Keywords

These keywords were added by machine and not by the authors. This process is experimental and the keywords may be updated as the learning algorithm improves.

1 Introduction

A typical example of Hermitian symmetry spaces of rank two is the complex two-plane Grassmannian \(G_2({\mathbb C}^{m+2})\) defined by the set of all complex two-dimensional linear subspaces in \({\mathbb C}^{m+2}\). Another one is complex hyperbolic two-plane Grassmannians \(SU_{2,m}/S(U_{2}{\cdot }U_{m})\) the set of all complex two-dimensional linear subspaces in indefinite complex Euclidean space \({\mathbb C}_2^{m+2}\).

Characterizing model spaces of real hypersurfaces under certain geometric conditions is one of our main interests in the classification theory in complex two-plane Grassmannians \(G_2({\mathbb C}^{m+2})\) or complex hyperbolic two-plane Grassmannians \(SU_{2,m}/S(U_{2}{\cdot }U_{m})\). In this paper, we use the same geometric condition on real hypersurfaces in \(SU_{2,m}/S(U_{2}{\cdot }U_{m})\) as used in \(G_2({\mathbb C}^{m+2})\) to compare the results.

\(G_2({\mathbb C}^{m+2})=SU_{2+m}/S(U_2{\cdot }U_m)\) has a compact transitive group \(SU_{2+m}\), however \(SU_{2,m}/S(U_{2}{\cdot }U_{m})\) has a noncompact indefinite transitive group \(SU_{2,m}\). This distinction gives various remarkable results. Riemannian symmetric space \(SU_{2,m}/S(U_{2}{\cdot }U_{m})\) has a remarkable geometrical structure. It is the unique noncompact, Kähler, irreducible, quaternionic Kähler manifold with negative curvature.

Suppose that M is a real hypersurface in \(G_2({\mathbb C}^{m+2})\) (or \(SU_{2,m}/S(U_{2}{\cdot }U_{m})\)). Let N be a local unit normal vector field of M in \(G_2({\mathbb C}^{m+2})\) (or \(SU_{2,m}/S(U_{2}{\cdot }U_{m})\)). Since \(G_2({\mathbb C}^{m+2})\)(or \(SU_{2,m}/S(U_{2}{\cdot }U_{m})\)) has the Kähler structure J, we may define the Reeb vector field \(\xi =-{ JN}\) and a one dimensional distribution \([\xi ]={\mathscr {C}}^{\bot }\) where \(\mathscr {C}\) denotes the orthogonal complement in \(T_xM\), \(x\in M\), of the Reeb vector field \(\xi \). The Reeb vector field \(\xi \) is said to be Hopf if \(\mathscr {C}\) (or \({\mathscr {C}}^{\bot }\)) is invariant under the shape operator A of M. The one dimensional foliation of M defined by the integral curves of \(\xi \) is said to be a Hopf foliation of M. We say that M is a Hopf hypersurface if and only if the Hopf foliation of M is totally geodesic. By the formulas in [5, Sect. 2], it can be checked that \(\xi \) is Hopf vector field if and only if M is Hopf hypersurface.

From the quaternionic Kähler structure \(\mathfrak J\) of \(G_2({\mathbb C}^{m+2})\) (or \(SU_{2,m}/S(U_{2}{\cdot }U_{m})\)), there naturally exist almost contact 3-structure vector fields \(\xi _{\nu }=-J_{\nu }N\), \(\nu =1,2,3\). Put \({\mathscr {Q}}^{\bot }= \text {Span}\{\,\xi _1, \xi _2, \xi _3\}\). It is a 3-dimensional distribution in the tangent bundle \({ TM}\) of M. In addition, we denoted by \(\mathscr {Q}\) the orthogonal complement of \({\mathscr {Q}}^{\bot }\) in \({ TM}\). It is the quaternionic maximal subbundle of \({ TM}\). Thus the tangent bundle of M is expressed as a direct sum of \(\mathscr {Q}\) and \({\mathscr {Q}}^{\bot }\).

For any geodesic \(\gamma \) in M, a (1,1) type tensor field T is said to be Killing if \(T\dot{\gamma }\) is parallel displacement along \(\gamma \), which gives \(0=\nabla _{\dot{\gamma }}(T\dot{\gamma })=(\nabla _{\dot{\gamma }}T)\dot{\gamma }+T(\nabla _{\dot{\gamma }}\dot{\gamma })=(\nabla _{\dot{\gamma }}T)\dot{\gamma }\). That is, \((\nabla _X T)X=0\) for any tangent vector field X on M (see [2]).

$$\begin{aligned} \begin{aligned} 0&=(\nabla _{X+Y}T)(X+Y)\\&=(\nabla _{X}T)X+(\nabla _{X}T)Y+(\nabla _{Y}T)X+(\nabla _{Y}T)Y\\&=(\nabla _{X}T)Y+(\nabla _{Y}T)X \end{aligned} \end{aligned}$$

for any vector fields X and Y on M.

Thus the Killing tensor field T is equivalent to \((\nabla _{X}T)Y+(\nabla _{Y}T)X=0\).

From this notion, in this paper we consider a new condition related to the shape operator A of M defined in such a way that

$$\begin{aligned} (\nabla _XA)Y+(\nabla _YA)X=0 \end{aligned}$$
(C-1)

for any vector fields X on M.

In this paper, we give a complete classification for real hypersurfaces in \(\bar{M}\) (\(G_2({\mathbb C}^{m+2})\) or \(SU_{2,m}/S(U_{2}{\cdot }U_{m})\)) with Killing shape operator. In order to do it, we consider a problem related to the following:

Theorem 1

There does not exist any real hypersurface in \(\bar{M}\) complex Grassmannians of rank two, \(m \ge 3\), with Killing shape operator.

Since the notion of Killing tensor field is weaker than the notion of parallel tensor field, by Theorem 1, we naturally have the following:

quotation There does not exist any real hypersurface in \(G_2({\mathbb C}^{m+2})\), \(m \ge 3\), with parallel shape operator (see [11]).

On the other hand, by virtue of Theorem 2 we can assert the following:

Corollary 1

There does not exist any hypersurface in \(SU_{2,m}/S(U_{2}{\cdot }U_{m})\), \(m \ge 3\) with parallel shape operator.

2 Riemannian Geometry of \(G_2({\mathbb C}^{m+2})\) and \(SU_{2,m}/S(U_{2}{\cdot }U_{m})\)

In this section we summarize basic material about \(G_2({\mathbb C}^{m+2})\), for details we refer to [5, 6, 11, 12]. The complex two-plane Grassmannian \(G_2({\mathbb C}^{m+2})\) is defined by 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 \(\bar{\nabla }\) 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

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

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

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

$$\begin{aligned} \begin{aligned} \widetilde{R}(X,Y)Z&= g(Y,Z)X - g(X,Z)Y + \ g(JY,Z)JX \\&\quad - g(J X,Z)JY- 2g(JX,Y)JZ \\&\quad + \ \sum _{\nu =1}^3\Big \{g(J_{\nu } Y,Z)J_\nu X - g(J_\nu X,Z)J_{\nu }Y - 2g(J_\nu X,Y)J_\nu Z\Big \}\\&\quad + \ \sum _{\nu =1}^3\Big \{g(J_\nu JY,Z)J_\nu JX- g(J_\nu JX,Z)J_{\nu }JY\Big \}, \end{aligned} \end{aligned}$$
(2.1)

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

Now we summarize basic material about complex hyperbolic two-plane Grassmann manifolds \(SU_{2,m}/S(U_{2}{\cdot }U_{m})\), for details we refer to [14, 16].

The Riemannian symmetric space \(SU_{2,m}/S(U_2{\cdot }U_m)\), which consists of all complex two-dimensional linear subspaces in indefinite complex Euclidean space \({\mathbb C}_2^{m+2}\), is a connected, simply connected, irreducible Riemannian symmetric space of noncompact type and with rank 2. Let \(G = SU_{2,m}\) and \(K = S(U_2{\cdot }U_m)\), and denote by \({\mathfrak {g}}\) and \({\mathfrak {k}}\) the corresponding Lie algebra of the Lie group G and K, respectively. Let B be the Killing form of \({\mathfrak {g}}\) and denote by \({\mathfrak {p}}\) the orthogonal complement of \({\mathfrak {k}}\) in \({\mathfrak {g}}\) with respect to B. The resulting decomposition \({\mathfrak {g}} = {\mathfrak {k}} \oplus {\mathfrak {p}}\) is a Cartan decomposition of \({\mathfrak {g}}\). The Cartan involution \(\theta \in {\text{ A }ut}({\mathfrak {g}})\) on \({\mathfrak {s}}{\mathfrak {u}}_{2,m}\) is given by \(\theta (A) = I_{2,m} A I_{2,m}\), where

$$\begin{aligned} \begin{aligned} I_{2,m} = \begin{pmatrix} -I_{2} &{} 0_{2,m} \\ 0_{m,2} &{} I_{m} \end{pmatrix} \end{aligned} \end{aligned}$$

\(I_2\) (resp., \(I_m\)) denotes the identity \(2 \times 2\)-matrix (resp., \(m \times m\)-matrix). Then \(< X , Y > = -B(X,\theta Y)\) becomes a positive definite \({\text{ A }d}(K)\)-invariant inner product on \({\mathfrak {g}}\). Its restriction to \({\mathfrak {p}}\) induces a metric g on \(SU_{2,m}/S(U_2{\cdot }U_m)\), which is also known as the Killing metric on \(SU_{2,m}/S(U_2{\cdot }U_m)\). Throughout this paper we consider \(SU_{2,m}/S(U_2{\cdot }U_m)\) together with this particular Riemannian metric g.

The Lie algebra \({\mathfrak {k}}\) decomposes orthogonally into \({\mathfrak {k}} = {\mathfrak {s}}{\mathfrak {u}}_2 \oplus {\mathfrak {s}}{\mathfrak {u}}_m \oplus {\mathfrak {u}}_1\), where \({\mathfrak {u}}_1\) is the one-dimensional center of \({\mathfrak {k}}\). The adjoint action of \({\mathfrak {s}}{\mathfrak {u}}_2\) on \({\mathfrak {p}}\) induces the quaternionic Kähler structure \({\mathfrak {J}}\) on \(SU_{2,m}/S(U_2{\cdot }U_m)\), and the adjoint action of

$$\begin{aligned} \begin{aligned} Z = \begin{pmatrix} \frac{mi}{m+2}I_2 &{} 0_{2,m} \\ 0_{m,2} &{} \frac{-2i}{m+2}I_m \end{pmatrix} \in {\mathfrak {u}}_1 \end{aligned} \end{aligned}$$

induces the Kähler structure J on \(SU_{2,m}/S(U_2{\cdot }U_m)\). By construction, J commutes with each almost Hermitian structure \(J_{\nu }\) in \({\mathfrak {J}}\) for \({\nu }=1,2,3\). Recall that a canonical local basis \(\{J_1,J_2,J_3\}\) of a quaternionic Kähler structure \({\mathfrak {J}}\) consists of three almost Hermitian structures \(J_1,J_2,J_3\) in \({\mathfrak {J}}\) such that \(J_{\nu } J_{{\nu }+1} = J_{{\nu } + 2} = - J_{{\nu }+1} J_{\nu }\), where the index \({\nu }\) is to be taken modulo 3. The tensor field \(JJ_{\nu }\), which is locally defined on \(SU_{2,m}/S(U_2{\cdot }U_m)\), is self-adjoint and satisfies \((JJ_{\nu })^2 = I\) and \({\text{ t }r}(JJ_{\nu }) = 0\), where I is the identity transformation. For a nonzero tangent vector X we define \({\mathbb R}X = \{\lambda X \vert \lambda \in {\mathbb R}\}\), \({\mathbb C}X = {\mathbb R}X \oplus {\mathbb R}JX\), and \({\mathbb H}X = {\mathbb R}X \oplus {\mathfrak {J}}X\).

We identify the tangent space \(T_oSU_{2,m}/S(U_2{\cdot }U_m)\) of \(SU_{2,m}/S(U_2{\cdot }U_m)\) at o with \({\mathfrak {p}}\) in the usual way. Let \({\mathfrak {a}}\) be a maximal abelian subspace of \({\mathfrak {p}}\). Since \(SU_{2,m}/S(U_2{\cdot }U_m)\) has rank 2, the dimension of any such subspace is two. Every nonzero tangent vector \(X \in T_oSU_{2,m}/S(U_2{\cdot }U_m) \cong {\mathfrak {p}}\) is contained in some maximal abelian subspace of \({\mathfrak {p}}\). Generically this subspace is uniquely determined by X, in which case X is called regular. If there exists more than one maximal abelian subspaces of \({\mathfrak {p}}\) containing X, then X is called singular. There is a simple and useful characterization of the singular tangent vectors: A nonzero tangent vector \(X \in {\mathfrak {p}}\) is singular if and only if \(JX \in {\mathfrak {J}}X\) or \(JX \perp {\mathfrak {J}}X\).

Up to scaling there exists a unique \(SU_{2,m}\)-invariant Riemannian metric g on complex hyperbolic two-plane Grassmannians \(SU_{2,m}/S(U_{2}{\cdot }U_{m})\). Equipped with this metric \(SU_{2,m}/S(U_{2}{\cdot }U_{m})\) is a Riemannian symmetric space of rank 2 which is both Kähler and quaternionic Kähler. For computational reasons we normalize g such that the minimal sectional curvature of \(({SU_{2,m}/S(U_{2}{\cdot }U_{m})},g)\) is \(-4\). The sectional curvature K of the noncompact symmetric space \(SU_{2,m}/S(U_2{\cdot }U_m)\) equipped with the Killing metric g is bounded by \(-4{\le }K{\le }0\). The sectional curvature \(-4\) is obtained for all 2-planes \({\mathbb C}X\) when X is a non-zero vector with \(JX \in {\mathfrak {J}}X\).

When \(m=1\), \(G_2^{*}({\mathbb C}^3)=SU_{1,2}/S(U_1{\cdot }U_2)\) is isometric to the two-dimensional complex hyperbolic space \({\mathbb C}H^2\) with constant holomorphic sectional curvature \(-4\).

When \(m=2\), we note that the isomorphism \(SO(4,2)\simeq SU_{2,2}\) yields an isometry between \(G_2^{*}({\mathbb C}^4)=SU_{2,2}/S(U_2{\cdot }U_2)\) and the indefinite real Grassmann manifold \(G_2^{*}({\mathbb R}_2^6)\) of oriented two-dimensional linear subspaces of an indefinite Euclidean space \({\mathbb R}_2^6\). For this reason we assume \(m \ge 3\) from now on, although many of the subsequent results also hold for \(m = 1,2\).

Hereafter X,Y and Z always stand for any tangent vector fields on M.

The Riemannian curvature tensor \(\bar{R}\) of \(SU_{2,m}/S(U_{2}{\cdot }U_{m})\) is locally given by

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

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

3 Basic Formulas

In this section we derive some basic formulas and the Codazzi equation for a real hypersurface in \(G_2({\mathbb C}^{m+2})\) (or \(SU_{2,m}/S(U_{2}{\cdot }U_{m})\)) (see [3, 5, 7, 10,11,12, 18]).

Let M be a real hypersurface in \(G_2({\mathbb C}^{m+2})\) (or \(SU_{2,m}/S(U_{2}{\cdot }U_{m})\)). The induced Riemannian metric on M will also be denoted by g, and \(\nabla \) denotes the Riemannian connection of (Mg). 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 \(G_2({\mathbb C}^{m+2})\), where N denotes a unit normal vector field of M in \(G_2({\mathbb C}^{m+2})\). From the Kähler structure J of \(G_2({\mathbb C}^{m+2})\) (or \(SU_{2,m}/S(U_{2}{\cdot }U_{m})\)) 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 \(G_2({\mathbb C}^{m+2})\) (or \(SU_{2,m}/S(U_{2}{\cdot }U_{m})\)), 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:

$$\begin{aligned} \begin{aligned}&\phi _{\nu }^{2}X=-X+\eta _{\nu }(X)\xi _{\nu },\quad \eta _{\nu }(\xi _{\nu })=1, \quad \phi _{\nu } \xi _{\nu } =0, \\&{\phi }_{\nu +1}{\xi }_{\nu }=-{\xi }_{{\nu }+2},\quad {\phi }_{\nu }{\xi }_{{\nu }+1}={\xi }_{{\nu }+2},\\&{\phi }_{\nu }{\phi }_{{\nu }+1}X = {\phi }_{{\nu }+2}X+{\eta }_{{\nu }+1}(X){\xi }_{\nu },\\&{\phi }_{{\nu }+1}{\phi }_{\nu }X=-{\phi }_{{\nu }+2}X+{\eta }_{\nu }(X) {\xi }_{{\nu }+1} \end{aligned} \end{aligned}$$
(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 almost contact metric structures \((\phi , \xi , \eta , g)\) and \((\phi _{\nu }, \xi _{\nu }, \eta _{\nu }, g)\), \(\nu =1,2,3\), can be given by

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

On the other hand, from the parallelism of Kähler structure J, that is, \(\bar{\nabla } J=0\) and the quaternionic Kähler structure \(\mathfrak J\), 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)
$$\begin{aligned} \begin{aligned} ({\nabla }_X{\phi }_{\nu })Y&=-q_{{\nu }+1}(X){\phi }_{{\nu }+2}Y+q_{{\nu }+2}(X) {\phi }_{{\nu }+1}Y +{\eta }_{\nu }(Y)AX-g(AX,Y){\xi }_{\nu }. \end{aligned} \end{aligned}$$
(3.7)

Combining these formulas, we find the following:

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

Using the above expression (2.1) for the curvature tensor \(\bar{R}\) of \(G_2({\mathbb C}^{m+2})\) (or \(SU_{2,m}/S(U_{2}{\cdot }U_{m})\)), the equations of Codazzi is given by

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

where in the case of \(G_2({\mathbb C}^{m+2})\) (resp., \(SU_{2,m}/S(U_{2}{\cdot }U_{m})\)), the constant \(k=1\) and \(SU_{2,m}/S(U_{2}{\cdot }U_{m})\) (resp., \(k=-2\)).

4 Proof of Theorems

In this section, we classify real hypersurfaces in \(\bar{M}\) (\(G_2({\mathbb C}^{m+2})\) or \(SU_{2,m}/S(U_{2}{\cdot }U_{m})\)) whose shape operator has Killing tensor field.

From (C-1) and the Codazzi equation (3.9), we have

$$\begin{aligned} \begin{aligned} -2k(\nabla _YA)X&= \eta (X)\phi Y - \eta (Y)\phi X - 2g(\phi X,Y)\xi \\&\quad + \sum _{\nu =1}^3 \Big \{\eta _\nu (X)\phi _\nu Y - \eta _\nu (Y)\phi _\nu X - 2g(\phi _\nu X,Y)\xi _\nu \Big \} \\&\quad + \sum _{\nu =1}^3 \Big \{\eta _\nu (\phi X)\phi _\nu \phi Y - \eta _\nu (\phi Y)\phi _\nu \phi X\Big \} \\&\quad + \sum _{\nu =1}^3 \Big \{\eta (X)\eta _\nu (\phi Y) - \eta (Y)\eta _\nu (\phi X)\Big \}\xi _\nu \end{aligned} \end{aligned}$$
(4.1)

Putting \(Y=\xi \) into (4.1),

$$\begin{aligned} -2k(\nabla _{\xi } A)X=-\phi X+\sum _{\nu =1}^3\big \{{\eta }_{\nu }(X)\phi _{\nu }\xi -{\eta }_{\nu }(\xi )\phi _{\nu }X-3{\eta }_{\nu }(\phi X)\xi _{\nu }\big \}. \end{aligned}$$
(4.2)

Lemma 1

Let M be a real hypersurface in complex Grassmannians of rank two \(\bar{M}\), \(m \ge 3\) with Killing shape operator. Then the Reeb vector field \(\xi \) belongs to either the distribution \(\mathscr {Q}\) or the distribution \({\mathscr {Q}}^{\bot }\).

Proof

Without loss of generality, \(\xi \) is written as

$$\begin{aligned} \xi = \eta (X_{0})X_{0}+ \eta ({\xi }_1){\xi }_1, {(**)} \end{aligned}$$

where \(X_{0}\) (resp., \({\xi }_1\)) is a unit vector in \(\mathscr {Q}\) (resp., \({\mathscr {Q}}^{\bot }\)).

Taking the inner product of (4.2) with \(\xi \), we have

$$\begin{aligned} -2kg\big ((\nabla _{\xi } A)X,\xi \big )=-4\sum _{\nu =1}^3{\eta }_{\nu }(\xi ){\eta }_{\nu }(\phi X). \end{aligned}$$
(4.3)

Since \((\nabla _{\xi } A)\) is self-adjoint, it follows from (C-1) that \(-4\sum _{\nu =1}^3{\eta }_{\nu }(\xi ){\eta }_{\nu }(\phi X)=0\). By putting \(X=\phi X_{0}\) and using (**), we have \(-4\eta _1^2(\xi )\eta (X_{0})=0\).

Thus we have only two cases: \(\xi \in {\mathscr {Q}}^{\bot }\) or \(\xi \in \mathscr {Q}\).

  • Case 1. \(\xi \in {\mathscr {Q}}^{\bot }\).

Without loss of generality, we may put \(\xi ={\xi }_1\in {\mathscr {Q}}^{\bot }\). Then (4.2) is reduced into

$$\begin{aligned} -2k(\nabla _{\xi } A)X=-\phi X-\phi _{1}X+2\eta _{3}(X)\xi _{2}-2\eta _{2}(X)\xi _{3}. \end{aligned}$$
(4.4)

The symmetric endomorphism of (4.4) with respect to the metric g, we have

$$\begin{aligned} -2k(\nabla _{\xi } A)X=\phi X+\phi _{1}X-2\eta _{3}(X)\xi _{2}+2\eta _{2}(X)\xi _{3}. \end{aligned}$$
(4.5)

Combining (4.4) with (4.5), we have \(\phi X+\phi _{1}X-2\eta _{3}(X)\xi _{2}+2\eta _{2}(X)\xi _{3}=0\). By putting \(X={\xi }_3\) into the equation above, we have \(2{\xi }_3=0\). This is a contradiction.

Thus, there does not exist any hypersurface in \(\bar{M}\), \(m\ge 3\), with Killing shape operator and \(\xi \in {\mathscr {Q}}^{\bot }\) everywhere.

  • Case 2. \(\xi \in \mathscr {Q}\).

Equation (4.2) becomes

$$\begin{aligned} -2k(\nabla _{\xi } A)X=-\phi X+\sum _{\nu =1}^3\big \{{\eta }_{\nu }(X)\phi _{\nu }\xi -3{\eta }_{\nu }(\phi X)\xi _{\nu }\big \}. \end{aligned}$$
(4.6)

The symmetric endomorphism of (4.6) with respect to the metric g, we have

$$\begin{aligned} -2k(\nabla _{\xi } A)X=\phi X+\sum _{\nu =1}^3\big \{-{\eta }_{\nu }(\phi X)\xi _{\nu }+3{\eta }_{\nu }(X)\phi \xi _{\nu }\big \}. \end{aligned}$$
(4.7)

Combining (4.6) with (4.7), we have \(2\phi X+2\sum _{\nu =1}^3\big \{{\eta }_{\nu }(X)\phi \xi _{\nu }+{\eta }_{\nu }(\phi X)\xi _{\nu }\big \}=0\). By putting \(X={\xi }_1\) into above equation, we have \(4\phi {\xi }_1=0\). This is a contradiction, too. Thus, there does not exist any hypersurface in \(\bar{M}\), \(m\ge 3\), with Killing shape operator and \(\xi \in \mathscr {Q}\) everywhere.

Accordingly, we complete the proof of Theorem 1 in the introduction.

Usually, the notion of parallel is stronger than the notion of Killing, we also have a non-existence of parallel hypersurface in \(SU_{2,m}/S(U_{2}{\cdot }U_{m})\), \(m\ge 3\). Then Corollary 1 in the introduction is naturally proved.