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

A Chern-Calabi flow on Hermitian manifolds

Xi Sisi Shen Department of Mathematics
Northwestern University
Evanston, IL USA 60208
[email protected]
Abstract.

We study an analogue of the Calabi flow in the non-Kähler setting for compact Hermitian manifolds with vanishing first Bott-Chern class. We prove a priori estimates for the evolving metric along the flow given a uniform bound on the Chern scalar curvature. If the Chern scalar curvature remains uniformly bounded for all time, we show that the flow converges smoothly to the unique Chern-Ricci-flat metric in the ¯\partial\bar{\partial}-class of the initial metric.

1. Introduction

The Calabi flow was introduced by Calabi for Kähler metrics in [6, 7] and is defined by

ωt=1¯R,ω(0)=ω0,\displaystyle\frac{\partial\omega}{\partial t}=\sqrt{-1}\partial\bar{\partial}\operatorname{R},\ \ \omega(0)=\omega_{0},

where R=trωRic(ω)=trω1¯logdetg\operatorname{R}=\operatorname{tr}_{\omega}\operatorname{Ric}(\omega)=\operatorname{tr}_{\omega}\sqrt{-1}\partial\bar{\partial}\log\det g is the scalar curvature of a Kähler metric gg, with associated Kähler form ω\omega, and the flow preserves the Kähler class of the metric. If we let ωφ(t)=ω0+1¯φ(t)\omega_{\varphi}(t)=\omega_{0}+\sqrt{-1}\partial\bar{\partial}\varphi(t) and normalize φC(X)\varphi\in C^{\infty}(X) such that Xφωφn=0\int_{X}\varphi\omega_{\varphi}^{n}=0, then the flow can be represented in terms of the potential function φ\varphi by

φt=RφR¯,\displaystyle\frac{\partial\varphi}{\partial t}=\operatorname{R}_{\varphi}-\underline{\operatorname{R}},

where Rφ\operatorname{R}_{\varphi} is the scalar curvature of ωφ\omega_{\varphi} and R¯=XRφωφnXωφn\underline{\operatorname{R}}=\frac{\int_{X}\operatorname{R}_{\varphi}\omega_{\varphi}^{n}}{\int_{X}\omega_{\varphi}^{n}} is the average scalar curvature of ωφ\omega_{\varphi} on XX which is independent of tt. Short-time existence of the Calabi flow follows from the fact that it is a fourth order quasilinear parabolic equation as shown by Chen-He [9]. They also proved the global existence of the flow under the assumption of a uniform Ricci curvature bound [9]. It was shown by Székelyhidi in [28] that if the curvature tensor is uniformly bounded along the Calabi flow and the Mabuchi energy is proper, then the flow converges to a constant scalar curvature Kähler (cscK) metric. Chen-Sun proved in [12] that if the Calabi flow starts at an initial metric that is sufficiently close to a cscK metric, then the flow exists and converges uniformly to the cscK metric. Additional literature on the Calabi flow can be found in [3, 4, 8, 13, 14, 15, 18, 19, 21, 25, 26, 31].

We define an analogue of the Calabi flow on the ¯\partial\bar{\partial}-class of Hermitian metrics when the first Bott-Chern class vanishes. Let (X,ω0)(X,\omega_{0}) be a Hermitian manifold and let g0g_{0} be the associated Hermitian metric to the real (1,1)(1,1)-form ω0\omega_{0}. The real (1,1)(1,1) Bott-Chern cohomology is defined as

HBC1,1(X,)={d-closed real (1,1)-forms}{1¯ψ,ψC(X,)}\displaystyle H^{1,1}_{\operatorname{BC}}(X,\mathbb{R})=\frac{\{\text{$d$-closed real (1,1)-forms}\}}{\{\sqrt{-1}\partial\bar{\partial}\psi,\ \psi\in C^{\infty}(X,\mathbb{R})\}}

and the first Bott-Chern class, denoted c1BC(X)c_{1}^{\operatorname{BC}}(X), is given by the ¯\partial\bar{\partial}-class of the Chern-Ricci form,

Ric(ω)=1¯logdetg,\operatorname{Ric}(\omega)=-\sqrt{-1}\partial\bar{\partial}\log\det g,

for any metric ω\omega on XX. Now, let us define the space of metrics

={ωφ:ωφ=ω0+1¯φ>0,φC(X,)}.\mathcal{H}=\{\omega_{\varphi}:\omega_{\varphi}=\omega_{0}+\sqrt{-1}\partial\bar{\partial}\varphi>0,\varphi\in C^{\infty}(X,\mathbb{R})\}.

In the setting of vanishing first Bott-Chern class, c1BC(X)=0c_{1}^{\operatorname{BC}}(X)=0, Tosatti-Weinkove [33] observed that one can define the Mabuchi energy Mabω0(ωφ):\operatorname{Mab}_{\omega_{0}}(\omega_{\varphi}):\mathcal{H}\rightarrow\mathbb{R} by

Mabω0(ωφ)=1VX(logdetgφdetg0F)ωφn+1VXFω0n,\displaystyle\operatorname{Mab}_{\omega_{0}}(\omega_{\varphi})=\frac{1}{V}\int_{X}\Big{(}\log\frac{\det g_{\varphi}}{\det g_{0}}-F\Big{)}\omega_{\varphi}^{n}+\frac{1}{V}\int_{X}F\omega_{0}^{n},

where FF is the Chern-Ricci potential of ω0\omega_{0}, that is Ric(ω0)=1¯F\operatorname{Ric}(\omega_{0})=\sqrt{-1}\partial\bar{\partial}F, normalized so that XeFω0n=Xω0n\int_{X}e^{F}\omega_{0}^{n}=\int_{X}\omega_{0}^{n}. This definition of Mabuchi energy agrees with the formula in the Kähler setting [29] (see also Section 9 of [33]). Similar to how there are several generalizations of the Kähler-Ricci flow to the non-Kähler setting [17, 27, 34, 35], the Chern-Calabi flow we consider in this paper may not be the only generalization of the Calabi flow evolving the ¯\partial\bar{\partial}-potential of the metric by its Chern scalar curvature. In [3], Bedulli-Vezzoni prove short-time existence of a Calabi-type flow evolving the potential function by its Chern scalar curvature within the ¯\partial\bar{\partial}-class of ωn1\omega^{n-1}, see also [2, 20]. Other flows of Hermitian metrics in the non-Kähler setting have been studied in [1, 5, 22].

Assume that our Hermitian metric ω0\omega_{0} satisfies ¯ω0k=0\partial\bar{\partial}\omega_{0}^{k}=0 for k=1,2k=1,2 and let ωφ=ω0+1¯φ\omega_{\varphi}=\omega_{0}+\sqrt{-1}\partial\bar{\partial}\varphi for a smooth function φ\varphi normalized so that Xφωφn=0\int_{X}\varphi\omega_{\varphi}^{n}=0. We consider a gradient flow of the Mabuchi energy defined above, starting at ω0\omega_{0}. This flow can be expressed in terms of the potential φ\varphi by

(1) φt=Rφ+2RetrTφ,(logdetgφdetg0F)φ,φ(0)=0,\displaystyle\frac{\partial\varphi}{\partial t}=\operatorname{R}_{\varphi}+2\operatorname{Re}\langle\operatorname{tr}T_{\varphi},\partial(\log\frac{\det g_{\varphi}}{\det g_{0}}-F)\rangle_{\varphi},\ \ \varphi(0)=0,

where Rφ=trωφRic(ωφ)\operatorname{R}_{\varphi}=\operatorname{tr}_{\omega_{\varphi}}\operatorname{Ric}(\omega_{\varphi}) is the Chern scalar curvature of ωφ\omega_{\varphi} and trTφ=(Tφ)pp\operatorname{tr}T_{\varphi}=(T_{\varphi})^{p}_{p\bullet} is the trace of the torsion of ωφ\omega_{\varphi}. We note that R¯=0\underline{\operatorname{R}}=0 since c1BC(X)=0c_{1}^{\operatorname{BC}}(X)=0. When the metric is Kähler, this flow agrees with the Calabi flow. We note that we immediately obtain short-time existence of this flow since its leading order term is a strictly elliptic fourth-order operator and so the equation is a fourth order quasilinear parabolic equation, following the same line of reasoning as in [9]. Given that the Mabuchi energy is decreasing along this flow, we prove a new a priori estimate on the evolving metric, building on work by Chen-Cheng [11] and the author [24], see Theorem 2. In this paper, we prove

Theorem 1.

Let (X,ω0)(X,\omega_{0}) be a Hermitian manifold with c1BC(X)=0c_{1}^{\operatorname{BC}}(X)=0 and ¯ω0k=0\partial\bar{\partial}\omega_{0}^{k}=0 for k=1,2k=1,2. A solution to the flow given by Equation (1) starting at ω0\omega_{0} exists as long as the Chern scalar curvature remains bounded along the flow. In addition, if the Chern sclaar curvature remains bounded for all time, then we have smooth convergence of the flow to the unique Chern-Ricci-flat metric of the form ω=ω0+1¯φ\omega_{\infty}=\omega_{0}+\sqrt{-1}\partial\bar{\partial}\varphi_{\infty} for some smooth function φ\varphi_{\infty} on XX.

Our assumption that ω0\omega_{0} satisfy ¯ω0k=0\partial\bar{\partial}\omega_{0}^{k}=0 for k=1,2k=1,2, is in fact equivalent to ¯ω0k=0\partial\bar{\partial}\omega_{0}^{k}=0 for k=1,2,,n1k=1,2,\ldots,n-1 and this condition is preserved by the flow. This assumption allows us to ensure that the volume V=XωφnV=\int_{X}\omega_{\varphi}^{n} remains unchanged along the flow and to obtain C3,αC^{3,\alpha} estimates for φ\varphi along the flow dependent on a Chern scalar curvature bound. In order to show long-time existence, we will need to assume that the Chern scalar curvature remains bounded for all time along with a smoothing property that allows to obtain all higher order estimates on φ\varphi, following the work of Chen-He [9] for the Calabi flow.

This paper is structured as follows:

  • In Section 2, we cover the notation and basic properties of Hermitian metrics that we will need in the subsequent sections.

  • In Section 3, we discuss properties of the flow. We first show that the flow is indeed a gradient flow of the Mabuchi energy. We then prove that its fixed points are precisely the constant Chern scalar curvature metrics, which when c1BC(X)=0c_{1}^{\operatorname{BC}}(X)=0, are in fact Chern-Ricci-flat metrics. Lastly, we discuss short-time existence of the flow.

  • In Section 4, we prove a priori estimates along the flow following the methods of Chen-Cheng in [11] and previous work by the author in [24]. Finally, we prove Theorem 1.

2. Preliminaries

In this section, we include several well-known identities that will be needed for computations in the next sections (see also Section 2 of [34]).

Let XX be a compact complex manifold of complex dimension nn. We will work in complex coordinates z1,,znz^{1},\ldots,z^{n} and write tensors in terms of this coordinate system. Let g=gij¯g=g_{i\bar{j}} be a Hermitian metric on XX with associated (1,1)(1,1)-form ω=1gij¯dzidzj¯\omega=\sqrt{-1}g_{i\bar{j}}dz^{i}\wedge d\overline{z^{j}} where all repeated indices are being summed from 11 to nn. We will also refer to ω\omega as a Hermitian metric.

Let \nabla be the Chern connection associated to ω\omega, defined for a (1,0)(1,0)-form a=akdzka=a_{k}dz^{k} by

(2) iak=iakΓikjaj,iak¯=iak¯\displaystyle\begin{split}&\nabla_{i}a_{k}=\partial_{i}a_{k}-\Gamma_{ik}^{j}a_{j}\ ,\ \ \ \nabla_{i}\overline{a_{k}}=\partial_{i}\overline{a_{k}}\end{split}

and for a vector field X=XkkX=X^{k}\partial_{k} by

iXk=iXk+ΓijkXj,iXk¯=iXk¯\displaystyle\nabla_{i}X^{k}=\partial_{i}X^{k}+\Gamma^{k}_{ij}X^{j}\ ,\ \ \ \nabla_{i}\overline{X^{k}}=\partial_{i}\overline{X^{k}}

where Γijk=gkp¯igjp¯\Gamma^{k}_{ij}=g^{k\bar{p}}\partial_{i}g_{j\bar{p}} is the Christoffel symbol of gg and gkp¯gip¯=δikg^{k\bar{p}}g_{i\bar{p}}=\delta_{ik}. The Chern connection is compatible with the metric gg, that is, kgij¯=0\nabla_{k}g_{i\bar{j}}=0 i,j,k\forall i,j,k. We define the trace of a real (1,1)(1,1)-form α=αij¯dzidzj¯\alpha=\alpha_{i\bar{j}}dz^{i}\wedge d\overline{z^{j}} with respect to ω\omega by

trωα=gij¯αij¯=nωn1αωn.\operatorname{tr}_{\omega}\alpha=g^{i\bar{j}}\alpha_{i\bar{j}}=\tfrac{n\omega^{n-1}\wedge\alpha}{\omega^{n}}.

The torsion of gg is defined by Tijk=ΓijkΓjik.T^{k}_{ij}=\Gamma^{k}_{ij}-\Gamma^{k}_{ji}. Let ωφ=ω+1¯φ\omega_{\varphi}=\omega+\sqrt{-1}\partial\bar{\partial}\varphi be another Hermitian metric on XX. From this definition, it is clear that

(ω)jk¯=(ωφ)jk¯(\partial\omega)_{jk\bar{\ell}}=(\partial\omega_{\varphi})_{jk\bar{\ell}}

where (ω)jk¯=jgk¯kgj¯(\partial\omega)_{jk\bar{\ell}}=\partial_{j}g_{k\bar{\ell}}-\partial_{k}g_{j\bar{\ell}}. Denoting the torsion of ωφ\omega_{\varphi} by T~\tilde{T}, it follows that

(3) Tjkpgp¯=(ω)jk¯=(ωφ)jk¯=T~jkqg~q¯,\displaystyle\begin{split}T^{p}_{jk}g_{p\bar{\ell}}=(\partial\omega)_{jk\bar{\ell}}&=(\partial\omega_{\varphi})_{jk\bar{\ell}}=\tilde{T}^{q}_{jk}\tilde{g}_{q\bar{\ell}},\end{split}

where g~ij¯\tilde{g}_{i\bar{j}} is the metric in coordinates for ωφ\omega_{\varphi}. For simplicity, we will use the notation

Tjk¯=Tjkqgq¯,T~jk¯=T~jkpg~p¯T_{jk\bar{\ell}}=T^{q}_{jk}g_{q\bar{\ell}}\ ,\ \ \tilde{T}_{jk\bar{\ell}}=\tilde{T}^{p}_{jk}\tilde{g}_{p\bar{\ell}}

and so the first equality in Equation (3) can be rewritten as Tjk¯=T~jk¯T_{jk\bar{\ell}}=\tilde{T}_{jk\bar{\ell}}. In addition, we will let (trT)j(\operatorname{tr}T)_{j} denote TpjpT^{p}_{pj}. The curvature tensor is defined by

Rij¯kp=j¯Γikp,Rij¯k¯=gp¯Rij¯kp\displaystyle R_{i\bar{j}k}^{\;\;\;\;\;p}=-\partial_{\bar{j}}\Gamma^{p}_{ik}\ ,\ \ \ R_{i\bar{j}k\bar{\ell}}=g_{p\bar{\ell}}R_{i\bar{j}k}^{\;\;\;\;\;p}

where we note that Rij¯k¯¯=Rji¯k¯\overline{R_{i\bar{j}k\bar{\ell}}}=R_{j\bar{i}\ell\bar{k}}. We can commute the indices of the curvature tensor as follows:

(4) Rij¯kpRkj¯ip=j¯Γkipj¯Γikp=j¯Tkip.\displaystyle\begin{split}R_{i\bar{j}k}^{\;\;\;\;\;p}-R_{k\bar{j}i}^{\;\;\;\;\;p}=\partial_{\bar{j}}\Gamma^{p}_{ki}-\partial_{\bar{j}}\Gamma^{p}_{ik}=\partial_{\bar{j}}T^{p}_{ki}.\end{split}

The Chern-Ricci curvature of ω\omega is defined by

Rij¯=gk¯Rij¯k¯=ij¯logdetg,R_{i\bar{j}}=g^{k\bar{\ell}}R_{i\bar{j}k\bar{\ell}}=-\partial_{i}\partial_{\bar{j}}\log\det g,

its associated form by

Ric(ω)=1Rij¯dzidzj¯\operatorname{Ric}(\omega)=\sqrt{-1}R_{i\bar{j}}dz^{i}\wedge d\overline{z^{j}}

and its Chern scalar curvature by

R(ω)=gij¯Rij¯=trωRic(ω).\operatorname{R}(\omega)=g^{i\bar{j}}R_{i\bar{j}}=\operatorname{tr}_{\omega}\operatorname{Ric}(\omega).

The following commutation formulae will be useful to us in the later sections. For a (1,0)(1,0)-form a=akdzka=a_{k}dz^{k},

(5) [i,j¯]ak=Rij¯ka,[i,j¯]al¯=Rij¯¯k¯ak¯[i,j]ak¯=Tijrrak¯,[i¯,j¯]ak=Tijr¯r¯ak\displaystyle\begin{split}[\nabla_{i},\nabla_{\bar{j}}]a_{k}&=-R_{i\bar{j}k\;}^{\;\;\;\;\;\ell}a_{\ell},\ \ \ \ \ \ [\nabla_{i},\nabla_{\bar{j}}]\overline{a_{l}}=R_{i\bar{j}\;\;\bar{\ell}}^{\;\;\;\bar{k}}\overline{a_{k}}\\ [\nabla_{i},\nabla_{j}]\overline{a_{k}}&=-T^{r}_{ij}\nabla_{r}\overline{a_{k}},\ \ \ \ \ [\nabla_{\bar{i}},\nabla_{\bar{j}}]a_{k}=-\overline{T^{r}_{ij}}\nabla_{\bar{r}}a_{k}\end{split}

and for a scalar function ff, we have

(6) [i,j]f=Tijrrf.\displaystyle\begin{split}[\nabla_{i},\nabla_{j}]f&=-T_{ij}^{r}\nabla_{r}f.\end{split}

The Chern Laplacian with respect to gg of a function ff is defined by

Δf=trω1¯f=gij¯ij¯f=gij¯ij¯f.\displaystyle\Delta f=\operatorname{tr}_{\omega}\sqrt{-1}\partial\bar{\partial}f=g^{i\bar{j}}\partial_{i}\partial_{\bar{j}}f=g^{i\bar{j}}\nabla_{i}\nabla_{\bar{j}}f.

For a complex manifold, if we assume that ¯ωk=0 for k=1,2,\partial\bar{\partial}\omega^{k}=0\ \text{ for }\ k=1,2, then in fact ¯ω\partial\bar{\partial}\omega vanishes for all k=1,,n1k=1,\ldots,n-1, which follows from a straightforward computation. Under this assumption, the volume of the metric remains unchanged up to addition of ¯\partial\bar{\partial} of a smooth function, that is,

(7) X(ω+1¯ψ)n=Xωn,\displaystyle\int_{X}(\omega+\sqrt{-1}\partial\bar{\partial}\psi)^{n}=\int_{X}\omega^{n},

for all ψ\psi such that ω+1¯ψ>0\omega+\sqrt{-1}\partial\bar{\partial}\psi>0, hence is preserved by the flow. The condition that ¯ωφ=0\partial\bar{\partial}\omega_{\varphi}=0 is also preserved by the flow as

ddt¯ωφ=¯(1¯(Rφ+2RetrTφ,(logdetgφdetg0F)φ))=0,\displaystyle\frac{d}{dt}\partial\bar{\partial}\omega_{\varphi}=\partial\bar{\partial}\big{(}\sqrt{-1}\partial\bar{\partial}(\operatorname{R}_{\varphi}+2\operatorname{Re}\langle\operatorname{tr}T_{\varphi},\partial(\log\frac{\det g_{\varphi}}{\det g_{0}}-F)\rangle_{\varphi})\big{)}=0,

and similarly ¯ωφ2=0\partial\bar{\partial}\omega_{\varphi}^{2}=0 is preserved as

ddt¯ωφ2=2¯ωφddt¯ωφ=0,\displaystyle\frac{d}{dt}\partial\bar{\partial}\omega_{\varphi}^{2}=2\partial\bar{\partial}\omega_{\varphi}\wedge\frac{d}{dt}\partial\bar{\partial}\omega_{\varphi}=0,

which together gives us that ¯ωφk=0\partial\bar{\partial}\omega_{\varphi}^{k}=0 along the flow if ¯ω0k=0\partial\bar{\partial}\omega_{0}^{k}=0.

The Gauduchon condition, ¯ωn1=0\partial\bar{\partial}\omega^{n-1}=0, ensures the vanishing of the integrals of Chern Laplacians of functions:

XΔfωn=nX1¯fωn1=nXf1¯ωn1=0.\displaystyle\int_{X}\Delta f\omega^{n}=n\int_{X}\sqrt{-1}\partial\bar{\partial}f\wedge\omega^{n-1}=n\int_{X}f\sqrt{-1}\partial\bar{\partial}\omega^{n-1}=0.

We will need the following divergence theorem in the non-Kähler setting (see Lemma 1 of [23]):

Lemma 1.

For any Hermitian metric ω\omega and VΓ(X,T1,0,X)V\in\Gamma(X,T^{1,0},X), we have that

XiViωn=X(trT)iViωn,\displaystyle\int_{X}\nabla_{i}V^{i}\omega^{n}=\int_{X}(\operatorname{tr}T)_{i}V^{i}\omega^{n},

where \nabla is the Chern connection with respect to ω\omega and (trT)i=Tpip(\operatorname{tr}T)_{i}=T^{p}_{pi} is the trace of the torsion of ω\omega.

Using this divergence theorem, we can show the following:

Lemma 2.

For any metric gg with associated (1,1)(1,1)-form ω\omega, under the assumption that ω\omega is Gauduchon, i.e. ¯ωn1=0\partial\bar{\partial}\omega^{n-1}=0, we have that

gjk¯(k¯(trT)j(trT)k¯(trT)j)=0,\displaystyle g^{j\bar{k}}\big{(}\nabla_{\bar{k}}(\operatorname{tr}T)_{j}-\overline{(\operatorname{tr}T)_{k}}(\operatorname{tr}T)_{j}\big{)}=0,

where \nabla is the Chern connection with respect to ω\omega and TT is the torsion of the Chern connection with respect to ω\omega.

Proof.

This identity follows directly from the Gauduchon condition. One way to see this simply is that for any smooth function uu on XX, it follows from the Gauduchon condition and Lemma 1 that

0\displaystyle 0 =XΔu2ωn=2XuΔuωn+2Xgjk¯juk¯uωn\displaystyle=\int_{X}\Delta u^{2}\omega^{n}=2\int_{X}u\Delta u\omega^{n}+2\int_{X}g^{j\bar{k}}\nabla_{j}u\nabla_{\bar{k}}u\omega^{n}
=2Xgjk¯juk¯uωn+2Xgjk¯(trT)juk¯uωn+2Xgjk¯juk¯uωn\displaystyle=-2\int_{X}g^{j\bar{k}}\nabla_{j}u\nabla_{\bar{k}}u\omega^{n}+2\int_{X}g^{j\bar{k}}(\operatorname{tr}T)_{j}u\nabla_{\bar{k}}u\omega^{n}+2\int_{X}g^{j\bar{k}}\nabla_{j}u\nabla_{\bar{k}}u\omega^{n}
=2Xgjk¯(trT)juk¯uωn\displaystyle=2\int_{X}g^{j\bar{k}}(\operatorname{tr}T)_{j}u\nabla_{\bar{k}}u\omega^{n}
=Xgjk¯(trT)jk¯u2ωn\displaystyle=\int_{X}g^{j\bar{k}}(\operatorname{tr}T)_{j}\nabla_{\bar{k}}u^{2}\omega^{n}
=X(gjk¯(k¯(trT)j(trT)k¯(trT)j)u2ωn.\displaystyle=-\int_{X}(g^{j\bar{k}}(\nabla_{\bar{k}}(\operatorname{tr}T)_{j}-\overline{(\operatorname{tr}T)_{k}}(\operatorname{tr}T)_{j})u^{2}\omega^{n}.

Since this holds for arbitrary uu, it follows that gjk¯(k¯(trT)j(trT)k¯(trT)j)=0g^{j\bar{k}}(\nabla_{\bar{k}}(\operatorname{tr}T)_{j}-\overline{(\operatorname{tr}T)_{k}}(\operatorname{tr}T)_{j})=0. We note that if XX were not compact, this identity would still hold since we could take the function uu to be compactly supported. ∎

This outlines the key identities and formulae that we will need for the computations in this paper. Note that throughout this paper, the constants may vary from one line to another.

3. Properties of the flow

In this section, we prove that the flow defined in Equation (1) is indeed a gradient flow of the Mabuchi energy, that fixed points of the flow are exactly those metrics that are Chern-Ricci-flat, and that we have short-time existence of the flow.

Let ω0\omega_{0} be Hermitian on XX satisfying ¯ωk=0\partial\bar{\partial}\omega^{k}=0 for k=1,2k=1,2. We will assume that ω0\omega_{0} is normalized such that V=Xω0n=1V=\int_{X}\omega_{0}^{n}=1 and we note that this integral remains unchanged along the flow, see Equation (7). Using the assumption that c1BC(X)=0c_{1}^{\operatorname{BC}}(X)=0, we have that

Ric(ω0)=1¯F,\operatorname{Ric}(\omega_{0})=\sqrt{-1}\partial\bar{\partial}F,

for a smooth function FF which we call the Chern-Ricci potential, normalized so that XeFω0n=Xω0n\int_{X}e^{F}\omega_{0}^{n}=\int_{X}\omega_{0}^{n}. Let us define

={ωφ:ωφ=ω0+1¯φ>0,φC(X)}.\mathcal{H}=\{\omega_{\varphi}:\omega_{\varphi}=\omega_{0}+\sqrt{-1}\partial\bar{\partial}\varphi>0,\varphi\in C^{\infty}(X)\}.

We will show that the flow defined in Equation (1) is a gradient flow of the Mabuchi energy Mabω0:\operatorname{Mab}_{\omega_{0}}:\mathcal{H}\rightarrow\mathbb{R} defined by

Mabω0(ωφ)=X(logωφnω0nF)ωφn+XFω0n.\displaystyle\operatorname{Mab}_{\omega_{0}}(\omega_{\varphi})=\int_{X}\Big{(}\log\frac{\omega_{\varphi}^{n}}{\omega_{0}^{n}}-F\Big{)}\omega_{\varphi}^{n}+\int_{X}F\omega_{0}^{n}.

One can check that this definition of Mabuchi energy agrees with the formula in the Kähler setting [29] (see also Section 9 of [33]).

To set some notation, let Ω:=eFω0n\Omega:=e^{F}\omega_{0}^{n} be a volume form on XX which satisfies

1¯logΩ=0\sqrt{-1}\partial\bar{\partial}\log\Omega=0

by the fact that FF is the Chern-Ricci potential for ω0\omega_{0}. In this way, we see that the Chern scalar curvature of ωφ\omega_{\varphi} can be expressed as

Rφ=ΔφlogωφnΩ.\displaystyle\operatorname{R}_{\varphi}=-\Delta_{\varphi}\log\frac{\omega_{\varphi}^{n}}{\Omega}.
Lemma 3.

The Mabuchi energy is decreasing along the flow defined in Equation (1).

Proof.

Taking the time derivative of Mabuchi energy, we have that

tMabω0(ωφ)\displaystyle\frac{\partial}{\partial t}\operatorname{Mab}_{\omega_{0}}(\omega_{\varphi}) =XΔφφ˙ωφn+X(logωφnω0nF)Δφφ˙ωφn\displaystyle=\int_{X}\Delta_{\varphi}\dot{\varphi}\omega_{\varphi}^{n}+\int_{X}\Big{(}\log\frac{\omega_{\varphi}^{n}}{\omega_{0}^{n}}-F\Big{)}\Delta_{\varphi}\dot{\varphi}\omega_{\varphi}^{n}
=XlogωφneFω0nΔφφ˙ωφn\displaystyle=\int_{X}\log\frac{\omega_{\varphi}^{n}}{e^{F}\omega_{0}^{n}}\Delta_{\varphi}\dot{\varphi}\omega_{\varphi}^{n}

Let ~\tilde{\nabla} be the Chern connection with respect to ωφ\omega_{\varphi}. Substituting Ω=eFω0n\Omega=e^{F}\omega_{0}^{n} and integrating by parts using Lemma 1, we have that

tMabω0(ωφ)\displaystyle\frac{\partial}{\partial t}\operatorname{Mab}_{\omega_{0}}(\omega_{\varphi}) =XlogωφnΩΔφφ˙ωφn\displaystyle=\int_{X}\log\frac{\omega_{\varphi}^{n}}{\Omega}\Delta_{\varphi}\dot{\varphi}\omega_{\varphi}^{n}
=X(gφ)jk¯~jlogωφnΩ~k¯φ˙ωφn+X(gφ)jk¯(trTφ)jlogωφnΩ~k¯φ˙ωφn\displaystyle=-\int_{X}(g_{\varphi})^{j\bar{k}}\tilde{\nabla}_{j}\log\frac{\omega_{\varphi}^{n}}{\Omega}\tilde{\nabla}_{\bar{k}}\dot{\varphi}\omega_{\varphi}^{n}+\int_{X}(g_{\varphi})^{j\bar{k}}(\operatorname{tr}T_{\varphi})_{j}\log\frac{\omega_{\varphi}^{n}}{\Omega}\tilde{\nabla}_{\bar{k}}\dot{\varphi}\omega_{\varphi}^{n}
=X(gφ)jk¯~j~k¯logωφnΩφ˙ωφnX(gφ)jk¯(trTφ)k¯~jlogωφnΩφ˙ωφn\displaystyle=\int_{X}(g_{\varphi})^{j\bar{k}}\tilde{\nabla}_{j}\tilde{\nabla}_{\bar{k}}\log\frac{\omega_{\varphi}^{n}}{\Omega}\dot{\varphi}\omega_{\varphi}^{n}-\int_{X}(g_{\varphi})^{j\bar{k}}\overline{(\operatorname{tr}T_{\varphi})_{k}}\tilde{\nabla}_{j}\log\frac{\omega_{\varphi}^{n}}{\Omega}\dot{\varphi}\omega_{\varphi}^{n}
X(gφ)jk¯~k¯((trTφ)jlogωφnΩ)φ˙ωφn+X(gφ)jk¯(trTφ)k¯(trTφ)jlogωφnΩφ˙ωφn\displaystyle\ \ \ \ -\int_{X}(g_{\varphi})^{j\bar{k}}\tilde{\nabla}_{\bar{k}}((\operatorname{tr}T_{\varphi})_{j}\log\frac{\omega_{\varphi}^{n}}{\Omega})\dot{\varphi}\omega_{\varphi}^{n}+\int_{X}(g_{\varphi})^{j\bar{k}}\overline{(\operatorname{tr}T_{\varphi})_{k}}(\operatorname{tr}T_{\varphi})_{j}\log\frac{\omega_{\varphi}^{n}}{\Omega}\dot{\varphi}\omega_{\varphi}^{n}
=Xφ˙(Rφ+2Re((gφ)jk¯(trTφ)j~k¯logωφnΩ))ωφn,\displaystyle=-\int_{X}\dot{\varphi}\Big{(}{\operatorname{R}_{\varphi}}+2\operatorname{Re}((g_{\varphi})^{j\bar{k}}(\operatorname{tr}T_{\varphi})_{j}\tilde{\nabla}_{\bar{k}}\log\frac{\omega_{\varphi}^{n}}{\Omega})\Big{)}\omega_{\varphi}^{n},

where the second last step follows from the fact that 1¯logΩ=0\sqrt{-1}\partial\bar{\partial}\log\Omega=0 and the last step uses Lemma 2. Since this flow is defined by the following evolution equation

φ˙\displaystyle\dot{\varphi} =Rφ+2Re((gφ)jk¯(trTφ)jk¯logωφnΩ),\displaystyle={\operatorname{R}_{\varphi}}+2\operatorname{Re}((g_{\varphi})^{j\bar{k}}(\operatorname{tr}T_{\varphi})_{j}\partial_{\bar{k}}\log\frac{\omega_{\varphi}^{n}}{\Omega}),

we immediately have that the Mabuchi energy is decreasing along this flow. ∎

We now check that all fixed points of this flow are of constant Chern scalar curvature which, in the setting that c1BC(X)=0c_{1}^{\operatorname{BC}}(X)=0, is equivalent to being Chern-Ricci-flat.

Lemma 4.

A metric is a fixed point of the flow defined by Equation (1) if and only if it is Chern-Ricci-flat.

Proof.

Assume that ωφ\omega_{\varphi} is Chern-Ricci-flat. Then we have that

0=Rφ=ΔφlogωφnΩ,0={\operatorname{R}_{\varphi}}=-{\Delta_{\varphi}}\log\frac{\omega_{\varphi}^{n}}{\Omega},

which implies that logωφnΩ=const.\log\frac{\omega_{\varphi}^{n}}{\Omega}=const. since we are working on a compact manifold and integration by parts gives

(8) 0=XΔu2ωn=2XuΔuωn+2Xgjk¯juk¯uωn=2X|u|g2ωnu=const.,\displaystyle 0=\int_{X}\Delta u^{2}\omega^{n}=2\int_{X}u\Delta u\omega^{n}+2\int_{X}g^{j\bar{k}}\nabla_{j}u\nabla_{\bar{k}}u\omega^{n}=2\int_{X}|\nabla u|^{2}_{g}\omega^{n}\Rightarrow u=const.,

where we used that ω\omega is Gauduchon in the first equality. Using that logωφnΩ=const\log\frac{\omega_{\varphi}^{n}}{\Omega}=const, we have that

φ˙\displaystyle\dot{\varphi} =Rφ+2Re((gφ)jk¯(trTφ)jk¯logωφnΩ)=0.\displaystyle={\operatorname{R}_{\varphi}}+2\operatorname{Re}((g_{\varphi})^{j\bar{k}}(\operatorname{tr}T_{\varphi})_{j}\partial_{\bar{k}}\log\frac{\omega_{\varphi}^{n}}{\Omega})=0.

Thus, we have showed that a Chern-Ricci-flat metric is a fixed point of this flow.

In order to show the reverse direction, let us now assume that ωφ\omega_{\varphi} is a fixed point of the flow, that is, φ˙=0\dot{\varphi}=0. This means that Rφ+2Re((gφ)jk¯(trTφ)jk¯logωφnΩ)=0{\operatorname{R}_{\varphi}}+2\operatorname{Re}((g_{\varphi})^{j\bar{k}}(\operatorname{tr}T_{\varphi})_{j}\partial_{\bar{k}}\log\frac{\omega_{\varphi}^{n}}{\Omega})=0. Let ~\tilde{\nabla} be the Chern connection with respect to ωφ\omega_{\varphi}. Integrating against an arbitary smooth function uu and using Lemma 1 and Lemma 2, we have that

0\displaystyle 0 =Xu(Rφ+2Re((gφ)jk¯(trTφ)j~k¯logωφnΩ))ωφn\displaystyle=\int_{X}u({\operatorname{R}_{\varphi}}+2\operatorname{Re}((g_{\varphi})^{j\bar{k}}(\operatorname{tr}T_{\varphi})_{j}\tilde{\nabla}_{\bar{k}}\log\frac{\omega_{\varphi}^{n}}{\Omega}))\omega_{\varphi}^{n}
=Xu(gφ)jk¯~j~k¯logωφnΩωφn+2XuRe((gφ)jk¯(trTφ)j~k¯logωφnΩ)ωφn\displaystyle=-\int_{X}u(g_{\varphi})^{j\bar{k}}\tilde{\nabla}_{j}\tilde{\nabla}_{\bar{k}}\log\frac{\omega_{\varphi}^{n}}{\Omega}\omega_{\varphi}^{n}+2\int_{X}u\operatorname{Re}((g_{\varphi})^{j\bar{k}}(\operatorname{tr}T_{\varphi})_{j}\tilde{\nabla}_{\bar{k}}\log\frac{\omega_{\varphi}^{n}}{\Omega})\omega_{\varphi}^{n}
=X~ju(gφ)jk¯~k¯logωφnΩωφn+Xu(gφ)jk¯(trTφ)k¯~jlogωφnΩωφn\displaystyle=\int_{X}\tilde{\nabla}_{j}u(g_{\varphi})^{j\bar{k}}\tilde{\nabla}_{\bar{k}}\log\frac{\omega_{\varphi}^{n}}{\Omega}\omega_{\varphi}^{n}+\int_{X}u(g_{\varphi})^{j\bar{k}}\overline{(\operatorname{tr}T_{\varphi})_{k}}\tilde{\nabla}_{j}\log\frac{\omega_{\varphi}^{n}}{\Omega}\omega_{\varphi}^{n}
=XΔφulogωφnΩωφn+X(gφ)jk¯~ju(trTφ)k¯logωφnΩωφn+Xu(gφ)jk¯(trTφ)k¯~jlogωφnΩωφn\displaystyle=-\int_{X}{\Delta_{\varphi}}u\log\frac{\omega_{\varphi}^{n}}{\Omega}\omega_{\varphi}^{n}+\int_{X}(g_{\varphi})^{j\bar{k}}\tilde{\nabla}_{j}u\overline{(\operatorname{tr}T_{\varphi})_{k}}\log\frac{\omega_{\varphi}^{n}}{\Omega}\omega_{\varphi}^{n}+\int_{X}u(g_{\varphi})^{j\bar{k}}\overline{(\operatorname{tr}T_{\varphi})_{k}}\tilde{\nabla}_{j}\log\frac{\omega_{\varphi}^{n}}{\Omega}\omega_{\varphi}^{n}
=XΔφulogωφnΩωφn.\displaystyle=-\int_{X}{\Delta_{\varphi}}u\log\frac{\omega_{\varphi}^{n}}{\Omega}\omega_{\varphi}^{n}.

Choosing u=logωφnΩu=\log\frac{\omega_{\varphi}^{n}}{\Omega}, we have by Equation (8) that

0=X|~(logωφnΩ)|φ2ωφn00=\int_{X}|\tilde{\nabla}(\log\frac{\omega_{\varphi}^{n}}{\Omega})|^{2}_{\varphi}\omega_{\varphi}^{n}\geq 0

and so logωφnΩ=const.\log\frac{\omega_{\varphi}^{n}}{\Omega}=const. which implies that Ric(ωφ)=1¯logωφnΩ=0\operatorname{Ric}(\omega_{\varphi})=-\sqrt{-1}\partial\bar{\partial}\log\frac{\omega_{\varphi}^{n}}{\Omega}=0 and so ωφ\omega_{\varphi} is Chern-Ricci-flat. ∎

We now discuss short-time existence of this flow, following from [9].

Lemma 5.

There exists a unique solution φ(t)\varphi(t) satisfying Equation (1) for t[0,T)t\in[0,T).

Proof.

The flow evolves the potential function φ\varphi by

φt\displaystyle\frac{\partial\varphi}{\partial t} =Rφ+2Re(trTφ),logdetgφeFdetg0φ\displaystyle=\operatorname{R}_{\varphi}+2\operatorname{Re}\langle(\operatorname{tr}T_{\varphi}),\partial\log\frac{\det g_{\varphi}}{e^{F}\det g_{0}}\rangle_{\varphi}
=A(φ,2φ)φ+f(φ,2φ,3φ).\displaystyle=-A(\nabla\varphi,\nabla^{2}\varphi)\varphi+f(\nabla\varphi,\nabla^{2}\varphi,\nabla^{3}\varphi).

We see that this flow differs from the Calabi flow only by addition of lower order terms in φ\varphi which can be bundled into the ff term, while AA remains a strictly elliptic operator with coefficients depending on first- and second-order derivatives of φ\varphi. Thus, the short-time existence of this flow follows the same way as for the Calabi flow, using standard parabolic theory, see [9] and references therein. ∎

4. Proof of the main theorem

Our goal in this section is to prove a priori estimates for φ\varphi along this flow, conditional on a bound on the Chern scalar curvature of the evolving metric. We then prove Theorem 1.

Let us begin with showing the a priori estimates:

Theorem 2.

Let (X,ω)(X,\omega) be a compact Hermitian manifold with c1BC(X)=0c_{1}^{\operatorname{BC}}(X)=0 and ω\omega satisfying ¯ωk=0\partial\bar{\partial}\omega^{k}=0 for k=1,2k=1,2. If ωφ=ω+1¯φ\omega_{\varphi}=\omega+\sqrt{-1}\partial\bar{\partial}\varphi for a smooth potential function φ\varphi on XX, then for any 0<α<10<\alpha<1, there exists CC depending only on (X,ω)(X,\omega), RφC0\|\operatorname{R}_{\varphi}\|_{C^{0}}, and an upper bound for Mabω(ωφ)\emph{Mab}_{\omega}(\omega_{\varphi}) such that φC3,α(X,ω)C.\|\varphi\|_{C^{3,\alpha}(X,\omega)}\leq C.

Proof.

The entropy quantity Ent(ω,ωφ)\operatorname{Ent}(\omega,\omega_{\varphi}) used in [24] is implied by a bound on the Mabuchi energy Mabω(ωφ)\operatorname{Mab}_{\omega}(\omega_{\varphi}) since

Ent(ω,ωφ):=Xlogωφnωnωφn\displaystyle\operatorname{Ent}(\omega,\omega_{\varphi}):=\int_{X}\log\frac{\omega_{\varphi}^{n}}{\omega^{n}}\omega_{\varphi}^{n} =Mabω(ωφ)XFωφn+XFωn\displaystyle=\operatorname{Mab}_{\omega}(\omega_{\varphi})-\int_{X}F\omega_{\varphi}^{n}+\int_{X}F\omega^{n}
Mabω(ωφ)+supX|F|Xωφn+C\displaystyle\leq\operatorname{Mab}_{\omega}(\omega_{\varphi})+\sup_{X}|F|\int_{X}\omega_{\varphi}^{n}+C
Mabω(ωφ)+C,\displaystyle\leq\operatorname{Mab}_{\omega}(\omega_{\varphi})+C,

where the inequalities follow from the fact that FF is the Chern-Ricci potential of the fixed metric ω\omega. In what follows, we will pass the dependencies of the constants on Ent(ω,ωφ)\operatorname{Ent}(\omega,\omega_{\varphi}) to Mabω(ωφ)\operatorname{Mab}_{\omega}(\omega_{\varphi}).

The proof follows in almost exactly the same manner as in Chen-Cheng [11] using what has already been proven in the non-Kähler setting in [24]. We include a proof here to show how we account for the torsion terms that result from handling the derivative terms of Rφ\operatorname{R}_{\varphi} in the LL^{\infty} estimate that are no longer assumed to vanish. We consider the coupled second order equations

(9) F=logωφnωnΔφF=Rφ+trωφRic(ω).\displaystyle\begin{split}&\ \ F=\log\tfrac{\omega_{\varphi}^{n}}{\omega^{n}}\\ {\Delta_{\varphi}}F&=-{\operatorname{R}_{\varphi}}+\operatorname{tr}_{\omega_{\varphi}}\operatorname{Ric}(\omega).\end{split}

For the C0C^{0}, C1C^{1} and LpL^{p} estimates, lifting the assumption that Rφ=const.\operatorname{R}_{\varphi}=const. and noting the fact that there is never an instance where we need to differentiate Rφ\operatorname{R}_{\varphi}, the estimates work out exactly the same way as in [24] where the new bounds will additionally depend on RφC0\|{\operatorname{R}_{\varphi}}\|_{C^{0}}. It is only in the proof of the LL^{\infty} estimate that we need to differentiate Rφ{\operatorname{R}_{\varphi}} and here is where the proof will somewhat differ from the constant Chern scalar curvature case (see Section 4 of [24]). Specifically, the term of the derivative of ΔφF\Delta_{\varphi}F arising from the computation of Δφ|F|ωφ2\Delta_{\varphi}|\partial F|^{2}_{\omega_{\varphi}} must be handled using integration by parts, and the torsion integral arising from this integration by parts will need to be controlled.

Lemma 6.

Let (φ,F)(\varphi,F) be a smooth solution to (LABEL:coupled_eqns). Then there exists a constant CC depending only on (X,ω),RφC0(X,\omega),\|{\operatorname{R}_{\varphi}}\|_{C^{0}} and Mabω(ωφ)\emph{Mab}_{\omega}(\omega_{\varphi}), such that

maxX(trωωφ)+maxX|F|ωφ2C.\displaystyle\max_{X}(\operatorname{tr}_{\omega}\omega_{\varphi})+\max_{X}|\partial F|^{2}_{\omega_{\varphi}}\leq C.
Proof.

For the purposes of simplifying notation, let ~\tilde{\nabla}, R~\tilde{R} and T~\tilde{T} denote, respectively, the Chern connection, curvature tensor and torsion with respect to g~\tilde{g}, the associated metric to ωφ\omega_{\varphi}. Commuting derivatives using the identities in (5) and (6), we have the following:

(10) Δφ(|F|ωφ2)=g~ij¯g~pq¯((~p~i~j¯F+T~pir~rFj¯)Fq¯+T~qjr¯~iFr¯Fp+~iT~qjr¯Fr¯Fp+(~q¯~i~j¯F+R~iq¯j¯¯F¯)Fp)+|~~F|2ωφ+|~~¯F|2ωφ=g~pq¯(ΔφF)pFq¯+2Re(g~ij¯g~pq¯T~pir~rFj¯Fq¯)+g~ij¯g~pq¯~iT~qjr¯Fr¯Fp+g~pq¯(ΔφF)q¯Fp+g~ij¯g~pq¯R~iq¯j¯¯F¯Fp+|~~F|ωφ2+|~~¯F|ωφ2=g~pq¯(ΔφF)pFq¯+2Re(g~ij¯g~pq¯g~rk¯T~pik¯~rFj¯Fq¯)+g~ij¯g~pq¯g~tr¯~i(T~qjt¯¯)Fr¯Fp+g~pq¯(ΔφF)q¯Fp+g~ij¯g~pq¯g~k¯R~iq¯kj¯F¯Fp+|~~F|ωφ2+|~~¯F|ωφ2\displaystyle\begin{split}{\Delta_{\varphi}}(|\partial F|^{2}_{\omega_{\varphi}})&=\tilde{g}^{i\bar{j}}\tilde{g}^{p\bar{q}}\big{(}(\tilde{\nabla}_{p}\tilde{\nabla}_{i}\tilde{\nabla}_{\bar{j}}F+\tilde{T}^{r}_{pi}\tilde{\nabla}_{r}F_{\bar{j}})F_{\bar{q}}+\overline{\tilde{T}^{r}_{qj}}\tilde{\nabla}_{i}F_{\bar{r}}F_{p}+\tilde{\nabla}_{i}\overline{\tilde{T}^{r}_{qj}}F_{\bar{r}}F_{p}\\ &\ \ \ \ +(\tilde{\nabla}_{\bar{q}}\tilde{\nabla}_{i}\tilde{\nabla}_{\bar{j}}F+\tilde{R}_{i\bar{q}\;\bar{j}}^{\;\;\;\bar{\ell}}F_{\bar{\ell}})F_{p}\big{)}+|\tilde{\nabla}\tilde{\nabla}F|^{2}_{\omega_{\varphi}}+|\tilde{\nabla}\bar{\tilde{\nabla}}F|^{2}_{\omega_{\varphi}}\\ &=\tilde{g}^{p\bar{q}}\big{(}{\Delta_{\varphi}}F)_{p}F_{\bar{q}}+2\text{Re}(\tilde{g}^{i\bar{j}}\tilde{g}^{p\bar{q}}\tilde{T}^{r}_{pi}\tilde{\nabla}_{r}F_{\bar{j}}F_{\bar{q}})+\tilde{g}^{i\bar{j}}\tilde{g}^{p\bar{q}}\tilde{\nabla}_{i}\overline{\tilde{T}^{r}_{qj}}F_{\bar{r}}F_{p}\\ &\ \ \ \ +\tilde{g}^{p\bar{q}}({\Delta_{\varphi}}F)_{\bar{q}}F_{p}+\tilde{g}^{i\bar{j}}\tilde{g}^{p\bar{q}}\tilde{R}_{i\bar{q}\;\bar{j}}^{\;\;\;\bar{\ell}}F_{\bar{\ell}}F_{p}+|\tilde{\nabla}\tilde{\nabla}F|^{2}_{\omega_{\varphi}}+|\tilde{\nabla}\bar{\tilde{\nabla}}F|^{2}_{\omega_{\varphi}}\\ &=\tilde{g}^{p\bar{q}}\big{(}{\Delta_{\varphi}}F)_{p}F_{\bar{q}}+2\text{Re}(\tilde{g}^{i\bar{j}}\tilde{g}^{p\bar{q}}\tilde{g}^{r\bar{k}}\tilde{T}_{pi\bar{k}}\tilde{\nabla}_{r}F_{\bar{j}}F_{\bar{q}})\\ &\ \ \ \ +\tilde{g}^{i\bar{j}}\tilde{g}^{p\bar{q}}\tilde{g}^{t\bar{r}}\tilde{\nabla}_{i}(\overline{\tilde{T}_{qj\bar{t}}})F_{\bar{r}}F_{p}+\tilde{g}^{p\bar{q}}({\Delta_{\varphi}}F)_{\bar{q}}F_{p}+\tilde{g}^{i\bar{j}}\tilde{g}^{p\bar{q}}\tilde{g}^{k\bar{\ell}}\tilde{R}_{i\bar{q}k\bar{j}}F_{\bar{\ell}}F_{p}\\ &\ \ \ \ +|\tilde{\nabla}\tilde{\nabla}F|^{2}_{\omega_{\varphi}}+|\tilde{\nabla}\bar{\tilde{\nabla}}F|^{2}_{\omega_{\varphi}}\\ \end{split}

We now commute indices of the curvature tensor as in 4 and trace to get that

(11) Δφ(|F|ωφ2)=2Re(g~pq¯(ΔφF)pFq¯)+2Re(g~ij¯g~pq¯g~rk¯T~pik¯~rFj¯Fq¯)+g~ij¯g~pq¯g~tr¯~i(T~qjt¯¯)Fr¯Fp+g~pq¯g~k¯R~kq¯F¯Fpg~pq¯g~k¯g~rs¯~q¯(T~rks¯)F¯Fp+|~~F|ωφ2+|~~¯F|ωφ2.\displaystyle\begin{split}{\Delta_{\varphi}}(|\partial F|^{2}_{\omega_{\varphi}})&=2\text{Re}(\tilde{g}^{p\bar{q}}\big{(}{\Delta_{\varphi}}F)_{p}F_{\bar{q}})+2\text{Re}(\tilde{g}^{i\bar{j}}\tilde{g}^{p\bar{q}}\tilde{g}^{r\bar{k}}\tilde{T}_{pi\bar{k}}\tilde{\nabla}_{r}F_{\bar{j}}F_{\bar{q}})\\ &\ \ \ \ +\tilde{g}^{i\bar{j}}\tilde{g}^{p\bar{q}}\tilde{g}^{t\bar{r}}\tilde{\nabla}_{i}(\overline{\tilde{T}_{qj\bar{t}}})F_{\bar{r}}F_{p}+\tilde{g}^{p\bar{q}}\tilde{g}^{k\bar{\ell}}\tilde{R}_{k\bar{q}}F_{\bar{\ell}}F_{p}\\ &\ \ \ \ -\tilde{g}^{p\bar{q}}\tilde{g}^{k\bar{\ell}}\tilde{g}^{r\bar{s}}\tilde{\nabla}_{\bar{q}}(\tilde{T}_{rk\bar{s}})F_{\bar{\ell}}F_{p}+|\tilde{\nabla}\tilde{\nabla}F|^{2}_{\omega_{\varphi}}+|\tilde{\nabla}\bar{\tilde{\nabla}}F|^{2}_{\omega_{\varphi}}.\end{split}

For a general real-valued function A(F)A(F),

eA(F)Δφ(eA(F)|F|ωφ2)\displaystyle e^{-A(F)}{\Delta_{\varphi}}(e^{A(F)}|\partial F|^{2}_{\omega_{\varphi}}) =Δφ(|F|ωφ2)+2ARe(g~ij¯g~k¯(FiFkF¯j¯+FiF¯Fkj¯))\displaystyle={\Delta_{\varphi}}(|\partial F|^{2}_{\omega_{\varphi}})+2A^{\prime}\text{Re}(\tilde{g}^{i\bar{j}}\tilde{g}^{k\bar{\ell}}(F_{i}F_{k}F_{\bar{\ell}\bar{j}}+F_{i}F_{\bar{\ell}}F_{k\bar{j}}))
+(A2+A′′)|F|ωφ4+AΔφF|F|ωφ2,\displaystyle\ \ \ \ +(A^{\prime 2}+A^{\prime\prime})|\partial F|^{4}_{\omega_{\varphi}}+A^{\prime}{\Delta_{\varphi}}F|\partial F|^{2}_{\omega_{\varphi}},

where F¯j¯F_{\bar{\ell}\bar{j}} denotes ~j¯~¯F\tilde{\nabla}_{\bar{j}}\tilde{\nabla}_{\bar{\ell}}F. Substituting (10) for the first term in the above equation and using the completed square:

A2|F|ωφ4+2ARe(g~ij¯g~k¯FiFkF¯j¯)+|~~F|ωφ20,\displaystyle A^{\prime 2}|\partial F|^{4}_{\omega_{\varphi}}+2A^{\prime}\text{Re}(\tilde{g}^{i\bar{j}}\tilde{g}^{k\bar{\ell}}F_{i}F_{k}F_{\bar{\ell}\bar{j}})+|\tilde{\nabla}\tilde{\nabla}F|^{2}_{\omega_{\varphi}}\geq 0,

we obtain that

eA(F)Δφ(eA(F)|F|ωφ2)\displaystyle e^{-A(F)}{\Delta_{\varphi}}(e^{A(F)}|\partial F|^{2}_{\omega_{\varphi}}) 2Re(g~pq¯(ΔφF)pFq¯)+2Re(g~ij¯g~pq¯g~rk¯T~pik¯~rFj¯Fq¯)\displaystyle\geq 2\text{Re}(\tilde{g}^{p\bar{q}}({\Delta_{\varphi}}F)_{p}F_{\bar{q}})+2\text{Re}(\tilde{g}^{i\bar{j}}\tilde{g}^{p\bar{q}}\tilde{g}^{r\bar{k}}\tilde{T}_{pi\bar{k}}\tilde{\nabla}_{r}F_{\bar{j}}F_{\bar{q}})
+g~ij¯g~pq¯g~tr¯~i(T~qjt¯¯)Fr¯Fp+g~pq¯g~k¯R~kq¯F¯Fp\displaystyle\ \ \ \ +\tilde{g}^{i\bar{j}}\tilde{g}^{p\bar{q}}\tilde{g}^{t\bar{r}}\tilde{\nabla}_{i}(\overline{\tilde{T}_{qj\bar{t}}})F_{\bar{r}}F_{p}+\tilde{g}^{p\bar{q}}\tilde{g}^{k\bar{\ell}}\tilde{R}_{k\bar{q}}F_{\bar{\ell}}F_{p}
g~pq¯g~k¯g~rs¯~q¯(T~rks¯)F¯Fp+|~~¯F|ωφ2+2Ag~ij¯g~k¯FiF¯Fkj¯\displaystyle\ \ \ \ -\tilde{g}^{p\bar{q}}\tilde{g}^{k\bar{\ell}}\tilde{g}^{r\bar{s}}\tilde{\nabla}_{\bar{q}}(\tilde{T}_{rk\bar{s}})F_{\bar{\ell}}F_{p}+|\tilde{\nabla}\bar{\tilde{\nabla}}F|^{2}_{\omega_{\varphi}}+2A^{\prime}\tilde{g}^{i\bar{j}}\tilde{g}^{k\bar{\ell}}F_{i}F_{\bar{\ell}}F_{k\bar{j}}
+A′′|F|ωφ4+AΔφF|F|ωφ2.\displaystyle\ \ \ \ +A^{\prime\prime}|\partial F|^{4}_{\omega_{\varphi}}+A^{\prime}{\Delta_{\varphi}}F|\partial F|^{2}_{\omega_{\varphi}}.

Applying the relation R~kq¯=Rkq¯Fkq¯\tilde{R}_{k\bar{q}}=R_{k\bar{q}}-F_{k\bar{q}} to switch the Ricci curvature of ωφ\omega_{\varphi} to that of ω\omega,

(12) eΔφA(F)(eA(F)|F|ωφ2)2Re(g~pq¯(ΔφF)pFq¯)+2Re(g~ij¯g~pq¯g~rk¯T~pik¯~rFj¯Fq¯)+g~ij¯g~pq¯g~tr¯~i(T~qjt¯¯)Fr¯Fp+g~pq¯g~k¯Rkq¯F¯Fpg~pq¯g~k¯g~rs¯~q¯(T~rks¯)F¯Fp+|~~¯F|ωφ2+(2A1)g~ij¯g~k¯FiF¯Fkj¯+A′′|F|ωφ4+AΔφF|F|ωφ22Re(g~pq¯(ΔφF)pFq¯)+2Re(g~ij¯g~pq¯g~rk¯T~pik¯~rFj¯Fq¯)+g~ij¯g~pq¯g~tr¯~i(T~qjt¯¯)Fr¯Fp+g~pq¯g~k¯Rkq¯F¯Fpg~pq¯g~k¯g~rs¯~q¯(T~rks¯)F¯Fp+(1(A12))|~~¯F|ωφ2+(A′′(A12))|F|ωφ4+AΔφF|F|ωφ2,\displaystyle\begin{split}e&{}^{-A(F)}{\Delta_{\varphi}}(e^{A(F)}|\partial F|^{2}_{\omega_{\varphi}})\\ &\geq 2\text{Re}(\tilde{g}^{p\bar{q}}({\Delta_{\varphi}}F)_{p}F_{\bar{q}})+2\text{Re}(\tilde{g}^{i\bar{j}}\tilde{g}^{p\bar{q}}\tilde{g}^{r\bar{k}}\tilde{T}_{pi\bar{k}}\tilde{\nabla}_{r}F_{\bar{j}}F_{\bar{q}})+\tilde{g}^{i\bar{j}}\tilde{g}^{p\bar{q}}\tilde{g}^{t\bar{r}}\tilde{\nabla}_{i}(\overline{\tilde{T}_{qj\bar{t}}})F_{\bar{r}}F_{p}\\ &\ \ \ \ +\tilde{g}^{p\bar{q}}\tilde{g}^{k\bar{\ell}}R_{k\bar{q}}F_{\bar{\ell}}F_{p}-\tilde{g}^{p\bar{q}}\tilde{g}^{k\bar{\ell}}\tilde{g}^{r\bar{s}}\tilde{\nabla}_{\bar{q}}(\tilde{T}_{rk\bar{s}})F_{\bar{\ell}}F_{p}+|\tilde{\nabla}\bar{\tilde{\nabla}}F|^{2}_{\omega_{\varphi}}\\ &\ \ \ \ +(2A^{\prime}-1)\tilde{g}^{i\bar{j}}\tilde{g}^{k\bar{\ell}}F_{i}F_{\bar{\ell}}F_{k\bar{j}}+A^{\prime\prime}|\partial F|^{4}_{\omega_{\varphi}}+A^{\prime}{\Delta_{\varphi}}F|\partial F|^{2}_{\omega_{\varphi}}\\ &\geq 2\text{Re}(\tilde{g}^{p\bar{q}}({\Delta_{\varphi}}F)_{p}F_{\bar{q}})+2\text{Re}(\tilde{g}^{i\bar{j}}\tilde{g}^{p\bar{q}}\tilde{g}^{r\bar{k}}\tilde{T}_{pi\bar{k}}\tilde{\nabla}_{r}F_{\bar{j}}F_{\bar{q}})+\tilde{g}^{i\bar{j}}\tilde{g}^{p\bar{q}}\tilde{g}^{t\bar{r}}\tilde{\nabla}_{i}(\overline{\tilde{T}_{qj\bar{t}}})F_{\bar{r}}F_{p}\\ &\ \ \ \ +\tilde{g}^{p\bar{q}}\tilde{g}^{k\bar{\ell}}R_{k\bar{q}}F_{\bar{\ell}}F_{p}-\tilde{g}^{p\bar{q}}\tilde{g}^{k\bar{\ell}}\tilde{g}^{r\bar{s}}\tilde{\nabla}_{\bar{q}}(\tilde{T}_{rk\bar{s}})F_{\bar{\ell}}F_{p}+(1-(A^{\prime}-\tfrac{1}{2}))|\tilde{\nabla}\bar{\tilde{\nabla}}F|^{2}_{\omega_{\varphi}}\\ &\ \ \ \ +(A^{\prime\prime}-(A^{\prime}-\tfrac{1}{2}))|\partial F|^{4}_{\omega_{\varphi}}+A^{\prime}{\Delta_{\varphi}}F|\partial F|^{2}_{\omega_{\varphi}},\end{split}

where we used the following Cauchy-Schwarz inequality:

(2A1)g~ij¯g~k¯FiF¯Fkj¯\displaystyle(2A^{\prime}-1)\tilde{g}^{i\bar{j}}\tilde{g}^{k\bar{\ell}}F_{i}F_{\bar{\ell}}F_{k\bar{j}} (A12)|F|ωφ4(A12)|~~¯F|ωφ2\displaystyle\geq-(A^{\prime}-\tfrac{1}{2})|\partial F|^{4}_{\omega_{\varphi}}-(A^{\prime}-\tfrac{1}{2})|\tilde{\nabla}\bar{\tilde{\nabla}}F|^{2}_{\omega_{\varphi}}

for A>12A^{\prime}>\tfrac{1}{2}.

In order to control the bad torsion terms (the second, third and fifth terms in the last line of (12)), we will need to strategically choose our function A(F)A(F) to ensure that 1(A12)>01-(A^{\prime}-\tfrac{1}{2})>0 and A′′(A12)>0A^{\prime\prime}-(A^{\prime}-\tfrac{1}{2})>0. We can accomplish this by choosing

A(F)=κeF+F(12ε),A(F)=\kappa e^{F}+F(\tfrac{1}{2}-\varepsilon),

so that A(F)=κeF+12εA^{\prime}(F)=\kappa e^{F}+\tfrac{1}{2}-\varepsilon and A′′(F)=κeFA^{\prime\prime}(F)=\kappa e^{F}. We can then pick ε,κ>0\varepsilon,\kappa>0 such that

0A′′ε=A1212{κeminXFε0κemaxXFε12.0\leq A^{\prime\prime}-\varepsilon=A^{\prime}-\tfrac{1}{2}\leq\tfrac{1}{2}\ \ \ \Leftrightarrow\ \ \ \begin{cases}\kappa e^{\min_{X}F}-\varepsilon\geq 0\\ \kappa e^{\max_{X}F}-\varepsilon\leq\tfrac{1}{2}\ .\end{cases}

We can first choose κ\kappa small enough such that κemaxXF12\kappa e^{\max_{X}F}\leq\tfrac{1}{2}. Then choose ε\varepsilon small enough such that κeminFε\kappa e^{\min F}\geq\varepsilon. This ensures that A(12,1)A^{\prime}\in(\tfrac{1}{2},1).

It follows that

e\displaystyle e ΔφA(F)(eA(F)|F|ωφ2){}^{-A(F)}{\Delta_{\varphi}}(e^{A(F)}|\partial F|_{\omega_{\varphi}}^{2})
2Re(g~pq¯(ΔφF)pFq¯)+2Re(g~ij¯g~pq¯g~rk¯T~pik¯~rFj¯Fq¯)+g~ij¯g~pq¯g~tr¯~i(T~qjt¯¯)Fr¯Fp\displaystyle\geq 2\text{Re}\big{(}\tilde{g}^{p\bar{q}}\big{(}{\Delta_{\varphi}}F)_{p}F_{\bar{q}}\big{)}+2\text{Re}(\tilde{g}^{i\bar{j}}\tilde{g}^{p\bar{q}}\tilde{g}^{r\bar{k}}\tilde{T}_{pi\bar{k}}\tilde{\nabla}_{r}F_{\bar{j}}F_{\bar{q}})+\tilde{g}^{i\bar{j}}\tilde{g}^{p\bar{q}}\tilde{g}^{t\bar{r}}\tilde{\nabla}_{i}(\overline{\tilde{T}_{qj\bar{t}}})F_{\bar{r}}F_{p}
+g~pq¯g~k¯Rkq¯FpF¯g~pq¯g~k¯g~rs¯~q¯(T~rks¯)F¯Fp+12|~~¯F|ωφ2+ε|F|ωφ4+AΔφF|F|ωφ2\displaystyle\ \ \ \ +\tilde{g}^{p\bar{q}}\tilde{g}^{k\bar{\ell}}R_{k\bar{q}}F_{p}F_{\bar{\ell}}-\tilde{g}^{p\bar{q}}\tilde{g}^{k\bar{\ell}}\tilde{g}^{r\bar{s}}\tilde{\nabla}_{\bar{q}}(\tilde{T}_{rk\bar{s}})F_{\bar{\ell}}F_{p}+\tfrac{1}{2}|\tilde{\nabla}\bar{\tilde{\nabla}}F|^{2}_{\omega_{\varphi}}+\varepsilon|\partial F|_{\omega_{\varphi}}^{4}+A^{\prime}{\Delta_{\varphi}}F|\partial F|_{\omega_{\varphi}}^{2}
2Re(g~pq¯(ΔφF)pFq¯)+2Re(g~ij¯g~pq¯g~rk¯Tpik¯~rFj¯Fq¯)+g~ij¯g~pq¯g~tr¯iTqjt¯¯Fr¯Fp\displaystyle\geq 2\text{Re}\big{(}\tilde{g}^{p\bar{q}}\big{(}{\Delta_{\varphi}}F)_{p}F_{\bar{q}}\big{)}+2\text{Re}(\tilde{g}^{i\bar{j}}\tilde{g}^{p\bar{q}}\tilde{g}^{r\bar{k}}T_{pi\bar{k}}\tilde{\nabla}_{r}F_{\bar{j}}F_{\bar{q}})+\tilde{g}^{i\bar{j}}\tilde{g}^{p\bar{q}}\tilde{g}^{t\bar{r}}\partial_{i}\overline{T_{qj\bar{t}}}F_{\bar{r}}F_{p}
g~ij¯g~pq¯g~tr¯g~sk¯ig~tk¯Tqjs¯¯Fr¯Fp+g~pq¯g~k¯Rkq¯FpF¯g~pq¯g~k¯g~rs¯q¯(Trks¯)F¯Fp\displaystyle\ \ \ \ -\tilde{g}^{i\bar{j}}\tilde{g}^{p\bar{q}}\tilde{g}^{t\bar{r}}\tilde{g}^{s\bar{k}}\partial_{i}\tilde{g}_{t\bar{k}}\overline{T_{qj\bar{s}}}F_{\bar{r}}F_{p}+\tilde{g}^{p\bar{q}}\tilde{g}^{k\bar{\ell}}R_{k\bar{q}}F_{p}F_{\bar{\ell}}-\tilde{g}^{p\bar{q}}\tilde{g}^{k\bar{\ell}}\tilde{g}^{r\bar{s}}\partial_{\bar{q}}(T_{rk\bar{s}})F_{\bar{\ell}}F_{p}
+g~pq¯g~k¯g~rs¯g~ij¯q¯g~is¯Trkj¯F¯Fp+12|~~¯F|ωφ2+ε|F|ωφ4|ΔφFF|ωφ2\displaystyle\ \ \ \ +\tilde{g}^{p\bar{q}}\tilde{g}^{k\bar{\ell}}\tilde{g}^{r\bar{s}}\tilde{g}^{i\bar{j}}\partial_{\bar{q}}\tilde{g}_{i\bar{s}}T_{rk\bar{j}}F_{\bar{\ell}}F_{p}+\tfrac{1}{2}|\tilde{\nabla}\bar{\tilde{\nabla}}F|^{2}_{\omega_{\varphi}}+\varepsilon|\partial F|^{4}_{\omega_{\varphi}}-|{\Delta_{\varphi}}F\|\partial F|^{2}_{\omega_{\varphi}}

where we converted the covariant derivatives to partial derivatives as in (2) and passed the torsion terms of g~\tilde{g} to those of gg as in (3).

Applying Young’s inequality and choosing BB to be at least 3(n1)3(n-1), where the factor of n1n-1 comes from the fact that trωωφC(trωφω)n1\operatorname{tr}_{\omega}\omega_{\varphi}\leq C(\operatorname{tr}_{\omega_{\varphi}}\omega)^{n-1}, we have

e\displaystyle e ΔφA(F)(eA(F)|F|ωφ2){}^{-A(F)}{\Delta_{\varphi}}(e^{A(F)}|\partial F|^{2}_{\omega_{\varphi}})
2Re(g~pq¯(ΔφF)pFq¯)+2Re(g~ij¯g~pq¯g~rk¯Tpik¯~rFj¯Fq¯)\displaystyle\geq 2\text{Re}(\tilde{g}^{p\bar{q}}\big{(}{\Delta_{\varphi}}F)_{p}F_{\bar{q}})+2\text{Re}(\tilde{g}^{i\bar{j}}\tilde{g}^{p\bar{q}}\tilde{g}^{r\bar{k}}T_{pi\bar{k}}\tilde{\nabla}_{r}F_{\bar{j}}F_{\bar{q}})
+g~ij¯g~pq¯g~tr¯iTqjt¯¯Fr¯Fpg~ij¯g~pq¯g~tr¯g~sk¯ig~tk¯Tqjs¯¯Fr¯Fp+g~pq¯g~k¯Rkq¯FpF¯\displaystyle\ \ \ \ +\tilde{g}^{i\bar{j}}\tilde{g}^{p\bar{q}}\tilde{g}^{t\bar{r}}\partial_{i}\overline{T_{qj\bar{t}}}F_{\bar{r}}F_{p}-\tilde{g}^{i\bar{j}}\tilde{g}^{p\bar{q}}\tilde{g}^{t\bar{r}}\tilde{g}^{s\bar{k}}\partial_{i}\tilde{g}_{t\bar{k}}\overline{T_{qj\bar{s}}}F_{\bar{r}}F_{p}+\tilde{g}^{p\bar{q}}\tilde{g}^{k\bar{\ell}}R_{k\bar{q}}F_{p}F_{\bar{\ell}}
g~pq¯g~k¯g~rs¯q¯(Trks¯)F¯Fp+g~pq¯g~k¯g~rs¯g~ij¯q¯g~is¯Trkj¯F¯Fp+12|~~¯F|ωφ2\displaystyle\ \ \ \ -\tilde{g}^{p\bar{q}}\tilde{g}^{k\bar{\ell}}\tilde{g}^{r\bar{s}}\partial_{\bar{q}}(T_{rk\bar{s}})F_{\bar{\ell}}F_{p}+\tilde{g}^{p\bar{q}}\tilde{g}^{k\bar{\ell}}\tilde{g}^{r\bar{s}}\tilde{g}^{i\bar{j}}\partial_{\bar{q}}\tilde{g}_{i\bar{s}}T_{rk\bar{j}}F_{\bar{\ell}}F_{p}+\tfrac{1}{2}|\tilde{\nabla}\bar{\tilde{\nabla}}F|^{2}_{\omega_{\varphi}}
+ε|F|ωφ4|ΔφFF|ωφ2\displaystyle\ \ \ \ +\varepsilon|\partial F|^{4}_{\omega_{\varphi}}-|{\Delta_{\varphi}}F\|\partial F|^{2}_{\omega_{\varphi}}
2Re(g~pq¯(ΔφF)pFq¯)C(trωωφ)Bgij¯g~k¯g~pq¯ig~kq¯j¯g~p¯+14|~~¯F|ωφ2\displaystyle\geq 2\text{Re}(\tilde{g}^{p\bar{q}}\big{(}{\Delta_{\varphi}}F)_{p}F_{\bar{q}})-C(\operatorname{tr}_{\omega}\omega_{\varphi})^{B}g^{i\bar{j}}\tilde{g}^{k\bar{\ell}}\tilde{g}^{p\bar{q}}\partial_{i}\tilde{g}_{k\bar{q}}\partial_{\bar{j}}\tilde{g}_{p\bar{\ell}}+\tfrac{1}{4}|\tilde{\nabla}\bar{\tilde{\nabla}}F|^{2}_{\omega_{\varphi}}
C(trωωφ)B|F|ωφ2C(trωωφ)B,\displaystyle\ \ \ \ -C(\operatorname{tr}_{\omega}\omega_{\varphi})^{B}|\partial F|^{2}_{\omega_{\varphi}}-C(\operatorname{tr}_{\omega}\omega_{\varphi})^{B},

where the constant in front of the fourth term depends on RφC0\|{\operatorname{R}_{\varphi}}\|_{C^{0}}. Now, we use the following computation in the proof of Equation (9.5) of [34] for Δφtrωωφ{\Delta_{\varphi}}\operatorname{tr}_{\omega}\omega_{\varphi}:

Δφtrωωφ\displaystyle{\Delta_{\varphi}}\operatorname{tr}_{\omega}\omega_{\varphi} =g~pj¯g~iq¯gk¯kg~ij¯¯g~pq¯+2Re(g~ij¯gk¯Tkip¯g~pj¯)+g~ij¯gk¯TikpTjq¯g~pq¯\displaystyle=\tilde{g}^{p\bar{j}}\tilde{g}^{i\bar{q}}g^{k\bar{\ell}}\nabla_{k}\tilde{g}_{i\bar{j}}\nabla_{\bar{\ell}}\tilde{g}_{p\bar{q}}+2\text{Re}(\tilde{g}^{i\bar{j}}g^{k\bar{\ell}}T^{p}_{ki}\nabla_{\bar{\ell}}\tilde{g}_{p\bar{j}})+\tilde{g}^{i\bar{j}}g^{k\bar{\ell}}T^{p}_{ik}\overline{T^{q}_{j\ell}}\tilde{g}_{p\bar{q}}
+gij¯Fij¯R+g~ij¯iTj¯+g~ij¯gk¯¯Tikpg~ij¯gk¯g~kq¯(iTjq¯Ri¯pj¯gpq¯)\displaystyle\ \ \ \ +g^{i\bar{j}}F_{i\bar{j}}-R+\tilde{g}^{i\bar{j}}\nabla_{i}\overline{T^{\ell}_{j\ell}}+\tilde{g}^{i\bar{j}}g^{k\bar{\ell}}\nabla_{\bar{\ell}}T^{p}_{ik}-\tilde{g}^{i\bar{j}}g^{k\bar{\ell}}\tilde{g}_{k\bar{q}}(\nabla_{i}\overline{T^{q}_{j\ell}}-R_{i\bar{\ell}p\bar{j}}g^{p\bar{q}})
g~ij¯gk¯TikpTjq¯gpq¯.\displaystyle\ \ \ \ -\tilde{g}^{i\bar{j}}g^{k\bar{\ell}}T^{p}_{ik}\overline{T^{q}_{j\ell}}g_{p\bar{q}}.

Converting the first term into covariant derivatives and applying Young’s inequality, we have

g~pj¯g~iq¯gk¯kg~ij¯¯g~pq¯g~pj¯g~iq¯gk¯kg~ij¯¯g~pq¯ε2g~pj¯g~iq¯gk¯kg~ij¯¯g~pq¯C(trωωφ)n.\displaystyle\tilde{g}^{p\bar{j}}\tilde{g}^{i\bar{q}}g^{k\bar{\ell}}\nabla_{k}\tilde{g}_{i\bar{j}}\nabla_{\bar{\ell}}\tilde{g}_{p\bar{q}}\geq\tilde{g}^{p\bar{j}}\tilde{g}^{i\bar{q}}g^{k\bar{\ell}}\partial_{k}\tilde{g}_{i\bar{j}}\partial_{\bar{\ell}}\tilde{g}_{p\bar{q}}-\tfrac{\varepsilon}{2}\tilde{g}^{p\bar{j}}\tilde{g}^{i\bar{q}}g^{k\bar{\ell}}\partial_{k}\tilde{g}_{i\bar{j}}\partial_{\bar{\ell}}\tilde{g}_{p\bar{q}}-C(\operatorname{tr}_{\omega}\omega_{\varphi})^{n}.

Likewise, the second term can be bounded below by

2Re(g~ij¯gk¯Tkip¯g~pj¯)ε2g~pj¯g~iq¯gk¯kg~ij¯¯g~pq¯C(trωωφ)n,\displaystyle 2\text{Re}(\tilde{g}^{i\bar{j}}g^{k\bar{\ell}}T^{p}_{ki}\nabla_{\bar{\ell}}\tilde{g}_{p\bar{j}})\geq-\tfrac{\varepsilon}{2}\tilde{g}^{p\bar{j}}\tilde{g}^{i\bar{q}}g^{k\bar{\ell}}\partial_{k}\tilde{g}_{i\bar{j}}\partial_{\bar{\ell}}\tilde{g}_{p\bar{q}}-C(\operatorname{tr}_{\omega}\omega_{\varphi})^{n},

and the fourth term by

gij¯Fij¯\displaystyle g^{i\bar{j}}F_{i\bar{j}} |~~¯F|ωφ2δCδ(trωωφ)2.\displaystyle\geq-\tfrac{|\tilde{\nabla}\bar{\tilde{\nabla}}F|_{\omega_{\varphi}}^{2}}{\delta}-C\delta(\operatorname{tr}_{\omega}\omega_{\varphi})^{2}.

It is straightforward to see that the remaining terms can be bounded below by C(trωωφ)n-C(\operatorname{tr}_{\omega}\omega_{\varphi})^{n}. Choosing BnB\geq n and δ=4eA(F)N(B+1)(trωωφ)B\delta=4e^{-A(F)}N(B+1)(\operatorname{tr}_{\omega}\omega_{\varphi})^{B}, we arrive at the following:

Δφtrωωφ\displaystyle{\Delta_{\varphi}}\operatorname{tr}_{\omega}\omega_{\varphi} (1ε)g~pj¯g~iq¯gk¯kg~ij¯¯g~pq¯eA(F)4N(B+1)(trωωφ)B|~~¯F|ωφ2C(trωωφ)B+2.\displaystyle\geq(1-\varepsilon)\tilde{g}^{p\bar{j}}\tilde{g}^{i\bar{q}}g^{k\bar{\ell}}\partial_{k}\tilde{g}_{i\bar{j}}\partial_{\bar{\ell}}\tilde{g}_{p\bar{q}}-\tfrac{e^{A(F)}}{4N(B+1)(\operatorname{tr}_{\omega}\omega_{\varphi})^{B}}|\tilde{\nabla}\bar{\tilde{\nabla}}F|^{2}_{\omega_{\varphi}}-C(\operatorname{tr}_{\omega}\omega_{\varphi})^{B+2}.

Observe that

(13) Δφ(trωωφ)B+1=(B+1)B(trωωφ)B1|trωωφ|ωφ2+(B+1)(trωωφ)BΔφtrωωφ(B+1)(trωωφ)BΔφtrωωφ.\displaystyle\begin{split}{\Delta_{\varphi}}(\operatorname{tr}_{\omega}\omega_{\varphi})^{B+1}&=(B+1)B(\operatorname{tr}_{\omega}\omega_{\varphi})^{B-1}|\partial\operatorname{tr}_{\omega}\omega_{\varphi}|^{2}_{\omega_{\varphi}}+(B+1)(\operatorname{tr}_{\omega}\omega_{\varphi})^{B}{\Delta_{\varphi}}\operatorname{tr}_{\omega}\omega_{\varphi}\\ &\geq(B+1)(\operatorname{tr}_{\omega}\omega_{\varphi})^{B}{\Delta_{\varphi}}\operatorname{tr}_{\omega}\omega_{\varphi}.\end{split}

Choosing NN sufficiently large and letting Q:=eA(F)|F|ωφ2+N(trωωφ)B+1Q:=e^{A(F)}|\partial F|^{2}_{\omega_{\varphi}}+N(\operatorname{tr}_{\omega}\omega_{\varphi})^{B+1}, we arrive at

ΔφQ\displaystyle{\Delta_{\varphi}}Q =Δφ(eA(F)|F|ωφ2+N(trωωφ)B+1)\displaystyle={\Delta_{\varphi}}(e^{A(F)}|\partial F|^{2}_{\omega_{\varphi}}+N(\operatorname{tr}_{\omega}\omega_{\varphi})^{B+1})
2eA(F)Re(g~pq¯(ΔφF)pFq¯)C(trωωφ)B|F|ωφ2C(trωωφ)2B+2\displaystyle\geq 2e^{A(F)}\text{Re}(\tilde{g}^{p\bar{q}}\big{(}{\Delta_{\varphi}}F)_{p}F_{\bar{q}})-C(\operatorname{tr}_{\omega}\omega_{\varphi})^{B}|\partial F|^{2}_{\omega_{\varphi}}-C(\operatorname{tr}_{\omega}\omega_{\varphi})^{2B+2}
2eA(F)Re(g~pq¯(ΔφF)pFq¯)C(trωωφ)B+1Q.\displaystyle\geq 2e^{A(F)}\text{Re}(\tilde{g}^{p\bar{q}}\big{(}{\Delta_{\varphi}}F)_{p}F_{\bar{q}})-C(\operatorname{tr}_{\omega}\omega_{\varphi})^{B+1}Q.

Now we demonstrate how to handle the non-vanishing first derivative of ΔφF\Delta_{\varphi}F in the first term on the right-hand side of the above inequality. We will use integration by parts on this term in order to bound it, this follows from the same argument used by Chen-Cheng in Section 3 of [11]. Let us compute

ΔφQ2p+1=(2p+1)2p|Q|ωφ2Q2p1+(2p+1)Q2pΔφQ.\displaystyle{\Delta_{\varphi}}Q^{2p+1}=(2p+1)2p|\nabla Q|^{2}_{\omega_{\varphi}}Q^{2p-1}+(2p+1)Q^{2p}{\Delta_{\varphi}}Q.

Integrating both sides and applying our bound from Equation 13, we have that

X2p|Q|ωφ2Q2p1ωφn\displaystyle\int_{X}2p|\nabla Q|^{2}_{\omega_{\varphi}}Q^{2p-1}\omega_{\varphi}^{n} =XQ2p(ΔφQ)ωφn\displaystyle=\int_{X}Q^{2p}(-{\Delta_{\varphi}}Q)\omega_{\varphi}^{n}
2XeA(F)Re(g~pq¯(ΔφF)pFq¯)Q2pωφn\displaystyle\leq-2\int_{X}e^{A(F)}\text{Re}(\tilde{g}^{p\bar{q}}\big{(}{\Delta_{\varphi}}F)_{p}F_{\bar{q}})Q^{2p}\omega_{\varphi}^{n}
+CX(trωωφ)B+1Q2p+1ωφn.\displaystyle\ \ \ \ +C\int_{X}(\operatorname{tr}_{\omega}\omega_{\varphi})^{B+1}Q^{2p+1}\omega_{\varphi}^{n}.

The first integral on the right-hand side can be integrated by parts using Lemma 1 in the following way

2XeA(F)Re(g~pq¯(ΔφF)pFq¯)Q2pωφn\displaystyle-2\int_{X}e^{A(F)}\text{Re}(\tilde{g}^{p\bar{q}}\big{(}{\Delta_{\varphi}}F)_{p}F_{\bar{q}})Q^{2p}\omega_{\varphi}^{n} =2XRe(g~pq¯p(eA(F)Fq¯Q2p))ΔφFωφn\displaystyle=2\int_{X}\text{Re}(\tilde{g}^{p\bar{q}}\nabla_{p}(e^{A(F)}F_{\bar{q}}Q^{2p})){\Delta_{\varphi}}F\omega_{\varphi}^{n}
2XRe(g~pq¯eA(F)Fq¯Q2p(trT~)p)ΔφFωφn\displaystyle\ \ \ \ -2\int_{X}\text{Re}(\tilde{g}^{p\bar{q}}e^{A(F)}F_{\bar{q}}Q^{2p}(\operatorname{tr}\tilde{T})_{p}){\Delta_{\varphi}}F\omega_{\varphi}^{n}

Now, the first term on the right-hand side of the above equation can be handled exactly as in Equations (3.14)-(3.18) of [11]. Let us demonstrate how we can handle the second integral with the torsion term.

Firstly, note that (trT~)p=T~sps=g~s¯Tsppgp¯(\operatorname{tr}\tilde{T})_{p}=\tilde{T}^{s}_{sp}=\tilde{g}^{s\bar{\ell}}T^{p}_{sp}g_{p\bar{\ell}} and that we can bound the ΔφF\Delta_{\varphi}F term by

|ΔφF|RφC0+trωφRicC0C(1+(trωωφ)n1).\displaystyle|{\Delta_{\varphi}}F|\leq\|{\operatorname{R}_{\varphi}}\|_{C^{0}}+\|\operatorname{tr}_{\omega_{\varphi}}\operatorname{Ric}\|_{C^{0}}\leq C(1+(\operatorname{tr}_{\omega}\omega_{\varphi})^{n-1}).

Computing, the integral involving torsion can be controlled as follows:

2XRe(g~pq¯eA(F)\displaystyle-2\int_{X}\text{Re}(\tilde{g}^{p\bar{q}}e^{A(F)} Fq¯Q2p(trT~)p)ΔφFωφn=2XRe(g~pq¯eA(F)Fq¯Q2pg~s¯Tsprgr¯)ΔφFωφn\displaystyle F_{\bar{q}}Q^{2p}(\operatorname{tr}\tilde{T})_{p}){\Delta_{\varphi}}F\omega_{\varphi}^{n}=-2\int_{X}\text{Re}(\tilde{g}^{p\bar{q}}e^{A(F)}F_{\bar{q}}Q^{2p}\tilde{g}^{s\bar{\ell}}T^{r}_{sp}g_{r\bar{\ell}}){\Delta_{\varphi}}F\omega_{\varphi}^{n}
2XeA(F)Q2p(trωφω)|~F|ωφ|Tspr|ωφ|ΔφF|ωφn\displaystyle\leq 2\int_{X}e^{A(F)}Q^{2p}(\operatorname{tr}_{\omega_{\varphi}}\omega)|\tilde{\nabla}F|_{\omega_{\varphi}}|T^{r}_{sp}|_{\omega_{\varphi}}|{\Delta_{\varphi}}F|\omega_{\varphi}^{n}
Xe2A(F)Q2p|~F|ωφ2(ΔφF)2ωφn+XQ2p(trωφω)2|Tspr|ωφ2ωφn\displaystyle\leq\int_{X}e^{2A(F)}Q^{2p}|\tilde{\nabla}F|^{2}_{\omega_{\varphi}}({\Delta_{\varphi}}F)^{2}\omega_{\varphi}^{n}+\int_{X}Q^{2p}(\operatorname{tr}_{\omega_{\varphi}}\omega)^{2}|T^{r}_{sp}|^{2}_{\omega_{\varphi}}\omega_{\varphi}^{n}
XeA(F)Q2p+1(ΔφF)2ωφn+XQ2p(trωφω)3|Tspr|ω2ωφn\displaystyle\leq\int_{X}e^{A(F)}Q^{2p+1}({\Delta_{\varphi}}F)^{2}\omega_{\varphi}^{n}+\int_{X}Q^{2p}(\operatorname{tr}_{\omega_{\varphi}}\omega)^{3}|T^{r}_{sp}|^{2}_{\omega}\omega_{\varphi}^{n}
CXQ2p+1(trωωφ)2n2ωφn+CXQ2p+1ωφn\displaystyle\leq C\int_{X}Q^{2p+1}(\operatorname{tr}_{\omega}\omega_{\varphi})^{2n-2}\omega_{\varphi}^{n}+C\int_{X}Q^{2p+1}\omega_{\varphi}^{n}
CXQ2p+1(trωωφ)2n2ωφn,\displaystyle\leq C\int_{X}Q^{2p+1}(\operatorname{tr}_{\omega}\omega_{\varphi})^{2n-2}\omega_{\varphi}^{n},

where we used Young’s inequality in the third line. Now, our bound on the torsion integral is a constant multiple of the bound of the remaining terms as seen in Equation (3.20) of [11]. The remainder of the proof for the LL^{\infty} bound on QQ follows from Moser iteration and several applications of the Hölder and Sobolev inequalities with respect to the reference metric ω\omega, see Section 3 of [11] and Section 4 of [10]. The constants and powers of the trace differ slightly from those in the Kähler case, but will not affect the iteration method. Lastly, to show an L1L^{1} bound on the quantity QQ follows immediately from the fact that |F|ωφ2|\partial F|^{2}_{\omega_{\varphi}} holds the same way as in (4.35) of [10] and the L1L^{1} bound on (trωωφ)B+1(\operatorname{tr}_{\omega}\omega_{\varphi})^{B+1} follows using the LB+1L^{B+1} norm we obtain from the LpL^{p} bound. ∎

Finally, using that we have an upper and lower bound on trωωφ\operatorname{tr}_{\omega}{\omega_{\varphi}} gives us the quasi-isometry of ω\omega and ωφ{\omega}_{\varphi}. Now, going back to the coupled equations

(14) F=logωφnωn\displaystyle F=\log\frac{\omega_{\varphi}^{n}}{\omega^{n}}
(15) ΔφF=\displaystyle\Delta_{\varphi}F= Rφ+trωφRic(ω),\displaystyle-\operatorname{R}_{\varphi}+\operatorname{tr}_{\omega_{\varphi}}\operatorname{Ric}(\omega),

by ellipticity and that the right-hand side of (15) is bounded in LpL^{p} for any p>0p>0, we obtain a W2,pW^{2,p} bound on FF for any p>0p>0 (see Theorem 9.11 in [16]). By Morrey’s inequality, this gives us C1,αC^{1,\alpha} bounds on FF for any α(0,1)\alpha\in(0,1). Using a result in the non-Kähler setting by [30] and the fact that we have CαC^{\alpha} bounds on FF and trωωφ\operatorname{tr}_{\omega}\omega_{\varphi}, we can then obtain a C2,αC^{2,\alpha} estimate on φ\varphi for any α(0,1)\alpha\in(0,1). Differentiating (14), we have that all the coefficients are bounded in CαC^{\alpha} for any α(0,1)\alpha\in(0,1), hence by Schauder estimates, we obtain C3,αC^{3,\alpha} bounds on φ\varphi for any α(0,1)\alpha\in(0,1). This completes the proof of the theorem. ∎

From this, the result of Chen-He for the Calabi flow ([9], Theorem 3.2 and Theorem 3.3) which states that a C3,αC^{3,\alpha} bound on φ\varphi implies bounds on all higher order derivatives of φ\varphi along the flow, can be applied here. The flow under consideration can be represented as

tφ=A(φ,2φ)φ+f(φ,2φ,3φ),\displaystyle\frac{\partial}{\partial t}\varphi=-A(\nabla\varphi,\nabla^{2}\varphi)\varphi+f(\nabla\varphi,\nabla^{2}\varphi,\nabla^{3}\varphi),

where AA is a strictly elliptic fourth-order operator with coefficients depending on first- and second-order derivatives of φ\varphi. The parabolic PDE theory used by Chen-He [9] does not rely on any Kähler assumptions and holds in this setting as well. It follows that

Lemma 7.

If the evolving metric ωφ=ω0+1¯φ\omega_{\varphi}=\omega_{0}+\sqrt{-1}\partial\bar{\partial}\varphi satisfies φC3,α(X,ω0)C||\varphi||_{C^{3,\alpha}(X,\omega_{0})}\leq C then in fact φCk,α(X,ω0)C(k)||\varphi||_{C^{k,\alpha}(X,\omega_{0})}\leq C(k) for all k>3k>3.

We now prove the main result of the paper:

Theorem 3.

Let (X,ω0)(X,\omega_{0}) be a Hermitian manifold with c1BC(X)=0c_{1}^{BC}(X)=0 and ¯ω0k=0\partial\bar{\partial}\omega_{0}^{k}=0 for k=1,2k=1,2. A solution to the flow defined in Equation (1) exists as long as the Chern scalar curvature of the evolving metric remains bounded. In addition, if the Chern scalar curvature remains bounded for all time, then the flow converges smoothly to the unique Chern-Ricci-flat metric in the ¯\partial\bar{\partial}-class of ω0\omega_{0}.

Proof.

Firstly, we know that along the flow, the Mabuchi energy of ωφ\omega_{\varphi} with respect to ω0\omega_{0} is decreasing, in other words,

tMabω0(ωφ)=X(Rφ+2Re(gφjk¯(trTφ)jk¯logωφneFω0n))2ωφn0.\displaystyle\frac{\partial}{\partial t}\operatorname{Mab}_{\omega_{0}}(\omega_{\varphi})=-\int_{X}\Big{(}{\operatorname{R}_{\varphi}}+2\operatorname{Re}(g_{\varphi}^{j\bar{k}}(\operatorname{tr}T_{\varphi})_{j}\partial_{\bar{k}}\log\frac{\omega_{\varphi}^{n}}{e^{F}\omega_{0}^{n}})\Big{)}^{2}\omega_{\varphi}^{n}\leq 0.

In addition, the Mabuchi energy is bounded from below since

Mabω0(ωφ)\displaystyle\operatorname{Mab}_{\omega_{0}}(\omega_{\varphi}) =Xlogωφnω0nωφnXF(ωφnω0n)\displaystyle=\int_{X}\log\frac{\omega_{\varphi}^{n}}{\omega_{0}^{n}}\omega_{\varphi}^{n}-\int_{X}F(\omega_{\varphi}^{n}-\omega_{0}^{n})
CsupX|F|XωφnC\displaystyle\geq-C-\sup_{X}|F|\int_{X}\omega_{\varphi}^{n}-C
C\displaystyle\geq-C

where the first integral is bounded using the fact that the map xxlogxx\mapsto x\log x has a lower bound for x>0x>0. This gives us a uniform bound on the Mabuchi energy along the flow.

Using the assumption that Rφ\operatorname{R}_{\varphi} remains uniformly bounded for all time t[0,)t\in[0,\infty), it follows by Theorem 2 that we have uniform C3,αC^{3,\alpha} estimates on φ\varphi for all time. By the smoothing argument, this gives us bounds on all higher order derivatives of φ\varphi. Combining the fact that we have short-time existence of the flow with these uniform bounds on the potential along the flow gives us long-time existence. Additionally, using these CC^{\infty} estimates on ωφ\omega_{\varphi}, we have that

tXφ˙2ωφn\displaystyle\frac{\partial}{\partial t}\int_{X}\dot{\varphi}^{2}\omega_{\varphi}^{n} =X2φ˙φ¨ωφn+Xφ˙2Δφφ˙ωφnC(Xφ˙2ωφn)1/2.\displaystyle=\int_{X}2\dot{\varphi}\ddot{\varphi}\omega_{\varphi}^{n}+\int_{X}\dot{\varphi}^{2}\Delta_{\varphi}\dot{\varphi}\omega_{\varphi}^{n}\leq C\Big{(}\int_{X}\dot{\varphi}^{2}\omega_{\varphi}^{n}\Big{)}^{1/2}.

Let us denote f(t)=(Xφ˙2ωφ)(t)f(t)=\Big{(}\int_{X}\dot{\varphi}^{2}\omega_{\varphi}\Big{)}(t). The above differential inequality is equivalent to

fCf1/2C(f+δ)1/2f(f+δ)1/2=((f+δ)1/2)C,\displaystyle f^{\prime}\leq Cf^{1/2}\leq C(f+\delta)^{1/2}\ \Rightarrow\ \frac{f^{\prime}}{(f+\delta)^{1/2}}=((f+\delta)^{1/2})^{\prime}\leq C,

for a constant δ>0\delta>0. For t>tt^{\prime}>t, integrating the above equation from tt to tt^{\prime} gives us that

(f(t)+δ)1/2(f(t)+δ)1/2C(tt).\displaystyle(f(t^{\prime})+\delta)^{1/2}-(f(t)+\delta)^{1/2}\leq C(t^{\prime}-t).

Since this holds for every δ>0\delta>0, we may take δ0\delta\rightarrow 0 and rearrange to obtain

f(t)(f(t)1/2+C(tt))2,\displaystyle f(t^{\prime})\leq(f(t)^{1/2}+C(t^{\prime}-t))^{2},

which shows that f(t)f(t) can grow at most quadratically. Since we know that the Mabuchi energy is uniformly bounded and tMabω0(ωφ)=f(t)\frac{\partial}{\partial t}\operatorname{Mab}_{\omega_{0}}(\omega_{\varphi})=-f(t), integration gives us that

0f(t)𝑑t<.\displaystyle\int_{0}^{\infty}f(t)dt<\infty.

Let us assume for contradiction that f(t)f(t) does not converge to 0 for all tt. Then, there exists an ε(0,1)\varepsilon\in(0,1) and a sequence of times (ti)i=1(t_{i})_{i=1}^{\infty}\rightarrow\infty such that f(ti)εf(t_{i})\geq\varepsilon. For each interval [ti1,ti)[t_{i}-1,t_{i}),

ti1tif(t)𝑑t\displaystyle\int_{t_{i}-1}^{t_{i}}f(t)dt max(ti1,{t:ε1/2C(tit)>0})ti(ε1/2C(tit))2𝑑t\displaystyle\geq\int_{\max(t_{i}-1,\{t\ :\ \varepsilon^{1/2}-C(t_{i}-t)>0\})}^{t_{i}}(\varepsilon^{1/2}-C(t_{i}-t))^{2}dt
=max(ti1,tiε1/2C)ti(ε1/2C(tit))2𝑑t\displaystyle=\int_{\max(t_{i}-1,t_{i}-\frac{\varepsilon^{1/2}}{C})}^{t_{i}}(\varepsilon^{1/2}-C(t_{i}-t))^{2}dt
=13C(ε3/2max(0,(ε1/2C)3))\displaystyle=\frac{1}{3C}\big{(}\varepsilon^{3/2}-\max(0,(\varepsilon^{1/2}-C)^{3})\big{)}
=13Cmin(ε3/2,3Cε+3Cε1/2+C3)\displaystyle=\frac{1}{3C}\min(\varepsilon^{3/2},3C\varepsilon+3C\varepsilon^{1/2}+C^{3})
Cε3/2.\displaystyle\geq C\varepsilon^{3/2}.

However, since there are infinitely many such tit_{i} where tit_{i}\rightarrow\infty, the integral of f(t)f(t) cannot be bounded, giving us the desired contradiction.

This implies that f(t)=Xφ˙2ωφn0f(t)=\int_{X}\dot{\varphi}^{2}\omega_{\varphi}^{n}\rightarrow 0 for all tt\rightarrow\infty which implies that φ˙(t)0\dot{\varphi}(t)\rightarrow 0 as tt\rightarrow\infty. Since we have CC^{\infty} estimates on φ\varphi, by the Arzela-Ascoli theorem, there exists a sequence of times (tj)j=1(t_{j})_{j=1}^{\infty} such that φ(tj)φ\varphi(t_{j})\rightarrow\varphi_{\infty} in CC^{\infty}, where φ\varphi_{\infty} is smooth. In fact, since Xφ˙2ωφ=0\int_{X}\dot{\varphi}_{\infty}^{2}\omega_{\varphi_{\infty}}=0, φ˙=0\dot{\varphi}_{\infty}=0 which by Lemma 4 implies that ω=ω0+1¯φ\omega_{\infty}=\omega_{0}+\sqrt{-1}\partial\bar{\partial}\varphi_{\infty} is Chern-Ricci-flat. By the uniqueness of Chern-Ricci-flat metrics in the ¯\partial\bar{\partial}-class of a metric [32], ω\omega_{\infty} is the unique Chern-Ricci-flat metric along the flow. We can show that we indeed have CC^{\infty} convergence without passing to a sequence. To see this, suppose not. Then there exists a kk\in\mathbb{N}, ε>0\varepsilon>0 and a sequence of times (tn)n=1(t_{n})_{n=1}^{\infty} such that for all nn,

φ(tn)φCk(X)ε.\displaystyle\|\varphi(t_{n})-\varphi_{\infty}\|_{C^{k}(X)}\geq\varepsilon.

Since we have Ck+1C^{k+1} bounds on φ(tn)\varphi(t_{n}), the Arzela-Ascoli theorem gives us that there exists a subsequence (tnj)j=1(t_{n_{j}})_{j=1}^{\infty} such that φ(tnj)\varphi(t_{n_{j}}) converges in CkC^{k} to a limit φ\varphi_{\infty}^{\prime}, with

φφCk(X)ε.\displaystyle\|\varphi_{\infty}^{\prime}-\varphi_{\infty}\|_{C^{k}(X)}\geq\varepsilon.

This implies that ωφωφ\omega_{\varphi}^{\prime}\neq\omega_{\varphi}, but by the above argument ωφ\omega_{\varphi}^{\prime} is also Chern-Ricci-flat, and so this contradicts the uniqueness of the Chern-Ricci-flat metric in the ¯\partial\bar{\partial}-class of ω0\omega_{0}. This completes the proof of the theorem.

Acknowledgements

The author is very grateful to her thesis advisor Ben Weinkove for his helpful suggestions and his continued support and encouragement. She would also like to thank Gregory Edwards, Antoine Song and Jonathan Zhu for some useful discussions, as well as the referee for many helpful constructive comments.

References

  • [1] D. Angella, S. Calamai, and C. Spotti, On the Chern-Yamabe problem, Math. Res. Lett. 24 (2017), no. 3, 645–677.
  • [2] L. Bedulli and L. Vezzoni, A parabolic flow of balanced metrics, J. Reine Angew. Math. 723 (2017), 79–99.
  • [3] by same author, A scalar Calabi-type flow in Hermitian Geometry: Short-time existence and stability, Ann. Sc. Norm. Super. Pisa Cl. Sci. XX (2020).
  • [4] R. J. Berman, T. Darvas, and C. H. Lu, Convexity of the extended K-energy and the large time behavior of the weak Calabi flow, Geometry & Topology 21 (2017), 2945–2988.
  • [5] R. Bryant and F. Xu, Laplacian flow for closed G2-structures: short time behavior, arXiv:1101.2004.
  • [6] E. Calabi, Extremal Kähler metrics. In Seminar on Differential Geometry, Ann. of Math. Stud. Princeton Univ. Press, Princeton, NJ. 102 (1982), 259–290.
  • [7] by same author, Extremal Kähler Metrics II. In Differential Geometry and Complex Analysis, pp. 95–114, Springer-Verlag, New York, 1985.
  • [8] S.-C. Chang, The 2-dimensional Calabi flow, Nagoya Math. J. 181 (2006), 63–73.
  • [9] X. Chen and W. Y He, On the Calabi flow, Amer. J. Math. 130 (2008), 539–570.
  • [10] X. X. Chen and J. Cheng, On the constant scalar curvature Kähler metrics, apriori estimates, arXiv:1712.06697 (2017).
  • [11] by same author, On the constant scalar curvature Kähler metrics, existence results, arXiv:1801.00656 (2018).
  • [12] X.X. Chen and S. Sun, Calabi flow, Geodesic rays, and uniqueness of constant scalar curvature Kähler metrics, Annals of Mathematics 180 (2014), 407–454.
  • [13] P. T. Chruściel, Semi-global existence and convergence of solutions of the Robinson-Trautman (2-dimensional Calabi) equation, Comm. Math. Phys. 137 (1991), 289–313.
  • [14] R. Feng and H. Huang, The global existence and convergence of the Calabi flow on n/n+i𝕏n\mathbb{C}^{n}/\mathbb{Z}^{n}+i\mathbb{X}^{n}, J, Funct. Anal. 263 (2012), 1129–1146.
  • [15] J. Fine, Calabi flow and projective embeddings, J. Differential Geom. 84 (2010), no. 3, 489–523.
  • [16] D. Gilbarg and N. S. Trudinger, Elliptic partial differential equations of second order, Berlin: Springer-Verlag, 1983.
  • [17] M. Gill, Convergence of the parabolic complex Monge-Ampère equation on compact Hermitian manifolds, Comm. Anal. Geom. 19 (2011), no. 1, 65–78.
  • [18] W. He, On the convergence of the Calabi flow, Proc. Amer. Math. Soc. 143 (2015), 1273–1281.
  • [19] H. Huang and K. Zheng, Stability of the Calabi flow near an extremal metric, Ann. Sc. Norm. Super. Pisa Cl. Sci. 11 (2012), 167–175.
  • [20] M. Kawamura, A scalar Calabi-type flow in the almost Hermitian geometry, Tsukuba J. Math. 43 (2019), no. 1, 37–54.
  • [21] H. Li, B. Wang, and K. Zheng, Regularity scales and convergence of the Calabi flow, J. Geom. Anal. 28 (2018), 2050–2101.
  • [22] D. H. Phong, S. Picard, and X. Zhang, New curvature flows in complex geometry, Surveys in Differential Geometry 22 (2017), 331–364.
  • [23] S. Picard, Calabi-Yau Manifolds with Torsion and Geometric Flows, CIME Summer School, http://people.math.harvard.edu/\simspicard/cetraro.pdf (2018).
  • [24] X. S. Shen, Estimates for metrics of constant Chern scalar curvature, arXiv preprint arXiv: 1909.13445 (2019).
  • [25] J. Streets, Long time existence of minimizing movement solutions of Calabi flow, Adv. in Math. 259 (2014), 688–729.
  • [26] by same author, The consistency and convergence of K-energy minimizing movements, Trans. Amer. Math. Soc. 368 (2016), 5075–5091.
  • [27] J. Streets and G. Tian, Hermitian curvature flow, J. Eur. Math. Soc. 13 (2011), 601–634.
  • [28] G. Székelyhidi, Remark on the Calabi flow with bounded curvature, Univ. Iagel. Acta Math. 50 (2013), 107–115.
  • [29] G. Tian, Canonical metrics in Kähler geometry, Notes taken by Meike Akveld. Lectures in Mathematics ETH Zürich, Birkhäuser Verlag, Basel, 2000.
  • [30] V. Tosatti, Y. Wang, B. Weinkove, and X. Yang, C2,αC^{2,\alpha} estimates for nonlinear elliptic equations in complex and almost complex geometry, Calc. Var. Partial Differential Equations 54 (2015), no. 1, 431–453.
  • [31] V. Tosatti and B. Weinkove, The Calabi flow with small initial energy, Math. Res. Lett. 40 (2007), 1033–1039.
  • [32] by same author, The complex Monge-Ampère equation on compact Hermitian manifolds, J. Amer. Math. Soc. 23 (2010), 1187–1195.
  • [33] by same author, The Chern-Ricci flow on complex surfaces, Comp. Math. 149 (2013), 2101–2138.
  • [34] by same author, On the evolution of a Hermitian metric by its Chern-Ricci form, J. Differential Geom. 99 (2015), no. 1, 125–163.
  • [35] Y. Ustinovskiy, Hermitian curvature flow on manifolds with non-negative Griffiths curvature, Amer. J. Math. 141 (2019), no. 6, 1751–1775.