This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Pluriclosed Flow and Hermitian-Symplectic Structures

Abstract.

We show pluriclosed flow preserves the Hermitian-symplectic structures. And we observe that it can actually become a flow of Hermitian-symplectic forms when an extra evolution equation determined by the Bismut-Ricci form is considered. Moreover, we get a topological obstruction to the long-time existence in arbitrary dimension.

1. Introduction

A Hermitian-symplectic structure, which is introduced by Tian and Streets in [17], is a symplectic form such that its (1,1)(1,1)-part ω\omega is positive. In fact, the (1,1)(1,1)-part ω\omega is pluriclosed. So we can consider its deformation along the pluriclosed flow, which is also introduced by Tian and Streets in [17].

Formally, the pluriclosed flow is

{tω=ρ1,1(ω)ω(0)=ω0\left\{\begin{aligned} &\frac{\partial}{\partial t}\omega=-\rho^{1,1}(\omega)\\ &\omega(0)=\omega_{0}\end{aligned}\right.

Here, ρ1,1\rho^{1,1} is the (1,1)(1,1)-part of Bismut-Ricci form ρ\rho which will be defined later.

We say a pluriclosed metric ω\omega is Hermitian-symplectic if it can be extended to a Hermitian-symplectic form as the (1,1)(1,1)-part. In other words, there exists a (2,0)(2,0) form φ\varphi such that Ωφ+ω+φ¯\Omega\triangleq\varphi+\omega+\bar{\varphi} is a Hermitian-symplectic form.

Now, we can state our first theorem.

Theorem 1.1.

Pluriclosed flow preserves Hermitian-symplectic structures.

That means if the initial metric ω0\omega_{0} is Hermitian-symplectic, then the solution ω(t)\omega(t) is Hermitian-symplectic at any time.

Notice that when ω\omega is Hermitian-symplectic, the selection of (2,0)(2,0) form φ\varphi is not unique. In fact, different selections differ by a closed (2,0)(2,0) form. The pluriclosed flow only gives the deformation of (1,1)(1,1)-part ω\omega. Although the theorem above tells us that the solution ω(t)\omega(t) is Hermitian-symplectic, there is not a canonical way to choose φ(t)\varphi(t), the smooth one-parameter family of (2,0)(2,0) forms. One approach is to determine the smooth one-parameter family φ(t)\varphi(t) using an evolution equation.

By an observation, we find the following equation satisfies the requirements.

(1.1) {tω=ρ1,1(ω)tφ=trg(¯φ)ω(0)=ω0,φ(0)=φ0.\left\{\begin{aligned} &\frac{\partial}{\partial t}\omega=-\rho^{1,1}(\omega)\\ &\frac{\partial}{\partial t}\varphi=-\partial\text{\rm tr}_{g}(\bar{\partial}\varphi)\\ &\omega(0)=\omega_{0},\quad\varphi(0)=\varphi_{0}.\end{aligned}\right.

For any selection of (2,0)(2,0) form φ0\varphi_{0}, it is a flow of Hermitian-symplectic forms with initial data Ω0=φ0+ω0+φ¯0\Omega_{0}=\varphi_{0}+\omega_{0}+\bar{\varphi}_{0}.

The idea of flowing a (2,0)(2,0) form by the (2,0)(2,0)-part of Bismut-Ricci form along the pluriclosed flow first appears in [13] (can also see [6, 7]). They consider the system consisting of pluriclosed flow and an evolution equation of (2,0)(2,0) form

tβ=ρ2,0(ω).\displaystyle\frac{\partial}{\partial t}\beta=-\rho^{2,0}(\omega).

And direct computation shows that the evolution equation of (2,0)(2,0)-part in (1.1) is exactly the evolution equation above under the Hermitian-symplectic assumption. So we can rewrite (1.1) in a simple form

tΩ=ρ(Ω1,1).\displaystyle\frac{\partial}{\partial t}\Omega=-\rho(\Omega^{1,1}).

In general, for a Hermitian-symplectic form Ω=φ+ω+φ¯\Omega=\varphi+\omega+\bar{\varphi}, we can define a positive quantity

𝒱(Ω)1(k!)2k0(φk,φk)ω.\displaystyle\mathcal{V}(\Omega)\triangleq\frac{1}{(k!)^{2}}\sum_{k\geq 0}(\varphi^{k},\varphi^{k})_{\omega}.

Here (,)ω(\cdot,\cdot)_{\omega} is the inner product induced by ω\omega. And considering the function 𝒱(t)𝒱(Ω(t))\mathcal{V}(t)\triangleq\mathcal{V}(\Omega(t)) along a solution to (1.1), we show that it is actually a polynomial about tt whose coeifficients are determined only by initial data and the underlying complex manifold.

More precisely, we have the following theorem.

Theorem 1.2.

For a solution to (1.1) with initial data Ω0\Omega_{0}, the function 𝒱(t)=𝒱(Ω(t))\mathcal{V}(t)=\mathcal{V}(\Omega(t)) is a polynomial of degree at most nn

𝒱(t)=i=0naiti\displaystyle\mathcal{V}(t)=\sum_{i=0}^{n}a_{i}t^{i}

with coefficients

ai=1i!k02k+in1(k!)2(n-2k-i)!Mφ0kφ¯0kω0n2ki(1¯logdetg0)i\displaystyle a_{i}=\frac{1}{i!}\sum_{\begin{subarray}{c}k\geq 0\\ 2k+i\leq n\end{subarray}}\frac{1}{(k!)^{2}(n\mathord{-}2k\mathord{-}i)!}\int_{M}\varphi_{0}^{k}\wedge\bar{\varphi}_{0}^{k}\wedge\omega^{n-2k-i}_{0}\wedge(\sqrt{-1}\partial\bar{\partial}\log\det g_{0})^{i}

Specially, we have

an=1n!(1)nMc1(M)n=1n!(1)nc1n\displaystyle a_{n}=\frac{1}{n!}(-1)^{n}\int_{M}c_{1}(M)^{n}=\frac{1}{n!}(-1)^{n}c_{1}^{n}

in which c1(M)c_{1}(M) is the first Chern class of (M2n,J)(M^{2n},J).

Thus 𝒱(t)\mathcal{V}(t) gives an obstruction to how long solutions exist for 𝒱(t)\mathcal{V}(t) is strictly positive. Especially, we get a topological obstruction to global solutions, (i.e. solutions exist on [0,+)[0,+\infty)), to pluriclosed flow with Hermitian-symplectic initial data.

Theorem 1.3.

If there exists a global solution to pluriclosed flow with Hermitian-symplectic initial data on (M2n,J)(M^{2n},J), then (1)nc1n(-1)^{n}c_{1}^{n} must be a nonnegative integer.

One of the most important motivation for introducing pluriclosed flow is to study non-Kähler manifolds, especially to classify Kodaira’s class VII surfaces (see [16, 17]). We hope a pluriclosed metric may deform to some canonical metric along this flow. A natural possible canonical metric is static metric, which is introduced in [17]. Recall that a metric is static if there exists a real number λ\lambda such that

ρ1,1(ω)=λω\displaystyle\rho^{1,1}(\omega)=\lambda\omega

Actually, static metric is Hermitian-symplectic when λ0\lambda\neq 0. Although Tian and Streets[17] have proven non-Kähler surfaces does not admit Hermitian-symplectic structures using the classification of compact complex surfaces. But it is still not clear for high dimensional cases.

Now we can have another approach to study the Hermitian-symplectic structures, which is to flow them along (1.1). In the case of compact complex surfaces, we prove that the function (φ(t),φ(t))ω(t)(\varphi(t),\varphi(t))_{\omega(t)} is monotonically decreasing. As a corollary, we show again that non-Kähler surfaces do not exist static metric with λ<0\lambda<0 without help of classification of compact complex surfaces.

Here is an outline of the rest of this paper. In section 2 we recall some basic notions about pluriclosed flow and Hermitian-symplectic structures. In section 3 we show that pluriclosed flow preserves Hermitian-symplectic structures and study the flow (1.1). In section 4 we give a topological obstruction to global solutions using the function 𝒱(t)\mathcal{V}(t) defined above. Finally in section 5 we do some discussion in surfaces case.

Acknowledgements. I would like to express my sincere gratitude to my advisor Professor Gang Tian, for his helpful suggestions and patient guidance. And thanks to Professor Jeffrey Streets for his helpful and detailed comments.

2. Preliminary

In this section, we review some notions about pluriclosed flow and Hermitian-symplectic structures.

2.1. Pluriclosed flow

Consider a Hermitian manifold (M2n,J,g)(M^{2n},J,g). Bismut[3] shows that there exists a unique real connection \nabla compatible with complex structure and Hermitian metric such that its torsion tensor is totally skew-symmetric. That means

J=g=0\nabla J=\nabla g=0

and

H(X,Y,Z)=H(Y,X,Z)=H(X,Z,Y).H(X,Y,Z)=-H(Y,X,Z)=-H(X,Z,Y).

Here

H(X,Y,Z)=g(XYYX[X,Y],Z)H(X,Y,Z)=g(\nabla_{X}Y-\nabla_{Y}X-[X,Y],Z)

is the tensor induced by torsion operator.

We denote ω\omega the fundamental form corresponding to gg. A metric is pluriclosed if its fundamental form is pluriclosed, i.e. ¯ω=0\partial\bar{\partial}\omega=0. The torsion tensor HH corresponding to Bismut connection is a real 3-form and we have the formula H=dcω=dω(J,J,J)H=-d^{c}\omega=d\omega(J\cdot,J\cdot,J\cdot), in which dc=1(¯)d^{c}=\sqrt{-1}(\bar{\partial}-\partial) is the conjugate differential operator. Then a metric ω\omega is pluriclosed if and only if its Bismut torsion tensor is closed for ddcω=21¯ω-dd^{c}\omega=-2\sqrt{-1}\partial\bar{\partial}\omega. Gauduchon[8] shows that every complex surface admits pluriclosed metric.
We denote the Riemann operator corresponding to Bismut connection by

R(X,Y)Z=XYZYXZ[X,Y]Z.R(X,Y)Z=\nabla_{X}\nabla_{Y}Z-\nabla_{Y}\nabla_{X}Z-\nabla_{[X,Y]}Z.

For Bismut connection, we can also define a Ricci-type form by

ρ(X,Y)=i=1ng(R(X,Y)ei,Jei)=1i=1ng(R(X,Y)Zi,Z¯i).\rho(X,Y)=\sum_{i=1}^{n}g(R(X,Y)e_{i},Je_{i})=\sqrt{-1}\sum_{i=1}^{n}g(R(X,Y)Z_{i},\bar{Z}_{i}).

Here, {ei,Jei}\{e_{i},Je_{i}\} is any local orthonormal real basis. And {Zi}\{Z_{i}\} and {Z¯i}\{\bar{Z}_{i}\} are unitary bases determined by

Zi=22(ei1Jei),Z¯i=22(ei+1Jei).Z_{i}=\frac{\sqrt{2}}{2}(e_{i}-\sqrt{-1}Je_{i}),\quad\bar{Z}_{i}=\frac{\sqrt{2}}{2}(e_{i}+\sqrt{-1}Je_{i}).

It is easy to check that the definition of ρ\rho is independent of choice of basis.

Now, we are in a position to define pluriclosed flow.

{tω(t)=ρ1,1(ω(t))ω(0)=ω0.\left\{\begin{aligned} &\frac{\partial}{\partial t}\omega(t)=-\rho^{1,1}(\omega(t))\\ &\omega(0)=\omega_{0}.\end{aligned}\right.

Here, ρ1,1\rho^{1,1} is the (1,1)-part of Ricci form, that is

ρ1,1(X,Y)=12(ρ(X,Y)+ρ(JX,JY)).\rho^{1,1}(X,Y)=\frac{1}{2}\left(\rho(X,Y)+\rho(JX,JY)\right).

Tian and Streets[17] show that this flow preserves pluriclosed condition. That means the solution is pluriclosed at any time if the initial metric is pluriclosed. Meanwhile, they[17, 18, 19] prove it is a strictly parabolic systems under the pluriclosed assumption and give the short time existence and some basic regularity results. More results about regularity and long-time existence can be found in [6, 9, 13, 14, 15].

Remark 2.1.

In local coordinates, we can express pluriclosed flow by Hodge operator as

(2.1) {tω=ω+¯¯ω+1¯logdetgω(0)=ω0.\left\{\begin{aligned} &\frac{\partial}{\partial t}\omega=\partial\partial^{*}\omega+\bar{\partial}\bar{\partial}^{*}\omega+\sqrt{-1}\partial\bar{\partial}\log\det g\\ &\omega(0)=\omega_{0}.\end{aligned}\right.

Here, \partial^{*} is the dual operator of \partial corresponding to the inner product induced by metric ω(t)\omega(t).

2.2. Hermitian-symplectic structure

In this subsection, we will review the Hermitian-symplectic structure introduced by Tian and Streets[17].
A Hermitian-symplectic form on a complex manifold (M2n,J)(M^{2n},J) is a real symplectic form such that its (1,1)-part is positive. In other words, a Hermitian-symplectic form is a real 2-form Ω\Omega such that dΩ=0d\Omega=0 and in local coordinates {zi}\{z_{i}\},

Ω1,1=1gij¯dzidz¯j.\Omega^{1,1}=\sqrt{-1}g_{i\bar{j}}dz^{i}d\bar{z}^{j}.

in which (gij¯)n×n(g_{i\bar{j}})_{n\times n} is a positive definite Hermitian matrix. Thus, the (1,1)-part of a Hermitian-symplectic form is the fundamental form corresponding to some metric.

A Kähler form is naturally a Hermitian-symplectic form. Tian and Streets[17] prove that a complex surface admits Hermitian-symplectic structures if and only if it is Kähler. But it is not clear if is it still true for dimension higher than two.

For Ω\Omega is real and closed, we have

{¯Ω2,0+Ω1,1=0Ω2,0=0.\left\{\begin{aligned} &\bar{\partial}\Omega^{2,0}+\partial\Omega^{1,1}=0\\ &\partial\Omega^{2,0}=0.\end{aligned}\right.

Here, Ω=Ω2,0+Ω1,1+Ω2,0¯\Omega=\Omega^{2,0}+\Omega^{1,1}+\overline{\Omega^{2,0}} and Ω1,1\Omega^{1,1} are real forms. Applying ¯\bar{\partial} to the first equation, we get the fact that the (1,1)-part of a Hermitian-symplectic form is pluriclosed.
On the other hand, given a pluriclosed metric ω\omega, consider the equations with respect to (2,0)-form φ\varphi

(2.2) {¯φ+ω=0φ=0.\left\{\begin{aligned} &\bar{\partial}\varphi+\partial\omega=0\\ &\partial\varphi=0.\end{aligned}\right.

If a (2,0)-form φ\varphi is a solution, then ω~φ+ω+φ¯\widetilde{\omega}\triangleq\varphi+\omega+\overline{\varphi} is a Hermitian-symplectic form. And in this case we call the pluriclosed form ω\omega can be extended to a Hermitian-symplectic form. Notice that if equations (2.2) has a solution φ\varphi, then φ+ϕ\varphi+\phi is also a solution for any closed (2,0)(2,0)-form ϕ\phi.

3. A parabolic flow of Hermitian-symplectic forms

In this section, we first directly prove that pluriclosed flow preserves Hermitian-symplectic structures. Then we study the flow of Hermitian-symplectic forms (1.1). We will show this flow is parabolic. And by the standard theory of parabolic equations, we get the short-time existence and uniqueness of solutions. As a corolarry, we prove again that pluriclosed flow preserves Hermitian-symplectic structures. Meanwhile, we will see that if the initial Hermitian-symplectic form is degenerate, i.e. its (1,1)-part is Kähler, then the flow of (1,1)-part will degenerate to Kähler-Ricci flow and the (2,0)-part is constant.

Firstly, we prove the next theorem.

Theorem 3.1.

Pluriclosed flow preserves Hermitian-symplectic structures.

Proof.

Assume the solution ω(t)\omega(t) with Hermitian-symplectic initial data exists on [0,T)[0,T). Applying \partial to the first equation of (2.1), we get a evolution equation of ω\partial\omega

tω=¯¯ω.\displaystyle\frac{\partial}{\partial t}\partial\omega=\partial\bar{\partial}\bar{\partial}^{*}\omega.

And we will use it to find the (2,0)(2,0)-part φ(t)\varphi(t) satisfying ω(t)=¯φ(t)\partial\omega(t)=-\bar{\partial}\varphi(t) for all t[0,T)t\in[0,T).
Choose a fixed background metric hh and denote (,)h(\cdot,\cdot)_{h} the induced inner product on space of differential forms.

For a test (2,1)(2,1)-form η\eta, derivative (ω(t),η)h(\partial\omega(t),\eta)_{h} with respect to tt

ddt(ω(t),η)h\displaystyle\frac{d}{dt}(\partial\omega(t),\eta)_{h} =(tω(t),η)h\displaystyle=(\frac{\partial}{\partial t}\partial\omega(t),\eta)_{h}
=(¯¯tω(t),η)h\displaystyle=(\partial\bar{\partial}\bar{\partial}^{*_{t}}\omega(t),\eta)_{h}
=(¯¯tω(t),η)h\displaystyle=(-\bar{\partial}\partial\bar{\partial}^{*_{t}}\omega(t),\eta)_{h}
=(¯tω(t),¯hη)h\displaystyle=(-\partial\bar{\partial}^{*_{t}}\omega(t),\bar{\partial}^{*_{h}}\eta)_{h}

By Newton-Leibniz formula,

(ω(t),η)h(ω(0),η)h\displaystyle(\partial\omega(t),\eta)_{h}-(\partial\omega(0),\eta)_{h} =0tdds(ω(s),η)h𝑑s\displaystyle=\int_{0}^{t}\frac{d}{ds}(\partial\omega(s),\eta)_{h}ds
=0t(¯sω(s),¯hη)h𝑑s\displaystyle=\int_{0}^{t}(-\partial\bar{\partial}^{*_{s}}\omega(s),\bar{\partial}^{*_{h}}\eta)_{h}ds
=(0t¯sω(s)ds,¯hη)h.\displaystyle=-(\int_{0}^{t}\partial\bar{\partial}^{*_{s}}\omega(s)ds,\bar{\partial}^{*_{h}}\eta)_{h}.

This last equal is because the order of integration can be exchanged.

Note ω0\omega_{0} can extend to a Hermitian-symplectic form. Thus there is a (2,0)(2,0)-form φ0\varphi_{0} satisfying ω0=¯φ0\partial\omega_{0}=-\bar{\partial}\varphi_{0} and ϕ0=0\partial\phi_{0}=0.
Then we have

(ω(t),η)0=(φ0+0t¯sω(s)ds,¯0η)0.\displaystyle(\partial\omega(t),\eta)_{0}=-(\varphi_{0}+\int_{0}^{t}\partial\bar{\partial}^{*_{s}}\omega(s)ds,\bar{\partial}^{*_{0}}\eta)_{0}.

Thus, we can define φ(t)\varphi(t) as

φ(t)=φ0+0t¯sω(s)ds.\displaystyle\varphi(t)=\varphi_{0}+\int_{0}^{t}\partial\bar{\partial}^{*_{s}}\omega(s)ds.

So ω(t)=¯φ(t)\partial\omega(t)=-\bar{\partial}\varphi(t) and the only thing is to show φ(t)=0\partial\varphi(t)=0. It is easy to check by noticing ¯φ(0)=¯φ0=0\bar{\partial}\varphi(0)=\bar{\partial}\varphi_{0}=0. Thus we complete our proof. ∎

Remark 3.2.

From the proof above we know that if φ0\varphi_{0} is \partial-exact, i.e. there exists a (1,0)(1,0)-form α0\alpha_{0} such that α0=φ0\partial\alpha_{0}=\varphi_{0}. Then φ(t)\varphi(t) is \partial-exact and actually we can choose α(t)\alpha(t) as

α(t)=α0+0t¯sω(s)𝑑s.\displaystyle\alpha(t)=\alpha_{0}+\int_{0}^{t}\bar{\partial}^{*_{s}}\omega(s)ds.

Meanwhile, we have ω(t)=¯α(t)\partial\omega(t)=\partial\bar{\partial}\alpha(t). In fact, it is the case when [ω0]=0HBC2,1(M,)[\partial\omega_{0}]=0\in\text{H}^{2,1}_{BC}(M,\mathbb{C}). Here HBC2,1(M,)\text{H}^{2,1}_{BC}(M,\mathbb{C}) is the Bott-Chern cohomology group.

By an observation, we find that the evolution equation of (2,0)(2,0)-part is exactly determined by the (2,0)(2,0)-part of Bismut-Ricci form. Thus, we can actually regard pluriclosed flow as a flow of Hermitian-symplectic forms by adding the extra evolution equation of (2,0)(2,0)-part.

Definition 3.3.

Let Ω0=φ0+ω0+φ¯0\Omega_{0}=\varphi_{0}+\omega_{0}+\bar{\varphi}_{0} be a Hermitian-symplectic form, we can define a flow with initial data Ω0\Omega_{0} as

(3.1) {tω=ω+¯¯ω+1¯logdetgtφ=trg(¯φ)ω(0)=ω0,φ(0)=φ0.\left\{\begin{aligned} &\frac{\partial}{\partial t}\omega=\partial\partial^{*}\omega+\bar{\partial}\bar{\partial}^{*}\omega+\sqrt{-1}\partial\bar{\partial}\log\det g\\ &\frac{\partial}{\partial t}\varphi=-\partial\text{\rm tr}_{g}(\bar{\partial}\varphi)\\ &\omega(0)=\omega_{0},\quad\varphi(0)=\varphi_{0}.\end{aligned}\right.

in which, trg(β)1Λωβ\text{\rm tr}_{g}(\beta)\triangleq\sqrt{-1}\Lambda_{\omega}\beta for arbitrary differential forms β\beta. And Λω\Lambda_{\omega} is the adjoint operator of Lefschetz operator Lω:βωβL_{\omega}:\beta\mapsto\omega\wedge\beta

We have a few things to say about this definition. Recall that Ω0\Omega_{0} is a Hermitian-symplectic form, so its (1,1)-part ω0\omega_{0} is pluriclosed and we denote g0g_{0} the metric corresponding to ω0\omega_{0}. The first equation is just the pluriclosed flow with initial metric ω0\omega_{0}, thus ω(t)\omega(t) exists uniquely for at least a short time. So the evolution equation of φ\varphi is well-defined when g(t)g(t) exists.

Then we show φ(t)\varphi(t) exists uniquely in short time. That is the lemma below.

Lemma 3.4.

Given a smooth one-parameter family of metrics g(t)g(t) in time interval [0,T)[0,T). Consider the flow of (p,0)-forms α\alpha with initial data α0\alpha_{0} satisfying α0=0\partial\alpha_{0}=0

{tα=trg(¯α)α(0)=α0\left\{\begin{aligned} &\frac{\partial}{\partial t}\alpha=-\partial\text{\rm tr}_{g}(\bar{\partial}\alpha)\\ &\alpha(0)=\alpha_{0}\end{aligned}\right.

Then, there exists a unique \partial-closed solution α(t)\alpha(t) in [0,ε)[0,\varepsilon), in which ε<T\varepsilon<T is a positive number depending on α0\alpha_{0}.

Proof.

We denote 𝒜p,q\mathcal{A}^{p,q} the sheaf of differential (p,q)(p,q)-forms and 𝒜p,q={α𝒜p,q|α=0}\mathcal{A}^{p,q}_{\partial}=\{\alpha\in\mathcal{A}^{p,q}|\partial\alpha=0\} the set of \partial-closed (p,q)(p,q)-forms. And notice the differential operator

Φ=trg(¯)|𝒜p,0:𝒜p,0𝒜p,0\displaystyle\Phi=-\partial\text{tr}_{g}(\bar{\partial}\cdot)\Big{|}_{\mathcal{A}_{\partial}^{p,0}}:\mathcal{A}_{\partial}^{p,0}\to\mathcal{A}_{\partial}^{p,0}

is a linear differential operator from 𝒜p,0\mathcal{A}_{\partial}^{p,0} to itself.

We claim that Φ\Phi is actually elliptic.

In local complex coordinates, we assume that

α\displaystyle\alpha =i1<<ipαi1ipdzi1dzip\displaystyle=\sum_{i_{1}<\cdots<i_{p}}\alpha_{i_{1}\cdots i_{p}}dz^{i_{1}}\wedge\cdots\wedge dz^{i_{p}}
=1p!αi1ipdzi1dzip\displaystyle=\frac{1}{p!}\alpha_{i_{1}\cdots i_{p}}dz^{i_{1}}\wedge\cdots\wedge dz^{i_{p}}

where the subscripts are antisymmetric. By direct computation, we have

Φ(α)\displaystyle\Phi(\alpha) =trg(¯α)\displaystyle=-\partial\text{tr}_{g}(\bar{\partial}\alpha)
=trg(1p!αi1ip,t¯dz¯tdzi1dzip)\displaystyle=-\partial\text{tr}_{g}(\frac{1}{p!}\alpha_{i_{1}\cdots i_{p},\bar{t}}d\bar{z}^{t}\wedge dz^{i_{1}}\wedge\cdots\wedge dz^{i_{p}})
=1(p1)!(gt¯sαsi2ip,t¯dzi2dzip)\displaystyle=\frac{1}{(p-1)!}\partial(g^{\bar{t}s}\alpha_{si_{2}\cdots i_{p},\bar{t}}dz^{i_{2}}\wedge\cdots\wedge dz^{i_{p}})
=1(p1)!(gt¯sαsi2ip,t¯),i1dzi1dzip\displaystyle=\frac{1}{(p-1)!}(g^{\bar{t}s}\alpha_{si_{2}\cdots i_{p},\bar{t}})_{,i_{1}}dz^{i_{1}}\wedge\cdots\wedge dz^{i_{p}}

The terms containing second order derivatives with respect to α\alpha are

2nd order =1(p1)!gt¯sαi1ip1s,ipt¯dzi1dzip\displaystyle=\frac{1}{(p-1)!}g^{\bar{t}s}\alpha_{i_{1}\cdots i_{p-1}s,i_{p}\bar{t}}dz^{i_{1}}\wedge\cdots\wedge dz^{i_{p}}
=i1<<ipgt¯sa=1p(1)paαi1ia^ips,iat¯dzi1dzip\displaystyle=\sum_{i_{1}<\cdots<i_{p}}g^{\bar{t}s}\sum_{a=1}^{p}(-1)^{p-a}\alpha_{i_{1}\cdots\widehat{i_{a}}\cdots i_{p}s,i_{a}\bar{t}}dz^{i_{1}}\wedge\cdots\wedge dz^{i_{p}}
=i1<<ipgt¯sαi1ip,st¯dzi1dzip\displaystyle=\sum_{i_{1}<\cdots<i_{p}}g^{\bar{t}s}\alpha_{i_{1}\cdots i_{p},s\bar{t}}dz^{i_{1}}\wedge\cdots\wedge dz^{i_{p}}

The last line uses the condition α=0\partial\alpha=0, which is

αi1ip,sa=1p(1)paαi1ia^ips,ia=0\displaystyle\alpha_{i_{1}\cdots i_{p},s}-\sum_{a=1}^{p}(-1)^{p-a}\alpha_{i_{1}\cdots\widehat{i_{a}}\cdots i_{p}s,i_{a}}=0

in local coordinates. So we prove Φ\Phi is elliptic. Then, by the standard theorem of parabolic equations, we get the short-time existence and uniqueness of solutions. ∎

Applying Lemma 3.4 in the case of p=2p=2 and note that φ0\varphi_{0} is \partial-closed for Ω0\Omega_{0} is closed. So we get the short-time existence and uniqueness of φ(t)\varphi(t).

To show the flow (3.1) preserves Hermitian-symplectic structures, we need the next lemma.

Lemma 3.5.

Given a smooth one-parameter family of metrics g(t)g(t) in time interval [0,T)[0,T). Consider the flow of (p,q)(p,q)-form ϕ\phi with initial data ϕ0\phi_{0} satisfying dϕ0=0d\phi_{0}=0

{tϕ=¯trg(ϕ)ϕ(0)=ϕ0\left\{\begin{aligned} &\frac{\partial}{\partial t}\phi=\partial\bar{\partial}\text{\rm tr}_{g}(\phi)\\ &\phi(0)=\phi_{0}\end{aligned}\right.

Then, there exists a unique closed solution ϕ(t)\phi(t) in [0,ε)[0,\varepsilon), in which ε<T\varepsilon<T is a positive number depending on ϕ0\phi_{0}.

Proof.

This proof is similar with Lemma 3.4. Denote 𝒜dp,q={ϕ𝒜p,q|dϕ=0}\mathcal{A}^{p,q}_{d}=\{\phi\in\mathcal{A}^{p,q}|d\phi=0\} the set of closed (p,q)(p,q)-forms. And consider the operator

Ψ=¯trg()|𝒜dp,q:𝒜dp,q𝒜dp,q\displaystyle\Psi=\partial\bar{\partial}\text{tr}_{g}(\cdot)\Big{|}_{\mathcal{A}^{p,q}_{d}}:\mathcal{A}^{p,q}_{d}\to\mathcal{A}^{p,q}_{d}

from 𝒜dp,q\mathcal{A}^{p,q}_{d} to itself.

We claim that Ψ\Psi is a linear elliptic operator.

In local complex coordinates, we assume

ϕ\displaystyle\phi =i1<<irj1<<jsϕi1irj1¯js¯dzi1dzirdz¯j1dz¯js\displaystyle=\sum_{\begin{subarray}{c}i_{1}<\cdots<i_{r}\\ j_{1}<\cdots<j_{s}\end{subarray}}\phi_{i_{1}\cdots i_{r}\bar{j_{1}}\cdots\bar{j_{s}}}dz^{i_{1}}\wedge\cdots\wedge dz^{i_{r}}\wedge d\bar{z}^{j_{1}}\wedge\cdots\wedge d\bar{z}^{j_{s}}
=1r!s!ϕi1irj1¯js¯dzi1dzirdz¯j1dz¯js\displaystyle=\frac{1}{r!s!}\phi_{i_{1}\cdots i_{r}\bar{j_{1}}\cdots\bar{j_{s}}}dz^{i_{1}}\wedge\cdots\wedge dz^{i_{r}}\wedge d\bar{z}^{j_{1}}\wedge\cdots\wedge d\bar{z}^{j_{s}}

where the subscripts are antisymmetric. Then

trg(ϕ)\displaystyle\text{tr}_{g}(\phi) =1(r1)!(s1)!gq¯pϕi1ir1pq¯j2¯js¯dzi1dzir1dz¯j2dz¯js\displaystyle=\frac{1}{(r-1)!(s-1)!}g^{\bar{q}p}\phi_{i_{1}\cdots i_{r-1}p\bar{q}\bar{j_{2}}\cdots\bar{j_{s}}}dz^{i_{1}}\wedge\cdots\wedge dz^{i_{r-1}}\wedge d\bar{z}^{j_{2}}\wedge\cdots\wedge d\bar{z}^{j_{s}}

And consider the second order term of ¯trg(ϕ)\partial\bar{\partial}\text{tr}_{g}(\phi)

2nd order =1(r1)!(s1)!gq¯pϕi1ir1pq¯j2¯js¯,irj1¯dzi1dzirdz¯j1dz¯js\displaystyle=\frac{1}{(r-1)!(s-1)!}g^{\bar{q}p}\phi_{i_{1}\cdots i_{r-1}p\bar{q}\bar{j_{2}}\cdots\bar{j_{s}},i_{r}\bar{j_{1}}}dz^{i_{1}}\wedge\cdots\wedge dz^{i_{r}}\wedge d\bar{z}^{j_{1}}\wedge\cdots\wedge d\bar{z}^{j_{s}}
=1(s1)!i1<<irgq¯pa=1r(1)raϕi1ia^ir1pq¯j2¯js¯,iaj1¯\displaystyle=\frac{1}{(s-1)!}\sum_{i_{1}<\cdots<i_{r}}g^{\bar{q}p}\sum_{a=1}^{r}(-1)^{r-a}\phi_{i_{1}\cdots\widehat{i_{a}}\cdots i_{r-1}p\bar{q}\bar{j_{2}}\cdots\bar{j_{s}},i_{a}\bar{j_{1}}}
dzi1dzirdz¯j1dz¯js\displaystyle\cdot dz^{i_{1}}\wedge\cdots\wedge dz^{i_{r}}\wedge d\bar{z}^{j_{1}}\wedge\cdots\wedge d\bar{z}^{j_{s}}
=1(s1)!i1<<irgq¯pϕi1irq¯j2¯js¯,pj1¯dzi1dzirdz¯j1dz¯js\displaystyle=\frac{1}{(s-1)!}\sum_{i_{1}<\cdots<i_{r}}g^{\bar{q}p}\phi_{i_{1}\cdots i_{r}\bar{q}\bar{j_{2}}\cdots\bar{j_{s}},p\bar{j_{1}}}dz^{i_{1}}\wedge\cdots\wedge dz^{i_{r}}\wedge d\bar{z}^{j_{1}}\wedge\cdots\wedge d\bar{z}^{j_{s}}

The last line is because ϕ=0\partial\phi=0, which if and only if

ϕi1irj1¯js¯,pa=1r(1)raϕi1ia^irpj1¯js¯,ia=0.\displaystyle\phi_{i_{1}\cdots i_{r}\bar{j_{1}}\cdots\bar{j_{s}},p}-\sum_{a=1}^{r}(-1)^{r-a}\phi_{i_{1}\cdots\widehat{i_{a}}\cdots i_{r}p\bar{j_{1}}\cdots\bar{j_{s}},i_{a}}=0.

Similarly, using the fact ¯ϕ=0\bar{\partial}\phi=0, the terms of second order are

2nd order =1(s1)!i1<<irgq¯pϕi1irj2¯js¯q¯,pj1¯dzi1dzirdz¯j2dz¯jsdz¯j1\displaystyle=\frac{1}{(s-1)!}\sum_{i_{1}<\cdots<i_{r}}g^{\bar{q}p}\phi_{i_{1}\cdots i_{r}\bar{j_{2}}\cdots\bar{j_{s}}\bar{q},p\bar{j_{1}}}dz^{i_{1}}\wedge\cdots\wedge dz^{i_{r}}\wedge d\bar{z}^{j_{2}}\wedge\cdots\wedge d\bar{z}^{j_{s}}\wedge d\bar{z}^{j_{1}}
=i1<<irj1<<jsgq¯pϕi1irj1¯js¯,pq¯dzi1dzirdz¯j1dz¯js\displaystyle=\sum_{\begin{subarray}{c}i_{1}<\cdots<i_{r}\\ j_{1}<\cdots<j_{s}\end{subarray}}g^{\bar{q}p}\phi_{i_{1}\cdots i_{r}\bar{j_{1}}\cdots\bar{j_{s}},p\bar{q}}dz^{i_{1}}\wedge\cdots\wedge dz^{i_{r}}\wedge d\bar{z}^{j_{1}}\wedge\cdots\wedge d\bar{z}^{j_{s}}

Thus Ψ\Psi is elliptic and we complete our proof. ∎

Now we show that the Hermitian-symplectic structures are preserved by flow (3.1). This is why we say it is a flow of Hermitian-symplectic forms.

Theorem 3.6.

Hermitian-symplectic structures are preserved by flow (3.1).

Proof.

Let Ω(t)=φ(t)+ω(t)+φ¯(t)\Omega(t)=\varphi(t)+\omega(t)+\bar{\varphi}(t) be a solution to flow (3.1) with initial data Ω0=φ0+ω0+φ¯0\Omega_{0}=\varphi_{0}+\omega_{0}+\bar{\varphi}_{0}. Recall the definition of Hermitian-symplectic forms. The only thing we need to check is that Ω(t)\Omega(t) is closed, which equivalents to equations (2.2). Applying Lemma 3.4 in the case of p=2p=2, we know that φ(t)\varphi(t) is \partial-closed. To complete our proof, we need to show ¯φ(t)+ω(t)=0\bar{\partial}\varphi(t)+\partial\omega(t)=0.

Applying \partial and ¯\bar{\partial} to both sides of the evolution equations of ω\omega and φ\varphi, respectively, we get

(3.2) tω=¯¯ω\displaystyle\frac{\partial}{\partial t}\partial\omega=\partial\bar{\partial}\bar{\partial}^{*}\omega
(3.3) t¯φ=¯trg(¯φ).\displaystyle\frac{\partial}{\partial t}\bar{\partial}\varphi=\partial\bar{\partial}\text{tr}_{g}(\bar{\partial}\varphi).

Firstly, we show that both equations (3.2) and (3.3) have the same form. By direct computation, we have

¯ω\displaystyle\bar{\partial}^{*}\omega =ω\displaystyle=-*\partial*\omega
=1(n1)!(ωn1)\displaystyle=-\frac{1}{(n-1)!}*\partial(\omega^{n-1})
=1(n2)!(ωn2ω)\displaystyle=-\frac{1}{(n-2)!}*(\omega^{n-2}\wedge\partial\omega)
=trg(ω)\displaystyle=\text{tr}_{g}(\partial\omega)

The last line uses Lemma 6.1 in the appendix, which can be proved by direct computation. Thus equations (3.2) and (3.3) both have the form in Lemma 3.5

(3.4) tϕ=¯trg(ϕ)\displaystyle\frac{\partial}{\partial t}\partial\phi=\partial\bar{\partial}\text{tr}_{g}(\phi)

for closed (2,1)(2,1) forms ϕ\phi.

Note that both ω(t)\partial\omega(t) and ¯φ(t)-\bar{\partial}\varphi(t) are solutions to (3.4) with initial data ω0\partial\omega_{0} and ¯φ0-\bar{\partial}\varphi_{0}, respectively. Meanwhile we have ω0=¯φ0\partial\omega_{0}=-\bar{\partial}\varphi_{0} for Ω0\Omega_{0} is a Hermitian-symplectic form. Applying Lemma 3.5, we get ω(t)=¯φ(t)\partial\omega(t)=-\bar{\partial}\varphi(t) by the uniqueness of solutions. Thus we complete our proof. ∎

Remark 3.7.

From the above proof and Lemma 6.1, we know that the evolution equation of φ\varphi in (3.1) can rewrite as

tφ=¯ω.\displaystyle\frac{\partial}{\partial t}\varphi=\partial\bar{\partial}^{*}\omega.

And by direct computation, we have ¯ω=ρ2,0(ω)\partial\bar{\partial}^{*}\omega=-\rho^{2,0}(\omega). The right-hand term is exactly the (2,0)(2,0)-part of Bismut-Ricci form. So we say the additional equation is natural with respect to pluriclosed flow.

4. A topological obstruction

In this section, we will introduce a positive function 𝒱:Ω𝒱(Ω)>0\mathcal{V}:\Omega\mapsto\mathcal{V}(\Omega)>0 related to Hermitian-symplectic forms. And show that when a Hermitian-symplectic form deform along the parabolic flow (3.1), the function 𝒱(t)𝒱(Ω(t))\mathcal{V}(t)\triangleq\mathcal{V}(\Omega(t)) is actually a polynomial with respect to time tt. Meanwhile, all coefficients of this polynomial are constants depending only on initial data Ω0\Omega_{0} and the underlying complex manifold. That will bring some obstructions to how long a solution to flow (3.1) exists. Especially, it brings a topological obstruction to global solutions to pluriclosed flow with Hermitian-symplectic initial data.

4.1. Definition of exponential-type volume function

Let’s begin with the definition of 𝒱\mathcal{V}. Given a complex manifold (M2n,J)(M^{2n},J). Denote 𝒜HS2(M)\mathcal{A}^{2}_{HS}(M) the set of Hermitian-symplectic forms on the manifold and we assume it is not empty.

Definition 4.1.

We define the exponential-type volume function 𝒱:𝒜HS2(M)(0,+)\mathcal{V}:\mathcal{A}^{2}_{HS}(M)\to(0,+\infty) by

(4.1) 𝒱(Ω)=k01(k!)2(φk,φk)ω\displaystyle\mathcal{V}(\Omega)=\sum_{k\geq 0}\frac{1}{(k!)^{2}}(\varphi^{k},\varphi^{k})_{\omega}

Here Ω=φ+ω+φ¯\Omega=\varphi+\omega+\bar{\varphi} and (,)ω(\cdot,\cdot)_{\omega} is the inner product induced by metric ω\omega.

Here we will give some explanations about this definition. Firstly, definition (4.1) is actually a sum of finite term for a given manifold has finite dimension. Secondly, notice that ω\omega is the (1,1)(1,1)-part of a Hermitian-symplectic form, so it is a metric and we can have the inner product (,)ω(\cdot,\cdot)_{\omega}. For this reason, 𝒱\mathcal{V} can not define for all symplectic forms. And when n=2n=2, 𝒱(Ω)=(1,1)ω+(φ,φ)ω=12(Ω,Ω)ω\mathcal{V}(\Omega)=(1,1)_{\omega}+(\varphi,\varphi)_{\omega}=\frac{1}{2}(\Omega,\Omega)_{\omega}, is the volume of Hermitian-symplectic form up to a constant. This is why we call 𝒱\mathcal{V} volume function.

Remark 4.2.

If we define the formal exponential map for differential forms by

exp(ϕ)=k0ϕkk!\displaystyle\exp(\phi)=\sum_{k\geq 0}\frac{\phi^{k}}{k!}

Then 𝒱(ω)=(exp(φ),exp(φ))ω\mathcal{V}(\omega)=(\exp(\varphi),\exp(\varphi))_{\omega} for the inner product of two forms with different bidegree is zero. This is why we call 𝒱\mathcal{V} exponential-type.

4.2. Exponential-type volume function along Hermitian-symplectic flow

In this subsection, we study the exponential-type volume function 𝒱(t)=𝒱(Ω(t))\mathcal{V}(t)=\mathcal{V}(\Omega(t)) along the flow (3.1). And we claim that 𝒱\mathcal{V} is actually a polynomial of tt and the coefficient of its highest degree term is a topological quantity of the manifold.

More precisely, we have

Theorem 4.3.

For a solution to (3.1) with initial data Ω0\Omega_{0}, the function 𝒱(t)\mathcal{V}(t) is a polynomial of degree at most nn

𝒱(t)=i=0naiti\displaystyle\mathcal{V}(t)=\sum_{i=0}^{n}a_{i}t^{i}

with coefficients

ai=1i!k02k+in1(k!)2(n-2k-i)!Mφ0kφ¯0kω0n2ki(1¯logdetg0)i\displaystyle a_{i}=\frac{1}{i!}\sum_{\begin{subarray}{c}k\geq 0\\ 2k+i\leq n\end{subarray}}\frac{1}{(k!)^{2}(n\mathord{-}2k\mathord{-}i)!}\int_{M}\varphi_{0}^{k}\wedge\bar{\varphi}_{0}^{k}\wedge\omega^{n-2k-i}_{0}\wedge(\sqrt{-1}\partial\bar{\partial}\log\det g_{0})^{i}

Specially, we have

an=1n!(1)nMc1(M)n=1n!(1)nc1n\displaystyle a_{n}=\frac{1}{n!}(-1)^{n}\int_{M}c_{1}(M)^{n}=\frac{1}{n!}(-1)^{n}c_{1}^{n}

in which c1(M)c_{1}(M) is the first Chern class of (M2n,J)(M^{2n},J).

We put the proof at the end of this section.

From Theorem 4.3, we know that a0=𝒱(0)=𝒱(Ω0)a_{0}=\mathcal{V}(0)=\mathcal{V}(\Omega_{0}) is a positive number. Other coefficients may be difficult to calculate in general case. The formula of 𝒱\mathcal{V} can bring us an obstruction to how long a solution to flow (3.1) exists.

Formally, we have

Corollary 4.4.

Assume a solution to flow (3.1) exists on [0,T)[0,T), then the polynomial of tt defined by Theorem 4.3 must have no roots on [0,T)[0,T).

Proof.

Just note that the polynomial is the exponential-type function and it is positive by definition. ∎

For global solutions, i.e. solutions exist on [0,+)[0,+\infty), we have a more useful obstruction. And we state it as a theorem.

Theorem 4.5.

If there exists a global solution to flow (3.1) on (M2n,J)(M^{2n},J), then (1)nc1n(-1)^{n}c_{1}^{n} must be nonnegative.

Proof.

If (1)nc1n(-1)^{n}c_{1}^{n} is negative, then 𝒱(t)\mathcal{V}(t) must have a root in [0,+)[0,+\infty). This is a contradiction to long-time existence of the solution. ∎

In Remark 3.2, we point out that pluriclosed flow with initial data ω0\omega_{0} satisfying [ω0]=0HBC2,1(M,)[\partial\omega_{0}]=0\in\text{H}^{2,1}_{BC}(M,\mathbb{C}) is a special case that can be extended to a Hermitian-symplectic form. So the sign of (1)nc1n(-1)^{n}c_{1}^{n} is also an obstruction to global solutions to pluriclosed flow under those assumption.

4.3. Proof of Theorem 4.3

In this subsection, we give the proof of Theorem 4.3.

Firstly, we need the following lemma.

Lemma 4.6.

Assume Ω(t)=φ(t)+ω(t)+φ¯(t)\Omega(t)=\varphi(t)+\omega(t)+\bar{\varphi}(t) is a solution to (3.1) with initial data Ω0\Omega_{0}. Define the auxiliary differential forms

(4.2) α[k,s](t)={φk(t)φ¯k(t)ωn2ks(t),s,k0 and 2k+sn0,otherwise\alpha[k,s](t)=\left\{\begin{aligned} &\varphi^{k}(t)\wedge\bar{\varphi}^{k}(t)\wedge\omega^{n-2k-s}(t),&&s,k\geq 0\text{\rm\ and }2k+s\leq n\\ &0,&&\text{otherwise}\end{aligned}\right.

and

(4.3) β[s](t)=k01(k!)2(n2ks)!α[k,s](t)\displaystyle\beta[s](t)=\sum_{k\geq 0}\frac{1}{(k!)^{2}(n-2k-s)!}\alpha[k,s](t)

Then β[s](t)\beta[s](t) is pluriclosed.

Proof.

By direct computation,

¯α[k,s]\displaystyle\partial\bar{\partial}\alpha[k,s] ={k¯φφk1φ¯kωn2ks+(n-2k-s)φkφ¯kωn2ks1¯ω}\displaystyle=\partial\left\{k\bar{\partial}\varphi\wedge\varphi^{k-1}\wedge\bar{\varphi}^{k}\wedge\omega^{n-2k-s}+(n\mathord{-}2k\mathord{-}s)\varphi^{k}\wedge\bar{\varphi}^{k}\wedge\omega^{n-2k-s-1}\wedge\bar{\partial}\omega\right\}
=k2¯φφk1φ¯k1φ¯ωn2ks\displaystyle=-k^{2}\bar{\partial}\varphi\wedge\varphi^{k-1}\wedge\bar{\varphi}^{k-1}\wedge\partial\bar{\varphi}\wedge\omega^{n-2k-s}
k(n-2k-s)¯φφk1φ¯kωωn2ks1\displaystyle\quad-k(n\mathord{-}2k\mathord{-}s)\bar{\partial}\varphi\wedge\varphi^{k-1}\wedge\bar{\varphi}^{k}\wedge\partial\omega\wedge\omega^{n-2k-s-1}
+k(n-2k-s)φφk1φ¯k¯ωωn2ks1\displaystyle\quad+k(n\mathord{-}2k\mathord{-}s)\partial\varphi\wedge\varphi^{k-1}\wedge\bar{\varphi}^{k}\wedge\bar{\partial}\omega\wedge\omega^{n-2k-s-1}
+(n-2k-s)(n-2k-s-1)φkφ¯kωn2ks2ω¯ω\displaystyle\quad+(n\mathord{-}2k\mathord{-}s)(n\mathord{-}2k\mathord{-}s\mathord{-}1)\varphi^{k}\wedge\bar{\varphi}^{k}\wedge\omega^{n-2k-s-2}\wedge\partial\omega\wedge\bar{\partial}\omega
={k2α[k-1,s+2]+(n-2k-s)(n-2k-s-1)α[k,s+2]}ω¯ω\displaystyle=\Big{\{}-k^{2}\alpha[k\mathord{-}1,s\mathord{+}2]+(n\mathord{-}2k\mathord{-}s)(n\mathord{-}2k\mathord{-}s\mathord{-}1)\alpha[k,s\mathord{+}2]\Big{\}}\wedge\partial\omega\wedge\bar{\partial}\omega

Note that ω=¯φ\partial\omega=-\bar{\partial}\varphi. So ¯φω=ωω=0\bar{\partial}\varphi\wedge\partial\omega=-\partial\omega\wedge\partial\omega=0 for ω\partial\omega is an odd form. Thus the third and fourth lines equal zero. By direct computation, it is easy to check that the formula above is also true when k=0k=0 and k=(ns)/2k=(n-s)/2.

Then we have

¯β[s]\displaystyle\partial\bar{\partial}\beta[s] =k01(k!)2(n-2k-s)!(k2α[k-1,s+2]+(n-2k-s)(n-2k-s-1)α[k,s+2])\displaystyle=\sum_{k\geq 0}\frac{1}{(k!)^{2}(n\mathord{-}2k\mathord{-}s)!}\Big{(}-k^{2}\alpha[k\mathord{-}1,s\mathord{+}2]+(n\mathord{-}2k\mathord{-}s)(n\mathord{-}2k\mathord{-}s\mathord{-}1)\alpha[k,s\mathord{+}2]\Big{)}
ω¯ω\displaystyle\qquad\wedge\partial\omega\wedge\bar{\partial}\omega
=k1α[k-1,s+2](k1!)2(n-2k-s)!+k0α[k,s+2](k!)2(n-2k-s-2)!\displaystyle=-\sum_{k\geq 1}\frac{\alpha[k\mathord{-}1,s\mathord{+}2]}{(k-1!)^{2}(n\mathord{-}2k\mathord{-}s)!}+\sum_{k\geq 0}\frac{\alpha[k,s\mathord{+}2]}{(k!)^{2}(n\mathord{-}2k\mathord{-}s\mathord{-}2)!}
=k0α[k,s+2](k!)2(n-2k-s-2)!+k0α[k,s+2](k!)2(n-2k-s-2)!\displaystyle=-\sum_{k\geq 0}\frac{\alpha[k,s\mathord{+}2]}{(k!)^{2}(n\mathord{-}2k\mathord{-}s\mathord{-}2)!}+\sum_{k\geq 0}\frac{\alpha[k,s\mathord{+}2]}{(k!)^{2}(n\mathord{-}2k\mathord{-}s\mathord{-}2)!}
=0\displaystyle=0

The second equal sign is because α[1,s]=0\alpha[-1,s]=0 and we just replace the index k1k-1 by kk in the third line. Thus we complete our proof. ∎

To compute the derivative of function 𝒱(t)=𝒱(Ω(t))\mathcal{V}(t)=\mathcal{V}(\Omega(t)) with respect to time tt. We need the next two lemmas.

Lemma 4.7.

Given a fixed real form ϕ\phi with degree 2s2s which is \partial and ¯\bar{\partial} closed. Define a function with respect to tt by

P[k,s;ϕ](t)=Mα[k,s](t)ϕ\displaystyle P[k,s;\phi](t)=\int_{M}\alpha[k,s](t)\wedge\phi

in which α\alpha is defined by (4.2). Then we have

(4.4) ddtP[k,s;ϕ]\displaystyle\frac{d}{dt}P[k,s;\phi] =k2(A[k,s]+A[k,s]¯)\displaystyle=-k^{2}(A[k,s]+\overline{A[k,s]})
+(n-2k-s)(n-2k-s-1)(B[k,s]+B[k,s]¯)\displaystyle\qquad+(n\mathord{-}2k\mathord{-}s)(n\mathord{-}2k\mathord{-}s\mathord{-}1)(B[k,s]+\overline{B[k,s]})
+(n-2k-s)Mα[k,s+1]1¯logdetgϕ\displaystyle\qquad+(n\mathord{-}2k\mathord{-}s)\int_{M}\alpha[k,s\mathord{+}1]\wedge\sqrt{-1}\partial\bar{\partial}\log\det g\wedge\phi

where

(4.5) A[k,s]=M¯ωα[k-1,s+2]¯ωϕ\displaystyle A[k,s]=\int_{M}\bar{\partial}^{*}\omega\wedge\alpha[k\mathord{-}1,s\mathord{+}2]\wedge\bar{\partial}\omega\wedge\phi

and

(4.6) B[k,s]=M¯ωα[k,s+2]¯ωϕ.\displaystyle B[k,s]=\int_{M}\bar{\partial}^{*}\omega\wedge\alpha[k,s\mathord{+}2]\wedge\bar{\partial}\omega\wedge\phi.
Proof.

Here we use a point overhead to represent the derivative with respect to time tt. Derivate PP with respect to tt and notice that ω(t)\omega(t) satisfies equations (3.2).

ddtP[k,s;ϕ]\displaystyle\frac{d}{dt}P[k,s;\phi] =ddtMφkφ¯kωn2ksϕ\displaystyle=\frac{d}{dt}\int_{M}\varphi^{k}\wedge\bar{\varphi}^{k}\wedge\omega^{n-2k-s}\wedge\phi
=k{Mφ˙φk1φ¯kωn2ksϕ+Mφkφ¯˙φ¯k1ωn2ksϕ}\displaystyle=k\left\{\int_{M}\dot{\varphi}\wedge\varphi^{k-1}\wedge\bar{\varphi}^{k}\wedge\omega^{n-2k-s}\wedge\phi+\int_{M}\varphi^{k}\wedge\dot{\bar{\varphi}}\wedge\bar{\varphi}^{k-1}\wedge\omega^{n-2k-s}\wedge\phi\right\}
+(n-2k-s)Mφkφ¯kωn2ks1¯¯ωϕ\displaystyle\quad+(n\mathord{-}2k\mathord{-}s)\int_{M}\varphi^{k}\wedge\bar{\varphi}^{k}\wedge\omega^{n-2k-s-1}\wedge\bar{\partial}\bar{\partial}^{*}\omega\wedge\phi
+(n-2k-s)Mφkφ¯kωn2ks1ωϕ\displaystyle\quad+(n\mathord{-}2k\mathord{-}s)\int_{M}\varphi^{k}\wedge\bar{\varphi}^{k}\wedge\omega^{n-2k-s-1}\wedge\partial\partial^{*}\omega\wedge\phi
+(n-2k-s)Mφkφ¯kωn2ks11¯logdetgϕ\displaystyle\quad+(n\mathord{-}2k\mathord{-}s)\int_{M}\varphi^{k}\wedge\bar{\varphi}^{k}\wedge\omega^{n-2k-s-1}\wedge\sqrt{-1}\partial\bar{\partial}\log\det g\wedge\phi

The second and the fourth terms are just the conjugation of the first and the third terms, respectively. Recall Remark 3.7 and note ϕ\phi is \partial-closed. Then the first term is

1st term =kM¯ωφk1φ¯kωn2ksϕ\displaystyle=k\int_{M}\partial\bar{\partial}^{*}\omega\wedge\varphi^{k-1}\wedge\bar{\varphi}^{k}\wedge\omega^{n-2k-s}\wedge\phi
=k2M¯ωφk1φ¯φ¯k1ωn2ksϕ\displaystyle=k^{2}\int_{M}\bar{\partial}^{*}\omega\wedge\varphi^{k-1}\wedge\partial\bar{\varphi}\wedge\bar{\varphi}^{k-1}\wedge\omega^{n-2k-s}\wedge\phi
+k(n-2k-s)M¯ωφk1φ¯kωωn2ks1ϕ\displaystyle\qquad+k(n\mathord{-}2k\mathord{-}s)\int_{M}\bar{\partial}^{*}\omega\wedge\varphi^{k-1}\wedge\bar{\varphi}^{k}\wedge\partial\omega\wedge\omega^{n-2k-s-1}\wedge\phi

Similarly,

3rd term =(n-2k-s)M¯¯ωφkφ¯kωn2ks1ϕ\displaystyle=(n\mathord{-}2k\mathord{-}s)\int_{M}\bar{\partial}\bar{\partial}^{*}\omega\wedge\varphi^{k}\wedge\bar{\varphi}^{k}\wedge\omega^{n-2k-s-1}\wedge\phi
=(n-2k-s)(n-2k-s-1)M¯ωφkφ¯k¯ωωn2ks2ϕ\displaystyle=(n\mathord{-}2k\mathord{-}s)(n\mathord{-}2k\mathord{-}s\mathord{-}1)\int_{M}\bar{\partial}^{*}\omega\wedge\varphi^{k}\wedge\bar{\varphi}^{k}\wedge\bar{\partial}\omega\wedge\omega^{n-2k-s-2}\wedge\phi
+k(n-2k-s)M¯ω¯φφk1φ¯kωn2ks1ϕ\displaystyle\qquad+k(n\mathord{-}2k\mathord{-}s)\int_{M}\bar{\partial}^{*}\omega\wedge\bar{\partial}\varphi\wedge\varphi^{k-1}\wedge\bar{\varphi}^{k}\wedge\omega^{n-2k-s-1}\wedge\phi

Here we use ¯ϕ=0\bar{\partial}\phi=0. The sum in last line of “1st term” and in last line of “3rd term” are zero for ω+¯φ=0\partial\omega+\bar{\partial}\varphi=0.

Summing the above and their conjugations, we get formula (4.4). ∎

Notice that we can also choose s=0s=0 in the above lemma. And now we give another lemma we need.

Lemma 4.8.

Given a fixed real form ϕ\phi with degree 2s2s which is \partial and ¯\bar{\partial} closed. Define a function with respect to tt by

(4.7) Q[s;ϕ](t)=Mβ[s](t)ϕ\displaystyle Q[s;\phi](t)=\int_{M}\beta[s](t)\wedge\phi

in which β\beta is defined by (4.3). Then we have

ddtQ[s;ϕ](t)\displaystyle\frac{d}{dt}Q[s;\phi](t) =Q[s+1;1¯logdetg(t)ϕ]\displaystyle=Q[s+1;\sqrt{-1}\partial\bar{\partial}\log\det g(t)\wedge\phi]
=Mβ[s+1](t)1¯logdetg(t)ϕ.\displaystyle=\int_{M}\beta[s+1](t)\wedge\sqrt{-1}\partial\bar{\partial}\log\det g(t)\wedge\phi.
Proof.

Recall the definition of β\beta and use Lemma 4.7.

ddtQ[s;ϕ]\displaystyle\frac{d}{dt}Q[s;\phi] =ddtMk01(k!)2(n-2k-s)!α[k,s](t)ϕ\displaystyle=\frac{d}{dt}\int_{M}\sum_{k\geq 0}\frac{1}{(k!)^{2}(n\mathord{-}2k\mathord{-}s)!}\alpha[k,s](t)\wedge\phi
=k01(k!)2(n-2k-s)!ddtP[k,s;ϕ]\displaystyle=\sum_{k\geq 0}\frac{1}{(k!)^{2}(n\mathord{-}2k\mathord{-}s)!}\frac{d}{dt}P[k,s;\phi]
=k01(k!)2(n-2k-s)!{k2(A[k,s]+A[k,s]¯)\displaystyle=\sum_{k\geq 0}\frac{1}{(k!)^{2}(n\mathord{-}2k\mathord{-}s)!}\Bigg{\{}-k^{2}(A[k,s]+\overline{A[k,s]})
+(n-2k-s)(n-2k-s-1)(B[k,s]+B[k,s]¯)}\displaystyle\qquad+(n\mathord{-}2k\mathord{-}s)(n\mathord{-}2k\mathord{-}s\mathord{-}1)(B[k,s]+\overline{B[k,s]})\Bigg{\}}
+k0n-2k-s(k!)2(n-2k-s)!Mα[k,s+1]1¯logdetgϕ\displaystyle\qquad+\sum_{k\geq 0}\frac{n\mathord{-}2k\mathord{-}s}{(k!)^{2}(n\mathord{-}2k\mathord{-}s)!}\int_{M}\alpha[k,s\mathord{+}1]\wedge\sqrt{-1}\partial\bar{\partial}\log\det g\wedge\phi

Here A[k,s]A[k,s] and B[k,s]B[k,s] are defined by (4.5) and (4.6), respectively.

Consider the term

k01(k!)2(n-2k-s)!(k2A[k,s]+(n-2k-s)(n-2k-s-1)B[k,s])\displaystyle\sum_{k\geq 0}\frac{1}{(k!)^{2}(n\mathord{-}2k\mathord{-}s)!}\Big{(}-k^{2}A[k,s]+(n\mathord{-}2k\mathord{-}s)(n\mathord{-}2k\mathord{-}s\mathord{-}1)B[k,s]\Big{)}
=\displaystyle= k11((k-1)!)2(n-2k-s)!M¯ωα[k-1,s+2]¯ωϕ\displaystyle-\sum_{k\geq 1}\frac{1}{((k\mathord{-}1)!)^{2}(n\mathord{-}2k\mathord{-}s)!}\int_{M}\bar{\partial}^{*}\omega\wedge\alpha[k\mathord{-}1,s\mathord{+}2]\wedge\bar{\partial}\omega\wedge\phi
+k01(k!)2(n-2k-s-2)!M¯ωα[k,s+2]¯ωϕ\displaystyle\qquad+\sum_{k\geq 0}\frac{1}{(k!)^{2}(n\mathord{-}2k\mathord{-}s\mathord{-}2)!}\int_{M}\bar{\partial}^{*}\omega\wedge\alpha[k,s\mathord{+}2]\wedge\bar{\partial}\omega\wedge\phi
=\displaystyle= k01(k!)2(n-2k-s-2)!M¯ωα[k,s+2]¯ωϕ\displaystyle-\sum_{k\geq 0}\frac{1}{(k!)^{2}(n\mathord{-}2k\mathord{-}s\mathord{-}2)!}\int_{M}\bar{\partial}^{*}\omega\wedge\alpha[k,s\mathord{+}2]\wedge\bar{\partial}\omega\wedge\phi
+M¯ωβ[s+2]¯ωϕ\displaystyle\qquad+\int_{M}\bar{\partial}^{*}\omega\wedge\beta[s\mathord{+}2]\wedge\bar{\partial}\omega\wedge\phi
=\displaystyle= M¯ωβ[s+2]¯ωϕ+M¯ωβ[s+2]¯ωϕ\displaystyle-\int_{M}\bar{\partial}^{*}\omega\wedge\beta[s\mathord{+}2]\wedge\bar{\partial}\omega\wedge\phi+\int_{M}\bar{\partial}^{*}\omega\wedge\beta[s\mathord{+}2]\wedge\bar{\partial}\omega\wedge\phi
=\displaystyle= 0\displaystyle 0

The first equal sign is because α[1,s+2]=0\alpha[-1,s+2]=0.
Thus,

ddtQ[s;ϕ]\displaystyle\frac{d}{dt}Q[s;\phi] =k01(k!)2(n-2k-s-1)!Mα[k,s+1]1¯logdetgϕ\displaystyle=\sum_{k\geq 0}\frac{1}{(k!)^{2}(n\mathord{-}2k\mathord{-}s\mathord{-}1)!}\int_{M}\alpha[k,s\mathord{+}1]\wedge\sqrt{-1}\partial\bar{\partial}\log\det g\wedge\phi
=Mβ[s+1]1¯logdetgϕ\displaystyle=\int_{M}\beta[s\mathord{+}1]\wedge\sqrt{-1}\partial\bar{\partial}\log\det g\wedge\phi
=Q[s+1;1¯logdetgϕ].\displaystyle=Q[s+1;\sqrt{-1}\partial\bar{\partial}\log\det g\wedge\phi].

So we complete our proof. ∎

Now we are in the position to prove Theorem 4.3.

Proof of Theorem 4.3.

Assume Ω(t)=φ(t)+ω(t)+φ¯(t)\Omega(t)=\varphi(t)\mathord{+}\omega(t)\mathord{+}\bar{\varphi}(t) is a solution to flow (3.1). We denote (,)t(\cdot,\cdot)_{t} the inner product induced by ω(t)\omega(t). Notice that φk\varphi^{k} is a prime form with respect to the adjoint Lefschetz operator defined by ω(t)\omega(t), for it is a (2k,0)(2k,0)-form. Then by the Lefschetz theory,

(φk,φk)t\displaystyle(\varphi^{k},\varphi^{k})_{t} =Mφkφ¯k=1(n2k)!Mφkφ¯kωn2k\displaystyle=\int_{M}\varphi^{k}\wedge*\bar{\varphi}^{k}=\frac{1}{(n-2k)!}\int_{M}\varphi^{k}\wedge\bar{\varphi}^{k}\wedge\omega^{n-2k}

So we have

𝒱(t)\displaystyle\mathcal{V}(t) =k01(k!)2(φk,φk)t\displaystyle=\sum_{k\geq 0}\frac{1}{(k!)^{2}}(\varphi^{k},\varphi^{k})_{t}
=k01(k!)2(n-2k)!Mφkφ¯kωn2k\displaystyle=\sum_{k\geq 0}\frac{1}{(k!)^{2}(n\mathord{-}2k)!}\int_{M}\varphi^{k}\wedge\bar{\varphi}^{k}\wedge\omega^{n-2k}
=Mβ[0]\displaystyle=\int_{M}\beta[0]
=Q[0;1](t).\displaystyle=Q[0;1](t).

in which β[s]\beta[s] and Q[s;ϕ]Q[s;\phi] are defined by (4.3) and (4.7), respectively.

Here we choose ϕ=1\phi=1, which is a real form of degree 0 and satisfies 1=0\partial 1=0 and ¯1=0\bar{\partial}1=0. Thus Lemma 4.8 can be applied.

Derivate 𝒱(t)\mathcal{V}(t) with respect to tt and apply Lemma 4.8.

ddt𝒱(t)\displaystyle\frac{d}{dt}\mathcal{V}(t) =ddtQ[0;1](t)\displaystyle=\frac{d}{dt}Q[0;1](t)
=Q[1;1¯logdetg](t)\displaystyle=Q[1;\sqrt{-1}\partial\bar{\partial}\log\det g](t)
=Mβ[1](t)1¯logdetg(t)\displaystyle=\int_{M}\beta[1](t)\wedge\sqrt{-1}\partial\bar{\partial}\log\det g(t)
=Mβ[1](t)¯logdeth+Mβ[1](t)1¯logdetg(t)deth\displaystyle=\int_{M}\beta[1](t)\wedge\partial\bar{\partial}\log\det h+\int_{M}\beta[1](t)\wedge\sqrt{-1}\partial\bar{\partial}\log\frac{\det g(t)}{\det h}
=Q[1;1¯logdeth]\displaystyle=Q[1;\sqrt{-1}\partial\bar{\partial}\log\det h]

where hh is an arbitrary fixed Hermitian metric. The last equal sign is because logdetgdeth\log\frac{\det g}{\det h} is a well-defined function on manifolds MM. Then applying Lemma 4.6 we know that β[1](t)\beta[1](t) is pluriclosed. So by integrating by parts, we have

Mβ[1](t)1¯logdetg(t)deth=M¯β[1](t)1logdetg(t)deth=0\displaystyle\int_{M}\beta[1](t)\wedge\sqrt{-1}\partial\bar{\partial}\log\frac{\det g(t)}{\det h}=\int_{M}\partial\bar{\partial}\beta[1](t)\wedge\sqrt{-1}\log\frac{\det g(t)}{\det h}=0

We complete the proof by induction. The case of first order derivative has been proved. Assume it is true for the mm-th order derivative. That is

dmdtm𝒱(t)\displaystyle\frac{d^{m}}{dt^{m}}\mathcal{V}(t) =k01(k!)2(n-2k-m)!Mφkφ¯kωn2km(1¯logdeth)m\displaystyle=\sum_{k\geq 0}\frac{1}{(k!)^{2}(n\mathord{-}2k\mathord{-}m)!}\int_{M}\varphi^{k}\wedge\bar{\varphi}^{k}\wedge\omega^{n-2k-m}\wedge(\sqrt{-1}\partial\bar{\partial}\log\det h)^{m}
=Mβ[m](1¯logdeth)m\displaystyle=\int_{M}\beta[m]\wedge(\sqrt{-1}\partial\bar{\partial}\log\det h)^{m}
=Q[m;(1¯logdeth)m]\displaystyle=Q[m;(\sqrt{-1}\partial\bar{\partial}\log\det h)^{m}]

Then derivate it with respect to tt and apply Lemma 4.6 and Lemma 4.8 again.

dm+1dtm+1𝒱(t)\displaystyle\frac{d^{m+1}}{dt^{m+1}}\mathcal{V}(t) =ddtQ[m;(1¯logdeth)m]\displaystyle=\frac{d}{dt}Q[m;(\sqrt{-1}\partial\bar{\partial}\log\det h)^{m}]
=Q[m+1;1¯logdetg(t)(1¯logdeth)m]\displaystyle=Q[m\mathord{+}1;\sqrt{-1}\partial\bar{\partial}\log\det g(t)\wedge(\sqrt{-1}\partial\bar{\partial}\log\det h)^{m}]
=Mβ[m+1](1¯logdeth)m1¯logdetg(t)\displaystyle=\int_{M}\beta[m\mathord{+}1]\wedge(\sqrt{-1}\partial\bar{\partial}\log\det h)^{m}\wedge\sqrt{-1}\partial\bar{\partial}\log\det g(t)
=Mβ[m+1](1¯logdeth)m+1\displaystyle=\int_{M}\beta[m\mathord{+}1]\wedge(\sqrt{-1}\partial\bar{\partial}\log\det h)^{m+1}
=k01(k!)2(n-2k-m-1)!Mφkφ¯kωn2km1(1¯logdeth)m+1\displaystyle=\sum_{k\geq 0}\frac{1}{(k!)^{2}(n\mathord{-}2k\mathord{-}m\mathord{-}1)!}\int_{M}\varphi^{k}\mathord{\wedge}\bar{\varphi}^{k}\mathord{\wedge}\omega^{n-2k-m-1}\mathord{\wedge}(\sqrt{-1}\partial\bar{\partial}\log\det h)^{m+1}

Notice that β[m]=0\beta[m]=0 for mn+1m\geq n+1, so the mm-th order derivative of 𝒱\mathcal{V} equals zero for all mn+1m\geq n+1. Specially, when m=nm=n, the dimension of manifold, we have

dndtn𝒱(t)\displaystyle\frac{d^{n}}{dt^{n}}\mathcal{V}(t) =Q[n,(1¯logdeth)n]\displaystyle=Q[n,(\sqrt{-1}\partial\bar{\partial}\log\det h)^{n}]
=M(1¯logdeth)n\displaystyle=\int_{M}(\sqrt{-1}\partial\bar{\partial}\log\det h)^{n}
=(1)nMc1(M)n\displaystyle=(-1)^{n}\int_{M}c_{1}(M)^{n}
=(1)nc1n,\displaystyle=(-1)^{n}c_{1}^{n},

in which c1(M)c_{1}(M) is the first Chern class of MM.

Thus 𝒱(t)=aiti\mathcal{V}(t)=\sum a_{i}t^{i} is a polynomial of degree at most nn and coefficients can be represented by

ai=1n!k01(k!)2(n-2k-i)!Mφ0kφ¯0kω0n2ki(1¯logdetg0)i\displaystyle a_{i}=\frac{1}{n!}\sum_{k\geq 0}\frac{1}{(k!)^{2}(n\mathord{-}2k\mathord{-}i)!}\int_{M}\varphi_{0}^{k}\wedge\bar{\varphi}_{0}^{k}\wedge\omega^{n-2k-i}_{0}\wedge(\sqrt{-1}\partial\bar{\partial}\log\det g_{0})^{i}

Here we choose h=g0h=g_{0}, the initial metric. So aia_{i} depends only on the complex manifold and initial data Ω0\Omega_{0}. Specially, an=(1)nc1n/n!a_{n}=(-1)^{n}c_{1}^{n}/n! is a topological quantity of the complex manifolds. ∎

5. More discussion

In this section we do some discussion in the case of compact complex surfaces. One of the most important motivation of pluriclosed flow is to classify Kodaira’s class VII surfaces. Inspired by Perelman’s great work on Ricci flow (see [10, 11, 12]), Tian and Streets use pluriclosed flow to find canonical metrics on non-Kähler manifolds in order to understand its geometry and topology. A natural possible canonical metric is static metric, which satisfies ω=λρ1,1\omega=\lambda\rho^{1,1} for some real number λ\lambda. Notice that the static metric with λ0\lambda\neq 0 is also Hermitian-symplectic. And this is our motivation to study Hermitian-symplectic structures. An important first step is that Tian and Streets[17] prove class VII+\text{VII}^{+} surfaces, (i.e. class VII surfaces with b2>0b_{2}>0), admit no static metrics. But it is not true for class VII surfaces with b2=0b_{2}=0 because the standard metric of Hopf surface is static with λ=0\lambda=0. Recall that Bogomolov[4, 5] proves a class VII surface with b2=0b_{2}=0 is biholomorphic to either a Hopf surface or an Inoue surface.

Tian and Streets[17] have shown that a compact complex surface admits Hermitian-symplectic structures must be Kähler using the classification of compact complex surfaces. As a corollary, non-Kähler surfaces admit no static metric with λ0\lambda\neq 0. Here we can give more information about the relationship between Hermitian-symplectic structures and Kähler structures in dimension two using flow (3.1).

We begin with an observation.

Lemma 5.1.

If Ω(t)=φ(t)+ω(t)+φ¯(t)\Omega(t)=\varphi(t)+\omega(t)+\bar{\varphi}(t) is a solution to (3.1) with initial data Ω0\Omega_{0} on a compact complex surface. Then the function F(t)(φ(t),φ(t))ω(t)F(t)\triangleq(\varphi(t),\varphi(t))_{\omega(t)} is monotonically decreasing. Moreover, if the initial data ω0\omega_{0} is non-Kähler, it is strictly monotonically decreasing.

Proof.

Consider the function F(t)=(φ(t),φ(t))ω(t)F(t)=(\varphi(t),\varphi(t))_{\omega(t)}. Derivate it with respect to tt and notice that φ\varphi satisfies (3.1).

ddtF\displaystyle\frac{d}{dt}F =ddtMφφ¯\displaystyle=\frac{d}{dt}\int_{M}\varphi\wedge\bar{\varphi}
=Mtφφ¯+conjugation\displaystyle=\int_{M}\frac{\partial}{\partial t}\varphi\wedge\bar{\varphi}+\text{conjugation}
=M¯ωφ¯+conjugation\displaystyle=\int_{M}\partial\bar{\partial}^{*}\omega\wedge\bar{\varphi}+\text{conjugation}

The first line uses the fact φ\varphi is a prime form with respect to the adjoint Lefschetz operator induced by ω(t)\omega(t). Using integration by parts and noting that ω+¯φ=0\partial\omega+\bar{\partial}\varphi=0, we have

M¯ωφ¯=\displaystyle\int_{M}\partial\bar{\partial}^{*}\omega\wedge\bar{\varphi}= M¯ω¯ω=(¯ω,¯ω)t\displaystyle-\int_{M}\bar{\partial}^{*}\omega\wedge\bar{\partial}\omega=-(\bar{\partial}\omega,\bar{\partial}\omega)_{t}

So

ddtF=2(¯ω,¯ω)t0.\displaystyle\frac{d}{dt}F=-2(\bar{\partial}\omega,\bar{\partial}\omega)_{t}\leq 0.

Thus (φ,φ)t(\varphi,\varphi)_{t} is monotonically decreasing.

Then we prove it is actually strictly monotonically decreasing when ω0\omega_{0} is non-Kähler. We just need to show that the solution ω(t)\omega(t) to pluriclosed flow is non-Kähler when initial metric is non-Kähler.

Assume ω~t\tilde{\omega}_{t} is a solution with non-Kähler initial metric. If ω~t0=0\partial\tilde{\omega}_{t_{0}}=0 for some t0>0t_{0}>0, then we can find t>0t_{*}>0 such that ωt=0\partial\omega_{t_{*}}=0 and ω~t0\partial\tilde{\omega}_{t}\neq 0 for t<tt<t_{*}. From the standard theorem of parabolic equations, we get the uniqueness of solutions to flow (2.1) with condition ω(t)=ω~t\omega(t_{*})=\tilde{\omega}_{t_{*}} in small neighborhood (tε,t+ε)(t_{*}-\varepsilon,t_{*}+\varepsilon). Tian and Streets[17] point out that pluriclosed flow degenerates to Kähler-Ricci flow when initial data is Kähler and so preserves Kähler condition. Then ω~t=0\partial\tilde{\omega}_{t}=0 in (tε,t+ε)(t_{*}-\varepsilon,t_{*}+\varepsilon) by uniqueness. This is a contradiction. Thus ω~t\tilde{\omega}_{t} will not be Kähler at any finite time when the solution exists. ∎

As an application, we show that a Hermitian-symplectic structure can deform to a Kähler structure, if (3.1) converges at infinity time. Notice that

(φ(t),φ(t))t(φ(0),φ(0))0=0t(ω(s),ω(s))s𝑑s\displaystyle(\varphi(t),\varphi(t))_{t}-(\varphi(0),\varphi(0))_{0}=-\int_{0}^{t}(\partial\omega(s),\partial\omega(s))_{s}ds

for all t>0t>0. So we have 0+(ω(s),ω(s))s𝑑s<\int_{0}^{+\infty}(\partial\omega(s),\partial\omega(s))_{s}ds<\infty for (φ(t),φ(t))t(\varphi(t),\varphi(t))_{t} is always positive. If ω(t)\omega(t) converges to ω\omega_{\infty} at infinity time, then (ω,ω)ω=0(\partial\omega_{\infty},\partial\omega_{\infty})_{\omega_{\infty}}=0. Thus the limitation metric is Kähler. In fact, this is true in every dimension. And the proof is due to Jeffrey Streets. We just need to make a slight adjustment to the proof of Proposition 4.2 in [1]. In short, the limitation of a pluriclosed flow must be a steady soliton, which satisfies d(efH)=0d^{*}(e^{-f}H_{\infty})=0 for some function ff. Here H=1(¯)ωH_{\infty}=\sqrt{-1}(\partial-\bar{\partial})\omega_{\infty} is the torsion form of ω\omega_{\infty}. Then we have ¯(efω)=0\bar{\partial}^{*}(e^{-f}\partial\omega_{\infty})=0. And note that ω\omega_{\infty} is Hermitian-symplectic. So

(efω,ω)=(efω,¯φ)=(¯(efω),φ)=0\displaystyle(e^{-f}\partial\omega_{\infty},\partial\omega_{\infty})_{\infty}=(e^{-f}\partial\omega_{\infty},-\bar{\partial}\varphi_{\infty})_{\infty}=(\bar{\partial}^{*}(e^{-f}\partial\omega_{\infty}),-\varphi_{\infty})_{\infty}=0

Thus ω=0\partial\omega_{\infty}=0 and the limitation is Kähler.

But not all solutions can exist for a long time. Let’s consider the function 𝒱(t)\mathcal{V}(t). In dimension two, it is just the volume of Hermitian-symplectic forms (Ω(t),Ω(t))ω(t)(\Omega(t),\Omega(t))_{\omega(t)} up to a constant. Applying Theorem 4.3, we know that 𝒱(t)\mathcal{V}(t) is a polynomial of degree at most 22 such that all coefficients are determined by initial data. So we can calculate all roots of 𝒱(t)\mathcal{V}(t) explicitly. By direct computation, we know that the coefficients are a0=12(Ω0,Ω0)0>0a_{0}=\frac{1}{2}(\Omega_{0},\Omega_{0})_{0}>0, a2=c12/2a_{2}=c_{1}^{2}/2 and

a1=Mω01¯logdetg0.\displaystyle a_{1}=\int_{M}\omega_{0}\wedge\sqrt{-1}\partial\bar{\partial}\log\det g_{0}.

All possibilities are listed in Table 1.

Table 1. All possibilities in dimension two
a2a_{2} a1a_{1} Δ=a124a2a0\Delta=a_{1}^{2}-4a_{2}a_{0} Minimum positive root Is an obstruction?
=0=0 0\geq 0 - - -
=0=0 <0<0 - a0/a1-a_{0}/a_{1} \surd
>0>0 0\geq 0 - - -
>0>0 <0<0 <0<0 - -
>0>0 <0<0 0\geq 0 (a1+Δ)/(2a2)-(a_{1}+\Delta)/(2a_{2}) \surd
<0<0 - - (a1+Δ)/(2a2)-(a_{1}+\Delta)/(2a_{2}) \surd

For example, by the classification of Kodaira, the ruled surfaces of genus f1f\geq 1 (may see Table 10 in chapter VI of [2], where uses gg to denote genus) have c12=8(1f)c_{1}^{2}=8(1-f). Thus there is no global solution to Hermitian-symplectic flow (3.1) on ruled surfaces of genus great that 11.

As the end of this section, we prove again that non-Kähler surfaces admit no static metric with λ<0\lambda<0 without help of classification of compact complex surfaces.

Theorem 5.2.

Non-Kähler compact complex surfaces admit no static metric with λ<0\lambda<0.

Proof.

Without loss of generality, we only need to consider the case of λ=1\lambda=-1.

Assume ω0\omega_{0} is a static metric such that ω0=ρ1,1(ω0)\omega_{0}=-\rho^{1,1}(\omega_{0}). Notice that ω0\omega_{0} can be extended to a Hermitian-symplectic form by selecting φ0=ρ2,0(ω0)\varphi_{0}=-\rho^{2,0}(\omega_{0}). And now ω(t)(1+t)ω0\omega(t)\triangleq(1+t)\omega_{0} is a global solution to flow (3.1) because ρ1,1\rho^{1,1} is a homogeneous operator with respect to metric, i.e. ρ1,1(Cω)=ρ1,1(ω)\rho^{1,1}(C\omega)=\rho^{1,1}(\omega) for any constant CC. In this situation, we have φ(t)=(1+t)φ0\varphi(t)=(1+t)\varphi_{0} and

F(t)=(φ(t),φ(t))ω(t)=Mφ(t)φ¯(t)=(1+t)2(φ0,φ0)ω0.\displaystyle F(t)=(\varphi(t),\varphi(t))_{\omega(t)}=\int_{M}\varphi(t)\wedge\bar{\varphi}(t)=(1+t)^{2}(\varphi_{0},\varphi_{0})_{\omega_{0}}.

From Lemma 5.1, we know that F(t)F(t) is monotonically decreasing. So we must have (φ0,φ0)ω0=0(\varphi_{0},\varphi_{0})_{\omega_{0}}=0. That means φ0=0\varphi_{0}=0 and ω0\omega_{0} is a Kähler metric. Thus we complete our proof. ∎

6. Appendix

We give a proof of the lemma used above.

Lemma 6.1.

Given a Hermitian manifold (M2n,g)(M^{2n},g) and denote ω\omega the fundamental form corresponding to metric gg. Then we have

(ωn2β)=(n2)!trg(β)\displaystyle*(\omega^{n-2}\wedge\beta)=-(n-2)!\text{\rm tr}_{g}(\beta)

for arbitrary (2,1)-forms β\beta.

Proof.

Because we can do all calculation in the tangent space at a fixed point, so we choose an orthonormal basis so that

ω=1pdzpdz¯p\displaystyle\omega=\sqrt{-1}\sum_{p}dz^{p}\wedge d\bar{z}^{p}

And the volume form is

1n!ωn=(1)ndz1dz¯1dzndz¯n\displaystyle\frac{1}{n!}\omega^{n}=(\sqrt{-1})^{n}dz^{1}\wedge d\bar{z}^{1}\wedge\cdots\wedge dz^{n}\wedge d\bar{z}^{n}

Recall that actions of the Hodge operator on (n,n1)(n,n-1)-forms are

(dz1dz¯1dzsdz¯s^dzndz¯n)=(1)ndzs\displaystyle*(dz^{1}\wedge d\bar{z}^{1}\wedge\cdots\wedge dz^{s}\wedge\widehat{d\bar{z}^{s}}\wedge\cdots\wedge dz^{n}\wedge d\bar{z}^{n})=(-\sqrt{-1})^{n}dz^{s}

Assume β=12βijk¯dzidzjdz¯k\beta=\frac{1}{2}\beta_{ij\bar{k}}dz^{i}\wedge dz^{j}\wedge d\bar{z}^{k} such that βijk¯+βjik¯=0\beta_{ij\bar{k}}+\beta_{ji\bar{k}}=0.

By direct calculation,

ωn2β\displaystyle\omega^{n-2}\wedge\beta =(1)n2(n2)!(s<tdz1dzsdz¯s^dzsdz¯s^dz¯n)\displaystyle=(\sqrt{-1})^{n-2}(n-2)!(\sum_{s<t}dz^{1}\wedge\cdots\wedge dz^{s}\wedge\widehat{d\bar{z}^{s}}\wedge\cdots\wedge dz^{s}\wedge\widehat{d\bar{z}^{s}}\wedge\cdots\wedge d\bar{z}^{n})
(12βijk¯dzidzjdz¯k)\displaystyle\qquad\qquad\wedge(\frac{1}{2}\beta_{ij\bar{k}}dz^{i}\wedge dz^{j}\wedge d\bar{z}^{k})
=(1)n2(n2)!kβskk¯dz1dzsdz¯s^dz¯n\displaystyle=(\sqrt{-1})^{n-2}(n-2)!\sum_{k}\beta_{sk\bar{k}}dz^{1}\wedge\cdots\wedge dz^{s}\wedge\widehat{d\bar{z}^{s}}\wedge\cdots\wedge d\bar{z}^{n}

So

(ωn2β)\displaystyle*(\omega^{n-2}\wedge\beta) =(n2)!kβskk¯dzs\displaystyle=-(n-2)!\sum_{k}\beta_{sk\bar{k}}dz^{s}
=(n2)!trg(β)\displaystyle=-(n-2)!\text{tr}_{g}(\beta)

The last equal uses the definition of trg()\text{tr}_{g}(\cdot) and note that we calculate under orthonormal basis. ∎

References

  • [1] Vestislav Apostolov, Jeffrey Streets, and Yury Ustinovskiy. Variational structure and uniqueness of generalized Kähler-Ricci solitons. arXiv:2109.10295, September 2021.
  • [2] Wolf P. Barth, Klaus Hulek, Chris A. M. Peters, and Antonius Van de Ven. Compact complex surfaces. Springer-Verlag, Berlin, second edition, 2004.
  • [3] Jean-Michel Bismut. A local index theorem for non-Kähler manifolds. Math. Ann., 284(4):681–699, 1989.
  • [4] F. A. Bogomolov. Classification of surfaces of class VII0 with b2=0b_{2}=0. Izv. Akad. Nauk SSSR Ser. Mat., 40(2):273–288, 469, 1976.
  • [5] F. A. Bogomolov. Surfaces of class VII0{\rm VII}_{0} and affine geometry. Izv. Akad. Nauk SSSR Ser. Mat., 46(4):710–761, 896, 1982.
  • [6] Mario Garcia-Fernandez, Joshua Jordan, and Jeffrey Streets. Non-Kähler Calabi-Yau geometry and pluriclosed flow. arXiv:2106.13716.
  • [7] Mario Garcia-Fernandez and Jeffrey Streets. Generalized Ricci Flow. arXiv e-prints, page arXiv:2008.07004, August 2020.
  • [8] Paul Gauduchon. Le théorème de l’excentricité nulle. C. R. Acad. Sci. Paris Sér. A-B, 285(5):A387–A390, 1977.
  • [9] Joshua Jordan and Jeffrey Streets. On a Calabi-type estimate for pluriclosed flow. Adv. Math., 366:107097, 18, 2020.
  • [10] Grisha Perelman. The entropy formula for the Ricci flow and its geometric applications. arXiv math/0211159, November 2002.
  • [11] Grisha Perelman. Finite extinction time for the solutions to the Ricci flow on certain three-manifolds. arXiv math/0307245, July 2003.
  • [12] Grisha Perelman. Ricci flow with surgery on three-manifolds. arXiv math/0303109, March 2003.
  • [13] Jeffrey Streets. Pluriclosed flow, Born-Infeld geometry, and rigidity results for generalized Kähler manifolds. Comm. Partial Differential Equations, 41(2):318–374, 2016.
  • [14] Jeffrey Streets. Pluriclosed flow on manifolds with globally generated bundles. Complex Manifolds, 3(1):222–230, 2016.
  • [15] Jeffrey Streets. Pluriclosed flow on generalized Kähler manifolds with split tangent bundle. J. Reine Angew. Math., 739:241–276, 2018.
  • [16] Jeffrey Streets. Pluriclosed flow and the geometrization of complex surfaces. In Geometric analysis—in honor of Gang Tian’s 60th birthday, volume 333 of Progr. Math., pages 471–510. Birkhäuser/Springer, Cham, [2020] ©2020.
  • [17] Jeffrey Streets and Gang Tian. A parabolic flow of pluriclosed metrics. Int. Math. Res. Not. IMRN, (16):3101–3133, 2010.
  • [18] Jeffrey Streets and Gang Tian. Hermitian curvature flow. J. Eur. Math. Soc. (JEMS), 13(3):601–634, 2011.
  • [19] Jeffrey Streets and Gang Tian. Regularity results for pluriclosed flow. Geom. Topol., 17(4):2389–2429, 2013.