1 Introduction

Let \(x: M^n\rightarrow {\mathbb {R}}^{n+p}\) be an n-dimensional submanifold in the \((n+p)\)-dimensional Euclidean space \({\mathbb {R}}^{n+p}\). Then x is called a self-shrinker (to the mean curvature flow) in \({\mathbb {R}}^{n+p}\) if its mean curvature vector field H satisfies

$$\begin{aligned} H+x^{\bot }=0, \end{aligned}$$
(1.1)

where \(x^\bot \) is the orthogonal projection of the position vector x to the normal space \(T^\bot M^n\) of x.

It is well known that the self-shrinker plays an important role in the study of the mean curvature flow. Not only self-shrinkers correspond to self-shrinking solutions to the mean curvature flow, but also they describe all possible Type I blow ups at a given singularity of the flow. Up to now, there have been a plenty of research papers on self-shrinkers among which are many that provide various results of classification or rigidity theorems. In particular, there are also interesting results about the Lagrangian self-shrinkers in the complex Euclidean n-space \({\mathbb C}^n\). For example, in [1], Anciaux gives new examples of self-shrinking and self-expanding Lagrangian solutions to the mean curvature flow. In [3], the authors classify all Hamiltonian stationary Lagrangian surfaces in the complex plane \({\mathbb C}^2\), which are self-similar solutions of the mean curvature flow and, in [4], several rigidity results for Lagrangian mean curvature flow are obtained. As we know, a canonical example of the compact Lagrangian self-shrinker in \({\mathbb {C}}^2\) is the Clifford torus \({\mathbb {S}}^{1}(1)\times {\mathbb {S}}^{1}(1)\).

Recently in [13], Li and Wang prove a rigidity theorem which improves a previous theorem by Castro and Lerma [4].

Theorem 1.1

(cf. [4, 13]). Let \(x:M^2\rightarrow {\mathbb C}^2\) be a compact oriented Lagrangian self-shrinker with h its second fundamental form. If \(|h|^{2}\le 2\), then \(|h|^{2}=2\) and \(x(M^{2})\) is the Clifford torus \({\mathbb {S}}^{1}(1)\times {\mathbb {S}}^{1}(1)\), up to a holomorphic isometry on \({\mathbb C}^2\).

Remark 1.1

Castro and Lerma also proved Theorem 1.1 in [4] under the additional condition that the Gauss curvature K of \(M^{2}\) is either non-negative or non-positive.

To make an extension of hypersurface self-shrinkers, Cheng and Wei recently introduce in [7] the definition of \(\lambda \)-hypersurface of weighted volume-preserving mean curvature flow in Euclidean space, and classify complete \(\lambda \)-hypersurfaces with polynomial area growth and \(H-\lambda \ge 0\), which are generalizations of the results due to Huisken [12] and Colding-Minicozzi [9]. According to [7], a hypersurface \(x: M^{n}\rightarrow {\mathbb {R}}^{n+1}\) is called a \(\lambda \)-hypersurface if its mean curvature \(H_0\) satisfies

$$\begin{aligned} H_{0}+\langle x, N\rangle =\lambda \end{aligned}$$
(1.2)

for some constant \(\lambda \), where N is the unit normal vector of x. Some rigidity or classification results for \(\lambda \)-hypersurfaces are obtained, for example, in [6, 8, 11]; for the rigidity theorems for space-like \(\lambda \)-hypersurfaces see [15].

As a natural generalization of both self-shrinkers and \(\lambda \)-hypersurfaces, we introduce the concept of \(\xi \)-submanifolds. Precisely, an immersed submanifold \(x: M^{n}\rightarrow {\mathbb {R}}^{n+p}\) is called a \(\xi \)-submanifold if there is a parallel normal vector field \(\xi \) such that the mean curvature vector field H satisfies

$$\begin{aligned} H+x^{\bot }=\xi . \end{aligned}$$
(1.3)

Obviously, the Clifford tori \({\mathbb {S}}^{1}(a)\times {\mathbb {S}}^{1}(b)\) with positive numbers a and b are examples of Lagrangian \(\xi \)-submanifold in \({\mathbb C}^2\). Similar examples in higher dimensions can be listed as those in [5] for self-shrinkers. In this paper, we focus on the rigidity of compact Lagrangian \(\xi \)-submanifolds in \({\mathbb {C}}^2\), and our main theorem is as follows:

Theorem 1.2

Let \(x: M^{2}\rightarrow {\mathbb {C}}^{2}\) be a compact oriented Lagrangian \(\xi \)-submanifold with the second fundamental form h and mean curvature vector H. Assume that

$$\begin{aligned} |h|^{2}+|H-\xi |^{2}\le |\xi |^{2}+4. \end{aligned}$$

Then \(|h|^{2}+|H-\xi |^{2}\equiv |\xi |^{2}+4\) and \(x(M^2)=T^2\) is a topological torus.

Furthermore, if \(\langle H, \xi \rangle \) is constant and one of the following four conditions holds:

$$\begin{aligned} (1)\ |h|^{2}\ge 2,\quad (2)\ |H|^{2}\ge 2,\quad (3)\ |h|^{2}\ge \langle H, H-\xi \rangle ,\quad (4)\ \langle H, \xi \rangle \ge 0, \end{aligned}$$
(1.4)

then, up to a holomorphic isometry on \({\mathbb {C}}^{2}\), \(x(M^{2})={\mathbb {S}}^{1}(a)\times {\mathbb {S}}^{1}(b)\) is a standard torus, where a and b are positive numbers satisfying \(a^{2}+b^{2}\ge 2a^{2}b^{2}\).

Corollary 1.3

Let \(x: M^{2}\rightarrow {\mathbb {C}}^{2}\) be a compact oriented Lagrangian self-shrinker. If

$$\begin{aligned} |h|^{2}+|H|^{2}\le 4, \end{aligned}$$

then \(|h|^{2}+|H|^{2}\equiv 4\) and \(x(M^{2})={\mathbb {S}}^{1}(1)\times {\mathbb {S}}^{1}(1)\) up to a holomorphic isometry on \({\mathbb {C}}^{2}\).

Clearly, Corollary 1.3 can be viewed as a different new version of Theorem 1.1.

Remark 1.2

We believe that the last condition (1.4) in Theorem 1.2 can be removed. On the other hand, the condition that \(\langle H,\xi \rangle \) is constant may also be removed. In fact, as suggested by the referee, we can use (3.11) and the compactness of \(M^2\) to show that \(|x|^2\) is constant when either \(\langle H, H-\xi \rangle \le 2\) or \(\langle H, H-\xi \rangle \ge 2\). Then by the argument at the end of the paper, we can simplify Theorem 1.2 as follows:

Theorem 1.4

Let \(x: M^{2}\rightarrow {\mathbb {C}}^{2}\) be a compact oriented Lagrangian \(\xi \)-submanifold with the second fundamental form h and mean curvature vector H. Assume that

$$\begin{aligned} |h|^{2}+|H-\xi |^{2}\le |\xi |^{2}+4. \end{aligned}$$

Then \(|h|^{2}+|H-\xi |^{2}\equiv |\xi |^{2}+4\) and \(x(M^2)=T^2\) is a topological torus.

Furthermore, if either \(\langle H, H-\xi \rangle \le 2\) or \(\langle H, H-\xi \rangle \ge 2\), then, up to a holomorphic isometry on \({\mathbb {C}}^{2}\), \(x(M^{2})={\mathbb {S}}^{1}(a)\times {\mathbb {S}}^{1}(b)\) is a standard torus for some \(a,b>0\).

Remark 1.3

Cheng and Wei have introduced in [7] a weighted area functional \({\mathcal {A}}\) and derived a related variation formula. Besides the relation between \(\lambda \)-hypersurfaces and the weighted volume preserving mean curvature flow, they also prove that \(\lambda \)-hypersurfaces are the critical points of the weighted area functional. Based on this, we believe that similar conclusions will be valid for the \(\xi \)-submanifolds defined above. Furthermore, We reasonably believe that, if self-shrinkers and \(\lambda \)-hypersurfaces take the places of minimal submanifolds and constant mean curvature hypersurfaces, respectively, then \(\xi \)-submanifolds must take the place of submanifolds of parallel mean curvature vector.

2 Lagrangian submanifolds in \({\mathbb C}^n\) and their Maslov class

Let \({\mathbb {C}}^{n}\) be the complex Euclidean n-space with the canonical complex structure J. Through out this paper, \(x:M^{n}\rightarrow {\mathbb C}^n\) always denotes an n-dimensional Lagrangian submanifold, and \(\nabla \), D, \(\nabla ^{\bot }\) denote, respectively, the Levi-Civita connections on \(M^{n}\), \({\mathbb {C}}^{n}\), and the normal connection on the normal boundle \(T^\bot M^n\). The formulas of Gauss and Weingarten are given by

$$\begin{aligned} D_{X}Y=\nabla _{X}Y+h(X, Y),\quad D_{X}\eta =-A_{\eta }X+\nabla ^{\bot }_{X}\eta , \end{aligned}$$

where X, Y are tangent vector fields on \(M^n\) and \(\eta \) is a normal vector field of x. The Lagrangian condition implies that

$$\begin{aligned} \nabla ^{\bot }_{X}JY=J\nabla _{X}Y, \quad A_{JX}Y=-Jh(X,Y)=A_{JY}X, \end{aligned}$$

where h and A are the second fundamental form and the shape operator of x, respectively. In particular, \(\langle h(X, Y), JZ \rangle \) is totally symmetric as a 3-form, namely

$$\begin{aligned} \langle h(X, Y), JZ\rangle = \langle h(X, Z), JY\rangle = \langle h(Y, Z), JX\rangle . \end{aligned}$$
(2.1)

From now on, we agree with the following convention on the ranges of indices:

$$\begin{aligned} 1\le i,j,\cdots \le n,\quad n+1\le \alpha ,\beta ,\cdots \le 2n,\quad 1\le A,B,\cdots \le 2n,\quad i^*=n+i. \end{aligned}$$

For a Lagrangian submanifold \(x:M^{n}\rightarrow {\mathbb {C}}^{n}\), there are orthonormal frame fields of the form \(\{e_i,e_{i^*}\}\) for \({\mathbb C}^n\) along x, where \(e_i\in TM^n\) and \(e_{i^*}=Je_i\). Such a frame is called an adapted Lagrangian frame field in the literature. The dual frame field is always denoted by \(\{\theta _i,\theta _{i^*}\}\), where \(\theta _{i^*}=-J\theta _i\). Write

$$\begin{aligned} h=\sum h^{k^*}_{ij}\theta _i\theta _je_{k^*},\quad \text{ where } h^{k^*}_{ij}=\langle h(e_i,e_j),e_{k^*}\rangle , \end{aligned}$$

or equivalently,

$$\begin{aligned} h(e_{i}, e_{j})=\sum _{k}h_{ij}^{k^{*}}e_{k^{*}},\quad \text{ for } \text{ all } e_i,e_j. \end{aligned}$$

Then (2.1) is equivalent to

$$\begin{aligned} h_{ij}^{k^{*}}=h_{kj}^{i^{*}}=h_{ik}^{j^{*}}, \quad 1\le i, j, k\le n. \end{aligned}$$
(2.2)

If \(\theta _{ij}\) and \(\theta _{i^*j^*}\) denote the connection forms of \(\nabla \) and \(\nabla ^\bot \), respectively, then the components \(h_{ij,l}^{k^{*}}\), \(h_{ij,lp}^{k^{*}}\) of the covariant derivatives of h are given respectively by

$$\begin{aligned} \sum _{l}h_{ij,l}^{k^{*}}\theta _{l}= & {} dh_{ij}^{k^{*}}+\sum _{l}h_{lj}^{k^{*}}\theta _{li} +\sum _{l}h_{il}^{k^{*}}\theta _{lj}+\sum _{m}h_{ij}^{m^{*}}\theta _{m^{*}k^{*}}; \end{aligned}$$
(2.3)
$$\begin{aligned} \sum _{p}h_{ij,lp}^{k^{*}}\theta _{p}= & {} dh_{ij,l}^{k^{*}}+\sum _{p}h_{pj,l}^{k^{*}}\theta _{pi} +\sum _{p}h_{ip,l}^{k^{*}}\theta _{pj}+\sum _{p}h_{ij,p}^{k^{*}}\theta _{pl} +\sum _{p}h_{ij,l}^{p^{*}}\theta _{p^{*}k^{*}}.\quad \qquad \end{aligned}$$
(2.4)

Moreover, the equations of motion are as follows:

$$\begin{aligned} dx= & {} \sum _{i}\theta _{i}e_{i},\quad de_{i}=\sum _{j}\theta _{ij}e_{j}+\sum _{k,j}h_{ij}^{k^{*}}\theta _{j}e_{k^{*}},\end{aligned}$$
(2.5)
$$\begin{aligned} de_{k^{*}}= & {} -\sum _{i,j}h_{ij}^{k^{*}}\theta _{j}e_{i}+\sum _{l}\theta _{k^{*}l^{*}}e_{l^{*}}. \end{aligned}$$
(2.6)

Let \(R_{ijkl}\) and \(R_{i^{*}j^{*}kl}\) denote the components of curvature operators of \(\nabla \) and \(\nabla ^{\bot }\), respectively. Then the equations of Gauss, Codazzi and Ricci are as follows:

$$\begin{aligned} R_{mijk}= & {} \sum _{l}(h_{mk}^{l^{*}}h_{ij}^{l^{*}}-h_{mj}^{l^{*}}h_{ik}^{l^{*}}), \quad 1\le m,i, j, k\le n, \end{aligned}$$
(2.7)
$$\begin{aligned} h_{ij,l}^{k^{*}}= & {} h_{il,j}^{k^{*}}, \quad 1\le i, j, k, l\le n, \end{aligned}$$
(2.8)
$$\begin{aligned} R_{i^{*}j^{*}kl}= & {} \sum _{m}(h_{ml}^{i^{*}}h_{mk}^{j^{*}}-h_{mk}^{i^{*}}h_{ml}^{j^{*}}), \quad 1\le i, j, k, l\le n. \end{aligned}$$
(2.9)

The scalar curvature of \(\nabla \) is

$$\begin{aligned} R=|H|^{2}-|h|^{2}\quad \text{ with } |H|^{2}=\sum _{k}\left( \sum _{i}h^{k^*}_{ii}\right) ^2,\quad |h|^{2}=\sum _{i,j,k}(h_{ij}^{k^{*}})^{2}, \end{aligned}$$
(2.10)

where the mean curvature vector field H is defined by

$$\begin{aligned} H=\sum _{k}H^{k^{*}}e_{k^{*}}=\sum _{i,k} h_{ii}^{k^{*}}e_{k^{*}}. \end{aligned}$$

Combining (2.2) and (2.8), we know that \(h_{ij,l}^{k^{*}}\) is totally symmetric, namely

$$\begin{aligned} h_{ij,l}^{k^{*}}=h_{jl,k}^{i^{*}}=h_{lk,i}^{j^{*}}=h_{ki,j}^{l^{*}},\quad 1\le i,j,k,l\le n, \end{aligned}$$
(2.11)

and the Ricci identities are as follows:

$$\begin{aligned} h_{ij,lp}^{k^{*}}-h_{ij,pl}^{k^{*}}=\sum _{m}h_{mj}^{k^{*}}R_{imlp} +\sum _{m}h_{im}^{k^{*}}R_{jmlp}+\sum _{m}h_{ij}^{m^{*}}R_{k^{*}m^{*}lp}. \end{aligned}$$
(2.12)

Note that, with respect to the adapted Lagrangian frame \(\{e_i,e_{i^*}\}\), the connection forms \(\theta _{i^*j^*}=\theta _{ij}\). It follows that

$$\begin{aligned} R_{m^{*}i^{*}jk}=R_{mijk},\quad \forall m,i,j,k. \end{aligned}$$
(2.13)

Furthermore, the first and second derivatives \(H^{k^*}_{,i}\), \(H^{k^*}_{,ij}\) of the mean curvature vector field H are given as

$$\begin{aligned} H^{k^*}_{,i}=\sum _jh^{k^*}_{jj,i},\quad H^{k^*}_{,ij}=\sum _lh^{k^*}_{ll,ij}. \end{aligned}$$
(2.14)

For any smooth function f on \(M^{n}\), the covariant derivatives \(f_{,i}\), \(f_{,ij}\) of f, the Laplacian of f are respectively defined as follows:

$$\begin{aligned} df=\sum _{i}f_{,i}\theta _{i}, \quad \sum _{j}f_{,ij}\theta _{j}=df_{,i}-\sum _{j}f_{,j}\theta _{ij},\quad \triangle f=\sum _{i}f_{,ii}. \end{aligned}$$
(2.15)

Finally, we also need to introduce the Lagrangian angles, Maslov form and Maslov class of a Lagrangian submanifold in \({\mathbb C}^n\) which we shall make use of later.

Let \((z^1,\ldots ,z^n)\) be the standard complex coordinates on \({\mathbb C}^n\). Then \(\Omega =dz^1\wedge \cdots \wedge dz^n\) is a globally defined holomorphic volume form which is clearly parallel. For a Lagrangian submanifold \(x:M^n\rightarrow {\mathbb C}^n\), the Lagrangian angle of x is by definition a multi-valued function \(\beta :M^n\rightarrow {\mathbb {R}}/2\pi {\mathbb {Z}}\) given by

$$\begin{aligned} \Omega _M:=x^*\Omega =e^{\sqrt{-1}\beta }dV_M. \end{aligned}$$

As one knows, although the Lagrangian angle \(\beta \) can not be determined globally in general, its gradient \(\nabla \beta \) is clearly a well-defined vector field on \(M^n\), or the same, \(\alpha :=d\beta \) is a globally defined 1-form which is called the Maslov form of x. Clearly, \(\alpha \) is closed and thus represents a cohomology class \([\alpha ]\in H^1(M^n)\) called the Maslov class.

In [16], the author proved an important formula by which the mean curvature and the Lagrangian angle of a Lagrangian submanifold are linked to each other; A. Arsie has extended this result in [2] to Lagrangian submanifolds in a general Calabi-Yau manifold.

Theorem 2.1

([16]) Let \(x:M^n\rightarrow {\mathbb C}^n\) be a Lagrangian submanifold and J be the canonical complex structure of \({\mathbb C}^n\). Then the mean curvature vector H and the Lagrangian angle \(\beta \) meet the following formula:

$$\begin{aligned} x_*(\nabla \beta )=-JH. \end{aligned}$$
(2.16)

Corollary 2.2

([4, 17]) Let \(x:M^n\rightarrow {\mathbb C}^n\) be a compact and oriented Lagrangian self-shrinkers. Then the Maslov class \([\alpha ]\) can not be trivial. In particular, there does not exist any Lagrangian self-shrinker in \({\mathbb C}^n\) with the topology of a sphere.

Remark 2.1

For our use in this paper, it is necessary to show that Corollary 2.2 is still true if we replace the self-shrinker by a \(\xi \)-submanifold. Precisely, we need

Proposition 2.3

Let \(x: M^{n}\rightarrow {\mathbb {C}}^{n}\) be a Lagrangian \(\xi \)-submanifold. If M is compact and orientable, then \([\alpha ]\ne 0\); Consequently, there does not exist any Lagrangian \(\xi \)-submanifold in \({\mathbb C}^n\) with the topology of a sphere.

Proof

By the definition of a \(\xi \)-submanifold, we have \(x=x^{\top }+\xi -H\). By Gauss and Weingarten formulas it follows that, for any \(v\in TM^n\),

$$\begin{aligned} A_{H}v= & {} -D_{v}H+\nabla ^{\bot }_{v}H =-D_{v}(\xi -x^{\bot })+\nabla ^{\bot }_{v}H\\= & {} D_{v}x^{\bot }-D_{v}\xi +\nabla ^{\bot }_{v}H =D_{v}x-D_{v}x^{\top }-D_{v}\xi +\nabla ^{\bot }_{v}H\\= & {} v-\nabla _{v}x^{\top }+A_\xi (v)-h(v,x^{\top })+\nabla ^{\bot }_{v}H, \end{aligned}$$

where \(A_{H}\) and \(A_\xi \) are Weingarten transformations with respect to H and \(\xi \), respectively. Thus

$$\begin{aligned} A_{H}v=v-\nabla _{v}x^{\top }+A_\xi (v),\quad \nabla _{v}^{\bot }H=h(v,x^{\top }). \end{aligned}$$

So that

$$\begin{aligned} \mathrm{div\,}JH= & {} \sum _{i}\langle \nabla _{e_{i}}JH,e_{i}\rangle =\sum _{i}\langle J\nabla _{e_{i}}JH, Je_{i} \rangle =\sum _{i}\langle -\nabla _{e_{i}}^{\bot }H, Je_{i}\rangle \nonumber \\= & {} \sum _{i}\langle -h(e_{i}, x^{\top }), Je_{i}\rangle =\sum _{i}-\langle h(e_{i},e_{i}), Jx^{\top }\rangle \nonumber \\= & {} \sum _{i}\langle Jh(e_{i},e_{i}), x^{\top }\rangle =\langle JH, x^{\top }\rangle , \end{aligned}$$
(2.17)

where \(\mathrm{div\,}\) is the divergence operator. By (2.16) and (2.17) we obtain

$$\begin{aligned} \Delta \beta =\langle \nabla \beta ,x^\top \rangle =\frac{1}{2}\langle \nabla \beta , \nabla |x|^{2}\rangle . \end{aligned}$$
(2.18)

If \([\alpha ]= 0\), then there exists a globally defined Lagrangian angle \(\beta \) such that \(\alpha =-d\beta \), implying (2.18) holds globally on \(M^n\). Then the compactness assumption and the maximum principle for a second linear elliptic partial equation (see [10], for example) assure that \(\beta \) must be constant. Hence \(H=x_*(J\nabla \beta )\equiv 0\), contradicting to the fact that there are no compact minimal submanifolds in Euclidean space. This contradiction proves that \([\alpha ]\ne 0\).

Since the first homology of a sphere \(S^n\) vanishes for \(n>1\), there can not be any Lagrangian \(\xi \)-submanifolds with the topology of a sphere. \(\square \)

3 Proof of the main theorem

Let \(x: M^n\rightarrow {\mathbb {C}}^n\) be a Lagrangian \(\xi \)-submanifold without boundary. Then, with respect to an orthonormal frame field \(\{e_i\}\), the defining equation (1.3) is equivalent to

$$\begin{aligned} H^{k^{*}}=-\langle x, e_{k^{*}}\rangle +\xi ^{k^{*}}, \quad 1 \le k \le n. \end{aligned}$$
(3.1)

where \(\xi =\sum \xi ^{k^*}e_{k^*}\) is a given parallel normal vector field. From now on, we always assume that \(n=2\) if no other specification is given.

We start with a well-known operator \({\mathcal {L}}\) acting on smooth functions defined by

$$\begin{aligned} {\mathcal {L}}=\triangle -\langle x, \nabla \cdot \rangle =e^{\frac{|x|^{2}}{2}}\mathrm{div\,}(e^{-\frac{|x|^{2}}{2}}\nabla \cdot ), \end{aligned}$$
(3.2)

which was first introduced by Colding and Minicozzi [9] to the study of self-shrinkers. Since then, the operator \({\mathcal {L}}\) has been one of the most effect tools adapted by many authors. In particular, the following is a fundamental lemma related to \({\mathcal {L}}\):

Lemma 3.1

([14]) Let \(x: M^n\rightarrow {\mathbb {R}}^{n+p}\) be a complete immersed submanifold. If u and v are \(C^2\)-smooth functions with

$$\begin{aligned} \int _M(|u\nabla v|+|\nabla u||\nabla v|+|u{\mathcal {L}} v|)e^{-\frac{|x|^2}{2}}dV_M< \infty , \end{aligned}$$

then it holds that

$$\begin{aligned} \int _M u{\mathcal {L}}v e^{-\frac{|x|^2}{2}}dV_M=-\int _M\langle \nabla u, \nabla v\rangle e^{-\frac{|x|^2}{2}}dV_M. \end{aligned}$$

Now, to make the whole argument more readable, we divide our proof into the following lemmas and propositions:

Lemma 3.2

(cf. [13]) Let \(x: M^{2}\rightarrow {\mathbb {C}}^{2}\) be a Lagrangian \(\xi \)-submanifold. Then

$$\begin{aligned} H_{,i}^{k^{*}}= & {} \sum _{j}h_{ij}^{k^{*}}\langle x, e_{j}\rangle , \quad 1\le i,k\le 2, \end{aligned}$$
(3.3)
$$\begin{aligned} H_{,ij}^{k^{*}}= & {} \sum _{m}h_{im,j}^{k^{*}}\langle x, e_{m}\rangle +h_{ij}^{k^{*}}-\sum _{m,p}(H-\xi )^{p^{*}}h_{im}^{k^{*}}h_{mj}^{p^{*}},\quad 1\le i,j,k\le 2. \end{aligned}$$
(3.4)

Lemma 3.3

It holds that

$$\begin{aligned} \frac{1}{2}{\mathcal {L}}(|h|^{2}+|H-\xi |^{2})= & {} |\nabla h|^{2}+|\nabla ^{\bot }H|^{2}+|h|^{2}\nonumber \\&-\,\frac{1}{2}(|h|^{2}-|H|^{2})(3|h|^{2}-2|H|^{2}+\langle H, H-\xi \rangle )\nonumber \\&+\,\langle H,H-\xi \rangle -\sum _{i,j,k,l}h_{ij}^{k^{*}}h_{ij}^{l^{*}}(H-\xi )^{k^{*}}(H-\xi )^{l^{*}}\nonumber \\&-\sum _{i,j,k,l}h_{ij}^{k^{*}}h_{ij}^{l^{*}}H^{k^{*}}(H-\xi )^{l^{*}}. \end{aligned}$$
(3.5)

Proof

By a direct computation using Lemma 3.2 we find (cf. [13])

$$\begin{aligned} \frac{1}{2}{\mathcal {L}}|h|^{2}= & {} |\nabla h|^{2}+|h|^{2}-\frac{3}{2}|h|^{4} +\frac{5}{2}|H|^{2}|h|^{2}-|H|^{4}\nonumber \\&+\frac{1}{2}\langle H, H-\xi \rangle (|H|^{2}-|h|^{2})- \sum _{i,j,k,l}H^{k^{*}}h_{ij}^{k^{*}}h_{ij}^{l^{*}}(H-\xi )^{l^{*}}; \end{aligned}$$
(3.6)
$$\begin{aligned} \frac{1}{2}{\mathcal {L}}(|H-\xi |^{2})= & {} \frac{1}{2}\triangle (|H-\xi |^{2}) -\frac{1}{2}\langle x, \nabla |H-\xi |^{2}\rangle \nonumber \\= & {} \sum _{i,k}(H-\xi )^{k^{*}}H_{,ii}^{k^{*}}+|\nabla ^{\bot }H|^{2} -\sum _{i.k}(H-\xi )^{k^{*}}H_{,i}^{k^{*}}\langle x, e_{i}\rangle \nonumber \\= & {} \langle H- \xi , H\rangle +|\nabla ^{\bot }H|^{2} -\sum _{i,j,k,l}(H-\xi )^{k^{*}}h_{ij}^{k^{*}}h_{ij}^{l^{*}}(H-\xi )^{l^{*}}. \end{aligned}$$
(3.7)

By taking the sum we obtain (3.5). \(\square \)

Lemma 3.4

It holds that

$$\begin{aligned} \frac{1}{2}\triangle (|x^{\top }|^{2})= & {} \sum _{i,j,k}h_{ij}^{k^{*}}\langle x,e_{i}\rangle \langle x, e_{j}\rangle (\xi -H)^{k^{*}}-\sum _{i,j,k,l}h_{il}^{k^{*}}h_{lj}^{k^{*}} \langle x, e_{i}\rangle \langle x, e_{j}\rangle \nonumber \\&+\,2-2\langle H,H-\xi \rangle +\sum _{i,j,k,l}h_{ij}^{k^{*}}h_{ij}^{l^{*}}(H-\xi )^{k^{*}}(H-\xi )^{l^{*}}. \end{aligned}$$
(3.8)

Proof

We find

$$\begin{aligned} \frac{1}{2}\triangle (|x^{\top }|^{2})= & {} \frac{1}{2}\sum _{i,j}\langle x, e_{j}\rangle ^{2}_{,ii}=\sum _{i,j}(\langle x, e_{j}\rangle \langle x, e_{j}\rangle _{i})_{,i}\\= & {} \sum _{i,j}\left( \langle x, e_{j}\rangle \langle x_{i}, e_{j}\rangle \right) _{,i}+\sum _{i,j,k}\left( \langle x, e_{j}\rangle \langle x, h_{ji}^{k^{*}}e_{k^{*}}\rangle \right) _{,i}\\= & {} 2+2\sum _{i,k}h_{ii}^{k^{*}}\langle x, e_{k^{*}}\rangle +\sum _{j,k}H_{,j}^{k^{*}}\langle x, e_{j}\rangle \langle x, e_{k^{*}}\rangle \\&+\sum _{i,j,k,l}h_{ij}^{k^{*}}h_{ij}^{l^{*}}\langle x, e_{l^{*}}\rangle \langle x, e_{k^{*}}\rangle -\sum _{i,j,k,l}h_{ij}^{k^{*}}h_{il}^{k^{*}}\langle x, e_{j}\rangle \langle x, e_{l}\rangle \\= & {} 2-2\langle H,H-\xi \rangle +\sum _{i,j,k}h_{ij}^{k^{*}}\langle x, e_{i}\rangle \langle x, e_{j}\rangle (\xi -H)^{k^{*}}\\&+\sum _{i,j,k,l}h_{ij}^{k^{*}}h_{ij}^{l^{*}}(H-\xi )^{l^{*}}(H-\xi )^{k^{*}} -\sum _{i,j,k,l}h_{il}^{k^{*}}h_{lj}^{k^{*}}\langle x, e_{i}\rangle \langle x, e_{j}\rangle , \end{aligned}$$

and the lemma is proved. \(\square \)

Lemma 3.5

It holds that

$$\begin{aligned} \triangle (\langle H, \xi \rangle )= & {} \sum _{i,j,k}h_{ij}^{k^{*}}\langle x, e_{i}\rangle \langle x, e_{j}\rangle \xi ^{k^{*}}+\langle H, \xi \rangle -\sum _{i,j,k,l}h_{ij}^{k^{*}}h_{ij}^{l^{*}}\xi ^{k^{*}}(H-\xi )^{l^{*}},\end{aligned}$$
(3.9)
$$\begin{aligned} {\mathcal {L}}(\langle H, \xi \rangle )= & {} \langle H, \xi \rangle -\sum _{i,j,k,l}h_{ij}^{k^{*}}h_{ij}^{l^{*}}\xi ^{k^{*}}(H-\xi )^{l^{*}}. \end{aligned}$$
(3.10)

Proof

By (3.3) and (3.4),

$$\begin{aligned} \triangle (\langle H, \xi \rangle )= & {} \sum _{i,k}(H^{k^{*}}\xi ^{k^{*}})_{,ii}=\sum _{i,k}H^{k^{*}}_{,ii}\xi ^{k^{*}}\\= & {} \sum _{i,k,l,m}(h_{im,i}^{k^{*}}\langle x, e_{m}\rangle +h_{ii}^{k^{*}}-(H-\xi )^{^{l^{*}}}h_{im}^{k^{*}}h_{mi}^{l^{*}})\xi ^{k^{*}}\\= & {} \sum _{i,k}H_{,i}^{k^{*}}\langle x, e_{i}\rangle \xi ^{k^{*}}+\langle H, \xi \rangle - \sum _{i,j,k,l}h_{ij}^{k^{*}}h_{ij}^{l^{*}}\xi ^{k^{*}}(H-\xi )^{l^{*}}\\= & {} \sum _{i,j,k}h_{ij}^{k^{*}}\langle x, e_{i}\rangle \langle x, e_{j}\rangle \xi ^{k^{*}}+\langle H, \xi \rangle -\sum _{i,j,k,l}h_{ij}^{k^{*}}h_{ij}^{l^{*}}\xi ^{k^{*}}(H-\xi )^{l^{*}};\\ \langle x, \nabla \langle H, \xi \rangle \rangle= & {} \sum _{i}\langle H, \xi \rangle _{,i}\langle x, e_{i}\rangle =\sum _{i,j,k}h_{ij}^{k^{*}}\langle x, e_{i}\rangle \langle x, e_{j}\rangle \xi ^{k^{*}}. \end{aligned}$$

Thus, by adding them up, we get (3.10). \(\square \)

Lemma 3.6

(cf. [5, 9]; also [13]) It holds that

$$\begin{aligned} \frac{1}{2}\triangle (|x|^{2})= & {} 2-\langle H, H-\xi \rangle , \end{aligned}$$
(3.11)
$$\begin{aligned} \frac{1}{2}{\mathcal {L}}(|x|^{2})= & {} |\xi |^{2}+2-(|x|^{2}+\langle H, \xi \rangle ). \end{aligned}$$
(3.12)

Proof

From (3.1), we find

$$\begin{aligned} \frac{1}{2}\triangle (|x|^{2})= & {} 2+\langle x, \triangle x\rangle =2+\sum _{k}H^{k^{*}}\langle x, e_{k^{*}}\rangle =2-\langle H,H-\xi \rangle ,\\ \frac{1}{2}{\mathcal {L}}(|x|^{2})= & {} \frac{1}{2}\triangle (|x|^{2})-\frac{1}{2}\langle x, \nabla |x|^{2}\rangle =2-|H|^{2}+\langle H,\xi \rangle -|x^{\top }|^{2}\\= & {} 2+|\xi |^{2}-(|x|^{2}+\langle H, \xi \rangle ). \end{aligned}$$

\(\square \)

Proposition 3.7

Let \(M^2\) be oriented and compact. If

$$\begin{aligned} |h|^{2}+|H-\xi |^{2}\le |\xi |^{2}+4, \end{aligned}$$

then

$$\begin{aligned} |h|^{2}+|H-\xi |^{2}\equiv |\xi |^{2}+4 \end{aligned}$$
(3.13)

and \(x(M^{2})\) is a topological torus.

Proof

By Lemma 3.6,

$$\begin{aligned} \int _{M}|H-\xi |^{2}dV_{M}= & {} \int _{M}(|\xi |^{2}+2(|H|^{2}-\langle H, \xi \rangle )-|H|^{2})dV_{M}\nonumber \\= & {} \int _{M}(|\xi |^{2}+4-|H|^{2})dV_{M}. \end{aligned}$$
(3.14)

Let K be the Gauss curvature of \(M^2\). Then the Gauss equation gives that

$$\begin{aligned} 2K= |H|^{2}-|h|^{2}. \end{aligned}$$

Denote by \(\mathrm{gen}(M^2)\) the genus of \(M^2\). Then from the Gauss-Bonnet theorem and (3.14) it follows that

$$\begin{aligned} 8\pi (1-\mathrm{gen}(M^2))= & {} 2\int _{M}KdV_{M}=\int _{M}(|H|^{2}-|h|^{2})dV_{M} \nonumber \\= & {} \int _{M}(|\xi |^{2}+4-(|h|^{2}+|H-\xi |^{2}))dV_{M}\ge 0, \end{aligned}$$
(3.15)

implying that \({\mathrm{gen}}(M^2)\le 1\). So \(M^2\) is topologically either a 2-sphere or a torus. But Proposition 2.3 excludes the first possibility. So \({\mathrm{gen}}(M^2)=1\) and (3.13) is proved. \(\square \)

Lemma 3.8

Let \(p_{0}\in M^2\) be a point where \(|x|^{2}\) attains its minimum on \(M^2\). If \(M^2\) is orientable, compact and

$$\begin{aligned} |h|^{2}+|H-\xi |^{2}=\mathrm{const}, \end{aligned}$$

then

$$\begin{aligned} \nabla ^{\bot }H(p_{0})=0, \quad (\nabla h)(p_{0})=0. \end{aligned}$$
(3.16)

Proof

Since \((|x|^{2})_{,j}=0,1\le j\le 2\) at \(p_{0}\), it holds that \(\langle x, e_{j}\rangle (p_{0})=0\), \(1\le j\le 2\). So by (3.3) we have

$$\begin{aligned} H^{k^{*}}_{,i}=0, \ 1\le i,k \le 2,\quad |H-\xi |^{2}_{,i}=2\sum _{k}(H-\xi )^{k^{*}}H_{,i}^{k^{*}} =0, \quad 1\le i\le 2 \quad \text{ at } \ p_{0} \nonumber \\ \end{aligned}$$
(3.17)

where the first set of equalities are exactly \(\nabla ^\bot H(p_0)=0\), which give

$$\begin{aligned} h_{11,1}^{1^{*}}+h_{22,1}^{1^{*}}=0,\quad h_{11,2}^{1^{*}}+h_{22,2}^{1^{*}}=0,\quad h_{11,1}^{2^{*}}+h_{22,1}^{2^{*}}=0,\quad h_{11,2}^{2^{*}}+h_{22,2}^{2^{*}}=0.\nonumber \\ \end{aligned}$$
(3.18)

On the other hand, from

$$\begin{aligned} |h|^{2}+|H-\xi |^{2}=\mathrm{const}, \end{aligned}$$
(3.19)

we obtain

$$\begin{aligned} |h|^{2}_{,k}+|H-\xi |^{2}_{,k}\equiv 0,\quad 1\le k\le 2, \end{aligned}$$
(3.20)

which with (3.17) implies that

$$\begin{aligned} (|h|^{2})_{,k}=0,\quad 1\le k\le 2\quad \text {at } p_{0}. \end{aligned}$$

Since

$$\begin{aligned} |h|^{2}=(h_{11}^{1^{*}})^{2}+2(h_{12}^{1^{*}})^{2}+(h_{22}^{1^{*}})^{2} +(h_{11}^{2^{*}})^{2}+2(h_{12}^{2^{*}})^{2}+(h_{22}^{2^{*}})^{2}, \end{aligned}$$

we find that

$$\begin{aligned}&h_{11}^{1^{*}}h_{11,1}^{1^{*}}+2h_{12}^{1^{*}}h_{12,1}^{1^{*}}+h_{22}^{1^{*}}h_{22,1}^{1^{*}} +h_{11}^{2^{*}}h_{11,1}^{2^{*}}+2h_{12}^{2^{*}}h_{12,1}^{2^{*}} +h_{22}^{2^{*}}h_{22,1}^{2^{*}}=0, \end{aligned}$$
(3.21)
$$\begin{aligned}&h_{11}^{1^{*}}h_{11,2}^{1^{*}}+2h_{12}^{1^{*}}h_{12,2}^{1^{*}}+h_{22}^{1^{*}}h_{22,2}^{1^{*}} +h_{11}^{2^{*}}h_{11,2}^{2^{*}}+2h_{12}^{2^{*}}h_{12,2}^{2^{*}} +h_{22}^{2^{*}}h_{22,2}^{2^{*}}=0 \end{aligned}$$
(3.22)

hold at \(p_{0}\). From (2.11) and (3.18) we get

$$\begin{aligned} h_{22,1}^{1^{*}}=-h_{11,1}^{1^{*}},\quad h_{22,2}^{1^{*}}=-h_{11,2}^{1^{*}},\quad h_{22,2}^{2^{*}}=h_{11,1}^{1^{*}}\quad \text {at }p_{0}. \end{aligned}$$
(3.23)

Since, by (2.2) and (2.11), both \(h_{ij}^{k^{*}}\) and \(h_{ij,l}^{k^{*}}\) are totally symmetric, we obtain by (3.23), (3.21) and (3.22) that

$$\begin{aligned}&(h_{11}^{1^{*}}-3h_{22}^{1^{*}})h_{11,1}^{1^{*}}- (h_{22}^{2^{*}}-3h_{11}^{2^{*}})h_{11,2}^{1^{*}}=0, \end{aligned}$$
(3.24)
$$\begin{aligned}&(h_{22}^{2^{*}}-3h_{11}^{2^{*}})h_{11,1}^{1^{*}}+ (h_{11}^{1^{*}}-3h_{22}^{1^{*}})h_{11,2}^{1^{*}}=0\quad \text {at }p_{0}. \end{aligned}$$
(3.25)

We claim that

$$\begin{aligned} (\nabla h)(p_{0})=0. \end{aligned}$$
(3.26)

Otherwise, we should have \((h_{11,1}^{1^{*}})^{2}+(h_{11,2}^{1^{*}})^{2}\ne 0\) at \(p_0\). Then from (3.24) and (3.25) it follows that

$$\begin{aligned} (h_{11}^{1^{*}}-3h_{22}^{1^{*}})^{2}+(h_{22}^{2^{*}}-3h_{11}^{2^{*}})^{2}=0\quad \text {at }p_{0}. \end{aligned}$$

Thus

$$\begin{aligned} |h|^{2}(p_{0})=\frac{4}{3}((h_{11}^{1^{*}})^{2}+(h_{22}^{2^{*}})^{2}),\quad |H|^{2}(p_{0})=\frac{16}{9}((h_{11}^{1^{*}})^{2}+(h_{22}^{2^{*}})^{2}). \end{aligned}$$
(3.27)

Now by the definition of \(p_0\) and Lemma 3.6,

$$\begin{aligned} 0\le \frac{1}{2}\triangle {|x|^{2}}(p_{0}) =2-\langle H, H-\xi \rangle (p_{0}). \end{aligned}$$

It follows that

$$\begin{aligned} |h|^{2}+|H-\xi |^{2}= & {} (|h|^{2}+|H-\xi |^{2})(p_{0})\nonumber \\= & {} \frac{3}{4}|H|^{2}(p_{0})+2\langle H, H-\xi \rangle (p_{0}) -|H|^{2}(p_{0})+|\xi |^{2}\nonumber \\= & {} |\xi |^{2}+2\langle H, H-\xi \rangle (p_{0})-\frac{1}{4}|H|^{2}(p_{0}) \end{aligned}$$
(3.28)
$$\begin{aligned}\le & {} |\xi |^{2}+4. \end{aligned}$$
(3.29)

Therefore, by Proposition 3.7, \(|h|^{2}+|H-\xi |^{2}=|\xi |^{2}+4\). But it is easy to see that the equality in (3.29) holds if and only if \(|H|^{2}(p_{0})=0\) and \(\langle H, H-\xi \rangle (p_{0})=2\), which is of course not possible! This contradiction proves the above claim and completes the proof of Lemma 3.8. \(\square \)

Remark 3.1

Our main observation here is that, if \(p_{0}\in M^{2}\) is a minimum point of \(|x|^{2}\) then

$$\begin{aligned} x^{\top }(p_{0})=\sum _{i}\langle x, e_{i}\rangle e_{i}(p_{0})=0, \end{aligned}$$

implying

$$\begin{aligned} \nabla ^{\bot }H(p_{0})=\nabla ^{\bot }(H-\xi )(p_{0})=0. \end{aligned}$$

In particular, \(p_{0}\) is also a minimum point of \(|x^{\top }|^{2}\).

Proposition 3.9

Let \(x: M^{2}\rightarrow {\mathbb {C}}^{2}\) be a compact and oriented Lagrangian \(\xi \)-submanifold. Suppose that

$$\begin{aligned} |h|^{2}+|H-\xi |^{2}=|\xi |^{2}+4 \end{aligned}$$

and \(\langle H, \xi \rangle \) is constant. If one of the followings holds,

$$\begin{aligned} (1)\ |h|^{2}\ge 2,\quad (2)\ |H|^{2}\ge 2,\quad (3)\ |h|^{2}\ge \langle H, H-\xi \rangle ,\quad (4)\ \langle H, \xi \rangle \ge 0, \end{aligned}$$
(3.30)

then \(|x|^{2}\) is a constant.

Proof

As above, let \(p_0\) be a minimum point of \(|x|^2\). Then, by Lemma 3.3 and Lemma 3.8, it holds at \(p_{0}\) that

$$\begin{aligned} 0= & {} \frac{1}{2}{\mathcal {L}}(|h|^{2}+|H-\xi |^{2})\nonumber \\= & {} |h|^{2}-\frac{1}{2}(|h|^{2}-|H|^{2})(3|h|^{2}-2|H|^{2} +\langle H,H-\xi \rangle )+\langle H,H-\xi \rangle \nonumber \\&-\sum _{i,j,k,l}h_{ij}^{k^{*}}h_{ij}^{l^{*}}(H-\xi )^{k^{*}}(H-\xi )^{l^{*}} -\sum _{i,j,k,l}h_{ij}^{k^{*}}h_{ij}^{l^{*}}H^{k^{*}}(H-\xi )^{l^{*}}. \end{aligned}$$
(3.31)

Furthermore, form Lemma 3.4 and Lemma 3.5 it follows that, at \(p_{0}\)

$$\begin{aligned} 0\le & {} \frac{1}{2}\triangle (|x^{\top }|^{2})=2-2\langle H,H-\xi \rangle + \sum _{i,j,k,l}h_{ij}^{k^{*}}h_{ij}^{l^{*}}(H-\xi )^{k^{*}}(H-\xi )^{l^{*}}, \end{aligned}$$
(3.32)
$$\begin{aligned} 0= & {} \frac{1}{2}{\mathcal {L}}(\langle H,\xi \rangle ) =\frac{1}{2}(\langle H,\xi \rangle - \sum _{i,j,k,l}h_{ij}^{k^{*}}h_{ij}^{l^{*}}\xi ^{k^{*}}(H-\xi )^{l^{*}}), \end{aligned}$$
(3.33)

implying

$$\begin{aligned} -\sum _{i,j,k,l}h_{ij}^{k^{*}}h_{ij}^{l^{*}}(H-\xi )^{k^{*}}(H-\xi )^{l^{*}} \le 2-2\langle H,H-\xi \rangle , \end{aligned}$$
(3.34)

and

$$\begin{aligned}&-\sum _{i,j,k,l}h_{ij}^{k^{*}}h_{ij}^{l^{*}}(H-\xi )^{l^{*}}H^{k^{*}}\\&\quad =-\sum _{i,j,k,l}h_{ij}^{k^{*}}h_{ij}^{l^{*}}(H-\xi )^{k^{*}}(H-\xi )^{l^{*}} -\sum _{i,j,k,l}h_{ij}^{k^{*}}h_{ij}^{l^{*}}\xi ^{k^{*}}(H-\xi )^{l^{*}} \\&\quad \le 2-2\langle H,H-\xi \rangle -\langle H,\xi \rangle =2-\langle H,H-\xi \rangle -|H|^{2}. \end{aligned}$$

Consequently, we have at \(p_{0}\)

$$\begin{aligned} 0= & {} \frac{1}{2}{\mathcal {L}}(|h|^{2}+|H-\xi |^{2})\\\le & {} -\frac{1}{2}(|h|^{2}-|H|^{2})(3|h|^{2}-2|H|^{2} +\langle H,H-\xi \rangle )\\&+\,|h|^{2}-|H|^{2}+2(2-\langle H,H-\xi \rangle ). \end{aligned}$$

On the other hand, from

$$\begin{aligned} |h|^{2}+|H-\xi |^{2}=|\xi |^{2}+4, \end{aligned}$$

we know that

$$\begin{aligned} |h|^{2}-|H|^{2}=2(2-\langle H,H-\xi \rangle )\ge 0\quad \text {at } p_{0}. \end{aligned}$$
(3.35)

Thus, if one of (3.30) holds, then at \(p_0\)

$$\begin{aligned} 0= & {} \frac{1}{2}{\mathcal {L}}(|h|^{2}+|H-\xi |^{2}) \\\le & {} -\frac{1}{2}(|h|^{2}-|H|^{2})(2|h|^{2}-|H|^{2}+4 -\langle H,H-\xi \rangle )+2(|h|^{2}-|H|^{2}) \\= & {} -\frac{1}{2}(|h|^{2}-|H|^{2})(2|h|^{2} -|H|^{2}-\langle H,H-\xi \rangle ) \\= & {} -\frac{1}{2}(|h|^{2}-|H|^{2})(|h|^{2}-|H|^{2}+|h|^{2} -\langle H,H-\xi \rangle ) \\= & {} -\frac{1}{2}(|h|^{2}-|H|^{2})(|h|^{2}-|H|^{2}+|h|^{2} -2+2-\langle H,H-\xi \rangle ) \\= & {} -\frac{1}{2}(|h|^{2}-|H|^{2})(2(|h|^{2}-|H|^{2}) +\langle H,\xi \rangle ) \\= & {} -\frac{1}{2}(|h|^{2}-|H|^{2})(2(|h|^{2}-|H|^{2})+|H|^{2} -2+2-\langle H,H-\xi \rangle ) \le 0. \end{aligned}$$

Consequently

$$\begin{aligned} |h|^{2}-|H|^{2}=2-\langle H,H-\xi \rangle =0\quad \text {at } p_{0}. \end{aligned}$$
(3.36)

It follows that

$$\begin{aligned} |x|^{2}+\langle H,\xi \rangle\ge & {} |x|^{2}(p_{0})+\langle H,\xi \rangle (p_{0})\\= & {} |H-\xi |^{2}(p_{0})+\langle H,\xi \rangle (p_{0})\\= & {} \langle H,H-\xi \rangle (p_{0})+|\xi |^{2}\\= & {} |\xi |^{2}+2. \end{aligned}$$

This together with Lemma 3.1 (for \(u=1\), \(v=|x|^2\)) and 3.6 gives that

$$\begin{aligned} 0=\int _{M}\frac{1}{2}{\mathcal {L}}(|x|^{2})e^{-\frac{|x|^{2}}{2}}dV_{M} =\int _{M}(|\xi |^{2}+2-(|x|^{2}+\langle H,\xi \rangle ) )e^{-\frac{|x|^{2}}{2}}dV_{M}\le 0 \end{aligned}$$

implying that \(|x|^{2}+\langle H,\xi \rangle =|\xi |^{2}+2\). In particular, \(|x|^{2}=\mathrm{const}\). \(\square \)

Proposition 3.10

Let \(x: M^{n}\rightarrow N^{n}\) be a Lagrangian submanifold in a \(K\ddot{a}hler\) manifold \(N^{n}\). If both \(M^{n}\) and \(N^{n}\) are flat, then around each point \(p\in M^{n}\), there exists some orthonormal frame field \(\{e_{i},e_{i^*}\}\) with \(e_{i^{*}}=Je_{i}\) (\(1\le i\le n\)), such that

$$\begin{aligned} h_{ij}^{k^{*}}:=\langle h(e_{i},e_{j}),e_{k^{*}}\rangle =\lambda _{i}^{k^{*}}\delta _{ij}, \quad 1\le i,j,k\le n. \end{aligned}$$

Proof

For \(p\in M^{n}\), we pick an orthonormal tangent frame \(\{\bar{e}_{i}\}\) and an orthonormal normal frame \(\{\bar{e}_{\alpha }\}_{n+1\le \alpha \le 2n}\). Define

$$\begin{aligned} \bar{h}_{ij}^{\alpha }=\langle h(\bar{e}_{i},\bar{e}_{j}),\bar{e}_{\alpha }\rangle . \end{aligned}$$

Since \(M^n\) is flat, x is Lagrangian and \(N^n\) is Kähler, \(T^{\bot }M^{n}\) is also flat with respect to the normal connection. By the Ricci equation and the flatness of \(N^n\),

$$\begin{aligned} 0=\langle R^{^{\bot }}(e_{i},e_{j})e_{\alpha },e_{\beta }\rangle =\sum _k\left( h_{ik}^{\beta }h_{jk}^{\alpha }-h_{ik}^{\alpha }h_{jk}^{\beta }\right) . \end{aligned}$$

Hence we can choose another orthonormal tangent frame \(\{e_{i}\}\) such that

$$\begin{aligned} h_{ij}^{\alpha }:=\langle h(e_{i},e_{j}),e_{\alpha }\rangle =\mu _{i}^{\alpha }\delta _{ij}. \end{aligned}$$

Write \(e_{k^{*}}=\sum _{\alpha }a_{k^{*}}^{\alpha }e_{\alpha }\). Then

$$\begin{aligned} h_{ij}^{k^{*}}= & {} \langle h(e_{i},e_{j}),e_{k^{*}}\rangle =\big \langle h(e_{i},e_{j}),\sum _{\alpha }a_{k^{*}}^{\alpha }e_{\alpha }\big \rangle =\sum _{\alpha }a_{k^{*}}^{\alpha }\langle h(e_{i},e_{j}),e_{\alpha }\rangle \nonumber \\= & {} \sum _{\alpha }a_{k^{*}}^{\alpha }h_{ij}^{\alpha } =\sum _{\alpha }a_{k^{*}}^{\alpha }\mu _{i}^{\alpha }\delta _{ij} =\lambda _{i}^{k^{*}}\delta _{ij}\ \text{ with } \lambda _{i}^{k^{*}}:=\sum _{\alpha }a_{k^{*}}^{\alpha }\mu _{i}^{\alpha }. \end{aligned}$$
(3.37)

Thus Proposition 3.10 is proved. \(\square \)

Proof of Theorem

Since \(|x|^2=\mathrm{const}\), \(|H|^{2}-|h|^{2}\le 0\) on \(M^2\) by (3.35). Then it follows from (3.15) that \(|H|^{2}-|h|^{2}\equiv 0\) which with the Gauss equation shows that \(M^{2}\) is flat. Therefore, due to Proposition 3.10, we can choose \(\{e_{1},e_{2}\}\) such that

$$\begin{aligned} h_{12}^{1^{*}}=h_{12}^{2^{*}}=0. \end{aligned}$$
(3.38)

It follows that

$$\begin{aligned} h_{22}^{1^{*}}=h_{11}^{2^{*}}=0. \end{aligned}$$
(3.39)

On the other hand, since \(\nabla h\equiv 0\), we have

$$\begin{aligned} 0=\sum h _{ijl}^{k^{*}}\theta _{l}=dh_{ij}^{k^{*}}-\sum h_{lj}^{k^{*}}\theta _{il} -\sum h_{il}^{k^{*}}\theta _{jl}+\sum h_{ij}^{p^{*}}\theta _{p^{*}k^{*}}. \end{aligned}$$

It follows that

  1. (1)

    \(i=j=1\), \(k=2\), we get \( h_{11}^{1^{*}}\theta _{1^{*}2^{*}}=0\),

  2. (2)

    \(i=j=2\), \(k=1\), we get \(h_{22}^{2^{*}}\theta _{2^{*}1^{*}}=0\),

  3. (3)

    \(i=j=k=1\), \(0=dh_{11}^{1^{*}}+h_{11}^{1^{*}}\theta _{1^{*}1^{*}}=dh_{11}^{1^{*}}\), we get \(h_{11}^{1^{*}}=\mathrm{const}\),

  4. (4)

    \(i=j=k=2\), \(0=dh_{22}^{2^{*}}+h_{22}^{2^{*}}\theta _{2^{*}2^{*}}=dh_{22}^{2^{*}}\), we get \(h_{22}^{2^{*}}=\mathrm{const}\).

Since x can not be totally geodesic, \((h_{11}^{1^{*}})^{2}+(h_{22}^{2^{*}})^{2}=|h|^2\ne 0\) by (3.38) and (3.39). Without loss of generality we can assume that \(h_{11}^{1^{*}}\ne 0\). Thus, by (1), we have \(\theta _{1^{*}2^{*}}=0\) which with \(\nabla J=0\) shows that \(\theta _{12}=0\). Now let

$$\begin{aligned} \tilde{e}_{1}=e_{1}\cos \theta -e_{2}\sin \theta , \quad \tilde{e}_{2}=e_{1}\sin \theta +e_{2}\cos \theta \end{aligned}$$

be another frame field such that

$$\begin{aligned} \tilde{h}_{ij}^{k^{*}}:=\langle h(\tilde{e}_{i},\tilde{e}_{j}),\tilde{e}_{k^{*}}\rangle =\tilde{\lambda }_{i}^{k^{*}}\delta _{ij}. \end{aligned}$$

Then a direct computation shows that

$$\begin{aligned} \sin \theta \cos \theta (h_{11}^{1^{*}}\cos \theta +h_{22}^{2^{*}}\sin \theta ) =\sin \theta \cos \theta (h_{11}^{1^{*}}\sin \theta -h_{22}^{2^{*}}\cos \theta )=0. \end{aligned}$$

Since \((h_{11}^{1^{*}})^{2}+(h_{22}^{2^{*}})^{2}\ne 0\), we have \(\sin 2\theta =0\), that is \(\theta =0,\) or \(\frac{\pi }{2}\) or \(\pi \). Clearly, by choosing \(\theta =\frac{\pi }{2}\), we can change the sign of \(h_{11}^{1^{*}}h_{22}^{2^{*}}\); while by choosing \(\theta =\pi \), we can change the sign of both \(h_{11}^{1^{*}}\) and \(h_{22}^{2^{*}}\). Thus we can always assume that \(h_{11}^{1^{*}}>0\) and \(h_{22}^{2^{*}}\ge 0\). It then follows that \(\{e_{1},e_{2}\} \) can be uniquely determined and, in particular, is globally defined.

Now we claim that \(h_{22}^{2^{*}}>0\). In fact, if \(h_{22}^{2^{*}}=0\), then \(\theta _{22^{*}}=0\). This with \(\theta _{12}=\theta _{21^{*}}=0\) shows that \(e_2\) is constant in \({\mathbb C}^2\) along M which means that M contains a family of parallel straight lines, contradicting the assumption that M is compact.

Define

$$\begin{aligned} V_{1}=\mathrm{Span\,}_{\mathbb R}\{e_{1},e_{1^*}\}=\mathrm{Span\,}_{{\mathbb {C}}}\{e_{1}\}, \quad V_{2}=\mathrm{Span\,}_{\mathbb R}\{e_2,e_{2^*}\}=\mathrm{Span\,}_{{\mathbb {C}}}\{e_{2}\}. \end{aligned}$$
(3.40)

Since

$$\begin{aligned} de_{1}= & {} \nabla e_{1}+\sum h_{1j}^{k^{*}}\theta _{j}e_{k^{*}}=h_{11}^{1^{*}}\theta _{1}e_{1^{*}}\in V_{1},\\ de_{1^{*}}= & {} Jde_{1}=h_{11}^{1^{*}}\theta _{1}Je_{1^{*}}=-h_{11}^{1^{*}}\theta _{1}e_{1}\in V_{1}, \end{aligned}$$

we know that \(V_{1}\) is a 1-dimensional constant complex subspace of \({\mathbb {C}}^{2}\).

Similarly, \(V_{2}\) is also a 1-dimensional constant complex subspace of \({\mathbb {C}}^{2}\). Furthermore, \(V_1\) and \(V_2\) are clearly orthogonal. So, up to a holomorphic isometry on \({\mathbb {C}}^{2}\), we can assume that \(V_{1}={\mathbb {C}}^{1}\), \(V_{2}={\mathbb {C}}^{1}\) so that \({\mathbb {C}}^{2}=V_{1}\times V_{2}\). Write

$$\begin{aligned} x=(x^1,x^2)\in V_1\times V_2\equiv {\mathbb C}^2. \end{aligned}$$

Then

$$\begin{aligned} 0=e_{i}(|x|^{2})=e_{i}(|x^1|^{2})+e_{i}(|x^2|^{2}),\quad i=1,2, \end{aligned}$$

which with the definitions (3.40) of \(V_1\) and \(V_2\) shows that

$$\begin{aligned} e_{1}(|x^{1}|^{2})=e_{2}(|x^{1}|^{2})=0,\quad e_{1}(|x^{2}|^{2})=e_{2}(|x^{2}|^{2})=0, \end{aligned}$$

that is,

$$\begin{aligned} |x^{1}|^{2}=\mathrm{const},\quad |x^{2}|^{2}=const. \end{aligned}$$

It is easily seen that both \(|x^{1}|^{2}\) and \(|x^{2}|^{2}\) are positive since x is non-degenerate. Thus we can write \(|x^{1}|^{2}=a^{2}>0\), \(|x^{2}|^{2}=b^{2}>0\). It then follows that \(M^{2}={\mathbb {S}}^{1}(a)\times {\mathbb {S}}^{1}(b)\).

Finally, by the assumption (1.4), \(|h|^{2}\ge 2\), it should holds that \(a^{2}+b^{2}\ge 2a^{2}b^{2}\). \(\square \)