g(\xi _{2}, \xi _{2})\nonumber \\&= -4m, \end{aligned}$$
(6.3)

together with \(\text {Tr}(\phi \phi _{\nu })=2\eta _{\nu }(\xi )\), \(\nu =1,2,3\) (see [13]).

On the other hand, we are able to calculate the following:

$$\begin{aligned} ||\phi A - A \phi ||^{2}&= \sum _{i=1}^{4m-1} g\big ((\phi A - A \phi )e_{i}, (\phi A - A \phi )e_{i}\big )\nonumber \\&= - \sum _{i=1}^{4m-1}g(A \phi ^{2} A e_{i}, e_{i}) + \sum _{i=1}^{4m-1} g(\phi A \phi A e_{i}, e_{i})\nonumber \\&+ \sum _{i=1}^{4m-1}g(A \phi A \phi e_{i}, e_{i}) - \sum _{i=1}^{4m-1} g(\phi A^{2} \phi e_{i}, e_{i})\nonumber \\&= \sum _{i=1}^{4m-1}g(A^{2} e_{i}, e_{i}) - \sum _{i=1}^{4m-1}\eta (Ae_{i})g(A \xi , e_{i})\nonumber \\&+ 2\sum _{i=1}^{4m-1}g(A \phi A \phi e_{i}, e_{i}) - \sum _{i=1}^{4m-1} g(\phi A^{2} \phi e_{i}, e_{i})\nonumber \\&= \text {Tr}A^{2} + 2 \text {Tr} (A \phi A \phi ) - \text {Tr}(\phi A^{2} \phi )\nonumber \\&= \text {Tr}A^{2} + 2 \text {Tr} (\phi A \phi A) - \text {Tr}(A^{2} \phi ^{2})\nonumber \\&= 2\text {Tr}A^{2} + 2 \text {Tr} (\phi A \phi A), \end{aligned}$$
(6.4)

using the facts, \(A\xi =0\) and \(\text {Tr}(AB)=\text {Tr}(BA)\) for any two matrices \(A, B\) with same size.

From this, together with (6.3) and our assumption for the squared norm of shape operator \(A\) of \(M\), the left side of (6.4) should vanish for a real hypersurface \(M\) in \(G_2({\mathbb C}^{m+2})\) with \(\alpha =0\) and \(\nabla _{\xi }A=0\). This gives that the shape operator \(A\) commutes with the structure tensor \(\phi \), that is, \(A\phi = \phi A\). According to the result due to Berndt and Suh [4], the Reeb flow on \(M\) becomes isometric. It completes a proof of our proposition. \(\square \)

Hence from Theorem \(\mathrm B\), we can assert that if a real hypersurface \(M\) in \(G_2({\mathbb C}^{m+2})\) satisfies the conditions in Proposition 6.1, then \(M\) becomes a model space of Type \((A)\) in Theorem \(\mathrm A\). To serve the convenience of notation, a model space of Type \((A)\) with radius \(r\) is denoted by \(M_{A}\) or \( M_{A}(r)\). From this let us now check if the model space \(M_{A}\) satisfies the assumptions in Proposition 6.1 or not.

First, we can state that \(M_{A}\) has Reeb parallel shape operator from the observation given in Sect. 4. Moreover, we see that a model space \(M_A\) becomes an open part of a tube around a totally geodesic \(G_2({\mathbb C}^{m+1})\) in \(G_2({\mathbb C}^{m+2})\) with radius \(r=\frac{\pi }{4\sqrt{2}}\), because the principal curvature \(\alpha \) of \(\xi \) on \(M_{A}\) must be zero. From this and Proposition \(\mathrm A\) in Sect. 4, we have the following three distinct principal curvatures and the corresponding multiplicities with respect to the eigenspaces of \(M_{A}(\frac{\pi }{4\sqrt{2}})\):

Principal curvature

Multiplicity

Eigenspace

\(\alpha =0\)

\(1\)

\(T_{\alpha }=\text {Span}\{\xi \}\)

\(\beta =\sqrt{2}\)

\(2\)

\(T_{\beta }=\text {Span}\{\xi _{2}, \xi _{3}\}\)

\(\lambda =-\sqrt{2}\)

\(2(m-1)\)

\(T_{\lambda }=\{X\,|\ X\bot \mathbb {H}\xi ,\ JX=J_{1}X \}\)

\(\mu =0\)

\(2(m-1)\)

\(T_{\mu }=\{X\,|\ X \bot \mathbb {H}\xi ,\ JX= -J_{1}X \}\)

By this table and a straightforward calculation we have the squared norm of the shape operator \(A\) on \(M_{A}(\frac{\pi }{4\sqrt{2}})\) as follows.

$$\begin{aligned} ||A^{2}||&= \sum _{i=1}^{4m-1} g(Ae_{i}, Ae_{i})\nonumber \\&= \sum _{i=1}^{2m-2} g(Ae_{i}, Ae_{i}) + \sum _{i=2m-1}^{4m-4} g(Ae_{i}, Ae_{i}) + g(A e_{4m-3}, A e_{4m-3})\nonumber \\&+ g(A e_{4m-2}, A e_{4m-2})+ g(A e_{4m-1}, A e_{4m-1})\nonumber \\&= \sum _{i=1}^{2m-2} \lambda ^{2}g(e_{i}, e_{i}) + \sum _{i=2m-1}^{4m-4} \mu ^{2}g(e_{i}, e_{i}) + g(A \xi , A \xi )\nonumber \\&+ g(A \xi _{2}, A \xi _{2})+ g(A \xi _{3}, A \xi _{3})\nonumber \\&= 2(m-1) \lambda ^{2} + 2(m-1) \mu ^{2} + \alpha ^{2} + 2 \beta ^{2}\nonumber \\&= 4(m-1) + 4\nonumber \\&= 4m, \end{aligned}$$

where \(e_{1}, e_{2}, \cdots , e_{2m-2}\in T_{\lambda }\), \(e_{2m-1}, \cdots , e_{4m-4}\in T_{\mu }\), \(e_{4m-3}=\xi ={\xi }_1\), \(e_{4m-2}=\xi _{2}\), \(e_{4m-1}=\xi _{3}\). From this calculation, we see that \(M_{A}(\frac{\pi }{4\sqrt{2}})\) also satisfies our assumption in Proposition 6.1.

Summing up these discussions, we obtain our Theorem 2 mentioned in the introduction. \(\square \)

Lastly, we will give a proof for our assertion given in the introduction as follows.

Lemma 6.2

Let \(M\) be a real hypersurface in \(G_2({\mathbb C}^{m+2})\), \(m \ge 3\), with vanishing geodesic Reeb flow. If the Reeb vector field \(\xi \) belongs to the distribution \(\mathfrak D^{\bot }\), then the shape operator \(A\) of \(M\) is Reeb parallel.

Proof

Using the equation of Codazzi (3.9), we obtain that

$$\begin{aligned} (\nabla _{\xi } A) Y -(\nabla _{Y}A)\xi = \phi Y + \sum _{\nu =1}^{3} \big \{ \eta _{\nu }(\xi ) \phi _{\nu } Y - \eta _{\nu }(Y) \phi _{\nu }\xi +3\eta _{\nu }(\phi Y) \xi _{\nu } \big \} \end{aligned}$$

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

From our assumptions, \(A\xi = 0\) and \(\xi = {\xi }_1\), we have

$$\begin{aligned} (\nabla _{\xi } A) Y + A\phi AY = \phi Y + \phi _{1}Y + 2 \eta _{2}(Y) \xi _{2}-2 \eta _{3}(Y)\xi _{3}. \end{aligned}$$
(6.5)

Moreover, since \(M\) is Hopf, (4.3) implies that

$$\begin{aligned} A\phi AY = \phi Y + \phi _{1}Y + 2 \eta _{2}(Y) \xi _{2}-2 \eta _{3}(Y)\xi _{3}, \end{aligned}$$
(6.6)

together with \(\alpha =0\) and \(\xi ={\xi }_1\).

From (6.5) and (6.6), we get \((\nabla _{\xi }A)Y=0\) for any tangent vector field \(Y\) on \(M\). That is, a real hypersurface \(M\) in \(G_2({\mathbb C}^{m+2})\) with vanishing geodesic Reeb flow, that is, \(\alpha =g(A\xi , \xi )=0\) and \(\xi \in \mathfrak D^{\bot }\) has automatically Reeb parallel shape operator.