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

Bismut Einstein Metrics on Compact Complex Manifolds

Abstract.

We observe that, for a Bismut Einstein metric, the (2,0)-part of Bismut Ricci form is an eigenvector of the Chern Laplacian. With the help of this observation, we prove that a Bismut Einstein metric with non-zero Einstein constant is Kähler Einstein. Additionally, for Bismut Einstein metrics with zero Einstein constant, we prove that they are actually Bismut Ricci flat.

1. Introduction

The pluriclosed flow introduced by Streets and Tian[12, 13] is a parabolic flow, which evolves a pluriclosed metric by its Bismut Ricci curvature. As Ricci flow and Kähler-Ricci flow, a Einstein-type metric (if exists) can generate a solution to the flow by a rescaling depended only on time. This is the simplest case of solitons (see e.g., [8, 9, 15]). So understanding the property of Bismut Einstein metrics is helpful to investigate the behavior of pluriclosed flow, which can be applied to classify some manifolds, especially the Kodaira’s class VII surfaces (see [11, 14]).

Firstly, we give the definition of Bismut Einstein metrics.

Definition 1.1.

A pluriclosed metric ω\omega is called Bismut Einstein with Einstein constant λ\lambda if

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

Here ρ1,1\rho^{1,1} is the (1,1)(1,1)-part of Bismut Ricci form that will be defined explicitly in Section 2.1. Since ρ1,1\rho^{1,1} is not elliptic in general, we add pluriclosed condition in the definition. But notice that in the case of λ0\lambda\neq 0, a solution to (1.1) is automatically pluriclosed for ρ1,1\rho^{1,1} is always pluriclosed. To see this, we rewrite the Bismut Ricci form as ρ=ρ1,1+ρ2,0+ρ0,2\rho=\rho^{1,1}+\rho^{2,0}+\rho^{0,2} by bi-degree. Since ρ\rho is real and closed, we obtain

ρ2,0=0,ρ1,1+¯ρ2,0=0.\displaystyle\partial\rho^{2,0}=0,\qquad\partial\rho^{1,1}+\bar{\partial}\rho^{2,0}=0.
Definition 1.2.

We say a manifold admits a Hermitian-symplectic (HS) structure if it admits a HS form, which is a real closed 2-form with positive definite (1,1)-part.

Kähler metrics are special examples of HS forms. And the (1,1)-part of any HS form is precisely pluriclosed. In fact, a Bismut Einstein metric with λ0\lambda\neq 0 can be extended to a HS form as the (1,1)-part (see [17]).

One motivation of this paper is a question asked by Streets and Tian in [12]. Using the classification of compact surfaces (see e.g., [2]), they show that only Kähler surfaces admit HS structures in dimension 2. And they ask is it valid in high dimensional cases?

Question 1.3 (Streets & Tian in [12]).

Is there a non-Kähler manifold admitting HS structures?

Authors of [17] find that HS forms are preserved by pluriclosed flow. They deform a HS form along pluriclosed flow and prove that the limitation (if exists) must be Kähler in dimension 2. This gives a way to consider Question 1.3. One can study the canonical HS forms obtained by deforming HS forms along pluriclosed flow. In this viewpoint, Bismut Einstein metric with λ0\lambda\neq 0 is an important class of canonical HS forms since it is the fixed point of pluriclosed flow up to a rescaling. Then a natural question is that whether Bismut Einstein metrics with λ0\lambda\neq 0 are Kähler? A classification of solitons given by Streets in [10] gives an affirmative answer in dimension 2. For high dimensional cases, this paper also gives an affirmative answer.

More precisely, we prove that

Theorem 1.4.

On compact complex manifolds, we have

(a) Bismut Einstein metrics with non-zero Einstein constant are Kähler Einstein;

(b) Bismut Einstein metrics with zero Einstein constant are Bismut Ricci flat.

In the case of λ=0\lambda=0, there are some non-Kähler examples (see Example 2.7-Example 2.10 in [5]). And all those examples are actually Bismut flat.

To prove Theorem 1.4, we establish a Bochner formula expressed in terms of Bismut Ricci curvature. Meanwhile, we obtain some vanishing results under certain conditions on Bismut Ricci curvature and Bismut scalar curvature (see Section 4). More analogues vanishing theorems about Bismut connection can be found in [1, 7].

Here is an outline of the rest paper. In section 2, we recall some basic notions that will be used later. In section 3, we collect some calculation results about Chern Laplacian and ¯\bar{\partial}-Laplacian, which are different in the non-Kähler case. In section 4, we establish a Bochner formula expressed in terms of Bismut Ricci curvature and obtain some vanishing results of Dolbeault cohomology. In section 5, we give an observation about Bismut Einstein metrics and use it to complete the proof of Theorem 1.4. In section 6, we obtain a stronger vanishing theorem on surfaces. As an application, we give another proof of Theorem 1.4 in the surface case.

Acknowledgments. I want to express my gratitude to my advisor, Professor Gang Tian, for his helpful suggestions and patient guidance. I also thanks Professor Jeffery Streets and Professor Mario Garcia-Fernandez for helpful comments and notifying me the paper [5]. And thanks Professor Stefan Ivanov and Giuseppe Barbaro for their helpful comments on an earlier version.

2. Preliminary

In this section, we give a quick review of some basic notions which will be used later.

2.1. Bismut connection and Bismut Ricci form

Given a complex manifold (M2n,J)(M^{2n},J) with a Hermitian metric gg. Bismut connection B\nabla^{\text{\rm B}} is the unique connection satisfying

Bg=0,BJ=0,B(x,y,z)+B(z,y,x)=0,\displaystyle\nabla^{\text{\rm B}}g=0,\qquad\nabla^{\text{\rm B}}J=0,\qquad B(x,y,z)+B(z,y,x)=0,

in which

B(x,y,z)=g(xByyBx[x,y],z)\displaystyle B(x,y,z)=g(\nabla^{\text{\rm B}}_{x}y-\nabla^{\text{\rm B}}_{y}x-[x,y],z)

is the tensor induced by torsion operator (see e.g., [3]). Notice that BB is a real 33-form, which is closed if and only if the metric is pluriclosed.

We denote ρ\rho the Ricci form of Bismut connection. It is well know that ρ\rho is a closed real 22 form (see e.g., [4]). If we rewrite is as ρ=ρ1,1+ρ2,0+ρ2,0¯\rho=\rho^{1,1}+\rho^{2,0}+\overline{\rho^{2,0}} by bidegree, then we have

ρ1,1(ω)\displaystyle\rho^{1,1}(\omega) =ω¯¯ω1¯logdetg\displaystyle=-\partial\partial^{*}\omega-\bar{\partial}\bar{\partial}^{*}\omega-\sqrt{-1}\partial\bar{\partial}\log\det g
ρ2,0(ω)\displaystyle\rho^{2,0}(\omega) =¯ω\displaystyle=-\partial\bar{\partial}^{*}\omega

where ω(,)=g(J,)\omega(\cdot,\cdot)=g(J\cdot,\cdot) is the fundamental form.

2.2. Chern connection without Kähler assumption

In this subsection we review some basic facts of Chern connection \nabla without Kähler assumption. In such a case, it is not Levi-Civita connection anymore. And the torsion tensor is

T(x,y,z)=g(xyyx[x,y],z).\displaystyle T(x,y,z)=g(\nabla_{x}y-\nabla_{y}x-[x,y],z).

In local coordinates, the Christoffel symbol of Chern connection is

Γijs=gt¯sigjt¯\displaystyle\Gamma_{ij}^{s}=g^{\bar{t}s}\partial_{i}g_{j\bar{t}}

Then

Tijt¯=T(i,j,t¯)=Tijsgst¯=igjt¯jgit¯.\displaystyle T_{ij\bar{t}}=T(\partial_{i},\partial_{j},\partial_{\bar{t}})=T_{ij}^{s}g_{s\bar{t}}=\partial_{i}g_{j\bar{t}}-\partial_{j}g_{i\bar{t}}.

Notice that

ω=1igjt¯dzidzjdz¯t=12Tijt¯dzidzjdz¯t\displaystyle\partial\omega=\sqrt{-1}\partial_{i}g_{j\bar{t}}dz^{i}\wedge dz^{j}\wedge d\bar{z}^{t}=\frac{\sqrt{-1}}{2}T_{ij\bar{t}}dz^{i}\wedge dz^{j}\wedge d\bar{z}^{t}

So if we regard ω\partial\omega as just a tensor, then we have ω=1T\partial\omega=\sqrt{-1}T.

Definition 2.1.

We can define the second Chern Ricci SS and a quadratic term QQ respectively by

Sij¯=gq¯pΩpq¯ij¯,Qij¯=gq¯pgt¯sTisq¯Tjtp¯¯=gq¯pgt¯sTisq¯Tj¯t¯p\displaystyle S_{i\bar{j}}=g^{\bar{q}p}\Omega_{p\bar{q}i\bar{j}},\quad Q_{i\bar{j}}=g^{\bar{q}p}g^{\bar{t}s}T_{is\bar{q}}\overline{T_{jt\bar{p}}}=g^{\bar{q}p}g^{\bar{t}s}T_{is\bar{q}}T_{\bar{j}\bar{t}p}

where Ωij¯pq¯\Omega_{i\bar{j}p\bar{q}} is the Chern curvature tensor.

Recall that the Chern Ricci is defined by Ricij¯=gq¯pΩij¯pq¯\text{\rm Ric}_{i\bar{j}}=g^{\bar{q}p}\Omega_{i\bar{j}p\bar{q}}. In Kähler case, SS is precisely the Chern Ricci. In general, they are different since Chern connection has torsion. An basic fact is that SS is alway a second order elliptic operator with respect to metrics.

For pluriclosed metrics, there is a relationship between SS and the (1,1)(1,1)-part of Bismut Ricci form ρ1,1\rho^{1,1} (see e.g., [12, 13]). More explicitly, if we assume

ρ1,1=1Pij¯dzidz¯j\displaystyle\rho^{1,1}=\sqrt{-1}P_{i\bar{j}}dz^{i}\wedge d\bar{z}^{j}

then we have

(2.1) P=SQ.\displaystyle P=S-Q.
Remark 2.2.

In fact, the quadratic term QQ is non-negative. To see this, we can choose a normal coordinates such that gij¯(x)=δijg_{i\bar{j}}(x)=\delta_{ij} and Q(x)Q(x) is diagonal at a fixed point xx. Then we have

Qii¯(x)=s,pTisp¯Ti¯s¯p=s,p|Tisp¯|20\displaystyle Q_{i\bar{i}}(x)=\sum_{s,p}T_{is\bar{p}}T_{\bar{i}\bar{s}p}=\sum_{s,p}|T_{is\bar{p}}|^{2}\geq 0

2.3. Lefschetz-type operator

For the convenience of use later, we recall the Lefschetz-type operator in this subsection.

Definition 2.3.

Let γ\gamma be a form. The Lefschetz-type operator LγL_{\gamma} is defined by

Lγα=γα.\displaystyle L_{\gamma}\alpha=\gamma\wedge\alpha.

And its conjugate adjoint LγL^{*}_{\gamma} is defined by

(Lγα,β)=(α,γ¯β).\displaystyle(L^{*}_{\gamma}\alpha,\beta)=(\alpha,\bar{\gamma}\wedge\beta).

where (,)=g(,¯)(\cdot,\cdot)=g(\cdot,\overline{\cdot}) is the pointwise Hermitian inner product.

Notice that in the case of γ=ω\gamma=\omega, those operators defined above are the classical Lefschetz operator LL and Λ\Lambda.

We give a local expression of the Lefschetz-type operator in a special case which will be used later.

Lemma 2.4.

Given a (p,1)(p,1)-form

α=1p!αi1ipk¯dzi1dzipdz¯k,αiuivk¯+αiviuk¯=0\displaystyle\alpha=\frac{1}{p!}\alpha_{i_{1}\cdots i_{p}\bar{k}}dz^{i_{1}}\wedge\cdots\wedge dz^{i_{p}}\wedge d\bar{z}^{k},\quad\alpha_{\cdots i_{u}\cdots i_{v}\cdots\bar{k}}+\alpha_{\cdots i_{v}\cdots i_{u}\cdots\bar{k}}=0

and a (1,0)(1,0)-form γ=γsdzs\gamma=\gamma_{s}dz^{s}. We have

Lγα=1p!(1)pgk¯lαi1ipk¯γldzi1dzip.\displaystyle L^{*}_{\gamma}\alpha=\frac{1}{p!}(-1)^{p}g^{\bar{k}l}\alpha_{i_{1}\cdots i_{p}\bar{k}}\gamma_{l}dz^{i_{1}}\wedge\cdots\wedge dz^{i_{p}}.
Proof.

Choose an arbitrary (p,0)(p,0)-form β=βj1jpdzj1dzjp\beta=\beta_{j_{1}\cdots j_{p}}dz^{j_{1}}\wedge\cdots\wedge dz^{j_{p}}. Assume

Lγα=1p!ηi1ipdzi1dzip.\displaystyle L^{*}_{\gamma}\alpha=\frac{1}{p!}\eta_{i_{1}\cdots i_{p}}dz^{i_{1}}\wedge\cdots\wedge dz^{i_{p}}.

By direct computation, we obtain

(Lγα,β)=gj1¯i1gjp¯ipηi1ipβj1jp¯\displaystyle(L^{*}_{\gamma}\alpha,\beta)=g^{\bar{j_{1}}i_{1}}\cdots g^{\bar{j_{p}}i_{p}}\eta_{i_{1}\cdots i_{p}}\overline{\beta_{j_{1}\cdots j_{p}}}

and

γ¯β=(1)pβj1jpγl¯dzj1dzjpdz¯l\displaystyle\bar{\gamma}\wedge\beta=(-1)^{p}\beta_{j_{1}\cdots j_{p}}\overline{\gamma_{l}}dz^{j_{1}}\wedge\cdots\wedge dz^{j_{p}}\wedge d\bar{z}^{l}

By definition,

(Lγα,β)\displaystyle(L^{*}_{\gamma}\alpha,\beta) =(α,γ¯β)=(1)pgj1¯i1gjp¯ipgk¯lαi1ipk¯γlβj1jp¯\displaystyle=(\alpha,\bar{\gamma}\wedge\beta)=(-1)^{p}g^{\bar{j_{1}}i_{1}}\cdots g^{\bar{j_{p}}i_{p}}g^{\bar{k}l}\alpha_{i_{1}\cdots i_{p}\bar{k}}\gamma_{l}\overline{\beta_{j_{1}\cdots j_{p}}}

Then we get

gj1¯i1gjp¯ipηi1ipβj1jp¯=(1)pgj1¯i1gjp¯ipgk¯lαi1ipk¯γlβj1jp¯\displaystyle g^{\bar{j_{1}}i_{1}}\cdots g^{\bar{j_{p}}i_{p}}\eta_{i_{1}\cdots i_{p}}\overline{\beta_{j_{1}\cdots j_{p}}}=(-1)^{p}g^{\bar{j_{1}}i_{1}}\cdots g^{\bar{j_{p}}i_{p}}g^{\bar{k}l}\alpha_{i_{1}\cdots i_{p}\bar{k}}\gamma_{l}\overline{\beta_{j_{1}\cdots j_{p}}}

which implies

ηi1ip=(1)pgk¯lαi1ipk¯γl\displaystyle\eta_{i_{1}\cdots i_{p}}=(-1)^{p}g^{\bar{k}l}\alpha_{i_{1}\cdots i_{p}\bar{k}}\gamma_{l}

3. Some Calculation Results of Laplacians

In this section, we collect some calculation results which will be used later. For easy of notations, let us start with some definitions.

Definition 3.1.

For a tensor AA, we denote its gradient by

A=iAdzi,¯A=i¯Adz¯i\displaystyle\nabla A=\nabla_{i}A_{\cdots}dz^{i}\otimes\cdots,\quad\overline{\nabla}A=\nabla_{\bar{i}}A_{\cdots}d\bar{z}^{i}\otimes\cdots

We denote the Chern Laplacian by

Δ=gq¯ppq¯,Δ¯=gq¯pq¯p.\displaystyle\Delta=g^{\bar{q}p}\nabla_{p}\nabla_{\bar{q}},\quad\overline{\Delta}=g^{\bar{q}p}\nabla_{\bar{q}}\nabla_{p}.

We denote the \partial-Laplacian and ¯\bar{\partial}-Laplacian by

Δ¯=¯¯+¯¯,Δ=+.\displaystyle\Delta_{\bar{\partial}}=\bar{\partial}^{*}\bar{\partial}+\bar{\partial}\bar{\partial}^{*},\qquad\Delta_{\partial}=\partial^{*}\partial+\partial\partial^{*}.
Definition 3.2.

Given two tensors Aij¯A_{\cdots i\bar{j}} and Bp1psq¯1q¯tB_{p_{1}\cdots p_{s}\bar{q}_{1}\cdots\bar{q}_{t}}. We can define a tensor ABA\circ B by

(AB)p1psq¯1q¯t=\displaystyle(A\circ B)_{\cdots p_{1}\cdots p_{s}\bar{q}_{1}\cdots\bar{q}_{t}}= m=1sgβ¯αApmβ¯Bpm1αpm+1\displaystyle-\sum_{m=1}^{s}g^{\bar{\beta}\alpha}A_{\cdots p_{m}\bar{\beta}}B_{\cdots p_{m-1}\alpha p_{m+1}\cdots}
+n=1tgβ¯αAαq¯nBq¯n1β¯q¯n+1\displaystyle+\sum_{n=1}^{t}g^{\bar{\beta}\alpha}A_{\cdots\alpha\bar{q}_{n}}B_{\cdots\bar{q}_{n-1}\bar{\beta}\bar{q}_{n+1}\cdots}
Remark 3.3.

Using the notation defined above, we can rewrite the Chern curvature operator as

2A(x,y)2A(y,x)=ΩxyA=(ΩA)(x,y)\displaystyle\nabla^{2}A(x,y)-\nabla^{2}A(y,x)=\Omega_{xy}\circ A=(\Omega\circ A)(x,y)

Taking trace, we obtain

ΔAΔ¯A=SA\displaystyle\Delta A-\overline{\Delta}A=S\circ A

where SS is the second Chern Ricci defined in Section 2.2.

Remark 3.4.

We would like to give a remark on the computation of ABA\circ B in a special case where B=Bi1ip(T1,0M)pB=B_{i_{1}\cdots i_{p}}\in(T^{*1,0}M)^{\otimes p} and A=Aij¯A=A_{i\bar{j}} is Hermitian, i.e., Aij¯¯=Aji¯\overline{A_{i\bar{j}}}=A_{j\bar{i}}. For convenience, we choose a normal coordinates for a fixed point xx such that gij¯(x)=δijg_{i\bar{j}}(x)=\delta_{ij} and Aij¯(x)=λiδijA_{i\bar{j}}(x)=\lambda_{i}\delta_{ij}. Then by definition

(AB)i1ip(x)=(m=1pλim)Bi1ip.\displaystyle(A\circ B)_{i_{1}\cdots i_{p}}(x)=-(\sum_{m=1}^{p}\lambda_{i_{m}})B_{i_{1}\cdots i_{p}}.

In particular, we have

gB=pB\displaystyle g\circ B=-pB

Moreover, if the subscripts of BB is skew-symmetric, then in the case of p=np=n we have

(AB)i1in(x)=(m=1nλm)Bi1in=(trgA)(x)Bi1in\displaystyle(A\circ B)_{i_{1}\cdots i_{n}}(x)=-(\sum_{m=1}^{n}\lambda_{m})B_{i_{1}\cdots i_{n}}=-(\text{\rm tr}_{g}A)(x)B_{i_{1}\cdots i_{n}}

Since xx is arbitrary, we get

AB=(trgA)B.\displaystyle A\circ B=-(\text{\rm tr}_{g}A)B.

We recall the formula of order exchange of Laplacian and gradient and leave the proof in the appendix.

Lemma 3.5.

Let AA be a tensor field. We have

kk¯l¯3Al¯kk¯3A=Ωkl¯k¯AkTk¯l¯s¯s¯ATk¯l¯s¯ks¯A\displaystyle\nabla^{3}_{k\bar{k}\bar{l}}A-\nabla^{3}_{\bar{l}k\bar{k}}A=\Omega_{k\bar{l}}\circ\nabla_{\bar{k}}A-\nabla_{k}T_{\bar{k}\bar{l}}^{\bar{s}}\nabla_{\bar{s}}A-T_{\bar{k}\bar{l}}^{\bar{s}}\nabla_{k}\nabla_{\bar{s}}A

Then we give two lemmas often used in the computation of ¯\bar{\partial}-Laplacian and \partial-Laplacian. And we leave the tedious calculations in the appendix.

Lemma 3.6.

Given a (p,q+1)(p,q+1)-form

α=1p!(q+1)!αi1ipj¯1j¯qk¯dzi1dzipdz¯j1dz¯jqdz¯k\displaystyle\alpha=\frac{1}{p!(q+1)!}\alpha_{i_{1}\cdots i_{p}\bar{j}_{1}\cdots\bar{j}_{q}\bar{k}}dz^{i_{1}}\wedge\cdots\wedge dz^{i_{p}}\wedge d\bar{z}^{j_{1}}\wedge\cdots\wedge d\bar{z}^{j_{q}}\wedge d\bar{z}^{k}

satisfying

αiuiv+αiviu=0,αj¯uj¯v+αj¯vj¯u=0.\displaystyle\alpha_{\cdots i_{u}\cdots i_{v}\cdots}+\alpha_{\cdots i_{v}\cdots i_{u}\cdots}=0,\quad\alpha_{\cdots\bar{j}_{u}\cdots\bar{j}_{v}\cdots}+\alpha_{\cdots\bar{j}_{v}\cdots\bar{j}_{u}\cdots}=0.

We have

¯α=1p!q!ηi1ipj¯1j¯qdzi1dzipdz¯j1dz¯jq\displaystyle\bar{\partial}^{*}\alpha=\frac{1}{p!q!}\eta_{i_{1}\cdots i_{p}\bar{j}_{1}\cdots\bar{j}_{q}}dz^{i_{1}}\wedge\cdots\wedge dz^{i_{p}}\wedge d\bar{z}^{j_{1}}\wedge\cdots\wedge d\bar{z}^{j_{q}}

in which

ηi1ipj¯1j¯q=(1)p+q+1gk¯l(lαi1ipj¯1j¯qk¯+αi1ipj¯1j¯qk¯Tlss+12gt¯sm=1qαi1ipj¯1t¯j¯qk¯Tslj¯m).\displaystyle\eta_{i_{1}\cdots i_{p}\bar{j}_{1}\cdots\bar{j}_{q}}=(-1)^{p+q+1}g^{\bar{k}l}\Big{(}\nabla_{l}\alpha_{i_{1}\cdots i_{p}\bar{j}_{1}\cdots\bar{j}_{q}\bar{k}}+\alpha_{i_{1}\cdots i_{p}\bar{j}_{1}\cdots\bar{j}_{q}\bar{k}}T_{ls}^{s}+\frac{1}{2}g^{\bar{t}s}\sum_{m=1}^{q}\alpha_{i_{1}\cdots i_{p}\bar{j}_{1}\cdots\bar{t}\cdots\bar{j}_{q}\bar{k}}T_{sl\bar{j}_{m}}\Big{)}.

Similarly, we have

Lemma 3.7.

Given a (p+1,q)(p+1,q)-form

α=1(p+1)!q!αli1ipj¯1j¯qdzldzi1dzipdz¯j1dz¯jq\displaystyle\alpha=\frac{1}{(p+1)!q!}\alpha_{li_{1}\cdots i_{p}\bar{j}_{1}\cdots\bar{j}_{q}}dz^{l}\wedge dz^{i_{1}}\wedge\cdots\wedge dz^{i_{p}}\wedge d\bar{z}^{j_{1}}\wedge\cdots\wedge d\bar{z}^{j_{q}}

satisfying

αiuiv+αiviu=0,αj¯uj¯v+αj¯vj¯u=0.\displaystyle\alpha_{\cdots i_{u}\cdots i_{v}\cdots}+\alpha_{\cdots i_{v}\cdots i_{u}\cdots}=0,\quad\alpha_{\cdots\bar{j}_{u}\cdots\bar{j}_{v}\cdots}+\alpha_{\cdots\bar{j}_{v}\cdots\bar{j}_{u}\cdots}=0.

We have

α=1p!q!ηi1ipj¯1j¯qdzi1dzipdz¯j1dz¯jq\displaystyle\partial^{*}\alpha=\frac{1}{p!q!}\eta_{i_{1}\cdots i_{p}\bar{j}_{1}\cdots\bar{j}_{q}}dz^{i_{1}}\wedge\cdots\wedge dz^{i_{p}}\wedge d\bar{z}^{j_{1}}\wedge\cdots\wedge d\bar{z}^{j_{q}}

in which

ηi1ipj¯1j¯q=gk¯l(k¯αli1ipj¯1j¯q+αli1ipj¯1j¯qTkss¯12gt¯sm=1qαli1sipj¯1j¯qTkti¯m¯).\displaystyle\eta_{i_{1}\cdots i_{p}\bar{j}_{1}\cdots\bar{j}_{q}}=-g^{\bar{k}l}\Big{(}\nabla_{\bar{k}}\alpha_{li_{1}\cdots i_{p}\bar{j}_{1}\cdots\bar{j}_{q}}+\alpha_{li_{1}\cdots i_{p}\bar{j}_{1}\cdots\bar{j}_{q}}\overline{T_{ks}^{s}}-\frac{1}{2}g^{\bar{t}s}\sum_{m=1}^{q}\alpha_{li_{1}\cdots s\cdots i_{p}\bar{j}_{1}\cdots\bar{j}_{q}}\overline{T_{kt\bar{i}_{m}}}\Big{)}.
Remark 3.8.

In Lemma 3.6, when q=0q=0, we do not have the last term in the expression of ηi1ipj¯1j¯q\eta_{i_{1}\cdots i_{p}\bar{j}_{1}\cdots\bar{j}_{q}}. In other words, the expression for (p,1)(p,1)-form is

ηi1ip=(1)p+1gk¯l(lαi1ipk¯+αi1ipk¯Tlss).\displaystyle\eta_{i_{1}\cdots i_{p}}=(-1)^{p+1}g^{\bar{k}l}\Big{(}\nabla_{l}\alpha_{i_{1}\cdots i_{p}\bar{k}}+\alpha_{i_{1}\cdots i_{p}\bar{k}}T_{ls}^{s}\Big{)}.

This is also valid for Lemma 3.7 in the case of p=0p=0.

Remark 3.9.

Applying Lemma 3.6 to ω\omega, we have

¯ω=1Tissdzi.\displaystyle\bar{\partial}^{*}\omega=\sqrt{-1}T_{is}^{s}dz^{i}.

3.1. Laplacians applied to functions

Next proposition gives the relationship between two Laplacians when applied to functions.

Proposition 3.10.

For a smooth function ff, we have

Δ¯f=Δf1(¯f,ω)\displaystyle-\Delta_{\bar{\partial}}f=\Delta f-\sqrt{-1}(\bar{\partial}f,\partial^{*}\omega)
Δf=Δf+1(f,¯ω)\displaystyle-\Delta_{\partial}f=\Delta f+\sqrt{-1}(\partial f,\bar{\partial}^{*}\omega)
Proof.

For Chern Laplacian, we have

Δf=trω(1¯f)=trω(1ij¯fdzidz¯j)=gq¯ppq¯f\displaystyle\Delta f=\text{\rm tr}_{\omega}(\sqrt{-1}\partial\bar{\partial}f)=\text{\rm tr}_{\omega}(\sqrt{-1}\partial_{i\bar{j}}fdz^{i}\wedge d\bar{z}^{j})=g^{\bar{q}p}\partial_{p\bar{q}}f

For ¯\bar{\partial}-Laplacian, we have

Δ¯f=¯¯f+¯¯f=¯¯f\displaystyle\Delta_{\bar{\partial}}f=\bar{\partial}\bar{\partial}^{*}f+\bar{\partial}^{*}\bar{\partial}f=\bar{\partial}^{*}\bar{\partial}f

Applying Lemma 3.6 to ¯f\bar{\partial}f and noticing Remark 3.9, we get

Δ¯f\displaystyle\Delta_{\bar{\partial}}f =gk¯ll(¯f)k¯gk¯l(¯f)k¯Tlss\displaystyle=-g^{\bar{k}l}\nabla_{l}(\bar{\partial}f)_{\bar{k}}-g^{\bar{k}l}(\bar{\partial}f)_{\bar{k}}T_{ls}^{s}
=gk¯llk¯f+1gk¯l(¯f)k¯(¯ω)l\displaystyle=-g^{\bar{k}l}\partial_{l\bar{k}}f+\sqrt{-1}g^{\bar{k}l}(\bar{\partial}f)_{\bar{k}}(\bar{\partial}^{*}\omega)_{l}
=Δf+1(¯f,ω)\displaystyle=-\Delta f+\sqrt{-1}(\bar{\partial}f,\partial^{*}\omega)

Similarly, applying Lemma 3.7, we get

Δf\displaystyle\Delta_{\partial}f =gk¯lk¯(f)lgk¯l(f)lTkss¯\displaystyle=-g^{\bar{k}l}\nabla_{\bar{k}}(\partial f)_{l}-g^{\bar{k}l}(\partial f)_{l}\overline{T_{ks}^{s}}
=gk¯llk¯f1gk¯l(f)l(ω)k¯\displaystyle=-g^{\bar{k}l}\partial_{l\bar{k}}f-\sqrt{-1}g^{\bar{k}l}(\partial f)_{l}(\partial^{*}\omega)_{\bar{k}}
=Δf1(f,¯ω)\displaystyle=-\Delta f-\sqrt{-1}(\partial f,\bar{\partial}^{*}\omega)

Remark 3.11.

From Proposition 3.10, we know that if ¯ω=0\bar{\partial}^{*}\omega=0, then Δ¯f=Δf=Δf-\Delta_{\bar{\partial}}f=-\Delta_{\partial}f=\Delta f for all functions. Notice that

¯ω=0ω=0ωn1=0\displaystyle\bar{\partial}^{*}\omega=0\iff-*\partial*\omega=0\iff\partial\omega^{n-1}=0

So Chern Laplacian and ¯\bar{\partial}-Laplacian are the same when applied to functions if and only if the metric is balanced, i.e., dωn1=0d\omega^{n-1}=0,

Recall that

MΔ¯f𝑑V=(Δ¯f,1)2=(¯¯f,1)2=(¯f,¯1)2=0\displaystyle\int_{M}\Delta_{\bar{\partial}}fdV=(\Delta_{\bar{\partial}}f,1)_{2}=(\bar{\partial}^{*}\bar{\partial}f,1)_{2}=(\bar{\partial}f,\bar{\partial}1)_{2}=0

where (,)2(\cdot,\cdot)_{2} is the L2L^{2} Hermitian inner product. But for Chern Laplacian, this is not valid in general. Actually, we have

Proposition 3.12.

If and only if the metric is Gauduchon (i.e., ¯ωn1=0\partial\bar{\partial}\omega^{n-1}=0), we have

MΔf𝑑V=0\displaystyle\int_{M}\Delta fdV=0

for all functions.

Proof.

By definition, we have

MΔf𝑑V\displaystyle\int_{M}\Delta fdV =Mtrω(1¯f)𝑑V=M(1¯f,ω)𝑑V\displaystyle=\int_{M}\text{\rm tr}_{\omega}(\sqrt{-1}\partial\bar{\partial}f)dV=\int_{M}(\sqrt{-1}\partial\bar{\partial}f,\omega)dV
=(1¯f,ω)2=(1f,¯ω)2\displaystyle=(\sqrt{-1}\partial\bar{\partial}f,\omega)_{2}=(\sqrt{-1}f,\bar{\partial}^{*}\partial^{*}\omega)_{2}

Notice

¯ω=0¯ω=0¯ωn1=0.\displaystyle\bar{\partial}^{*}\partial^{*}\omega=0\iff*\partial\bar{\partial}*\omega=0\iff\partial\bar{\partial}\omega^{n-1}=0.

So we complete our proof. ∎

Remark 3.13.

Gauduchon[6] proves that any complex manifold admits metrics satisfying ¯ωn1=0\partial\bar{\partial}\omega^{n-1}=0. And in the case of surface, Gauduchon condition is precisely pluriclosed condition.

3.2. Laplacians applied to (p,0)(p,0)-forms

For the purpose of use later, we give the relationship between those Laplacians when applied on (p,0)(p,0)-forms.

Proposition 3.14.

For a (p,0)(p,0)-form α\alpha, we have

Δ¯α=Δα+(1)p1L¯ω¯α\displaystyle\Delta_{\bar{\partial}}\alpha=-\Delta\alpha+(-1)^{p}\sqrt{-1}L^{*}_{\bar{\partial}^{*}\omega}\bar{\partial}\alpha

where L()()L^{*}_{(\cdot)}(\cdot) is the Lefschetz-type operator defined in section 2.3.

Proof.

In local coordinates, we assume

α=1p!αi1ipdzi1dzip,αiuiv+αiviu=0.\displaystyle\alpha=\frac{1}{p!}\alpha_{i_{1}\cdots i_{p}}dz^{i_{1}}\wedge\cdots\wedge dz^{i_{p}},\qquad\alpha_{\cdots i_{u}\cdots i_{v}\cdots}+\alpha_{\cdots i_{v}\cdots i_{u}\cdots}=0.

By definition,

¯α=1p!(1)pk¯αi1ipdzi1dzipdz¯k.\displaystyle\bar{\partial}\alpha=\frac{1}{p!}(-1)^{p}\partial_{\bar{k}}\alpha_{i_{1}\cdots i_{p}}dz^{i_{1}}\wedge\cdots\wedge dz^{i_{p}}\wedge d\bar{z}^{k}.

Applying Lemma 3.6 to ¯α\bar{\partial}\alpha, we obtain

¯¯α\displaystyle\bar{\partial}^{*}\bar{\partial}\alpha =1p!(gk¯llk¯αi1ip+gk¯lk¯αi1ipTlss)dzi1dzip\displaystyle=-\frac{1}{p!}(g^{\bar{k}l}\nabla_{l}\partial_{\bar{k}}\alpha_{i_{1}\cdots i_{p}}+g^{\bar{k}l}\partial_{\bar{k}}\alpha_{i_{1}\cdots i_{p}}T_{ls}^{s})dz^{i_{1}}\wedge\cdots\wedge dz^{i_{p}}

From Lemma 2.4 and Remark 3.9, we know

1p!gk¯lk¯αi1ipTlssdzi1dzip\displaystyle-\frac{1}{p!}g^{\bar{k}l}\partial_{\bar{k}}\alpha_{i_{1}\cdots i_{p}}T_{ls}^{s}dz^{i_{1}}\wedge\cdots\wedge dz^{i_{p}} =11p!gk¯lk¯αi1ip(¯ω)ldzi1dzip\displaystyle=\sqrt{-1}\frac{1}{p!}g^{\bar{k}l}\partial_{\bar{k}}\alpha_{i_{1}\cdots i_{p}}(\bar{\partial}^{*}\omega)_{l}dz^{i_{1}}\wedge\cdots\wedge dz^{i_{p}}
=(1)p1L¯ω¯α.\displaystyle=(-1)^{p}\sqrt{-1}L^{*}_{\bar{\partial}^{*}\omega}\bar{\partial}\alpha.

This implies

Δ¯α=Δα+(1)p1L¯ω¯α.\displaystyle\Delta_{\bar{\partial}}\alpha=-\Delta\alpha+(-1)^{p}\sqrt{-1}L^{*}_{\bar{\partial}^{*}\omega}\bar{\partial}\alpha.

4. A Bochner Formula

In this section, we give a Bochner formula for pluriclosed metrics in terms of Bismut Ricci curvature. As an application, we can obtain some vanishing results on the the Dolbeault cohomology H¯p,0(M;)H_{\bar{\partial}}^{p,0}(M;\mathbb{C}).

4.1. Bochner formula

For ease of notations, we start with some definition.

Definition 4.1.

Given a tensor A=Aij¯A=A_{i\bar{j}} satisfying Aij¯¯=Aji¯\overline{A_{i\bar{j}}}=A_{j\bar{i}}. We define the first eigenvalue function λ(A)\lambda_{*}(A) by

λ(A)(x)=min0ξTx1,0MA(x)(ξ,ξ¯)|ξ|2.\displaystyle\lambda_{*}(A)(x)=\min_{0\neq\xi\in T^{1,0}_{x}M}\frac{A(x)(\xi,\overline{\xi})}{|\xi|^{2}}.
Remark 4.2.

We choose a normal coordinates at a fixed point xx such that gij¯(x)=δijg_{i\bar{j}}(x)=\delta_{ij} and Aij¯(x)=λiδijA_{i\bar{j}}(x)=\lambda_{i}\delta_{ij}. It is easy to see that λ(A)(x)=min{λ1,,λn}\lambda_{*}(A)(x)=\min\{\lambda_{1},\cdots,\lambda_{n}\}. Thus AA is non-negative (resp. positive) if and only if λ(A)\lambda_{*}(A) is a non-negative (resp. positive) function.

Now we can state the Bochner formula.

Theorem 4.3.

For any tensor AA, we have

Δ|A|2=(ΔA,A)+(A,ΔA)+|A|2+|¯A|2(A,SA)\displaystyle\Delta|A|^{2}=(\Delta A,A)+(A,\Delta A)+|\nabla A|^{2}+|\overline{\nabla}A|^{2}-(A,S\circ A)

where SS is the second Chern Ricci and (,)(\cdot,\cdot) denotes the Hermitian inner product.

Proof.

By definition and noticing that Chern connection is compatible with metric, we obtain

Δ|A|2\displaystyle\Delta|A|^{2} =gq¯ppq¯g(A,A¯)\displaystyle=g^{\bar{q}p}\nabla_{p}\nabla_{\bar{q}}g(A,\overline{A})
=gq¯p{g(pq¯A,A¯)+g(A,pq¯A¯)+g(pA,q¯A¯)+g(q¯A,pA¯)}\displaystyle=g^{\bar{q}p}\Big{\{}g(\nabla_{p}\nabla_{\bar{q}}A,\overline{A})+g(A,\nabla_{p}\nabla_{\bar{q}}\overline{A})+g(\nabla_{p}A,\nabla_{\bar{q}}\overline{A})+g(\nabla_{\bar{q}}A,\nabla_{p}\overline{A})\Big{\}}
=(ΔA,A)+(A,Δ¯A)+|A|2+|¯A|2\displaystyle=(\Delta A,A)+(A,\overline{\Delta}A)+|\nabla A|^{2}+|\overline{\nabla}A|^{2}

Remark 3.3 tells us that

Δ¯A=ΔASA,\displaystyle\overline{\Delta}A=\Delta A-S\circ A,

which completes this proof. ∎

From equation (2.1), we can directly obtain

Corollary 4.4.

If the metric is pluriclosed, then we have

Δ|A|2=(ΔA,A)+(A,ΔA)+|A|2+|¯A|2(A,PA)(A,QA)\displaystyle\Delta|A|^{2}=(\Delta A,A)+(A,\Delta A)+|\nabla A|^{2}+|\overline{\nabla}A|^{2}-(A,P\circ A)-(A,Q\circ A)

4.2. Some vanishing results

In this subsection, we apply the Bochner formula established in Section 4.1 to get some vanishing results. Firstly, we state a lemma that will be used repeatedly.

Lemma 4.5.

If α\alpha is a (p,0)(p,0)-form and A=Aij¯A=A_{i\bar{j}} is Hermitian, then we have

(α,Aα)λ(A)|α|2\displaystyle-(\alpha,A\circ\alpha)\geq\lambda_{*}(A)|\alpha|^{2}
Proof.

We check it in a normal coordinates at a point xx such that gij¯(x)=δijg_{i\bar{j}}(x)=\delta_{ij} and Aij¯(x)=λiδijA_{i\bar{j}}(x)=\lambda_{i}\delta_{ij}. From Remark 3.4 and Remark 4.2, we know that

(α,Aα)(x)\displaystyle-(\alpha,A\circ\alpha)(x) =i1,,ipαi1ip(Aα)i1ip¯=i1,,ip(m=1pλim)αi1ipαi1ip¯\displaystyle=-\sum_{i_{1},\cdots,i_{p}}\alpha_{i_{1}\cdots i_{p}}\overline{(A\circ\alpha)_{i_{1}\cdots i_{p}}}=\sum_{i_{1},\cdots,i_{p}}(\sum_{m=1}^{p}\lambda_{i_{m}})\alpha_{i_{1}\cdots i_{p}}\overline{\alpha_{i_{1}\cdots i_{p}}}
λ(A)(x)|α|2(x).\displaystyle\geq\lambda_{*}(A)(x)\cdot|\alpha|^{2}(x).

Proposition 4.6.

Given a compact complex manifold (M2n,J)(M^{2n},J). If MM admits a metric such that the second Chern Ricci SS is positive, then the Dolbeault cohomology H¯p,0(M;)H_{\bar{\partial}}^{p,0}(M;\mathbb{C}) is trivial for 1pn1\leq p\leq n.

Proof.

Given a ¯\bar{\partial}-harmonic (p,0)(p,0)-form α\alpha. We have ¯α=¯α=0\bar{\partial}\alpha=\bar{\partial}^{*}\alpha=0. From Proposition 3.14, we obtain

Δα=Δ¯α+(1)p1L¯ω¯α=0.\displaystyle\Delta\alpha=-\Delta_{\bar{\partial}}\alpha+(-1)^{p}\sqrt{-1}L^{*}_{\bar{\partial}^{*}\omega}\bar{\partial}\alpha=0.

Applying Theorem 4.3 to α\alpha, we get

(4.1) Δ|α|2=|α|2+|¯α|2(α,Sα).\displaystyle\Delta|\alpha|^{2}=|\nabla\alpha|^{2}+|\overline{\nabla}\alpha|^{2}-(\alpha,S\circ\alpha).

By Lemma 4.5, we have

(α,Sα)λ(S)|α|2\displaystyle-(\alpha,S\circ\alpha)\geq\lambda_{*}(S)|\alpha|^{2}

Assume |α|2|\alpha|^{2} achieves the maximum at point xMx_{M}. By the maximum principle, we have

0Δ|α|2(xM)\displaystyle 0\geq\Delta|\alpha|^{2}(x_{M}) =|α|2(xM)+|¯α|2(xM)(α,Sα)(xM)\displaystyle=|\nabla\alpha|^{2}(x_{M})+|\overline{\nabla}\alpha|^{2}(x_{M})-(\alpha,S\circ\alpha)(x_{M})
|α|2(xM)+|¯α|2(xM)+λ(S)(xM)|α|2(xM)\displaystyle\geq|\nabla\alpha|^{2}(x_{M})+|\overline{\nabla}\alpha|^{2}(x_{M})+\lambda_{*}(S)(x_{M})\cdot|\alpha|^{2}(x_{M})
0\displaystyle\geq 0

The last equality uses the positive assumption on SS. So we obtain |α|2(xM)=0|\alpha|^{2}(x_{M})=0, which implies α=0\alpha=0. Since H¯p,0(M;)kerΔ¯|Λp,0H_{\bar{\partial}}^{p,0}(M;\mathbb{C})\simeq\ker\Delta_{\bar{\partial}}\Big{|}_{\Lambda^{p,0}}, we complete this proof. ∎

In particular, for H¯n,0(M;)H_{\bar{\partial}}^{n,0}(M;\mathbb{C}), we only need the assumption that Chern scalar curvature s=trωSs=\text{\rm tr}_{\omega}S is positive.

Proposition 4.7.

Given a compact complex manifold (M2n,J)(M^{2n},J). If MM admits a metric such that the Chern scalar curvature s=trωSs=\text{\rm tr}_{\omega}S is positive, then the Dolbeault cohomology H¯n,0(M;)H_{\bar{\partial}}^{n,0}(M;\mathbb{C}) is trivial.

Proof.

Remark 3.4 tells that

(α,Sα)=(α,trgSα)=s|α|2\displaystyle-(\alpha,S\circ\alpha)=(\alpha,\text{\rm tr}_{g}S\cdot\alpha)=s|\alpha|^{2}

for any (n,0)(n,0)-form α\alpha. And the rest argument is similar to Proposition 4.6. ∎

For Gauduchon metric, we can weaken the assumption slightly.

Proposition 4.8.

Given a compact complex manifold (M2n,J)(M^{2n},J). If MM admits a Gauduchon metric ω\omega such that SS is non-negative and is strictly positive at one point, then the Dolbeault cohomology H¯p,0(M;)H_{\bar{\partial}}^{p,0}(M;\mathbb{C}) is trivial for 1pn1\leq p\leq n.

Proof.

Let α\alpha be a ¯\bar{\partial}-harmonic (p,0)(p,0)-form. Integrating equation (4.1) on MM and applying Proposition 3.12, we obtain

0=α2+¯α2+M(α,Sα)dV\displaystyle 0=\|\nabla\alpha\|^{2}+\|\overline{\nabla}\alpha\|^{2}+\int_{M}-(\alpha,S\circ\alpha)dV

where \|\cdot\| denotes the L2L^{2} norm. Since (α,Sα)-(\alpha,S\circ\alpha) is a non-negative function, we get

α2=¯α2=M(α,Sα)dV=0,\displaystyle\|\nabla\alpha\|^{2}=\|\overline{\nabla}\alpha\|^{2}=\int_{M}-(\alpha,S\circ\alpha)dV=0,

which means α=¯α=0\nabla\alpha=\overline{\nabla}\alpha=0. We claim that |α|2|\alpha|^{2} is a constant function on MM. To see this, we take derivatives to |α|2|\alpha|^{2}.

i|α|2=i(α,α)=(iα,α)+(α,i¯α)=0\displaystyle\partial_{i}|\alpha|^{2}=\nabla_{i}(\alpha,\alpha)=(\nabla_{i}\alpha,\alpha)+(\alpha,\nabla_{\bar{i}}\alpha)=0

Similarly, we have i¯|α|2=0\partial_{\bar{i}}|\alpha|^{2}=0.

On the other hand, we have

0\displaystyle 0 =M(α,Sα)dVMλ(S)|α|2𝑑V=|α|2Mλ(S)𝑑V\displaystyle=\int_{M}-(\alpha,S\circ\alpha)dV\geq\int_{M}\lambda_{*}(S)|\alpha|^{2}dV=|\alpha|^{2}\int_{M}\lambda_{*}(S)dV

The integral on the right hand side is a positive number for λ(S)\lambda_{*}(S) is strictly positive at one point. So we get α=0\alpha=0 and complete this proof. ∎

Similarly, for H¯n,0(M;)H^{n,0}_{\bar{\partial}}(M;\mathbb{C}) we only need the assumption on the Chern scalar curvature ss.

Proposition 4.9.

Given a compact complex manifold (M2n,J)(M^{2n},J). If MM admits a Gauduchon metric such that the Chern scalar curvature ss is non-negative and is strictly positive at one point, then the Dolbeault cohomology H¯n,0(M;)H_{\bar{\partial}}^{n,0}(M;\mathbb{C}) is trivial.

Proof.

The argument is similar to Proposition 4.8. Firstly, we prove that |α|2|\alpha|^{2} is a constant function by integration. Then use the equation (α,Sα)=s|α|2-(\alpha,S\circ\alpha)=s|\alpha|^{2} to show that |α|2|\alpha|^{2} is actually 0. ∎

For pluriclosed metric, the second Chern Ricci can be represented by Bismut Ricci. Thus we can state the vanishing result in terms of Bismut Ricci. In other words, the positivity of the (1,1)(1,1)-part of Bismut Ricci form can give obstructions to the Dolbeault cohomology H¯p,0(M;)H^{p,0}_{\bar{\partial}}(M;\mathbb{C}).

Theorem 4.10.

Given a compact complex manifold with a pluriclosed metric (M2n,J,ω)(M^{2n},J,\omega).

(a) If ρ1,1\rho^{1,1} is strictly positive definite, then H¯p,0(M;)H_{\bar{\partial}}^{p,0}(M;\mathbb{C}) is trivial for 1pn1\leq p\leq n;

(b) If Bismut scalar curvature r=trωρ1,1r=\text{\rm tr}_{\omega}\rho^{1,1} is positive, then H¯n,0(M;)H_{\bar{\partial}}^{n,0}(M;\mathbb{C}) is trivial.

Proof.

Since equation (2.1) and Q0Q\geq 0, we can obtain (a) and (b) from Proposition 4.6 and Proposition 4.7, respectively. ∎

5. Bismut Einstein Metrics on Compact Complex Manifolds

In this section, we will discuss the Bismut Einstein metric on compact complex manifolds and use the Bochner formula established in Section 4.1 to give a proof of Theorem 1.4. Firstly, we introduce an observation that plays an important role in the proof of the main theorem.

We assume in local coordinates

ρ2,0=12ϕijdzidzj\displaystyle\rho^{2,0}=\frac{\sqrt{-1}}{2}\phi_{ij}dz^{i}\wedge dz^{j}

where ϕij+ϕji=0\phi_{ij}+\phi_{ji}=0. For convenience, we will also use ϕ=1ρ2,0\phi=-\sqrt{-1}\rho^{2,0} to denote the (2,0)(2,0)-part of Bismut Ricci in the following.

5.1. An observation of Bismut Einstein metrics

Let us begin with a formula given by authors of [5].

Lemma 5.1 (Proposition 3.24 of [5]).

The (2,0)(2,0)-part of Bismut Ricci can be represented by

ϕ=divT,\displaystyle\phi=-\text{\rm div}T,

where (divT)ij=gk¯llTijk¯(\text{\rm div}T)_{ij}=-g^{\bar{k}l}\nabla_{l}T_{ij\bar{k}}.

Now we give an observation about Bismut Einstein metrics.

Proposition 5.2.

If ω\omega is a Bismut Einstein metric with Einstein constant λ\lambda, then we have

Δϕ=λϕ.\displaystyle\Delta\phi=-\lambda\phi.

Moreover, in the case of λ0\lambda\neq 0 we have

λT+¯ϕ=0.\displaystyle\lambda T+\overline{\nabla}\phi=0.
Proof.

In the case of λ=0\lambda=0. Since ρ=ρ2,0+ρ2,0¯\rho=\rho^{2,0}+\overline{\rho^{2,0}} is closed, we obtain that ρ2,0\rho^{2,0} is closed. Applying Lemma 3.14, we get

Δρ2,0=Δ¯ρ2,0=0.\displaystyle\Delta\rho^{2,0}=-\Delta_{\bar{\partial}}\rho^{2,0}=0.

Recall that ϕ=1ρ2,0\phi=-\sqrt{-1}\rho^{2,0}. So we obtain Δϕ=0\Delta\phi=0.

From now on we assume λ0\lambda\neq 0. We have

ω=1λρ1,1=1λ¯ρ2,0\displaystyle\partial\omega=\frac{1}{\lambda}\partial\rho^{1,1}=-\frac{1}{\lambda}\bar{\partial}\rho^{2,0}

In local coordinates, the equation above becomes

12Tijk¯dzidzjdz¯k=1λ12k¯ϕijdzidzjdz¯k\displaystyle\frac{\sqrt{-1}}{2}T_{ij\bar{k}}dz^{i}\wedge dz^{j}\wedge d\bar{z}^{k}=-\frac{1}{\lambda}\frac{\sqrt{-1}}{2}\partial_{\bar{k}}\phi_{ij}dz^{i}\wedge dz^{j}\wedge d\bar{z}^{k}

which implies

(5.1) Tijk¯=1λk¯ϕij=1λk¯ϕij.\displaystyle T_{ij\bar{k}}=-\frac{1}{\lambda}\partial_{\bar{k}}\phi_{ij}=-\frac{1}{\lambda}\nabla_{\bar{k}}\phi_{ij}.

Applying Lemma 5.1, we obtain

ϕij=gk¯llTijk¯=1λgk¯llk¯ϕij=1λΔϕij\displaystyle\phi_{ij}=g^{\bar{k}l}\nabla_{l}T_{ij\bar{k}}=-\frac{1}{\lambda}g^{\bar{k}l}\nabla_{l}\nabla_{\bar{k}}\phi_{ij}=-\frac{1}{\lambda}\Delta\phi_{ij}

This completes the proof. ∎

Next, we state a differential equation which will be used later.

Lemma 5.3.

Assume ω\omega is a Bismut Einstein metric with Einstein constant λ\lambda. Then we have

(5.2) Δ|ϕ|2=|ϕ|2+|¯ϕ|2(ϕ,Qϕ)\displaystyle\Delta|\phi|^{2}=|\nabla\phi|^{2}+|\overline{\nabla}\phi|^{2}-(\phi,Q\circ\phi)

where ϕ=1ρ2,0\phi=-\sqrt{-1}\rho^{2,0}.

Proof.

Applying Corollary 4.4 to ϕ\phi, we obtain

Δ|ϕ|2=(Δϕ,ϕ)+(ϕ,Δϕ)+|ϕ|2+|¯ϕ|2(ϕ,Pϕ)(ϕ,Qϕ).\displaystyle\Delta|\phi|^{2}=(\Delta\phi,\phi)+(\phi,\Delta\phi)+|\nabla\phi|^{2}+|\overline{\nabla}\phi|^{2}-(\phi,P\circ\phi)-(\phi,Q\circ\phi).

From Remark 3.4 we know

(ϕ,Pϕ)=(ϕ,(λg)ϕ)=2λ|ϕ|2\displaystyle-(\phi,P\circ\phi)=-(\phi,(\lambda g)\circ\phi)=2\lambda|\phi|^{2}

Then applying Proposition 5.2, we get

Δ|ϕ|2\displaystyle\Delta|\phi|^{2} =(λϕ,ϕ)+(ϕ,λϕ)+|ϕ|2+|¯ϕ|2+2λ|ϕ|2(ϕ,Qϕ)\displaystyle=(-\lambda\phi,\phi)+(\phi,-\lambda\phi)+|\nabla\phi|^{2}+|\overline{\nabla}\phi|^{2}+2\lambda|\phi|^{2}-(\phi,Q\circ\phi)
=|ϕ|2+|¯ϕ|2(ϕ,Qϕ)\displaystyle=|\nabla\phi|^{2}+|\overline{\nabla}\phi|^{2}-(\phi,Q\circ\phi)

5.2. Proof of Theorem 1.4

Proof of Theorem 1.4.

From Remark 2.2 and Lemma 4.5, we know that

(ϕ,Qϕ)0.\displaystyle-(\phi,Q\circ\phi)\geq 0.

So Lemma 5.3 shows that

Δ|ϕ|20\displaystyle\Delta|\phi|^{2}\geq 0

Notice that the manifold is compact. Applying the strong maximum principle to |ϕ|2|\phi|^{2}, we obtain that |ϕ|2|\phi|^{2} is a constant function. Thus equation (5.2) becomes

0=|ϕ|2+|¯ϕ|2(ϕ,Qϕ)\displaystyle 0=|\nabla\phi|^{2}+|\overline{\nabla}\phi|^{2}-(\phi,Q\circ\phi)

Since the three terms on the right hand side are non-negative, we obtain

ϕ=¯ϕ=0,(ϕ,Qϕ)=0.\displaystyle\nabla\phi=\overline{\nabla}\phi=0,\quad-(\phi,Q\circ\phi)=0.

Case I λ0\lambda\neq 0. From Proposition 5.2 we know that

T=1λ¯ϕ=0,\displaystyle T=-\frac{1}{\lambda}\overline{\nabla}\phi=0,

This shows that the metric is Kähler and thus Kähler-Einstein.

Case II λ=0\lambda=0. By definition 3.2 and definition 2.1, we have

(ϕ,Qϕ)\displaystyle-(\phi,Q\circ\phi) =gs¯igt¯jϕijgb¯a(Qsb¯ϕat+Qtb¯ϕsa)¯=2gs¯igt¯jga¯bQbs¯ϕijϕa¯t¯\displaystyle=g^{\bar{s}i}g^{\bar{t}j}\phi_{ij}\overline{g^{\bar{b}a}(Q_{s\bar{b}}\phi_{at}+Q_{t\bar{b}}\phi_{sa})}=2g^{\bar{s}i}g^{\bar{t}j}g^{\bar{a}b}Q_{b\bar{s}}\phi_{ij}\phi_{\bar{a}\bar{t}}
=2gs¯igt¯jga¯bgq¯pgl¯kTbpl¯Ts¯q¯kϕijϕa¯t¯\displaystyle=2g^{\bar{s}i}g^{\bar{t}j}g^{\bar{a}b}g^{\bar{q}p}g^{\bar{l}k}T_{bp\bar{l}}T_{\bar{s}\bar{q}k}\phi_{ij}\phi_{\bar{a}\bar{t}}
=2gt¯jgq¯pgl¯k(ga¯bTbpl¯ϕa¯t¯)(gs¯iTs¯q¯kϕij)\displaystyle=2g^{\bar{t}j}g^{\bar{q}p}g^{\bar{l}k}(g^{\bar{a}b}T_{bp\bar{l}}\phi_{\bar{a}\bar{t}})(g^{\bar{s}i}T_{\bar{s}\bar{q}k}\phi_{ij})
=2gt¯jgq¯pgl¯k(ga¯bTbpl¯ϕa¯t¯)(gi¯sTsqk¯ϕi¯j¯¯)\displaystyle=2g^{\bar{t}j}g^{\bar{q}p}g^{\bar{l}k}(g^{\bar{a}b}T_{bp\bar{l}}\phi_{\bar{a}\bar{t}})(\overline{g^{\bar{i}s}T_{sq\bar{k}}\phi_{\bar{i}\bar{j}}})

Notice that the right hand side is the norm of the tensor Apl¯t¯=ga¯bTbpl¯ϕa¯t¯A_{p\bar{l}\bar{t}}=g^{\bar{a}b}T_{bp\bar{l}}\phi_{\bar{a}\bar{t}}. Thus we obtain

ga¯bTbpl¯ϕa¯t¯=0.\displaystyle g^{\bar{a}b}T_{bp\bar{l}}\phi_{\bar{a}\bar{t}}=0.

Differentiating both sides and noticing ¯ϕ=¯ϕ=0\overline{\nabla}\phi=\overline{\nabla}\phi=0, we get

0=k(ga¯bTbpl¯ϕa¯t¯)=ga¯bkTbpl¯ϕa¯t¯\displaystyle 0=\nabla_{k}(g^{\bar{a}b}T_{bp\bar{l}}\phi_{\bar{a}\bar{t}})=g^{\bar{a}b}\nabla_{k}T_{bp\bar{l}}\phi_{\bar{a}\bar{t}}

Taking trace and applying Lemma 5.1, we get

ga¯bϕbpϕa¯t¯=0\displaystyle g^{\bar{a}b}\phi_{bp}\phi_{\bar{a}\bar{t}}=0

So we obtain |ϕ|2=0|\phi|^{2}=0 by taking trace again. The proof is completed since ρ2,0=1ϕ\rho^{2,0}=\sqrt{-1}\phi. ∎

6. The Case of Surface

In section 4.2, we show that for Gauduchon metrics, those vanishing results still hold under a slightly weaker conditions. And in dimension 2, Gauduchon condition is precisely the pluriclosed condition. So we can obtain a slightly stronger vanishing result in the case of surfaces. As an application, we can give another proof of Theorem 1.4 in the case of surfaces.

6.1. A vanishing theorem for surfaces

Let us begin with an observation in dimension 2.

Lemma 6.1.

In dimension 2, we have

Q=12|T|2g\displaystyle Q=\frac{1}{2}|T|^{2}g
Proof.

We prove it in a normal coordinates such that gij¯(x)=δijg_{i\bar{j}}(x)=\delta_{ij} and Qij¯(x)=λiδijQ_{i\bar{j}}(x)=\lambda_{i}\delta_{ij}. By definition, we have

Q11¯(x)=i,jT1ij¯T1¯i¯j=j|T12j¯|2\displaystyle Q_{1\bar{1}}(x)=\sum_{i,j}T_{1i\bar{j}}T_{\bar{1}\bar{i}j}=\sum_{j}|T_{12\bar{j}}|^{2}

Here we use the fact that the first two subscripts of TT are skew-symmetric. Similarly, we have

Q22¯(x)=i,jT2ij¯T2¯i¯j=j|T21j¯|2\displaystyle Q_{2\bar{2}}(x)=\sum_{i,j}T_{2i\bar{j}}T_{\bar{2}\bar{i}j}=\sum_{j}|T_{21\bar{j}}|^{2}

Thus we complete this proof. ∎

Now we state a vanishing theorem for compact surfaces.

Theorem 6.2.

Given a compact complex surface with a pluriclosed metric (M4,J,ω)(M^{4},J,\omega).

(a) If ρ1,1\rho^{1,1} is non-negative definite, then either ω\omega is Kähler or H¯p,0(M;)H_{\bar{\partial}}^{p,0}(M;\mathbb{C}) is trivial for 1p21\leq p\leq 2;

(b) If Bismut scalar curvature r=trωρ1,1r=\text{\rm tr}_{\omega}\rho^{1,1} is non-negative, then either ω\omega is Kähler or H¯2,0(M;)H_{\bar{\partial}}^{2,0}(M;\mathbb{C}) is trivial.

Proof.

Notice that in dimension 2, pluriclosed metrics are Gauduchon metrics. And from Lemma 6.1 we know that

S=P+Q=P+12|T|2g\displaystyle S=P+Q=P+\frac{1}{2}|T|^{2}g

If ω\omega is non-Kähler, then there exists a point xx such that |T|2(x)>0|T|^{2}(x)>0. So part (a) and part (b) are obtained from Proposition 4.8 and Proposition 4.9, respectively. ∎

6.2. Bismut Einstein metrics on surfaces

We first give an proposition that is similar to Proposition 5.2.

Proposition 6.3.

Given a Bismut Einstein metric with Einstein constant λ\lambda on a complex surface. We have

Δ¯ρ2,0=λρ2,0\displaystyle\Delta_{\bar{\partial}}\rho^{2,0}=\lambda\rho^{2,0}
Proof.

If λ=0\lambda=0, then we have ρ=ρ2,0+ρ2,0¯\rho=\rho^{2,0}+\overline{\rho^{2,0}}. We obtain that ρ2,0\rho^{2,0} is closed for ρ\rho is closed. Thus ρ2,0\rho^{2,0} is ¯\bar{\partial}-harmonic.

If λ0\lambda\neq 0, then we have

(6.1) 0=ρ1,1+¯ρ2,0=λω+¯ρ2,0\displaystyle 0=\partial\rho^{1,1}+\bar{\partial}\rho^{2,0}=\lambda\partial\omega+\bar{\partial}\rho^{2,0}

and ρ2,0=0\partial\rho^{2,0}=0. On the other hand, we have (see Section 2.1)

ρ2,0=¯ω,\displaystyle\rho^{2,0}=-\partial\bar{\partial}^{*}\omega,

which implies

ρ2,0=ω=ω=1λ¯ρ2,0=1λ¯ρ2,0=1λ¯¯ρ2,0=1λΔ¯ρ2,0\displaystyle\rho^{2,0}=\partial*\partial*\omega=\partial*\partial\omega=-\frac{1}{\lambda}\partial*\bar{\partial}\rho^{2,0}=-\frac{1}{\lambda}*\partial*\bar{\partial}\rho^{2,0}=\frac{1}{\lambda}\bar{\partial}^{*}\bar{\partial}\rho^{2,0}=\frac{1}{\lambda}\Delta_{\bar{\partial}}\rho^{2,0}

The second equality is because of ω=ω*\omega=\omega in dimension 2. The third equality uses equation (6.1). And the fourth equality uses the fact that η=η*\eta=\eta for arbitrary (2,0)(2,0)-form in dimension 2, which is an application of Lefschetz theorem to primitive form (see e.g., [16]). ∎

Another proof of Theorem 1.4 in the case of surfaces..

For convenience, we denote φ=ρ2,0\varphi=\rho^{2,0}.

Case I λ<0\lambda<0. Using Proposition 6.3, we have

(φ,φ)2=(λΔ¯φ,φ)2=λ(¯¯φ,φ)2=λ(¯φ,¯φ)2.\displaystyle(\varphi,\varphi)_{2}=(\lambda\Delta_{\bar{\partial}}\varphi,\varphi)_{2}=\lambda(\bar{\partial}^{*}\bar{\partial}\varphi,\varphi)_{2}=\lambda(\bar{\partial}\varphi,\bar{\partial}\varphi)_{2}.

where (,)2(\cdot,\cdot)_{2} is the L2L^{2} Hermitian inner product. So we obtain φ=0\varphi=0 for λ\lambda is a negative number. Notice λω=ρ1,1=¯φ=0\lambda\partial\omega=\partial\rho^{1,1}=-\bar{\partial}\varphi=0, which means that ω\omega is Kähler.

Case II λ=0\lambda=0. From Proposition 6.3 we know that φ\varphi is a ¯\bar{\partial}-harmonic (2,0)(2,0)-form. Theorem 6.2 tell us that ω\omega is Kähler or φ=0\varphi=0. In fact, in both cases we have φ=0\varphi=0 because for Kähler metrics, the Bismut Ricci form is precisely the Kähler Ricci form which only has (1,1)(1,1)-part.

Case III λ>0\lambda>0. Applying Corollary 4.4 to φ\varphi, we get

(6.2) Δ|φ|2=(Δφ,φ)+(φ,Δφ)+|φ|2+|¯φ|2(φ,Pφ)(φ,Qφ)=(Δφ,φ)+(φ,Δφ)+|φ|2+|¯φ|2+(2λ+|T|2)|φ|2\begin{split}\Delta|\varphi|^{2}&=(\Delta\varphi,\varphi)+(\varphi,\Delta\varphi)+|\nabla\varphi|^{2}+|\overline{\nabla}\varphi|^{2}-(\varphi,P\circ\varphi)-(\varphi,Q\circ\varphi)\\ &=(\Delta\varphi,\varphi)+(\varphi,\Delta\varphi)+|\nabla\varphi|^{2}+|\overline{\nabla}\varphi|^{2}+(2\lambda+|T|^{2})|\varphi|^{2}\end{split}

The second row is because of Proposition 6.3 and Lemma 6.1.

Using Proposition 3.14, we get

(Δφ,φ)2=(Δ¯φ,φ)2+1(L¯ω¯φ,φ)2\displaystyle(\Delta\varphi,\varphi)_{2}=-(\Delta_{\bar{\partial}}\varphi,\varphi)_{2}+\sqrt{-1}(L^{*}_{\bar{\partial}^{*}\omega}\bar{\partial}\varphi,\varphi)_{2}

By the definition of L()()L^{*}_{(\cdot)}(\cdot) (see Section 2.3),

(L¯ω¯φ,φ)2\displaystyle(L^{*}_{\bar{\partial}^{*}\omega}\bar{\partial}\varphi,\varphi)_{2} =M(L¯ω¯φ,φ)𝑑V=M(¯φ,ωφ)𝑑V=(¯φ,ωφ)2\displaystyle=\int_{M}(L^{*}_{\bar{\partial}^{*}\omega}\bar{\partial}\varphi,\varphi)dV=\int_{M}(\bar{\partial}\varphi,\partial^{*}\omega\wedge\varphi)dV=(\bar{\partial}\varphi,\partial^{*}\omega\wedge\varphi)_{2}

By direct computation, we obtain

(¯φ,ωφ)2\displaystyle(\bar{\partial}\varphi,\partial^{*}\omega\wedge\varphi)_{2} =(¯ωφ¯,φ¯)2=M¯ωφ¯¯φ\displaystyle=(\bar{\partial}^{*}\omega\wedge\bar{\varphi},\partial\bar{\varphi})_{2}=\int_{M}\bar{\partial}^{*}\omega\wedge\bar{\varphi}\wedge*\bar{\partial}\varphi
=λM¯ωφ¯ω=λM¯ωφ¯ω\displaystyle=-\lambda\int_{M}\bar{\partial}^{*}\omega\wedge\bar{\varphi}\wedge*\partial\omega=-\lambda\int_{M}\bar{\partial}^{*}\omega\wedge\bar{\varphi}\wedge*\partial*\omega
=λM¯ωφ¯¯ω=λM¯ω¯ωφ¯\displaystyle=\lambda\int_{M}\bar{\partial}^{*}\omega\wedge\bar{\varphi}\wedge\bar{\partial}^{*}\omega=\lambda\int_{M}\bar{\partial}^{*}\omega\wedge\bar{\partial}^{*}\omega\wedge\bar{\varphi}
=0\displaystyle=0

The second row is because of λω=ρ1,1=¯φ\lambda\partial\omega=\partial\rho^{1,1}=-\bar{\partial}\varphi. And the last row uses the fact that ¯ω¯ω=0\bar{\partial}^{*}\omega\wedge\bar{\partial}^{*}\omega=0, which holds because ¯ω\bar{\partial}^{*}\omega is an odd degree form. So we obtain

(Δφ,φ)2=(Δ¯φ,φ)2=λφ2.\displaystyle(\Delta\varphi,\varphi)_{2}=-(\Delta_{\bar{\partial}}\varphi,\varphi)_{2}=-\lambda\|\varphi\|^{2}.

Similarly, we have

(φ,Δφ)2=λφ2.\displaystyle(\varphi,\Delta\varphi)_{2}=-\lambda\|\varphi\|^{2}.

Integrating equation (6.2) on the manifold and noticing Proposition 3.12, we have

0\displaystyle 0 =φ2+¯φ2+M|T|2|φ|2𝑑V\displaystyle=\|\nabla\varphi\|^{2}+\|\overline{\nabla}\varphi\|^{2}+\int_{M}|T|^{2}|\varphi|^{2}dV

which implies ¯φ=0\overline{\nabla}\varphi=0. Since λω=ρ1,1=¯φ=¯φ=0,\lambda\partial\omega=\partial\rho^{1,1}=-\bar{\partial}\varphi=-\overline{\nabla}\varphi=0, the metric ω\omega is Kähler and φ=0\varphi=0. ∎

7. Appendix

In this appendix, we give the proof of some lemmas used earlier.

7.1. Proof of Lemma 3.5

Proof.

Firstly, we recall that for a connection DD with torsion HH and curvature tensor Rm, we have

(7.1) Dxy2ADyx2A=RmxyADH(x,y)A\displaystyle D^{2}_{xy}A-D^{2}_{yx}A=\text{\rm Rm}_{xy}\circ A-D_{H(x,y)}A

Applying it to Chern connection, we get

kk¯l¯3A\displaystyle\nabla^{3}_{k\bar{k}\bar{l}}A =k(2A)k¯l¯=k(k¯l¯2A)\displaystyle=\nabla_{k}(\nabla^{2}A)_{\bar{k}\bar{l}}=\nabla_{k}(\nabla^{2}_{\bar{k}\bar{l}}A)
=k(l¯k¯2ATk¯l¯s¯s¯A)\displaystyle=\nabla_{k}(\nabla^{2}_{\bar{l}\bar{k}}A-T_{\bar{k}\bar{l}}^{\bar{s}}\nabla_{\bar{s}}A)
=kl¯k¯3AkTk¯l¯s¯s¯ATk¯l¯s¯ks¯A\displaystyle=\nabla^{3}_{k\bar{l}\bar{k}}A-\nabla_{k}T_{\bar{k}\bar{l}}^{\bar{s}}\nabla_{\bar{s}}A-T_{\bar{k}\bar{l}}^{\bar{s}}\nabla_{k}\nabla_{\bar{s}}A

The second row is because of Ωk¯l¯=0\Omega_{\bar{k}\bar{l}\cdots}=0.

Similarly, applying formula (7.1) to ¯A\overline{\nabla}A, we get

kl¯k¯3A=kl¯2(¯A)k¯=l¯kk¯3A+Ωkl¯k¯A\displaystyle\nabla^{3}_{k\bar{l}\bar{k}}A=\nabla^{2}_{k\bar{l}}(\overline{\nabla}A)_{\bar{k}}=\nabla^{3}_{\bar{l}k\bar{k}}A+\Omega_{k\bar{l}}\circ\nabla_{\bar{k}}A

Combining above together, we obtain

kk¯l¯3Al¯kk¯3A=Ωkl¯k¯AkTk¯l¯s¯s¯ATk¯l¯s¯ks¯A\displaystyle\nabla^{3}_{k\bar{k}\bar{l}}A-\nabla^{3}_{\bar{l}k\bar{k}}A=\Omega_{k\bar{l}}\circ\nabla_{\bar{k}}A-\nabla_{k}T_{\bar{k}\bar{l}}^{\bar{s}}\nabla_{\bar{s}}A-T_{\bar{k}\bar{l}}^{\bar{s}}\nabla_{k}\nabla_{\bar{s}}A

And we complete the proof. ∎

7.2. Proof of Lemma 3.6

Proof.

Since the test form β\beta can be chosen that is supported in a coordinates neighborhood, we can do calculation locally. Assume

β=βa1apb¯1b¯qdza1dzapdz¯b1dz¯bq.\displaystyle\beta=\beta_{a_{1}\cdots a_{p}\bar{b}_{1}\cdots\bar{b}_{q}}dz^{a_{1}}\wedge\cdots\wedge dz^{a_{p}}\wedge d\bar{z}^{b_{1}}\wedge\cdots\wedge d\bar{z}^{b_{q}}.

We have

(¯α,β)2\displaystyle(\bar{\partial}^{*}\alpha,\beta)_{2} =M(¯α,β)detg\displaystyle=\int_{M}(\bar{\partial}^{*}\alpha,\beta)\det g
=Mηi1ipj¯1j¯qβa1apb¯1b¯q¯ga¯1i1ga¯pipgj¯1b1gj¯qbqdetg.\displaystyle=\int_{M}\eta_{i_{1}\cdots i_{p}\bar{j}_{1}\cdots\bar{j}_{q}}\overline{\beta_{a_{1}\cdots a_{p}\bar{b}_{1}\cdots\bar{b}_{q}}}g^{\bar{a}_{1}i_{1}}\cdots g^{\bar{a}_{p}i_{p}}g^{\bar{j}_{1}b_{1}}\cdots g^{\bar{j}_{q}b_{q}}\det g.

Here we use (,)2(\cdot,\cdot)_{2} and (,)(\cdot,\cdot) to denote L2L^{2} Hermitian inner product and pointwise Hermitian inner product, respectively. By direct computation, we get

¯β=(1)p+ql¯βa1apb¯1b¯qdza1dzapdz¯b1dz¯bqdz¯l\displaystyle\bar{\partial}\beta=(-1)^{p+q}\partial_{\bar{l}}\beta_{a_{1}\cdots a_{p}\bar{b}_{1}\cdots\bar{b}_{q}}dz^{a_{1}}\wedge\cdots\wedge dz^{a_{p}}\wedge d\bar{z}^{b_{1}}\wedge\cdots\wedge d\bar{z}^{b_{q}}\wedge d\bar{z}^{l}

Then

(¯α,β)2\displaystyle(\bar{\partial}^{*}\alpha,\beta)_{2} =(α,¯β)2=M(α,¯β)detg\displaystyle=(\alpha,\bar{\partial}\beta)_{2}=\int_{M}(\alpha,\bar{\partial}\beta)\det g
=(1)p+qMαi1ipj¯1j¯qk¯lβa1apb¯1b¯q¯ga¯1i1ga¯pipgj¯1b1gj¯qbqgk¯ldetg\displaystyle=(-1)^{p+q}\int_{M}\alpha_{i_{1}\cdots i_{p}\bar{j}_{1}\cdots\bar{j}_{q}\bar{k}}\partial_{l}\overline{\beta_{a_{1}\cdots a_{p}\bar{b}_{1}\cdots\bar{b}_{q}}}g^{\bar{a}_{1}i_{1}}\cdots g^{\bar{a}_{p}i_{p}}g^{\bar{j}_{1}b_{1}}\cdots g^{\bar{j}_{q}b_{q}}g^{\bar{k}l}\det g
=(1)p+q+1Mβa1apb¯1b¯q¯l(αi1ipj¯1j¯qk¯ga¯1i1ga¯pipgj¯1b1gj¯qbqgk¯ldetg).\displaystyle=(-1)^{p+q+1}\int_{M}\overline{\beta_{a_{1}\cdots a_{p}\bar{b}_{1}\cdots\bar{b}_{q}}}\partial_{l}\Big{(}\alpha_{i_{1}\cdots i_{p}\bar{j}_{1}\cdots\bar{j}_{q}\bar{k}}g^{\bar{a}_{1}i_{1}}\cdots g^{\bar{a}_{p}i_{p}}g^{\bar{j}_{1}b_{1}}\cdots g^{\bar{j}_{q}b_{q}}g^{\bar{k}l}\det g\Big{)}.

The arbitrariness of β\beta implies

ηi1ipj¯1j¯qga¯1i1ga¯pipgj¯1b1gj¯qbqdetg\displaystyle\eta_{i_{1}\cdots i_{p}\bar{j}_{1}\cdots\bar{j}_{q}}g^{\bar{a}_{1}i_{1}}\cdots g^{\bar{a}_{p}i_{p}}g^{\bar{j}_{1}b_{1}}\cdots g^{\bar{j}_{q}b_{q}}\det g =(1)p+q+1l(αi1ipj¯1j¯qk¯ga¯1i1ga¯pip\displaystyle=(-1)^{p+q+1}\partial_{l}\Big{(}\alpha_{i_{1}\cdots i_{p}\bar{j}_{1}\cdots\bar{j}_{q}\bar{k}}g^{\bar{a}_{1}i_{1}}\cdots g^{\bar{a}_{p}i_{p}}
gj¯1b1gj¯qbqgk¯ldetg)\displaystyle\qquad\qquad g^{\bar{j}_{1}b_{1}}\cdots g^{\bar{j}_{q}b_{q}}g^{\bar{k}l}\det g\Big{)}

which gives

(1)p+q+1ηc1cpd¯1d¯q\displaystyle(-1)^{p+q+1}\eta_{c_{1}\cdots c_{p}\bar{d}_{1}\cdots\bar{d}_{q}} =gk¯llαc1cpd¯1d¯qk¯+gk¯lm=1pαc1imcpd¯1d¯qk¯gcma¯mlga¯mim\displaystyle=g^{\bar{k}l}\partial_{l}\alpha_{c_{1}\cdots c_{p}\bar{d}_{1}\cdots\bar{d}_{q}\bar{k}}+g^{\bar{k}l}\sum_{m=1}^{p}\alpha_{c_{1}\cdots i_{m}\cdots c_{p}\bar{d}_{1}\cdots\bar{d}_{q}\bar{k}}g_{c_{m}\bar{a}_{m}}\partial_{l}g^{\bar{a}_{m}i_{m}}
+gk¯lm=1qαc1cpd¯1j¯md¯qk¯gbmd¯mlgj¯mbm\displaystyle\qquad+g^{\bar{k}l}\sum_{m=1}^{q}\alpha_{c_{1}\cdots c_{p}\bar{d}_{1}\cdots\bar{j}_{m}\cdots\bar{d}_{q}\bar{k}}g_{b_{m}\bar{d}_{m}}\partial_{l}g^{\bar{j}_{m}b_{m}}
+αc1cpd¯1d¯qk¯lgk¯l+gk¯lαc1cpd¯1d¯qk¯gt¯slgst¯\displaystyle\qquad+\alpha_{c_{1}\cdots c_{p}\bar{d}_{1}\cdots\bar{d}_{q}\bar{k}}\partial_{l}g^{\bar{k}l}+g^{\bar{k}l}\alpha_{c_{1}\cdots c_{p}\bar{d}_{1}\cdots\bar{d}_{q}\bar{k}}g^{\bar{t}s}\partial_{l}g_{s\bar{t}}

Recall the derivative formula of inverse matrix

lgj¯i=gj¯pgq¯ilgpq¯=gj¯pΓlpi\displaystyle\partial_{l}g^{\bar{j}i}=-g^{\bar{j}p}g^{\bar{q}i}\partial_{l}g_{p\bar{q}}=-g^{\bar{j}p}\Gamma^{i}_{lp}

Changing some subscripts, we get

(1)p+q+1ηc1cpd¯1d¯q\displaystyle(-1)^{p+q+1}\eta_{c_{1}\cdots c_{p}\bar{d}_{1}\cdots\bar{d}_{q}} =gk¯llαc1cpd¯1d¯qk¯gk¯lm=1pαc1scpd¯1d¯qk¯Γlcms\displaystyle=g^{\bar{k}l}\partial_{l}\alpha_{c_{1}\cdots c_{p}\bar{d}_{1}\cdots\bar{d}_{q}\bar{k}}-g^{\bar{k}l}\sum_{m=1}^{p}\alpha_{c_{1}\cdots s\cdots c_{p}\bar{d}_{1}\cdots\bar{d}_{q}\bar{k}}\Gamma^{s}_{lc_{m}}
gk¯lgt¯sm=1qαc1cpd¯1t¯d¯qk¯lgsd¯m\displaystyle\qquad-g^{\bar{k}l}g^{\bar{t}s}\sum_{m=1}^{q}\alpha_{c_{1}\cdots c_{p}\bar{d}_{1}\cdots\bar{t}\cdots\bar{d}_{q}\bar{k}}\partial_{l}g_{s\bar{d}_{m}}
αc1cpd¯1d¯qk¯gk¯lgt¯ssglt¯+αc1cpd¯1d¯qk¯gk¯lgt¯slgst¯\displaystyle\qquad-\alpha_{c_{1}\cdots c_{p}\bar{d}_{1}\cdots\bar{d}_{q}\bar{k}}g^{\bar{k}l}g^{\bar{t}s}\partial_{s}g_{l\bar{t}}+\alpha_{c_{1}\cdots c_{p}\bar{d}_{1}\cdots\bar{d}_{q}\bar{k}}g^{\bar{k}l}g^{\bar{t}s}\partial_{l}g_{s\bar{t}}

Notice that

gk¯lgt¯sm=1qαc1cpd¯1t¯d¯qk¯lgsd¯m\displaystyle-g^{\bar{k}l}g^{\bar{t}s}\sum_{m=1}^{q}\alpha_{c_{1}\cdots c_{p}\bar{d}_{1}\cdots\bar{t}\cdots\bar{d}_{q}\bar{k}}\partial_{l}g_{s\bar{d}_{m}} =12gk¯lgt¯sm=1qαc1cpd¯1t¯d¯qk¯(lgsd¯msgld¯m)\displaystyle=-\frac{1}{2}g^{\bar{k}l}g^{\bar{t}s}\sum_{m=1}^{q}\alpha_{c_{1}\cdots c_{p}\bar{d}_{1}\cdots\bar{t}\cdots\bar{d}_{q}\bar{k}}(\partial_{l}g_{s\bar{d}_{m}}-\partial_{s}g_{l\bar{d}_{m}})
=12gk¯lgt¯sm=1qαc1cpd¯1t¯d¯qk¯Tsld¯m\displaystyle=\frac{1}{2}g^{\bar{k}l}g^{\bar{t}s}\sum_{m=1}^{q}\alpha_{c_{1}\cdots c_{p}\bar{d}_{1}\cdots\bar{t}\cdots\bar{d}_{q}\bar{k}}T_{sl\bar{d}_{m}}

Then we obtain

(1)p+q+1ηc1cpd¯1d¯q\displaystyle(-1)^{p+q+1}\eta_{c_{1}\cdots c_{p}\bar{d}_{1}\cdots\bar{d}_{q}} =gk¯llαc1cpd¯1d¯qk¯+12gk¯lgt¯sm=1qαc1cpd¯1t¯d¯qk¯Tsld¯m\displaystyle=g^{\bar{k}l}\nabla_{l}\alpha_{c_{1}\cdots c_{p}\bar{d}_{1}\cdots\bar{d}_{q}\bar{k}}+\frac{1}{2}g^{\bar{k}l}g^{\bar{t}s}\sum_{m=1}^{q}\alpha_{c_{1}\cdots c_{p}\bar{d}_{1}\cdots\bar{t}\cdots\bar{d}_{q}\bar{k}}T_{sl\bar{d}_{m}}
+gk¯lαc1cpd¯1d¯qk¯Tlss\displaystyle\qquad+g^{\bar{k}l}\alpha_{c_{1}\cdots c_{p}\bar{d}_{1}\cdots\bar{d}_{q}\bar{k}}T_{ls}^{s}

References

  • [1] Bogdan Alexandrov and Stefan Ivanov. Vanishing theorems on Hermitian manifolds. Differential Geom. Appl., 14(3):251–265, 2001.
  • [2] Wolf P. Barth, Klaus Hulek, Chris A. M. Peters, and Antonius Van de Ven. Compact complex surfaces, volume 4 of Ergebnisse der Mathematik und ihrer Grenzgebiete. 3. Folge. A Series of Modern Surveys in Mathematics [Results in Mathematics and Related Areas. 3rd Series. A Series of Modern Surveys in Mathematics]. 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] Shiing-shen Chern. Characteristic classes of Hermitian manifolds. Ann. of Math. (2), 47:85–121, 1946.
  • [5] Mario Garcia-Fernandez, Joshua Jordan, and Jeffrey Streets. Non-Kähler Calabi-Yau geometry and pluriclosed flow. arXiv e-prints, page arXiv:2106.13716, June 2021.
  • [6] Paul Gauduchon. Le théorème de l’excentricité nulle. C. R. Acad. Sci. Paris Sér. A-B, 285(5):A387–A390, 1977.
  • [7] S. Ivanov and G. Papadopoulos. Vanishing theorems and string backgrounds. Classical Quantum Gravity, 18(6):1089–1110, 2001.
  • [8] John Morgan and Gang Tian. The geometrization conjecture, volume 5 of Clay Mathematics Monographs. American Mathematical Society, Providence, RI; Clay Mathematics Institute, Cambridge, MA, 2014.
  • [9] Grisha Perelman. The entropy formula for the Ricci flow and its geometric applications. arXiv Mathematics e-prints, page math/0211159, November 2002.
  • [10] Jeffrey Streets. Classification of solitons for pluriclosed flow on complex surfaces. Math. Ann., 375(3-4):1555–1595, 2019.
  • [11] 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.
  • [12] Jeffrey Streets and Gang Tian. A parabolic flow of pluriclosed metrics. Int. Math. Res. Not. IMRN, (16):3101–3133, 2010.
  • [13] Jeffrey Streets and Gang Tian. Hermitian curvature flow. J. Eur. Math. Soc. (JEMS), 13(3):601–634, 2011.
  • [14] Jeffrey Streets and Gang Tian. Regularity theory for pluriclosed flow. C. R. Math. Acad. Sci. Paris, 349(1-2):1–4, 2011.
  • [15] Gang Tian and Xiaohua Zhu. Uniqueness of Kähler-Ricci solitons. Acta Math., 184(2):271–305, 2000.
  • [16] Raymond O. Wells, Jr. Differential analysis on complex manifolds, volume 65 of Graduate Texts in Mathematics. Springer, New York, third edition, 2008. With a new appendix by Oscar Garcia-Prada.
  • [17] Yanan Ye. Pluriclosed flow and Hermitian-symplectic structures. arXiv e-prints, page arXiv:2207.12643, July 2022.