1 Introduction

Let M be a compact manifold of dimension n. Given a Riemannian metric g on M, the volume \(\text{ Vol }(M,g)\) and the Laplace-Beltrami operator \(\Delta _{g}\) are defined. Let \(0 < \lambda _{1}(g) \le \lambda _{2}(g) \le \cdots \lambda _{k}(g) \le \cdots \) be the eigenvalues of \(\Delta _{g}\). The quantity \(\lambda _{k}(g) \text{ Vol }(M,g)^{2/n}\) is invariant under scaling of the metric g. Hersch [14] proved that on a 2-dimensional sphere \(S^{2}\), the scale-invariant quantity \(\lambda _{1}(g)\text{ Area }(g)\) is maximized when g is a round metric. Inspired by the work, Berger [3] asked whether

$$\begin{aligned} \Lambda _{1}(M) := \sup _{g} \lambda _{1}(g) \text{ Vol }(M,g)^{2/n} \end{aligned}$$

is finite for a compact manifold M of dimension n. For a surface M, \(\Lambda _{1}(M)\) is bounded by a constant depending on the genus [15, 36]. Berger [3] also conjectured that for a 2-dimensional torus \(T^{2}\), the flat equilateral torus attains \(\Lambda _{1}(T^{2})\). Nadirashvili [24] settled Berger’s conjecture affirmatively. In the same paper, he proved a theorem that a metric g on a given surface M is extremal for the functional \(\lambda _{1}: g \mapsto \lambda _{1}(g)\) with respect to all the volume-preserving deformations of the metric if and only if there exists a finite collection of \(\lambda _{1}(g)\)-eigenfunctions \(\{f_{j}\}_{j=1}^{N}\) such that \(F:= (f_{1}, \cdots , f_{N}): (M,g) \rightarrow \textbf{R}^{N}\) is an isometric minimal immersion into a sphere in \(\textbf{R}^{N}\). Later, El Soufi-Ilias [11] simplified the proof of the theorem and generalized it to a compact manifold M of any dimension. More explicitly, they proved the following:

Theorem 1.1

([11, 24]) Let (Mg) be a compact m-dimensional Riemannian manifold. The metric g is extremal for the functional \(\lambda _{1}\) with respect to all the volume-preserving deformations of the metric if and only if there exists a finite collection of \(\lambda _{1}(g)\)-eigenfunctions \(\{f_{1}, \ldots , f_{N} \}\) such that \(F:= (f_{1}, \cdots , f_{N}): (M,g) \rightarrow \textbf{R}^{N}\) is an isometric minimal immersion into \(S^{N-1}(\sqrt{m/\lambda _{1}(g)}) \subset \textbf{R}^{N}\).

In particular, they showed that the canonical metric on a compact isotropy irreducible homogeneous manifold is extremal for the functional \(\lambda _{1}\). For recent remarkable progress in study of \(\Lambda _{1}(M)\) for a surface M, see [16, 17, 26,27,28], for example.

On the other hand, on any manifold M with \(n \ge 3\), one can construct a 1-parameter family \((g_{t})_{t>0}\) such that the quantity \(\lambda _{1}(g_{t}) \text{ Vol }(M,g_{t})^{2/n}\) diverges to infinity as t goes to infinity [8]. (See also [6, 22, 23, 32, 33].) However, for a given Riemannian metric g on M, the restriction of the functional \(\lambda _{1}\) to metrics in the conformal class with fixed volume is bounded [10, 19]. El Soufi-Ilias [12] proved that a metric g is extremal for the functional \(\lambda _{1}\) among such metrics if and only if there exists a finite collection of eigenfunctions \(\{f_{j}\}_{j=1}^{N}\) such that \(F:= (f_{1}, \cdots , f_{N}): (M,g) \rightarrow \textbf{R}^{N}\) is a harmonic map into a unit sphere in \(\textbf{R}^{N}\) with constant energy density \(|dF|^{2} =\lambda _{1}(g)\).

Bourguignon et al. [7] proved the following result:

Theorem 1.2

([7]) Let (MJ) be a compact complex n-dimensional manifold admitting a full holomorphic immersion \(\Phi : (M, J) \rightarrow \textbf{C}P^{N}\). Let \(\sigma _{FS}\) be the Fubini-Study form on \(\textbf{C}P^{N}\) with constant holomorphic sectional curvature 1. Then, for any Kähler form \(\omega \) on (MJ), the first eigenvalue \(\lambda _{1}(\omega )\) satisfies

$$\begin{aligned} \lambda _{1}(\omega ) \le n\frac{N+1}{N} \frac{\int _{M}\Phi ^{*}\sigma _{FS} \wedge \omega ^{n-1}}{\int _{M} \omega ^{n}}. \end{aligned}$$

The divergence theorem implies that \(\lambda _{1}(\omega )\) is bounded by a constant depending on only n, N, \(\Phi \) and the Kähler class \([\omega ]\). The above theorem implies that the Fubini-Study metric on \(\textbf{C}P^{N}\) is a \(\lambda _{1}\)-maximizer in its Kähler class. Biliotti-Ghigi [5] generalized the result and showed that the canonical Kähler-Einstein metric on a Hermitian symmetric space of compact type is a \(\lambda _{1}\)-maximizer in its Kähler class. (See also [2].) Motivated by these results, Apostolov et al. [1] proved that the metric g on a compact Kähler manifold is extremal for the functional \(\lambda _{1}\) within its Kähler class if and only if there exists a finite collection of eigenfunctions \(\{f_{j}\}_{j=1}^{N}\) such that the equation

$$\begin{aligned} \lambda _{1}(g)^{2}\left( \sum _{j=1}^{N} f_{j}^{2} \right) -2\lambda _{1}(g) \left( \sum _{j=1}^{N} |\nabla f_{j}|^{2} \right) + \sum _{j=1}^{N} |dd^{c}f_{j}|^{2} = 0 \end{aligned}$$
(1.1)

holds. Using this equation, they showed that the metric g of a compact homogeneous Kähler-Einstein manifold \((M, J, g, \omega )\) of positive scalar curvature is extremal for the functional \(\lambda _{1}\) within its Kähler class. However, compared to the aforementioned theorems due to Nadirashvili and El Soufi-Ilias, the geometric meaning of the equation (1.1) is not clear.

Let (MJ) be a compact complex manifold satisfying the assumption of Theorem 1.2. Let \(H^{1,1}(M,J;\textbf{R}):= H^{1,1}(M,J) \cap H^{2}_{dR}(M)\). Then the map

$$\begin{aligned} H^{1,1}(M,J;\textbf{R}) \rightarrow \textbf{R}, \quad [\omega ] \mapsto \int _{M} \Phi ^{*}\sigma _{FS} \wedge \omega ^{n-1} \end{aligned}$$

is a well-defined continuous function. Thus this is bounded on the compact subset \(\{ [\omega ] \in H^{1,1}(M,J;\textbf{R}) \mid \int _{M} \omega ^{n}=1 \}\). In other words, the functional \(\lambda _{1}\) is bounded on the set of Kähler metrics with fixed volume on (MJ). However, the property of the \(\lambda _{1}\)-maximizing Kähler metrics has not been studied.

In this paper, on a compact complex manifold (MJ), we introduce the notion of \(\lambda _{k}\)-extremal Kähler metric by considering all volume-preserving deformations of the Kähler metric. Be cautioned that we fix the complex structure J and consider only J-compatible Kähler metrics. (See Definition 2.1 for the precise definition of the \(\lambda _{k}\)-extremality.) The notion of \(\lambda _{k}\)-extremality in this paper is stronger than that in the above theorem due to Apostolov et al. [1]. We prove that the Kähler metric g is \(\lambda _{1}\)-extremal if and only if there exists a finite collection of eigenfunctions \(\{f_{j}\}_{j=1}^{N}\) such that the equations

$$\begin{aligned} \left\{ \, \begin{aligned}&H \left( \sum _{j=1}^{N} f_{j}dd^{c}f_{j} \right) = -\omega , \\&\lambda _{1}(g)^{2}\left( \sum _{j=1}^{N} f_{j}^{2} \right) -2\lambda _{1}(g) \left( \sum _{j=1}^{N} |\nabla f_{j}|^{2} \right) + \sum _{j=1}^{N} |dd^{c}f_{j}|^{2} = 0 \\ \end{aligned} \right. \end{aligned}$$
(1.2)

hold (Theorem 2.9). Here H is the harmonic projector, which is defined due to the Hodge decomposition on a compact Kähler manifold. It is obvious that the equation (1.2) implies (1.1). In addition, we can obtain a result that is similar to the aforementioned results due to El Soufi-Ilias [11] and Apostolov et al [1]. That is, the metric on a compact isotropy irreducible homogeneous Kähler manifold is \(\lambda _{1}\)-extremal in our sense (Proposition 2.12). We also give an example of a Kähler metric that is \(\lambda _{1}\)-extremal within its Kähler class, but not so for all volume-preserving deformations of the Kähler metric (Example 2.17).

In the final section of this paper, we consider flat complex tori. We prove that the metric on a flat complex torus is \(\lambda _{1}\)-extremal within its Kähler class. Montiel and Ros [21] showed that among all the real 2-dimensional tori, the only square torus \(\textbf{R}^{2}/\textbf{Z}^{2}\) admits an isometric minimal immersion into a 3-dimensional Euclidean sphere by the first eigenfunctions. Later, using Theorem 1.1, El Soufi and Ilias [11] improved the result. That is, they proved that a real 2-dimensional torus admits isometric minimal immersion into a Euclidean sphere of some dimension by the first eigenfunctions if and only if the torus is the square torus or the equilateral torus. Very recently, Lü et al. [20] classified all the 3-dimensional flat tori and 4-dimensional flat tori that admit an isometric minimal immersion into a Euclidean sphere of some dimension by the first eigenfunctions. Furthermore, they constructed new examples of flat tori that admit an isometric minimal immersion into a Euclidean sphere by the first eigenfunctions. In Theorem 3.2, we give a necessary and sufficient condition for the metric on a flat complex torus to be \(\lambda _{1}\)-extremal in our sense. The notion of \(\lambda _{1}\)-extremality introduced in this paper is stronger than that in Theorem 1.1. Hence Theorem 3.2 gives a necessary condition for a flat complex torus to admit an isometric minimal immersion into a Euclidean sphere by the first eigenfunctions. As far as the author knows, this is the only currently known necessary condition for a flat complex torus to admit an isometric minimal immersion into a Euclidean sphere by the first eigenfunctions. We also show that if the multiplicity of the first eigenvalue of a flat complex torus is 2, then the flat metric on the torus is not \(\lambda _{1}\)-extremal in our sense (Corollary 3.3). In addition, we show that an m-dimensional flat torus \(\textbf{R}^{m}/D_{m}\), where \(D_{m}\) is an m-dimensional checkerboard lattice, admits an isometric minimal immersion into a Euclidean sphere by the first eigenfunctions (Proposition 3.6). This fact has been already proved by Lü et al. [20] for \(m=3,4\).

2 A \(\lambda _{k}\)-extremal Kähler metric

Let (MJg) be a compact Kähler manifold (without boundary) of complex dimension n. By scaling the metric, we may assume that the volume Vol(M, g) is 1. Let \(\omega \) be the Kähler form and \(d\mu \) the volume form. It is well known that \(d\mu \) is given by \(d\mu = \omega ^{n}/n!\). Below we follow the conventions in [1]. We define the exterior differential d by \(d = \partial + \overline{\partial }\) and twisted exterior differential \(d^{c}\) by \(d^{c} = i(\overline{\partial }-\partial )\). Both d and \(d^{c}\) are real operators. We clearly have \(dd^{c} = 2i\partial \overline{\partial }\). The Kähler metric g induces a pointwise hermitian inner product on (1, 1)-forms, which we also denote by g. Note that g is symmetric for a pair of real (1, 1)-forms. We have

$$\begin{aligned} |\omega |^{2} = g(\omega , \omega ) = n. \end{aligned}$$
(2.1)

Let \(\delta \) and \(\delta ^{c}\) be the \(L^{2}(g)\)-adjoints of d and \(d^{c}\) respectively. We define the Laplacian \(\Delta _{g}\) by \(\Delta _{g} =d\delta + \delta d\). Eigenvalues of the Laplacian \(\Delta _{g}\) acting on functions are nonnegative and we denote them by \(0 < \lambda _{1}(g) \le \lambda _{2}(g) \le \cdots \lambda _{k}(g) \le \cdots \). For any \(k \in \textbf{N}\), let \(E_{k}(g)\) be the vector space of real-valued eigenfunctions of \(\Delta _{g}\) corresponding to \(\lambda _{k}(g)\). That is, \(E_{k}(g)\) is given by \(E_{k}(g) = \text{ Ker }(\Delta _{g}-\lambda _{k}(g)I)\), where I is the identity map acting on functions. We have

$$\begin{aligned} \Delta _{g} \phi = -2 g^{j\overline{k}} \frac{\partial ^{2}\phi }{\partial z^{j} \partial {\overline{z}^{k}}} = -2g(\omega , i\partial \overline{\partial }\phi ) = -g(\omega , dd^{c}\phi ) \end{aligned}$$
(2.2)

for a function \(\phi \). We also have

$$\begin{aligned} *\omega = \frac{1}{(n-1)!}\omega ^{n-1}, \end{aligned}$$

where \(*\) is the Hodge star operator. This implies that the equation

$$\begin{aligned} \alpha \wedge \omega ^{n-1} = (n-1)!\alpha \wedge * \omega = (n-1)! g(\alpha , \omega )\frac{\omega ^{n}}{n!} = \frac{1}{n} g(\alpha , \omega ) \omega ^{n} \end{aligned}$$
(2.3)

holds for a (1, 1)-form \(\alpha \). In particular, the equations (2.2) and (2.3) imply

$$\begin{aligned} ndd^{c}\phi \wedge \omega ^{n-1} = g(dd^{c}\phi , \omega )\omega ^{n} = -(\Delta _{g} \phi ) \omega ^{n}. \end{aligned}$$
(2.4)

Let \(Z^{1,1}(M;\textbf{R})\) be the real vector space of d-closed real (1, 1)-forms on M. Let \(Z^{1,1}_{0}(M;\textbf{R})\) be the subspace defined by

$$\begin{aligned} Z^{1,1}_{0}(M;\textbf{R}) := \left\{ \alpha \in Z^{1,1}(M;\textbf{R}) \mid \int _{M} g(\alpha , \omega ) d\mu = 0 \right\} . \end{aligned}$$

Fix an arbitrary element \(\alpha \in Z^{1,1}_{0}(M;\textbf{R})\). The (1, 1)-form

$$\begin{aligned} \widetilde{\omega }_{t} := \omega + t\alpha \end{aligned}$$
(2.5)

is a Kähler form for a sufficiently small t. In particular, if we consider \(\alpha = dd^{c}\psi \) for a real-valued function \(\psi \), then \(\alpha \) satisfies

$$\begin{aligned} \int _{M} g(\alpha , \omega ) d\mu = -\int _{M} \Delta _{g} \psi d\mu = 0, \end{aligned}$$

and so we have \(\alpha \in Z^{1,1}_{0}(M;\textbf{R})\). The 1-parameter family \(\widetilde{\omega }_{t} = \omega + tdd^{c}\psi \) is a deformation of \(\omega \) in its Kähler class \([\omega ]\), which was studied by Apostolov et al. [1]. Let \(\widetilde{g}_{t}\) be the Kähler metric corresponding to the Kähler form \(\widetilde{\omega }_{t}\) in (2.5). Set

$$\begin{aligned} g_{t} := \text{ Vol }(M, \tilde{g}_{t})^{-1/n}\widetilde{g_{t}}, \quad \omega _{t} := \text{ Vol }(M, \tilde{g}_{t})^{-1/n}\widetilde{\omega _{t}}. \end{aligned}$$
(2.6)

Then we have \(g_{0} = g\) and \(\omega _{0} = \omega \). We also see that \((g_{t})_{t}\) is a volume-preserving 1-parameter family of Kähler metrics that depends analytically on t, and \(\omega _{t}\) is the Kähler form associated with \(g_{t}\). Moreover, we can verify that \(\left. \frac{d}{dt} \right| _{t=0} \omega _{t} = \alpha \). (See (2.9) below.)

Definition 2.1

The Kähler metric g on a compact Kähler manifold \((M, J, g, \omega )\) is called \(\lambda _{k}\)-extremal (for all the volume-preserving deformations of the Kähler metric) if the inequality

$$\begin{aligned} \left( \left. \frac{d}{dt}\right| _{t=0^{-}} \lambda _{k}(g_{t}) \right) \cdot \left( \left. \frac{d}{dt}\right| _{t=0^{+}} \lambda _{k}(g_{t}) \right) \le 0 \end{aligned}$$

holds for any 1-parameter family of volume-preserving Kähler metrics \((g_{t})_{t}\) that depends real analytically on t.

Remark 2.2

When we consider whether a Kähler metric g on (MJ) is \(\lambda _{k}\)-extremal, we may rescale the metric so that \(\text{ Vol }(M,g) =1\). Let \((g_{t})_{t}\) be a 1-parameter family of volume-preserving Kähler metrics that depends real analytically on t. Let \(\omega _{t}\) be the Kähler form associated with \(g_{t}\). El Soufi-Ilias [13] showed that \(\left. \frac{d}{dt}\right| _{t=0^{-}} \lambda _{k}(g_{t}) \) and \(\left. \frac{d}{dt}\right| _{t=0^{+}} \lambda _{k}(g_{t})\) depend on only \(\omega \) and \(\left. \frac{d}{dt}\right| _{t=0}\omega _{t}\). Since \((\omega _{t})_{t}\) is volume-preserving, we have \(\left. \frac{d}{dt}\right| _{t=0}\omega _{t} \in Z^{1,1}_{0}(M; \textbf{R})\). Hence it suffices to consider \((\omega _{t})_{t}\) given by (2.6). Thus a Kähler metric g on (MJ) with \(\text{ Vol }(M,g) =1\) is \(\lambda _{k}\)-extremal if and only if for any \(\alpha \in Z^{1,1}_{0}(M;\textbf{R})\), the associated volume-preserving 1-parameter family of Kähler metrics \((g_{t})_{t}\) defined by (2.6) satisfies

$$\begin{aligned} \left( \left. \frac{d}{dt}\right| _{t=0^{-}} \lambda _{k}(g_{t}) \right) \cdot \left( \left. \frac{d}{dt}\right| _{t=0^{+}} \lambda _{k}(g_{t}) \right) \le 0. \end{aligned}$$

We quote the following theorem due to El Soufi-Ilias [13]:

Theorem 2.3

([13]) Let (Mg) be a compact Riemannian manifold and \((g_{t})_{t}\) be a 1-parameter family of Riemannian metrics that depends real-analytically on t with \(g_{0}=g\). Let \(\Pi _{k}: L^{2}(M,g) \rightarrow E_{k}(g)\) be the orthogonal projection onto \(E_{k}(g)\). Define the operator \(P_{k}: E_{k}(g) \rightarrow E_{k}(g)\) by

$$\begin{aligned} P_{k}(f) := \Pi _{k} \left( \left. \frac{d}{dt}\right| _{t=0} \Delta _{g_{t}}f \right) . \end{aligned}$$
(2.7)

Then the following hold:

  1. (1)

    \(\left. \frac{d}{dt}\right| _{t=0^{-}} \lambda _{k}(g_{t})\) and \(\left. \frac{d}{dt}\right| _{t=0^{+}} \lambda _{k}(g_{t})\) are eigenvalues of \(P_{k}\).

  2. (2)

    If \(\lambda _{k}(g) > \lambda _{k-1}(g)\), then \(\left. \frac{d}{dt}\right| _{t=0^{-}} \lambda _{k}(g_{t})\) and \(\left. \frac{d}{dt}\right| _{t=0^{+}} \lambda _{k}(g_{t})\) are the greatest and the least eigenvalues of \(P_{k}\).

  3. (3)

    If \(\lambda _{k}(g) < \lambda _{k+1}(g)\), then \(\left. \frac{d}{dt}\right| _{t=0^{-}} \lambda _{k}(g_{t})\) and \(\left. \frac{d}{dt}\right| _{t=0^{+}} \lambda _{k}(g_{t})\) are the least and the greatest eigenvalues of \(P_{k}\).

For any \(\alpha \in Z^{1,1}_{0}(M;\textbf{R})\), the associated volume-preserving 1-parameter family of Kähler metrics \((g_{t})_{t}\) given by (2.6) defines the associated operator \(P_{k, \alpha }: E_{k}(g) \rightarrow E_{k}(g)\) by (2.7). Since \(P_{k, \alpha }\) is symmetric with respect to the \(L^{2}(g)\)-inner product induced by g, one can consider the corresponding quadratic form on \(E_{k}(g)\), given by

$$\begin{aligned} Q_{\alpha }(f) := \int _{M} f P_{k, \alpha }(f) d\mu = \int _{M} f \left( \left. \frac{d}{dt}\right| _{t=0} \Delta _{g_{t}}f \right) d\mu . \end{aligned}$$
(2.8)

The following proposition is an immediate consequence of Theorem 2.3 (1):

Proposition 2.4

Let \((M, J, g, \omega )\) be a compact Kähler manifold. If the metric g of \((M, J, g, \omega )\) is \(\lambda _{k}\)-extremal, then the quadratic form \(Q_{\alpha }\), defined in (2.8), is indefinite on \(E_{k}(g)\) for any \(\alpha \in Z^{1,1}_{0}(M;\textbf{R})\).

We also have the following proposition, which follows form Theorem 2.3 (2) and (3):

Proposition 2.5

Let \((M, J, g, \omega )\) be a compact Kähler manifold. Suppose that \(\lambda _{k}(g) > \lambda _{k-1}(g)\) or \(\lambda _{k}(g) < \lambda _{k+1}(g)\) holds. Then the metric g of (MJg) is \(\lambda _{k}\)-extremal if and only if the quadratic form \(Q_{\alpha }\) is indefinite on \(E_{k}(g)\) for any \(\alpha \in Z^{1,1}_{0}(M;\textbf{R})\).

The next corollary immediately follows.

Corollary 2.6

Let \((M, J, g, \omega )\) be a compact Kähler manifold. Then the metric g is \(\lambda _{1}\)-extremal if and only if the quadratic form \(Q_{\alpha }\) is indefinite on \(E_{1}(g)\) for any \(\alpha \in Z^{1,1}_{0}(M;\textbf{R})\).

Theorem 2.7

Let \((M, J, g, \omega )\) be a compact Kähler manifold of complex dimension n with \(\text{ Vol }(M,g) =1\). For any \(\alpha \in Z^{1,1}_{0}(M;\textbf{R})\), the quadratic form \(Q_{\alpha }\), given by (2.8), can be expressed as

$$\begin{aligned} Q_{\alpha }(f) = \int _{M}g(fdd^{c}f, \alpha ) d\mu . \end{aligned}$$

Proof

First we calculate \(\left. \frac{d}{dt}\right| _{t=0}\text{ Vol }(M, \widetilde{g_{t}})\) and \(\left. \frac{d}{dt}\right| _{t=0} \omega _{t}\). The volume form \(\widetilde{d\mu }_{t}\), determined by \(\widetilde{\omega }_{t}\), can be written as

$$\begin{aligned} \widetilde{d\mu }_{t} = \frac{1}{n!} \widetilde{\omega }_{t}^{n} = \frac{1}{n!}\left( \omega ^{n} + tn\alpha \wedge \omega ^{n-1} \right) + O(t^{2}) = [1+tg(\alpha , \omega )] d\mu + O(t^{2}). \end{aligned}$$

Hence one obtains

$$\begin{aligned} \text{ Vol }(M, \widetilde{g}_{t}) = \int _{M} \widetilde{d\mu }_{t} = 1 + t\int _{M}g(\alpha , \omega ) d\mu +O(t^{2}) = 1+O(t^{2}), \end{aligned}$$

where in the last equality we have used the the assumptions that \(\text{ Vol }(M, \widetilde{g}) = 1\) and \(\alpha \in Z^{1,1}_{0}(M;\textbf{R})\). Hence this implies \(\left. \frac{d}{dt}\right| _{t=0} \text{ Vol }(M, \tilde{g_{t}}) = 0\). Thus one obtains

$$\begin{aligned} \left. \frac{d}{dt} \right| _{t=0} \omega _{t} = \frac{d}{dt} \left( \text{ Vol }(M, \widetilde{g}_{t})^{-1/n}) \tilde{\omega }_{t} \right) = \text{ Vol }(M, \tilde{g}_{0})^{-1/n} \left. \frac{d}{dt} \right| _{t=0} \widetilde{\omega }_{t} = \alpha .\qquad \end{aligned}$$
(2.9)

Next we differentiate

$$\begin{aligned} ndd^{c}f \wedge \omega _{t}^{n-1} = -(\Delta _{g_{t}} f) \omega _{t}^{n}, \end{aligned}$$

which comes from (2.4). Differentiating the left hand side at \(t=0\), one obtains

$$\begin{aligned} n(n-1)dd^{c}f \wedge \left( \left. \frac{d}{dt}\right| _{t=0} \omega _{t}\right) \wedge \omega ^{n-2} = n (n-1) dd^{c}f \wedge \alpha \wedge \omega ^{n-2}, \end{aligned}$$

where (2.9) is used. On the other hand, differentiating the right hand side at \(t=0\), one obtains

$$\begin{aligned}&-\left( \left. \frac{d}{dt}\right| _{t=0} \Delta _{g_{t}}f \right) \omega ^{n} -n(\Delta _{g}f)\left( \left. \frac{d}{dt}\right| _{t=0} \omega _{t} \right) \wedge \omega ^{n-1} \\&\quad = -\left( \left. \frac{d}{dt}\right| _{t=0} \Delta _{g_{t}}f \right) \omega ^{n} - n(\Delta _{g}f)\alpha \wedge \omega ^{n-1} \\&\quad = -\left( \left. \frac{d}{dt}\right| _{t=0} \Delta _{g_{t}}f \right) \omega ^{n} - (\Delta _{g}f) g(\alpha , \omega ) \omega ^{n}, \end{aligned}$$

where (2.3) is used for the last equality. Thus one obtains

$$\begin{aligned} \left( \left. \frac{d}{dt}\right| _{t=0} \Delta _{g_{t}}f \right) \omega ^{n} = -n(n-1)dd^{c}f \wedge \alpha \wedge \omega ^{n-2} - g(\alpha , \omega )\omega ^{n}. \end{aligned}$$
(2.10)

It is known [1] that the equation

$$\begin{aligned} g(\phi , \psi ) \omega ^{n} = g(\phi , \omega ) g(\psi , \omega )\omega ^{n} - n(n-1)\phi \wedge \psi \wedge \omega ^{n-2} \end{aligned}$$

holds for a pair of real (1, 1)-forms \(\phi \) and \(\psi \). Substituting \(\phi = dd^{c}f\) and \(\psi = \alpha \) into this equation and combining with (2.10) and (2.2), one obtains

$$\begin{aligned} \left( \left. \frac{d}{dt}\right| _{t=0} \Delta _{g_{t}}f \right) \omega ^{n} = g(dd^{c}f, \alpha )\omega ^{n}. \end{aligned}$$

Hence one concludes

$$\begin{aligned} \left. \frac{d}{dt}\right| _{t=0} \Delta _{g_{t}}f = g(dd^{c}f, \alpha ) \end{aligned}$$

and thus the assertion follows. \(\square \)

For a compact Kähler manifold \((M, J, g, \omega )\), \(\mathcal {H}^{1,1}(M)\), the vector space of \(\textbf{C}\)-valued harmonic (1, 1)-forms, is known to be finite dimensional. In particular, \(\mathcal {H}^{1,1}(M)\) is a closed subspace of \(\Omega ^{1,1}(M)\) and hence we can define the \(\textbf{C}\)-linear \(L^{2}(g)\)-orthogonal projection \(H: \Omega ^{1,1}(M) \rightarrow \mathcal {H}^{1,1}(M)\). For a real (1, 1)-form \(\eta \), \(H(\eta )\) is a real harmonic (1, 1)-form. Let \(\mathcal {H}^{1,1}(M; \textbf{R})\) be the vector space of real harmonic (1, 1)-forms. Set

$$\begin{aligned} \mathcal {H}^{1,1}_{0}(M; \textbf{R}) := \left\{ \alpha \in \mathcal {H}^{1,1}(M;\textbf{R}) \mid \int _{M} g(\alpha , \omega ) d\mu = 0 \right\} \end{aligned}$$

and

$$\begin{aligned} C^{\infty }_{0}(M; \textbf{R}) := \left\{ \varphi \in C^{\infty }(M;{\textbf {R}}) \mid \int _{M} \varphi d\mu = 0 \right\} . \end{aligned}$$

By the \(dd^{c}\)-lemma, \(\alpha \in Z^{1,1}_{0}(M;\textbf{R})\) can be decomposed as

$$\begin{aligned} \alpha = H(\alpha ) + dd^{c} \varphi . \end{aligned}$$

This decomposition gives an \(\textbf{R}\)-linear bijection between \(Z^{1,1}_{0}(M;\textbf{R})\) and \(\mathcal {H}^{1,1}_{0}(M;\textbf{R}) \times C^{\infty }_{0}(M; \textbf{R})\).

Apostolov et al. [1] introduced the fourth order differential operator \(L(f):= \delta ^{c}\delta (fdd^{c}f)\), where \(\delta \) and \(\delta ^{c}\) are the \(L^{2}\)-adjoints of d and \(d^{c}\) respectively. They proved that the equation

$$\begin{aligned} L(f) = \lambda _{k}(g) f^{2} -2\lambda _{k}(g) |\nabla f|^{2} + |dd^{c} f|^{2} \end{aligned}$$

holds for an eigenfunction \(f \in E_{k}(g)\).

We have the following theorem:

Theorem 2.8

Let \((M, J, g, \omega )\) be a compact Kähler manifold of complex dimension n. The following are equivalent:

  1. (1)

    For any \(\alpha \in Z^{1,1}_{0}(M;\textbf{R})\), the quadratic form \(Q_{\alpha }\) given by (2.8) is indefinite on the eigenspace \(E_{k}(g)\).

  2. (2)

    There exists a finite subset \(\{ f_{1}, \cdots , f_{N} \} \subset E_{k}(g)\) such that the following equations hold:

    $$\begin{aligned} \left\{ \, \begin{aligned}&H \left( \sum _{j=1}^{N} f_{j}dd^{c}f_{j} \right) = -\omega \\&\sum _{j=1}^{N}L(f_{j} ) = \lambda _{k}(g)^{2}\left( \sum _{j=1}^{N} f_{j}^{2} \right) -2\lambda _{k}(g) \left( \sum _{j=1}^{N} |\nabla f_{j}|^{2} \right) + \sum _{j=1}^{N} |dd^{c}f_{j}|^{2} = 0 \\ \end{aligned} \right. \end{aligned}$$
    (2.11)

The proof of this theorem is inspired by that of Lemma 3.1 in [13].

Proof

We may assume that \(\text{ Vol }(M,g) =1\). We assume the condition (1). Let K be the convex hull of \(\{ \left( H(fdd^{c}f), L(f) \right) \mid f \in E_{k}(g) \}\) in \(\mathcal {H}^{1,1}(M; \textbf{R}) \times C^{\infty }(M; \textbf{R})\). Since \(E_{k}(g)\) is finite dimensional, K is contained in a finite dimensional subspace of \(\mathcal {H}^{1,1}(M; \textbf{R}) \times C^{\infty }(M; \textbf{R})\). We consider the product \(L^{2}\)-inner metric on the subspace. We assume that \((-\omega , 0) \notin K\). Then the hyperplane separation theorem implies that there exists \(\left( \widetilde{\alpha }_{H}, \widetilde{\varphi } \right) \in \mathcal {H}^{1,1}(M; \textbf{R}) \times C^{\infty }(M; \textbf{R})\) such that the inequalities

$$\begin{aligned} \int _{M} g(-\omega , \widetilde{\alpha }_{H}) d\mu < 0 \quad \text{ and } \quad \int _{M} g(\eta , \widetilde{\alpha }_{H}) d\mu + \int _{M}s\varphi d\mu \ge 0 \end{aligned}$$
(2.12)

hold for all \((\eta , s) \in K\setminus {\{0\}}\). Consider \(\alpha _{H} \in \mathcal {H}^{1,1}_{0}(M;\textbf{R})\) and \(\varphi \in C^{\infty }_{0}(M; \textbf{R})\) respectively defined by

$$\begin{aligned} & \alpha _{H}:= \widetilde{\alpha }_{H}- \frac{1}{n} \left( \int _{M} g(\omega , \widetilde{\alpha }_{H}) d\mu \right) \omega . \end{aligned}$$
(2.13)
$$\begin{aligned} & \varphi := \widetilde{\varphi }- \int _{M} \widetilde{\varphi } d\mu . \end{aligned}$$
(2.14)

Set

$$\begin{aligned} \alpha := \alpha _{H} + dd^{c}\varphi = \alpha _{H} + dd^{c}\widetilde{\varphi }. \end{aligned}$$

Then for any \(f \in E_{k}(g)\setminus {\{0\}}\), one has

$$\begin{aligned} Q_{\alpha }(f)&= \int _{M} g(fdd^{c}f, \alpha ) d\mu \\&= \int _{M} g(fdd^{c}f, \alpha _{H}) d\mu + \int _{M}g(fdd^{c}f, dd^{c}\widetilde{\varphi }) d\mu \\&= \int _{M} g(H(fdd^{c}f), \widetilde{\alpha }_{H}) d\mu - \frac{1}{n} \left[ \int _{M}g(fdd^{c}f, \omega ) d\mu \right] \left[ \int _{M} g(\omega , \widetilde{\alpha }_{H} ) d\mu \right] \\&\quad + \int _{M}L(f)\widetilde{\varphi } d\mu \\&= \int _{M} g(H(fdd^{c}f), \widetilde{\alpha }_{H}) d\mu + \int _{M}L(f)\widetilde{\varphi } d\mu + \frac{\lambda _{k}(g) }{n} \left[ \int _{M} f^{2} d\mu \right] \\&\quad \left[ \int _{M}g(\omega , \widetilde{\alpha }_{H}) d\mu \right] \\&>0. \end{aligned}$$

This contradicts the condition (1) and hence one concludes that \((-\omega , 0) \in K\). This implies that \((1) \Rightarrow (2)\).

Conversely, we assume that there exists a finite subset \(\{ f_{1}, \cdots , f_{N} \} \subset E_{k}(g)\) satisfying (2.11). Take any \(\alpha _{H} \in \mathcal {H}^{1,1}_{0}(M; \textbf{R})\) and \(\varphi \in C^{\infty }_{0}(M; \textbf{R})\) and consider \(\alpha \in Z^{1,1}_{0}(M; \textbf{R})\) defined by \(\alpha := \alpha _{H} + dd^{c} \varphi \). Then one has

$$\begin{aligned} \sum _{j=1}^{N}Q_{\alpha }(f_{j})&= \sum _{j=1}^{N} \int _{M} g (f_{j}dd^{c}f_{j}, \alpha ) d\mu \\&= - \int _{M} g\left( \sum _{j=1}^{N}f_{j}dd^{c}f_{j}, \alpha _{H} \right) d\mu + \sum _{j=1}^{N} \int _{M}g(f_{j}dd^{c}f_{j}, dd^{c}\varphi ) d\mu \\&= \int _{M}g(-\omega , \alpha _{H}) d\mu + \sum _{j=1}^{N}\int _{M} L(f_{j}) \varphi d\mu \\&=0. \end{aligned}$$

This implies that \(Q_{\alpha }\) is indefinite on \(E_{k}(g)\). This completes the proof. \(\square \)

Combining Corollary 2.6 and Theorem 2.8, we conclude the following:

Theorem 2.9

Let \((M, J, g, \omega )\) be a compact Kähler manifold. Suppose that the Kähler metric g is \(\lambda _{k}\)-extremal. Then there exists a finite subset \(\{ f_{1}, \cdots , f_{N} \} \subset E_{k}(g)\) satisfying (2.11). For \(k=1\), the existence of such a finite collection of eigenfunctions is a necessary and sufficient condition for the Kähler metric g to be \(\lambda _{1}\)-extremal.

We remark the relationship between Theorem 2.9 and Theorem 2.1 in [1], which we quote as follows:

Theorem 2.10

([1]) Let (MJg) be a compact Kähler manifold. The Kähler metric g is \(\lambda _{k}\)-extremal for all the deformations of the Kähler metric within its Kähler class. Then there exists a finite subset \(\{ f_{1}, \cdots , f_{N} \} \subset E_{k}(g)\) such that the equation

$$\begin{aligned} \sum _{j=1}^{N}L(f_{j}) = \lambda _{k}(g)^{2}\left( \sum _{j=1}^{N} f_{j}^{2} \right) -2\lambda _{k}(g) \left( \sum _{j=1}^{N} |\nabla f_{j}|^{2} \right) + \sum _{j=1}^{N} |dd^{c}f_{j}|^{2} = 0 \end{aligned}$$
(2.15)

holds. For \(k=1\), the existence of such a finite collection of eigenfunctions is a necessary and sufficient condition for the Kähler metric g to be \(\lambda _{1}\)-extremal for all the deformations of the Kähler metric within its Kähler class.

Obviously, the condition (2.11) is stronger than the condition (2.15). This fact is natural since our \(\lambda _{k}\)-extremality is stronger than that in [1]. Using Theorem 2.10 above, one can show that if the Kähler metric g is extremal for the functional \(\lambda _{k}\) within its Kähler class, then the eigenvalue \(\lambda _{k}(g)\) is multiple [1]. Hence we obtain the same conclusion for our \(\lambda _{k}\)-extremality:

Corollary 2.11

Let \((M, J, g, \omega )\) be a compact Kähler manifold. Suppose that g is \(\lambda _{k}\)-extremal for all the volume-preserving deformations of the Kähler metric. Then the eigenvalue \(\lambda _{k}(g)\) is multiple.

El Soufi-Ilias[11] showed that the metric on a compact isotropy irreducible homogeneous space is \(\lambda _{1}\)-extremal with respect to all the volume-preserving deformations of the Riemannian metric. Since their \(\lambda _{1}\)-exremality is stronger than ours, we have the following proposition:

Proposition 2.12

Let G/K be a compact isotropy irreducible homogeneous Kähler manifold. Then the metric is \(\lambda _{1}\)-extremal for all the volume-preserving deformations of the Kähler metric.

We remark that one can also prove this proposition directly from Theorem 2.9, using a similar discussion to that in [30, Section 3]. It is known that the metric of a compact isotropy irreducible homogeneous Kähler manifold is Einstein[35]. An irreducible Hermitian symmetric space of compact type is a compact isotropy irreducible homogeneous Kähler manifold. In fact, the converse also holds. That is, a compact isotropy irreducible homogeneous Kähler manifold is an irreducible Hermitian symmetric space of compact type [18, 34]. Apotolov et al. [1] proved that the metric on a compact homogeneous Käher-Einstein manifold of positive scalar curvature is \(\lambda _{1}\)-extremal within its Kähler class.

Before stating the corollaries of Theorem 2.9, we recall basic facts about the first eigenvalue of the Laplacian on a product of Riemannian manifolds. Let (Mg) and \((M', g')\) be Riemannian manifolds. For a function \(f \in C^{\infty }(M; \textbf{R})\), we define the function \(f \times 1\) on \(M \times M'\) by

$$\begin{aligned} (f \times 1)(x, y) := f(x), \quad (x,y) \in M \times M'. \end{aligned}$$

For a function \(h \in C^{\infty }(M'; \textbf{R})\), we define the function \(1 \times h\) on \(M \times M'\) in a similar manner. Suppose that \(\lambda _{1}(g) \le \lambda _{1}(g')\). Then the first eigenvalue \(\lambda _{1}(g\times g')\) of the product Riemannian manifold \((M \times M', g\times g')\) is equal to \(\lambda _{1}(g)\). \(E_{1}(g \times g')\), the space of \(\lambda _{1}(g \times g')\)-eigenfunctions on \(M \times M'\) is given by

$$\begin{aligned} E_{1}(g \times g') = \left\{ \begin{array}{ll} \text{ span }\{f \times 1, 1 \times h \mid f \in E_{1}(g), h \in E_{1}(g')\} & (\text{ if } \lambda _{1}(g) = \lambda _{1}(g') )\\ \text{ span }\{f \times 1 \mid f \in E_{1}(g)\} & (\text{ if } \lambda _{1}(g) < \lambda _{1}(g') ).\\ \end{array}\right. \end{aligned}$$

The following corollary follows from Theorem 2.9:

Corollary 2.13

Let \((M, J, g, \omega )\) and \((M', J', g', \omega ')\) be compact Kähler manifolds. Assume that \(\lambda _{1}(g) = \lambda _{1}(g')\) and that g and \(g'\) are both \(\lambda _{1}\)-extremal for all the volume-preserving deformations of the Kähler metrics. Then the product Kähler metric \(g \times g'\) on \((M, J) \times (M, J)\) is \(\lambda _{1}\)-extremal for all the volume-preserving deformations of the Kähler metric.

Proof

By hypothesis, there exist finite subsets \(\{f_{j}\} \subset E_{1}(g)\) and \(\{h_{k} \} \subset E_{1}(g')\) such that

$$\begin{aligned} \left\{ \, \begin{aligned}&\sum _{j}H_{M} \left( f_{j}dd^{c}f_{j} \right) = -\omega , \quad \sum _{j}L_{M}(f_{j}) = 0 \\&\sum _{k}H_{M'} \left( h_{k}dd^{c}h_{k} \right) = -\omega ', \quad \sum _{k}L_{M'}(h_{k}) = 0. \\ \end{aligned} \right. \end{aligned}$$

Then one has

$$\begin{aligned}&\sum _{j}L_{M\times M'} (f_{j} \times 1) + \sum _{k} L_{M\times M'} (1 \times h_{k}) \\&\quad = \sum _{j} L_{M}(f_{j}dd^{c}f_{j}) + \sum _{k} L_{M'}(h_{k}) \\&\quad =0. \end{aligned}$$

We also have

$$\begin{aligned}&\sum _{j}H_{M\times M'} \left( (f_{j}\times 1)d_{M\times M'}d^{c}_{M\times M'}(f_{j}\times 1) \right) \\&\quad = \sum _{j} H_{M\times M'} (f_{j}d_{M}d^{c}_{M}f_{j} \oplus 0) \\&\quad = \sum _{j} H_{M}(f_{j}d_{M}d^{c}_{M}f_{j}) \oplus 0 \\&\quad = -\omega \oplus 0. \end{aligned}$$

Similarly, we obtain

$$\begin{aligned} \sum _{k} H_{M\times M'} \left( (1\times h_{k} )d_{M\times M'}d^{c}_{M\times M'}(1 \times h_{k} ) \right) = 0 \oplus -\omega '. \end{aligned}$$

Hence we obtain

$$\begin{aligned} \sum _{j} H\left( (f_{j}\times 1)dd^{c}(f_{j}\times 1) \right) + \sum _{k} H \left( (1\times h_{k} )dd^{c}(1 \times h_{k} ) \right) = -\omega \oplus -\omega ', \end{aligned}$$

where we omit the subscript \(M \times M'\). Thus one concludes that \(\{ f_{j} \times 1 \}_{j} \cup \{1 \times h_{k} \}_{k}\) satisfy the equation (2.11). The proof is completed. \(\square \)

From the above proof, one can immediately obtain the following corollary:

Corollary 2.14

Let \((M, J, g, \omega )\) and \((M', J', g', \omega ')\) be compact Kähler manifolds. Assume that \(\lambda _{1}(g) \ne \lambda _{1}(g')\). Then the product Kähler metric \(g \times g'\) on \((M, J) \times (M', J')\) is not \(\lambda _{1}\)-extremal for all the volume-preserving deformations of the Kähler metric.

The above corollaries also correspond to the Corollary 2.3 in [1], which we quote in the following:

Proposition 2.15

([1]) Let \((M, J, g, \omega )\) and \((M', J', g', \omega ')\) be compact Kähler manifolds. Suppose that g is \(\lambda _{1}\)-extremal for all the deformations of the Kähler metric in its Kähler class and that the inequality \(\lambda _{1}(g) \le \lambda _{1}(g')\) holds. Then the product metric \(g \times g'\) is \(\lambda _{1}\)-extremal for all the deformations of the Kähler metric in its Kähler class on \((M, J) \times (M', J')\).

We can see that \(\lambda _{1}\)-extremality in Corollary 2.13 is stronger than that in Proposition 2.15. To clarify the difference, let us look at the following example:

Example 2.16

Let \((M, J, g, \omega )\), \((M', J', g', \omega ')\) be irreducible Hermitian symmetric spaces of compact type with \(\rho = c \, \omega \) and \(\rho ' = c'\omega '\) for some \(c, c'>0\), where \(\rho \) and \(\rho '\) are the Ricci forms on M and \(M'\) respectively. By Example 2.12, g is \(\lambda _{1}\)-extremal for all the volume-preserving deformations of the Kähler metric and so is \(g'\). The result due to Nagano [25] shows that \(\lambda _{1}(g) = 2c\) and \(\lambda _{1}(g') =2c'\) (see also [31]). By Proposition 2.15, the product Kähler metric \(g\times g'\) on \((M, J) \times (M', J')\) is \(\lambda _{1}\)-extremal within its Kähler class. However, Corollary 2.13 and 2.14 imply that the metric \(g\times g'\) is \(\lambda _{1}\)-extremal for all the volume-preserving deformations of the Kähler metric if and only if \(c=c'\).

The simplest case of this example is the following:

Example 2.17

Let \(g_{FS}\) be the Fubini-Study metric on the complex projective space \(\textbf{C}P^{n}\). Take any \(c>0\). Then the product Kähler metric \(g_{FS} \times c g_{FS}\) on \(\textbf{C}P^{n} \times \textbf{C}P^{n}\) is \(\lambda _{1}\)-extremal in its Kähler class. However, \(g_{FS} \times c g_{FS}\) is \(\lambda _{1}\)-extremal for all the volume-preserving deformations of the Kähler metric if and only if \(c=1\).

El Soufi-Ilias [12] showed that if a metric g of a Riemannian manifold (Mg) is \(\lambda _{k}\)-extremal for volume-preserving conformal deformations, then there exists a finite subset \(\{f_{1}, \cdots , f_{N} \} \subset E_{k}(g)\) such that \(F:=(f_{1}, \cdots f_{N}): M \rightarrow \textbf{R}^{N}\) is a harmonic map to a sphere with constant energy density. For a Riemann surface \((M, J, g, \omega )\), conformal deformations and deformations within the Kähler class are equivalent, so it is natural to expect the following result due to Apostolov-Jakobson-Kokarev [1]:

Proposition 2.18

([1]) Let \((M, J, g, \omega )\) be a compact Riemann surface without boundary. Let \(\{f_{1}, \cdots , f_{N} \} \subset E_{k}(g)\) be eigenfunctions associated with \(\lambda _{k}(g)\). The following two conditions are equivalent:

  1. (1)

    \(F:=(f_{1}, \cdots f_{N}): M \rightarrow \textbf{R}^{N}\) gives a harmonic map to \(S^{N-1}(c)\) with \(|dF|^{2} = c\lambda _{k}(g)\) for some \(c>0\), where \(S^{N-1}(c)\) is the unit sphere of radius c.

  2. (2)

    \(\sum _{j=1}^{N} L(f_{j}) = 0.\)

To prove the result, they used the identity \(dd^{c}\psi = - (\Delta _{g} \psi ) \omega \) and showed that the equation

$$\begin{aligned} \sum _{j} L(f_{j}) =2\lambda _{k}(g) \left( \lambda _{k}(g)\sum _{j}f_{j}^{2}-\sum _{j}|\nabla f_{j}|^{2}\right) = \lambda _{k}(g) \Delta _{g}\left( \sum _{j}f_{j}^{2}\right) \end{aligned}$$

holds for any finite subset \(\{f_{j} \} \subset E_{k}(g)\). Since we have

$$\begin{aligned} H\left( \sum _{j}f_{j}dd^{c}f_{j} \right) = \lambda _{k}(g)H\left( (\sum _{j} f_{j}^{2}) \omega \right) , \end{aligned}$$

one can see that if \(\{f_{j}\} \subset E_{k}(g)\) satisfies \(\sum _{j}L(f_{j})=0\), then the equation \(H(\sum _{j}f_{j}dd^{c}f_{j}) = -\omega \) holds after an appropriate rescaling of \(\{f_{j}\}\). Hence Theorem 2.9 and Theorem 2.10 are equivalent for a Riemann surface. This is also a natural result since deformations within the Kähler class and general Kähler deformations are equivalent for a compact Riemann surface.

3 Complex tori

In view of Theorem 2.9, the harmonic projector H and information of the space of eigenfunctions are important. However, it is hard to find an explicit formula for the harmonic projector H on a general Kähler manifold. It is also hard to determine the space of eigenfunctions explicitly on a general Riemannian manifold. However, H and the eigenfunctions can be written explicitly for a complex torus. Since it is hard to compute (2.11) for arbitrary collection of eigenfunctions, we compute it for the \(L^{2}\)-orthonormal basis of the space of eigenfunctions. From the calculations, we see that the metric on any flat complex torus is \(\lambda _{1}\)-extremal within its Kähler class. In addition, we deduce a condition for the flat metric to be \(\lambda _{1}\)-extremal for all the volume-preserving deformations of the Kähler metric.

Let \(\gamma _{1}, \cdots , \gamma _{2n}\) be vectors in \(\textbf{C}^{n}\) that are linearly independent over \(\textbf{R}\). We denote by \(\Gamma \) the lattice in \(\textbf{C}^{n}\) with basis \(\{ \gamma _{1}, \cdots , \gamma _{2n} \}\). Let \(z_{1}, \cdots , z_{n}\) be the standard complex coordinates of \(\textbf{C}^{n}\) and \(x^{1}, \cdots , x^{2n}\) the real coordinates defined by

$$\begin{aligned} z^{k} = x^{2k-1} + ix^{2k} \quad (k=1, \cdots , n). \end{aligned}$$

The lattice \(\Gamma \) naturally acts on \(\textbf{C}^{n}\) by translation. Then the quotient space \(T_{\Gamma }^{n}:= \textbf{C}^{n}/\Gamma \) becomes a complex manifold in a natural way. \(T_{\Gamma }^{n}\) is called a complex torus. The standard metric on \(\textbf{C}^{n}\) is given by \(\sum _{j=1}^{n} dz^{j} \otimes d\overline{z}^{j}\). This defines a Kähler form

$$\begin{aligned} \widetilde{\omega } := \frac{i}{2} \sum _{j=1}^{n} dz^{j} \wedge d\overline{z}^{j} \end{aligned}$$

on \(\textbf{C}^{n}\). The canonical holomorphic projection \(\textbf{C}^{n} \rightarrow \textbf{C}^{n}/\Gamma = T_{\Gamma }^{n}\) induces a Kähler metric g and a Kähler form \(\omega \) on \(T_{\Gamma }^{n}\). If we express \(w_{1}\), \(w_{2} \in \textbf{C}^{n}\) as

$$\begin{aligned} w_{k} = (w_{k}^{1}, \cdots ,w_{k}^{n}), \quad w_{k} = (u_{k}^{1}, \cdots , u_{k}^{2n}), \quad w_{k}^{j} = u_{k}^{2j-1}+iu_{k}^{2j} \quad (k=1, 2) \end{aligned}$$

in the complex coordinates \((z^{1}, \cdots z^{n})\) and the real coordinates \((x^{1}, \cdots , x^{2n})\) respectively, then the standard inner product of \(w_{1}\) and \(w_{2}\) is given by

$$\begin{aligned} \sum _{j=1}^{2n}u_{1}^{j}u_{2}^{j} = \frac{1}{2}(w_{1}\cdot \overline{w}_{2} + \overline{w}_{1}\cdot w_{2}). \end{aligned}$$

We define the dual lattice of \(\Gamma \), which is denoted by \(\Gamma ^{*}\), by

$$\begin{aligned} \Gamma ^{*}&= \{ w \in \textbf{C}^{n} \mid \frac{1}{2}(v \cdot \overline{w} + \overline{v} \cdot w) \in \textbf{Z} \quad \text{ for } \text{ all } v \in \Gamma \} \\&= \{ w \in \textbf{C}^{n} \mid \text{ exp }\left( \pi i (v \cdot \overline{w} + \overline{v} \cdot w) \right) = 1 \quad \text{ for } \text{ all } v \in \Gamma \}. \end{aligned}$$

For any \(w \in \Gamma ^{*}\), we define a function \(\Phi _{w}: T_{\Gamma }^{n} \rightarrow \textbf{C}\) by

$$\begin{aligned} \Phi _{w}(z) : = \text{ exp }\left( \pi i (z \cdot \overline{w} + \overline{z} \cdot w) \right) . \end{aligned}$$

Then \(\Phi _{w}\) is actually a well-defined function on the complex torus \(T_{\Gamma }^{n}\). It is known that \(\lambda \) is an eigenvalue of the Laplacian on \((T_{\Gamma }^{n}, g)\) if and only if there exists \(w \in \Gamma ^{*} \) such that \(\lambda = 4\pi ^{2}|w|^{2}\), where \(|w|^{2} = w \cdot \overline{w}\). We set

$$\begin{aligned} S(\lambda ) := \{ w \in \Gamma ^{*} \mid \lambda = 4\pi ^{2}|w|^{2} \}. \end{aligned}$$

Then the multiplicity of \(\lambda \) is given by \(\# S(\lambda )\). For \(\lambda \ne 0\), the number \(\# S(\lambda )\) is an even integer and \(S(\lambda )\) can be written as

$$\begin{aligned} S(\lambda ) = \{ \pm w_{1}, \pm w_{2}, \cdots , \pm w_{l(\lambda )} \}, \end{aligned}$$

where each \(w_{\nu }\) \((\nu =1, \cdots ,l(\lambda ))\) is an element of \(\Gamma ^{*}\) with \(\lambda = 4\pi ^{2}|w_{\nu }|^{2}\). Set

$$\begin{aligned} & \varphi _{w}(z):= \sqrt{\frac{2}{\text{ Vol }(T^{n}_{\Gamma })}} \text{ Re }\left( \Phi _{w}(z)\right) = \sqrt{\frac{2}{\text{ Vol }(T^{n}_{\Gamma })}} \cos \left( 2\pi \sum _{k=1}^{2n}x^{k}u^{k} \right) , \\ & \psi _{w}(z):= \sqrt{\frac{2}{\text{ Vol }(T^{n}_{\Gamma })}} \text{ Im }\left( \Phi _{w}(z)\right) = \sqrt{\frac{2}{\text{ Vol }(T^{n}_{\Gamma })}} \sin \left( 2\pi \sum _{k=1}^{2n}x^{k}u^{k}\right) , \end{aligned}$$

where we use the real coordinates \(z= (x^{1}, \cdots , x^{2n})\) and \(w= (u^{1}, \cdots , u^{2n})\). One can verify that \(\{ \varphi _{w_{\nu }}, \psi _{w_{\nu }} \mid \nu =1, \cdots ,l(\lambda ) \}\) is an \(L^{2}\)-orthonormal basis of the real eigenspace \(E(\lambda )\). (For details of the aforementioned spectral property of flat tori, see [29, pp.272-273], for instance.)

Apostolov-Jakobson-Kokarev[1] proved that the metric on a compact homogeneous Kähler-Einstein manifold of positive scalar curvature is \(\lambda _{1}\)-extremal within its Kähler class. We show that the metric on a flat complex torus is also \(\lambda _{1}\)-extremal within its Kähler class.

Proposition 3.1

Let \((T^{n}_{\Gamma }, g)\) be a flat complex torus. Then the metric g is \(\lambda _{1}\)-extremal within its Kähler class.

Proof

In the proof, we use the notations introduced above. By Theorem 2.10, it suffices to show that the \(L^{2}\)-orthonormal basis \(\{ \varphi _{w_{\nu }}, \psi _{w_{\nu }} \mid \nu =1, \cdots ,l(\lambda ) \}\) of \(E_{1}(g)\) satisfy

$$\begin{aligned} \sum _{\nu =1}^{l(\lambda _{1}(g) )} L(\varphi _{w_{\nu }}) + L(\psi _{w_{\nu }}) = 0. \end{aligned}$$
(3.1)

By a straightforward calculation, we have

$$\begin{aligned} |\nabla \varphi _{w_{\nu }} |^{2} = \frac{2}{\text{ Vol }(T^{n}_{\Gamma }) } \left[ 4\pi ^{2}\sum _{j=1}^{2n}(u_{\nu }^{j})^{2} \right] \sin ^{2}\left( 2\pi \sum _{k=1}^{2n}x^{k}u_{\nu }^{k}\right) = \lambda _{1}(g)\psi _{w}^{2}. \end{aligned}$$

Similarly, we have

$$\begin{aligned} |\nabla \psi _{w_{\nu }} |^{2} = \lambda _{1}(g)\varphi _{w}^{2}. \end{aligned}$$

On the other hand, it is easy to obtain

$$\begin{aligned} dd^{c}\varphi _{w_{\nu }} = -2\pi ^{2}i \varphi _{w_{\nu }} \sum _{\alpha , \beta =1}^{n} \overline{w}_{\nu }^{\alpha } w_{\nu }^{\beta } dz^{\alpha } \wedge d\overline{z}^{\beta }. \end{aligned}$$
(3.2)

Hence we obtain

$$\begin{aligned} |dd^{c}\varphi _{w_{\nu }}|^{2}&= 4\pi ^{4}\varphi _{w_{\nu }}^{2} \sum _{\alpha , \beta , \gamma , \zeta } g^{\alpha \overline{\zeta }} g^{\overline{\beta }\gamma } \overline{w}_{\nu }^{\alpha } w_{\nu }^{\beta } \overline{w}_{\nu }^{\gamma } w_{\nu }^{\zeta } \\&= 16\pi ^{4} \varphi _{w_{\nu }}^{2} \sum _{\alpha , \beta } |w_{\nu }^{\alpha }|^{2}|w_{\nu }^{\beta }|^{2} \\&=\lambda _{1}(g)^{2}\varphi _{w_{\nu }}^{2}. \\ \end{aligned}$$

Similarly, we have

$$\begin{aligned} dd^{c}\psi _{w_{\nu }} = -2\pi ^{2}i \psi _{w_{\nu }} \sum _{\alpha , \beta =1}^{n} \overline{w}_{\nu }^{\alpha } w_{\nu }^{\beta } dz^{\alpha } \wedge d\overline{z}^{\beta }. \end{aligned}$$
(3.3)

and

$$\begin{aligned} |dd^{c}\psi _{w_{\nu }}|^{2} =\lambda _{1}(g)^{2}\psi _{w_{\nu }}^{2}. \end{aligned}$$

Thus for each \(\nu \), we have

$$\begin{aligned}&L(\varphi _{w_{\nu }}) + L(\psi _{w_{\nu }}) \nonumber \\&\quad = \left( \lambda _{1}(g)^{2}\varphi _{w_{\nu }} - 2\lambda _{1}(g)^{2}\psi _{w_{\nu }}^{2} + \lambda _{1}(g)^{2}\varphi _{w_{\nu }}^{2} \right) \nonumber \\&\qquad + \left( \lambda _{1}(g)^{2}\psi _{w_{\nu }} - 2\lambda _{1}(g)^{2}\varphi _{w_{\nu }}^{2} + \lambda _{1}(g)^{2}\psi _{w_{\nu }}^{2} \right) \nonumber \\&\quad =0. \end{aligned}$$
(3.4)

Thus (3.1) is proved. \(\square \)

The harmonic projector \(H: \Omega ^{1,1}(T_{\Gamma }^{n}) \rightarrow \mathcal {H}^{1,1}(T_{\Gamma }^{n})\) is given by

$$\begin{aligned} H(\phi ) =\frac{1}{\text{ Vol }(T_{\Gamma }^{n})} \sum _{\alpha ,\beta =1}^{n} \left( \int _{T_{\Gamma }^{n}} \phi _{\alpha \overline{\beta }} d\mu \right) dz^{\alpha } \wedge d\overline{z}^{\beta } \end{aligned}$$

for a (1, 1)-form \(\phi = \sum _{\alpha , \beta =1}^{n} \phi _{\alpha \overline{\beta }} dz^{\alpha } \wedge d\overline{z}^{\beta }\), where \(d\mu \) is the volume form of \((T_{\Gamma }^{n}, g)\). (For details of the Hodge decomposition on a complex torus, see [4, Section 1.4].) Using this fact and Theorem 2.9, we prove the following:

Theorem 3.2

Let \((T_{\Gamma }^{n},g)\) be a flat n-dimensional complex torus. Let \(\{w_{\nu } \}_{\nu =1}^{l(\lambda _{k}(g))}\) be linearly independent vectors in \(\Gamma ^{*}\) satisfying \(\lambda _{k} (g)= 4\pi ^{2}|w_{\nu }|^{2}\). If the flat metric g is \(\lambda _{k}\)-extremal for all the volume-preserving deformations of the Kähler metric, then there exists \(\{R_{\nu } \ge 0\}_{\nu =1}^{l(\lambda _{k}(g))}\) such that the following equations hold:

$$\begin{aligned} \left\{ \, \begin{aligned}&\sum _{\nu =1}^{l(\lambda _{k}(g))} R_{\nu } \overline{w}_{\nu }^{\alpha }w_{\nu }^{\beta } = 0 \quad \text{ for } \quad 1\le \alpha \ne \beta \le n, \\&\sum _{\nu =1}^{l(\lambda _{k}(g))} R_{\nu } |w_{\nu }^{\alpha }|^{2} = 1 \quad \text{ for } \quad 1 \le \alpha \le n.\\ \end{aligned} \right. \end{aligned}$$
(3.5)

For \(k=1\), the existence of such \(\{R_{\nu } \ge 0\}_{\nu =1}^{l(\lambda _{1}(g))}\) is also a sufficient condition for the metric g to be \(\lambda _{1}\)-extremal for all the volume-preserving deformations of the Kähler metric.

Proof

First we prove the fist half of the assertion. By theorem 2.9, we must have

$$\begin{aligned} H \left( \sum _{j=1}^{N} f_{j}dd^{c}f_{j} \right) = -\frac{i}{2} \sum _{\alpha =1}^{n} dz^{\alpha } \wedge d\overline{z}^{\alpha } \end{aligned}$$
(3.6)

for some finite collection of eigenfunctions \(\{f_{j} \}_{j=1}^{N} \subset E_{k}(g)\). Each eigenfunction \(f_{j}\) is of the form

$$\begin{aligned} f_{j}(z) = \sum _{\nu =1}^{l(\lambda _{k}(g))} a_{j\nu } \varphi _{w_{\nu }} + b_{j\nu } \psi _{w_{\nu }} \quad (a_{j\nu }, b_{j\nu } \in \textbf{R}). \end{aligned}$$

Using (3.2) and (3.3), we obtain

$$\begin{aligned} f_{j}dd^{c}f_{j}&= -2\pi ^{2}i \sum _{\alpha , \beta =1}^{n} \sum _{\nu , \tau =1}^{l(\lambda _{k}(g))} \overline{w}_{\nu }^{\alpha }w_{\nu }^{\beta } [ a_{j\nu }a_{j\tau } \varphi _{w_{\nu }} \varphi _{w_{\tau }} + a_{j\nu }b_{j\tau } \varphi _{w_{\nu }} \psi _{w_{\tau }} \\&\quad + a_{j\tau }b_{j\nu } \varphi _{w_{\tau }} \psi _{w_{\nu }} + b_{j\nu } b_{j\tau } \psi _{w_{\nu }} \psi _{w_{\tau }} ] dz^{\alpha } \wedge d\overline{z}^{\beta }. \end{aligned}$$

Hence one obtains

$$\begin{aligned} H(f_{j}dd^{c}f_{j}) = \frac{-2\pi ^{2}i}{\text{ Vol }(T_{\Gamma }^{n})} \sum _{\alpha ,\beta =1}^{n} \sum _{\nu =1}^{l(\lambda _{k}(g))} \overline{w}_{\nu }^{\alpha }w_{\nu }^{\beta }( a_{j\nu }^{2} + b_{j\nu }^{2} ) dz^{\alpha } \wedge d\overline{z}^{\beta }. \end{aligned}$$

Thus one obtains

$$\begin{aligned} \sum _{j=1}^{N} H (f_{j}dd^{c}f_{j}) = \frac{-2\pi ^{2}i}{\text{ Vol }(T_{\Gamma }^{n})} \sum _{\alpha ,\beta =1}^{n} \sum _{\nu =1}^{l(\lambda _{k}(g))} \overline{w}_{\nu }^{\alpha }w_{\nu }^{\beta }(\sum _{j=1}^{N} a_{j\nu }^{2} + b_{j\nu }^{2} ) dz^{\alpha } \wedge d\overline{z}^{\beta }. \end{aligned}$$
(3.7)

Then the equations (3.6) and (3.7) imply that we must have

$$\begin{aligned} \left\{ \, \begin{aligned}&\sum _{\nu =1}^{l(\lambda _{k}(g))} \overline{w}_{\nu }^{\alpha }w_{\nu }^{\beta }(\sum _{j=1}^{N} a_{j\nu }^{2} + b_{j\nu }^{2} ) = 0 \quad \text{ for } \quad 1\le \alpha \ne \beta \le n, \\&\sum _{\nu =1}^{l(\lambda _{k}(g))} |\overline{w}_{\nu }^{\alpha }|^{2}(\sum _{j=1}^{N} a_{j\nu }^{2} + b_{j\nu }^{2} ) = \frac{ \text{ Vol }(T_{\Gamma }^{n}) }{4\pi ^{2}} \quad \text{ for } \quad 1 \le \alpha \le n.\\ \end{aligned} \right. \end{aligned}$$

Setting \(R_{\nu }:= 4\pi ^{2}(\sum _{j=1}^{N} a_{j\nu }^{2} + b_{j\nu }^{2} )/\text{Vol }(T_{\Gamma }^{n}) \), one concludes the first half of the assertion.

Next we prove the second half. We assume the existence of \(\{R_{\nu } \ge 0\}_{\nu =1}^{l(\lambda _{1}(g))}\) satisfying (3.5). We use Theorem 2.9 to prove the proposition. \(\{ \varphi _{w_{\nu }}, \psi _{w_{\nu }} \mid \nu =1, \cdots ,l(\lambda _{1}(g)) \}\) is an \(L^{2}\)-orthonormal basis of \(E(\lambda _{1}(g))\). By (3.4), we immediately have

$$\begin{aligned} L(\sqrt{R_{\nu }} \varphi _{w_{\nu }} ) + L(\sqrt{R_{\nu } }\psi _{w_{\nu }}) = 0 \end{aligned}$$

for each \(\nu \). Hence we have

$$\begin{aligned} \sum _{\nu =1}^{l(\lambda _{1}(g)} L(\sqrt{R_{\nu }} \varphi _{w_{\nu }} ) + L(\sqrt{R_{\nu } }\psi _{w_{\nu }}) = 0. \end{aligned}$$

Thus it suffices to prove

$$\begin{aligned} \sum _{\nu =1}^{l(\lambda _{1}(g))} R_{\nu } \left[ H(\varphi _{w_{\nu }}dd^{c}\varphi _{w_{\nu }}) + H(\psi _{w_{\nu }}dd^{c}\psi _{w_{\nu }}) \right] = -a\omega \end{aligned}$$

for some \(a>0\). Using the equations (3.2) and (3.3), one obtains

$$\begin{aligned}&\sum _{\nu =1}^{l(\lambda _{1}(g))} R_{\nu }\left[ H(\varphi _{w_{\nu }}dd^{c}\varphi _{w_{\nu }}) + H(\psi _{w_{\nu }}dd^{c}\psi _{w_{\nu }}) \right] \\&\quad = \frac{-2\pi ^{2}i}{\text{ Vol }(T_{\Gamma }^{n})} \sum _{\nu =1}^{l(\lambda _{1}(g))} \sum _{\alpha ,\beta =1}^{n} R_{\nu } \left( \int _{T_{\Gamma }^{n}} (\phi _{w_{\nu }}^{2}+\psi _{w_{\nu }}^{2})d\mu \right) \overline{w}_{\nu }^{\alpha } w_{\nu }^{\beta } dz^{\alpha } \wedge d\overline{z}^{\beta }\\&\quad = \frac{-4\pi ^{2}i}{\text{ Vol }(T_{\Gamma }^{n})} \sum _{\nu =1}^{l(\lambda _{1}(g))} \sum _{\alpha ,\beta =1}^{n} R_{\nu } \overline{w}_{\nu }^{\alpha } w_{\nu }^{\beta } dz^{\alpha } \wedge d\overline{z}^{\beta }. \end{aligned}$$

By hypothesis, the proof is completed. \(\square \)

The implication of this theorem is not clear, so we consider simple cases in what follows. First we consider the case where \(\text{ dim }E_{k}(g) =2\), that is, \(l(\lambda _{k}(g)) =1\). Then we have the following corollary:

Corollary 3.3

Let \((T_{\Gamma }^{n},g)\) be a flat n-dimensional complex torus. Suppose that \(\text{ dim }E_{k}(g) =2\) holds for some k. Then the metric g is not \(\lambda _{k}\)-extremal.

Proof

We prove the assertion by contradiction. Assume that the metric g is \(\lambda _{k}\)-extremal. By hypothesis, there exists a pair w, \(-w \in \Gamma ^{*} \subset \textbf{C}^{n}\) uniquely up to sign such that \(\lambda _{k}(g) = 4\pi ^{2}|w|^{2}\). First we show that the vector w is of the form \(w = (0, \cdots , \xi , \cdots , 0)\) for some \(\xi \in \textbf{C}\). The first equation in (3.5) implies \(\overline{w}^{\alpha }w^{\beta }=0\) for any pair of distinct integers \((\alpha ,\beta )\). Since \(w = (w^{1}, \cdots , w^{n})\) is a nonzero vector, we have \(w^{\alpha } \ne 0\) for some \(\alpha \). Let \(w^{j} = u^{2j-1}+iu^{2j}\) for each \(1 \le j \le n\). Then for any \(\beta \ne \alpha \), we have

$$\begin{aligned} u^{2\alpha -1}u^{2\beta -1} = -u^{2\alpha }u^{2\beta } \end{aligned}$$
(3.8)

and

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

Assume that \(u^{2\alpha } \ne 0\) and \(u^{2\beta } \ne 0\). Then by (3.9), there exists \(c \in \textbf{R}\) such that \(u^{2\alpha -1} = cu^{2\alpha }\) and \(u^{2\beta -1} = cu^{2\beta }\). Substituting these for (3.8), one obtains

$$\begin{aligned} c^{2}u^{2\alpha }u^{2\beta } = -u^{2\alpha }u^{2\beta }. \end{aligned}$$

This is a contradiction and so we have \(u^{2\alpha }=0\) or \(u^{2\beta }=0\).

If we have \(u^{2\alpha }=0\), then we must have \(u^{2\alpha -1} \ne 0\) since we now assume \(w^{\alpha } \ne 0\). Hence by (3.9), we have \(u^{2\beta } = 0\). Then (3.8) immediately implies \(u^{2\beta -1} = 0\). Thus we have \(u^{2\beta -1} = u^{2\beta } = 0\), that is \(w^{\beta }=0\).

If we have \(u^{2\beta }=0\), then (3.9) implies that we have \(u^{2\alpha }=0\) or \(u^{2\beta -1}=0\). We have already considered the case where \(u^{2\alpha }=0\). Hence we consider the case where \(u^{2\beta -1}=0\), but this immediately implies \(w^{\beta }=0\).

Thus we conclude that w is of the form \(w = (0, \cdots , \xi , \cdots , 0)\). However, this contradicts with the second equation in (3.5). The proof is completed. \(\square \)

Example 3.4

\({{\hbox {The standard lattice: }} \Gamma = \textbf{Z}^{2n}}\). Consider the standard complex torus \({\textbf {C}}^{n}/\textbf{Z}^{2n}\) with the flat metric g. Let \(\{e_{j} \}_{j=1}^{n}\) be the standard orthonormal basis of \({\textbf {C}}^{n}\). Set

$$\begin{aligned} w_{2k-1} := e_{k}, \quad w_{2k} := ie_{k} \end{aligned}$$

for every \(1 \le k \le n\). Then we have \(S(\lambda _{1}(g)) = \{\pm w_{\nu } \}_{\nu =1}^{2n}\) and \(l(\lambda _{1}(g)) = 2n\). It is clear that (3.5) is equivalent to the condition where \(R_{2k-1}+R_{2k} =1\) for any \(1\le k \le n\) and so the torus \({\textbf {C}}^{n}/\textbf{Z}^{2n}\) satisfies the assumption of Proposition 3.2. Hence the metric g is \(\lambda _{1}\)-extremal for all the volume-preserving deformations of the Kähler metric. This fact is not new since the metric is \(\lambda _{1}\)-extremal for all the volume-preserving metric deformations. This can be seen from Theorem 1.1 and the classical fact that the standard torus admits an isometric minimal immersion into a unit sphere by first eigenfunctions as follows:

$$\begin{aligned}&\textbf{C}^{n}/\textbf{Z}^{2n} \rightarrow S^{4n-1}\left( \sqrt{\frac{n}{2\pi ^{2}} }\right) , \\&(x^{1}, \ldots , x^{2n}) \mapsto \\&\quad \left( \frac{1}{2\pi }\cos (2\pi x^{1}), \frac{1}{2\pi }\sin (2\pi x^{1}), \ldots , \frac{1}{2\pi }\cos (2\pi x^{2n}), \frac{1}{2\pi }\sin (2\pi x^{2n}) \right) . \end{aligned}$$

Example 3.5

The checkerboard lattice. First we consider the (real) 4-dimensional checkerboard lattice \(D_{4}\), which is defined by

$$\begin{aligned} D_{4} := \{ (x^{1}, \ldots , x^{4}) \in \textbf{Z}^{4} \mid x^{1} + \cdots +x^{4} \in 2\textbf{Z} \}. \end{aligned}$$

\(D_{4}\) is self-dual, i.e. \(D_{4} \cong D^{*}_{4}\). The dual lattice \(D_{4}^{*} (\cong D_{4})\) is known to be the lattice in \(\textbf{C}^{2}\) with the basis (1, 0), (0, 1), (i, 0), \(\left( \frac{1+i}{2}, \frac{1+i}{2}\right) \). (See [9, pp.117-120], for instance.) Set

$$\begin{aligned}&w_{1} := (1,0), \quad w_{2}:= (0,1), \quad w_{3} := (i, 0), \quad w_{4} := (0, i), \\&w_{5} :=\left( \frac{1+i}{2}, \frac{1+i}{2} \right) , \quad w_{6}:= \left( \frac{1-i}{2}, \frac{1-i}{2} \right) , \\&w_{7}:= \left( \frac{1+i}{2}, -\frac{1+i}{2} \right) , \quad w_{8} := \left( \frac{1-i}{2} , -\frac{1-i}{2} \right) , \\&w_{9} := \left( \frac{1+i}{2}, \frac{1-i}{2} \right) , \quad w_{10} := \left( \frac{1+i}{2}, - \frac{1-i}{2} \right) , \\&w_{11} := \left( \frac{1-i}{2}, \frac{1+i}{2}\right) , \quad w_{12} := \left( \frac{1-i}{2}, -\frac{1+i}{2}\right) . \end{aligned}$$

If we set \(R_{1} = \cdots =R_{4} = 1/4\), \(R_{5} = \cdots =R_{12} = 1/8\), then it is elementary to check that the equations (3.5) hold for \(k=1\). Hence by Theorem 3.2, the flat metric g on the 2-dimensional complex torus \(\textbf{C}^{2}/D_{4}\) is \(\lambda _{1}\)-extremal for all the volume-preserving deformations of the Kähler metric. This fact is not new since the metric is \(\lambda _{1}\)-extremal for all the volume-preserving metric deformations. This follows from the fact that Lü et al. [20] recently found a 2-parameter family of isometric minimal immersion by the first eigenfunctions from \(\textbf{C}^{2}/D_{4}\) into the unit sphere \(S^{23} \subset \textbf{R}^{24}\). (See Example 1.1 in [20].)

In fact, for any \(m \ge 3\), the checkerboard lattice \(D_{m}\) is defined as a lattice in \(\textbf{R}^{m}\) by

$$\begin{aligned} D_{m} := \{ (x^{1}, \ldots , x^{m}) \in \textbf{Z}^{m} \mid x^{1} + \cdots +x^{m} \in 2\textbf{Z} \}. \end{aligned}$$

We show the following:

Proposition 3.6

For any \(m \ge 3\), the flat torus \(\textbf{R}^{m}/D_{m}\) admits an isometric minimal immersion into a Euclidean sphere.

Proof

The property of \(D_{m}\) should be considered separately for the case \(m=3\), \(m=4\) and \(m \ge 5\). For \(m=3\) and \(m=4\), the assertion has been proved by Lü et al. [20]. (See Example 4.3 in [20] for \(m=3\) and Example 1.1 in [20] for \(m=4\).) Hence it suffices to consider the case where \(m\ge 5\). For \(m \ge 5\), \(D_{m}^{*}\) is a lattice with the basis \(\{e_{j} \}_{j=1}^{m-1} \cup \{ \frac{1}{2}\sum _{k=1}^{m} e_{k} \}\), where \(\{ e_{j} \}_{j=1}^{m}\) is the standard basis in \(\textbf{R}^{m}\). (See [9, p.120], for instance.) The shortest vectors are exactly the 2m vectors \(\{ \pm e_{j} \}_{j=1}^{m}\). Thus we have \(\lambda _{1}(g) = 4\pi ^{2}\) and \(E_{1}(g)\) is spanned by \(\{ \cos (2\pi x^{j}), \sin (2\pi x^{j}) \}_{j=1}^{m}\), where \(\{x^{j} \}_{j=1}^{m}\) is the standard coordinate in \(\textbf{R}^{m}\). It is obvious that the map

$$\begin{aligned}&\textbf{R}^{m}/D_{m} \rightarrow S^{2m-1}\left( \frac{ \sqrt{m} }{2\pi }\right) , \\&(x^{1}, \ldots , x^{m}) \mapsto \\&\quad \left( \frac{1}{2\pi }\cos (2\pi x^{1}), \frac{1}{2\pi }\sin (2\pi x^{1}), \ldots , \frac{1}{2\pi }\cos (2\pi x^{m}), \frac{1}{2\pi }\sin (2\pi x^{m}) \right) \end{aligned}$$

is an isometric minimal immersion. \(\square \)

For a long time, only the standard torus (Example 3.4) had been an example of higher dimensional tori that admit an isometric minimal immersion into a Euclidean sphere by the first eigenfunctions. Very recently, Lü et al. [20] constructed new examples of higher dimensional flat tori that admit an isometric minimal immersion into a Euclidean sphere by the first eigenfunctions. The flat torus \(\textbf{R}^{m}/D_{m}\) \((m\ge 3)\) is a new example.

Example 3.7

For \(a, b \in [1, \infty )\), consider the lattice \(\Gamma _{a,b}\) in \(\textbf{C}^{2}\) with the lattice basis (1, 0), \((a^{-1}\sqrt{-1}, 0)\), (0, 1), \((0, b^{-1}\sqrt{-1})\). Let \(T^{2}_{a,b}\) be the 2-dimensional complex torus determined by \(\Gamma _{a,b}\) with the flat metric \(g_{a,b}\). Let \(\Gamma _{a} \subset \textbf{C}\) be the lattice with the lattice basis (1, 0), \((a^{-1}\sqrt{-1}, 0)\) and \((T^{1}_{a}, h_{a})\) the flat 1-dimensional complex torus determined by \(\Gamma _{a}\). Then \((T^{2}_{a,b}, g_{a,b})\) is the product of \((T^{1}_{a}, h_{a})\) and \((T^{1}_{b}, h_{b})\). We have \(\lambda _{1}(T^{1}_{a}, h_{a}) = 1 = \lambda _{1}(T^{1}_{b}, h_{b})\). Hence Proposition 3.1 and Corollary 2.13 imply that the metric \(g_{a,b}\) on \(T^{2}_{a,b}\) is \(\lambda _{1}\)-extremal for all the volume-preserving deformations of the Kähler metric. However, since we have \(E_{1}(g_{a,b}) = \text{ span }\{ \cos (2\pi x^{1}), \sin (2\pi x^{1}), \cos (2\pi x^{3}), \sin (2\pi x^{3}) \}\), the flat torus \((T^{2}_{a,b}, g_{a,b})\) does not admit an isometric minimal immersion into a Euclidean sphere by first eigenfunctions. Thus \(g_{a,b}\) is not \(\lambda _{1}\)-extremal for all the volume-preserving metric deformations.

Example 3.8

For \(a, b \in [1, \infty )\), consider the lattice \(\widetilde{\Gamma }_{a,b}\) in \(\textbf{C}^{2}\) with the lattice basis (1, 0), \((\sqrt{-1}, 0)\), \((0,a^{-1})\), \((0, b^{-1}\sqrt{-1})\). Let \(\widetilde{T}^{2}_{a,b}\) be the 2-dimensional complex torus determined by \(\widetilde{\Gamma }_{a,b}\) with the flat metric \(\widetilde{g}_{a,b}\). Let \((T^{1}_{std}, h_{std})\) be the flat 1-dimensional complex torus determined by the lattice with the lattice basis (1, 0), \((\sqrt{-1}, 0)\). Let \((T^{1}_{a,b}, h_{a,b})\) be the flat 1-dimensional complex torus determined by the lattice with the lattice basis \((0,a^{-1})\), \((0, b^{-1}\sqrt{-1})\). Then \((\widetilde{T}^{2}_{a,b}, \widetilde{g}_{a,b})\) is the product of \((T^{1}_{std}, h_{std})\) and \((T^{1}_{a,b}, h_{a,b})\). We have \(\lambda _{1}(T^{1}_{std}, h_{std}) = 1\) and \(\lambda _{1}(T^{1}_{a,b}, h_{a,b}) = \text{ min }\{a,b\}\). Hence if we have \(a=1\) or \(b=1\), then Proposition 3.1 and Corollary 2.13 imply that the metric \(\widetilde{g}_{a,b}\) on \(\widetilde{T}^{2}_{a,b}\) is \(\lambda _{1}\)-extremal for all the volume-preserving deformations of the Kähler metric. On the other hand, if we have \(a>1\) and \(b>1\), then by Corollary 2.14, \(\widetilde{g}_{a,b}\) is not \(\lambda _{1}\)-extremal for all the volume-preserving deformations of the Kähler metric.

In Example 3.7 and Example 3.8, if we ignore the complex structure on \(\textbf{C}^{2}\) and regard \(\textbf{C}^{2}\) as \(\textbf{R}^{4}\), then we have \(\Gamma _{a,b} \cong \widetilde{\Gamma }_{a,b}\). However, whether the flat metric is \(\lambda _{1}\)-extremal is different in Example 3.7 and Example 3.8. Hence Example 3.7 and Example 3.8 show that the notion of \(\lambda _{1}\)-extremality actually depends on the complex structure.

Finally we give a 1-parameter family of 2-dimensional complex tori whose flat metrics are not \(\lambda _{1}\)-extremal for all the volume-preserving deformations of the Kähler metric.

Example 3.9

For \(\pi /3< \theta <\pi /2\), we consider the lattice \(\Gamma _{\theta } \subset \textbf{C}^{2}\) with the lattice basis (1, 0), \((\cos \theta , \sin \theta )\), \((\sqrt{-1}, 0)\), \((\sqrt{-1}\cos \theta , \sqrt{-1}\sin \theta )\). Let \(g_{\theta }\) be the flat metric on \(\textbf{C}^{2}/\Gamma _{\theta }\). It is straightforward to check that the dual lattice \(\Gamma _{\theta }^{*}\) is the lattice with the basis \(w_{1}:= (1, -\cos \theta /\sin \theta )\), \(w_{2}:= (0, 1/\sin \theta )\), \(w_{3}:=(\sqrt{-1}, -\cos \theta /\sin \theta )\), \(w_{4}:=(0, \sqrt{-1}/\sin \theta )\). Then we have \(S(\lambda _{1}(g_{\theta })) = \{ \pm w_{\nu } \}_{\nu =1}^{4}\) and so the multiplicity of \(\lambda _{1}(g_{\theta })\) is 8. If g is \(\lambda _{1}\)-extremal, then Theorem 3.2 implies that there exists \(\{ R_{\nu } \}_{\nu =1}^{4}\) such that

$$\begin{aligned} -\frac{\cos \theta }{\sin \theta }(R_{1}+R_{3}) = 0, \quad R_{1}+R_{3}= 1, \quad \frac{1}{\sin ^{2}\theta }(R_{2}+R_{4}) =1. \end{aligned}$$

Since we have \(\pi /3< \theta <\pi /2\), this is a contradiction. Hence \(g_{\theta }\) is not \(\lambda _{1}\)-extremal. \(\textbf{C}^{2}/\Gamma _{\theta }\) is not a product of 1-dimensional flat complex tori. In fact, assume that \(\textbf{C}^{2}/\Gamma _{\theta }\) is a product of \((T_{1},h_{1})\) and \((T_{2}, h_{2})\), where each is a 1-dimensional flat complex torus. If we had \(\lambda _{1}(h_{1}) = \lambda _{1}(h_{2})\), then by Proposition 3.1 and Corollary 2.13, \(g_{\theta }\) would be \(\lambda _{1}\)-extremal. Hence we have \(\lambda _{1}(h_{1}) \ne \lambda _{1}(h_{2})\). We may assume \(\lambda _{1}(h_{1}) < \lambda _{1}(h_{2})\). Then the multiplicity of of \(\lambda _{1}(h_{1})\) is equal to that of \(\lambda _{1}(g_{\theta })\), that is, 8. This is a contradiction since the multiplicity of the first eigenvalue of a 1-dimensional flat complex torus is at most 6 (see [11], for example). Thus \(\textbf{C}^{2}/\Gamma _{\theta }\) is not a product of 1-dimensional complex tori. Let \(\widetilde{\Gamma }_{\theta } \subset \textbf{R}^{2}\) be the lattice with the lattice basis (1, 0), \((\cos \theta , \sin \theta )\). Let \((\textbf{R}^{2}/\widetilde{\Gamma }_{\theta }, h_{ \theta })\) be the flat real 2-dimensional torus. If we ignore the complex structure on \(\textbf{C}^{2}\) and regard it as \(\textbf{R}^{4}\), then \((\textbf{R}^{4}/\Gamma _{\theta }, g_{\theta })\) is a Riemannian product of \((\textbf{R}^{2}/\widetilde{\Gamma }_{\theta }, h_{ \theta })\) and \((\textbf{R}^{2}/\widetilde{\Gamma }_{\theta }, h_{ \theta })\).