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

Entropy and Heat Kernel on Generalized Ricci Flow

Xilun Li SMS, Peking University, Beijing 100871, China [email protected]
Abstract.

We introduce analogous geometric quantities and prove some geometric and analytic bounds in [Bam20a] to generalized Ricci flow.

1. Introduction

Generalized Ricci flow is a triple (M,gt,Ht)(M,g_{t},H_{t}), where MM is a smooth manifold, gtg_{t}, HtH_{t} is a one-parameter family of metrics and closed three-forms on MM, which satisfies the equation

(1.1) {tg=2Ric+12H2,tH=ddgH,\displaystyle\left\{\begin{aligned} \frac{\partial}{\partial t}g&=-2\mathrm{Ric}+\frac{1}{2}H^{2},\\ \frac{\partial}{\partial t}H&=-dd^{*}_{g}H,\end{aligned}\right.

where H2(X,Y):=iXH,iYHH^{2}(X,Y):=\left<i_{X}H,i_{Y}H\right>.

This flow arises in many aspects, such as generalized geometry [GFS21], complex geometry and mathematical physics [Str22a]. In complex geometry, G. Tian and J. Streets introduce the pluriclosed flow in [ST10] to study the geometry of non-Kähler manifold as extensions of the Kähler Ricci flow. Let (M2n,J,ω)(M^{2n},J,\omega) be a complex manifold, the pluriclosed flow is

(1.2) tω=ρB1,1,\frac{\partial}{\partial t}\omega=-\rho_{B}^{1,1},

where ρB\rho_{B} is the Ricci form of Bismut connection. In [ST13], regarding MM as a real manifold, pluriclosed flow is just the generalized Ricci flow up to a smooth family of diffeomorphism induced by the Lee form θ=Jdω\theta=-Jd^{*}\omega, where the 3-form HH is the torsion of Bismut connection H=dcω=i(¯)ωH=d^{c}\omega=i(\bar{\partial}-\partial)\omega. Thus the closeness of HH is equivalent to the pluriclosed condition, i.e. ¯ω=0\partial\bar{\partial}\omega=0.

The main motivation of pluriclosed flow is to provide a geometric-analytic approach to classify the non-Kähler complex surfaces, especially the class VII surfaces, as what Ricci flow does in Geometrization Conjectures and Kähler-Ricci flow does in Kähler manifolds. The pluriclosed flow preserves the pluriclosed condition. According to Gauduchon’s results[Gau77], each complex surfaces admit a pluriclosed metric. For this reason, we should mainly concern about the flows in complex dimension 2, i.e. real dimension 4.

A symmetry reduction of the generalized Ricci flow will give another interesting flow, which we called Ricci-Yang-Mills flow, see [Str22a]. When the manifold is the total space of a principal bundle, the phenomenon of the flow will become more clear. The Ricci-Yang-Mills flow not only is a symmetric solution of the generalized Ricci flow, but also has arisen in mathematical physics literature and as a tool to understand the geometry of principal bundles.

The classical Ricci flow is just the special case where H0H\equiv 0. In general, since H2(X,X)=|iXH|20H^{2}(X,X)=|i_{X}H|^{2}\geqslant 0, the generalized Ricci flow is actually a kind of super Ricci flow, i.e.

(1.3) tg2Ric.\frac{\partial}{\partial t}g\geqslant-2\mathrm{Ric}.

R. H. Bamler developed a compactness and regularity theory of super Ricci flow in [Bam20a, Bam20b, Bam20c], which states that any sequence of pointed super Ricci flows of bounded dimensions subsequentially converges to metric flow in a certain sense, which he called the 𝔽\mathbb{F}-limit. Moreover, the noncollapsed 𝔽\mathbb{F}-limits of Ricci flows is smooth away a set of codimension at least 4 in parabolic sense. Since the super Ricci flow is a quite big class of flows, it seems unlikely to get structure theory for general super Ricci flow. However, for some particular cases, such as generalized Ricci flow, it’s hopeful to have similar structure results as Ricci flow. A key step in the study of pluriclosed flow is the existence conjecture, see [ST13, Conjecture 5.2]. The structure theory, if exists, will be useful to deal with the conjecture.

In this paper, we will prove some theorems on generalized Ricci flow, which are analogous to the results on Ricci flow in [Bam20a]. The paper will be organized as follows.

In Section 2, we introduce the generalized definition of the pointed Nash and Perelman entropy, which has monotonicity along generalized Ricci flow. Then we will show that the lower bound of generalized point Nash entropy implies the weighted noncollapsed condition.

In Section 3, we will show the upper bounds of the heat kernel and its gradient.

In Section 4, we show an ε\varepsilon-regularity theorem, which roughly states that the generalized Nash entropy is close to that in Euclidean implies the geometry is close to Euclidean.

In Section 5, we show some estimates on HH, particularly in lower dimensional cases, i.e. n4n\leqslant 4.

Acknowledgements. I am grateful to my advisor Professor Gang Tian for his helpful guidance. I thank Yanan Ye and Shengxuan Zhou for inspiring discussions.

2. Generalized Entropy

2.1. Generalized Nash Entropy and Perelman Entropy

J. Streets[Str22b] introduced several concepts in generalized Ricci flow, we list them here for convenience.

Definition 2.1 ([Str22b, Definition 2.2]).

Given (M,gt,Ht)(M,g_{t},H_{t}) a solution to generalized Ricci flow, a one-parameter family ϕt\phi_{t} is called the dilaton flow if

(2.1) ϕ=16|H|2,\square\phi=\frac{1}{6}|H|^{2},

where :=tΔg(t)\square:=\partial_{t}-\Delta_{g(t)} is the heat operator coupled with the flow.

Definition 2.2 ([Str22b, Definition 2.1]).

Given a smooth manifold MM, a triple (g,H,f)(g,H,f) of a Riemannian metric, closed three-form and function, generalized scalar curvature is

(2.2) RH,f:=R112|H|2+2Δf|f|2,R^{H,f}:=R-\frac{1}{12}|H|^{2}+2\Delta f-|\nabla f|^{2},

generalized Ricci curvature, or twisted Bakry-Emery curvature is

(2.3) RicH,f:=Ric14H2+2f12(dgH+ifH).\mathrm{Ric}^{H,f}:=\mathrm{Ric}-\frac{1}{4}H^{2}+\nabla^{2}f-\frac{1}{2}(d^{*}_{g}H+i_{\nabla f}H).
Proposition 2.3 ([Str22b, Proposition 2.4]).

Let (M,g,H)(M,g,H) be the generalized Ricci flow, ϕt\phi_{t} dilaton flow,

(2.4) RH,ϕ=2|RicH,ϕ|2.\square R^{H,\phi}=2|\mathrm{Ric}^{H,\phi}|^{2}.
Remark 2.4.

Note that the generalized scalar curvature is not equal to the trace of generalized Ricci curvature, so it may not be bounded from below by a constant only depends on time and dimension as Ricci flow.

Corollary 2.5.

Let (M,g,H)(M,g,H) be the generalized Ricci flow, ϕt\phi_{t} dilaton flow,

(2.5) 112|H|2+|ϕ|2R+2ΔϕR0ϕ,\frac{1}{12}|H|^{2}+|\nabla\phi|^{2}\leqslant R+2\Delta\phi-R^{\phi}_{0},

where R0ϕ:=minxRH,ϕ(x,0)R^{\phi}_{0}:=\min_{x}R^{H,\phi}(x,0). Integrate the inequality, we have

(2.6) M(112|H|2+|ϕ|2)dgtM(RR0ϕ)dgt.\int_{M}\left(\frac{1}{12}|H|^{2}+|\nabla\phi|^{2}\right)\mathrm{d}g_{t}\leqslant\int_{M}\left(R-R^{\phi}_{0}\right)\mathrm{d}g_{t}.
Proof.

This follows from maximum principle applied to (2.4). ∎

Definition 2.6 ([Str22b, Definition 3.1]).

Let (M,gt,Ht,ϕt)(M,g_{t},H_{t},\phi_{t}) be a solution to generalized Ricci flow. Define the conjuate heat operator

:=tΔg(t)+R14|H|2.\square^{*}:=-\partial_{t}-\Delta_{g(t)}+R-\frac{1}{4}|H|^{2}.

Also define the weighted conjugate heat operator

ϕu:=eϕ(ueϕ)=tuΔg(t)u+2ϕ,u+RH,ϕu.\square_{\phi}^{*}u:=e^{\phi}\square^{*}\left(ue^{-\phi}\right)=-\partial_{t}u-\Delta_{g(t)}u+2\left<\nabla\phi,\nabla u\right>+R^{H,\phi}u.

These are conjugate operators in the following sense:

t1t2M(u)vdgtdt\displaystyle\int_{t_{1}}^{t_{2}}\int_{M}(\square u)v\mathrm{d}g_{t}\mathrm{d}t =Muvdgt|t1t2+t1t2Mu(v)dgtdt,\displaystyle=\left.\int_{M}uv\mathrm{d}g_{t}\right|_{t_{1}}^{t_{2}}+\int_{t_{1}}^{t_{2}}\int_{M}u(\square^{*}v)\mathrm{d}g_{t}\mathrm{d}t,
t1t2M(u)veϕdgtdt\displaystyle\int_{t_{1}}^{t_{2}}\int_{M}(\square u)ve^{-\phi}\mathrm{d}g_{t}\mathrm{d}t =Muveϕdgt|t1t2+t1t2Mu(v)eϕdgtdt.\displaystyle=\left.\int_{M}uve^{-\phi}\mathrm{d}g_{t}\right|_{t_{1}}^{t_{2}}+\int_{t_{1}}^{t_{2}}\int_{M}u(\square^{*}v)e^{-\phi}\mathrm{d}g_{t}\mathrm{d}t.
Definition 2.7 ([Str22b, Definition 4.1]).

Assume v=(4πτ)n/2efeϕv=(4\pi\tau)^{-n/2}e^{-f}e^{-\phi}, dν=vdg\mathrm{d}\nu=v\mathrm{d}g, Mdν=1\int_{M}\mathrm{d}\nu=1, define the weighted Nash entropy by

𝒩ϕ[g,f,τ]:=fdνn2.\mathcal{N}^{\phi}[g,f,\tau]:=\int f\mathrm{d}\nu-\frac{n}{2}.

And define the weighted Perelman entropy by

𝒲ϕ[g,f,τ]:=(τRf+ϕ+fn)dν=[τ(Rϕ+|f|2)+fn]dν.\mathcal{W}^{\phi}[g,f,\tau]:=\int(\tau R^{f+\phi}+f-n)\mathrm{d}\nu=\int[\tau(R^{\phi}+|\nabla f|^{2})+f-n]\mathrm{d}\nu.
Proposition 2.8.

Suppose v=0\square^{*}v=0, ddtτ=1\frac{\mathrm{d}}{\mathrm{d}t}\tau=-1, (M,g,H,ϕ)(M,g,H,\phi) is generalized Ricci flow, then

ddτ(τ𝒩ϕ)=𝒲ϕ,ddτ𝒲ϕ=2τ|Ricf+ϕ12τg|2dν+13|H|2dν.\frac{d}{d\tau}(\tau\mathcal{N}^{\phi})=\mathcal{W}^{\phi},\ \ \ \frac{d}{d\tau}\mathcal{W}^{\phi}=-2\tau\int|\mathrm{Ric}^{f+\phi}-\frac{1}{2\tau}g|^{2}\mathrm{d}\nu+\frac{1}{3}\int|H|^{2}\mathrm{d}\nu.
Proof.

Since f=logvϕn2log(4πτ)f=-\log v-\phi-\frac{n}{2}\log(4\pi\tau), v=0\square^{*}v=0, then

f=2v1Δv|(f+ϕ)|2R+112|H|2+n2τ,\square f=2v^{-1}\Delta v-|\nabla(f+\phi)|^{2}-R+\frac{1}{12}|H|^{2}+\frac{n}{2\tau},
ddt𝒩=fdν=(|(f+ϕ)|2+R112|H|2n2τ)dν.\frac{\mathrm{d}}{\mathrm{d}t}\mathcal{N}=\int\square f\mathrm{d}\nu=-\int\left(|\nabla(f+\phi)|^{2}+R-\frac{1}{12}|H|^{2}-\frac{n}{2\tau}\right)\mathrm{d}\nu.

Note that integration by part, we have |(f+ϕ)|2dν=Δ(f+ϕ)dν\int|\nabla(f+\phi)|^{2}\mathrm{d}\nu=\int\Delta(f+\phi)\mathrm{d}\nu, then

ddτ(τ𝒩)=𝒩τddt𝒩=[τ(R112|H|2+|(f+ϕ)|2)+fn]dν=𝒲.\frac{d}{d\tau}(\tau\mathcal{N})=\mathcal{N}-\tau\frac{\mathrm{d}}{\mathrm{d}t}\mathcal{N}=\int\left[\tau\left(R-\frac{1}{12}|H|^{2}+|\nabla(f+\phi)|^{2}\right)+f-n\right]\mathrm{d}\nu=\mathcal{W}.

Let u:=eϕvu:=e^{\phi}v, then ϕu=eϕ(ueϕ)=0\square^{*}_{\phi}u=e^{\phi}\square^{*}(ue^{-\phi})=0. Denote by WH,F:=τRH,F+FnW^{H,F}:=\tau R^{H,F}+F-n, then by the local computation in [Str22b, Proposition 3.12],

ddτ𝒲\displaystyle\frac{d}{d\tau}\mathcal{W} =ddτ(Wϕ)ueϕdg=(ϕ(Wu)+ϕuϕϕu)eϕdg\displaystyle=\frac{d}{d\tau}\int(W-\phi)ue^{-\phi}\mathrm{d}g=\int\left(\square^{*}_{\phi}(Wu)+\square\phi u-\phi\square^{*}_{\phi}u\right)e^{-\phi}\mathrm{d}g
=2τ|Ricf+ϕ12τg|2dν+13|H|2dν.\displaystyle=-2\tau\int|\mathrm{Ric}^{f+\phi}-\frac{1}{2\tau}g|^{2}\mathrm{d}\nu+\frac{1}{3}\int|H|^{2}\mathrm{d}\nu.

Proposition 2.9 ([Gue02, Theorem 2.1]).

There exists the unique heat kernel K(x,t;y,s)C({x,yM,t>s};+)K(x,t;y,s)\in C^{\infty}(\{x,y\in M,t>s\};\mathbb{R}_{+}) satisfies

{x,tK(x,t;y,s)=0,limtsK(x,t;y,s)=δy(x).\displaystyle\left\{\begin{aligned} &\square_{x,t}K(x,t;y,s)=0,\\ &\lim_{t\searrow s}K(x,t;y,s)=\delta_{y}(x).\end{aligned}\right.

Moreover, the conjugate heat kernel is exactly the same function:

{y,sK(x,t;y,s)=0,limstK(x,t;y,s)=δx(y).\displaystyle\left\{\begin{aligned} &\square^{*}_{y,s}K(x,t;y,s)=0,\\ &\lim_{s\nearrow t}K(x,t;y,s)=\delta_{x}(y).\end{aligned}\right.

As in [Bam20a, Definition 5.1], we can define the pointed entropy, i.e. taking the conjugate heat kernel as the function vv :

Definition 2.10.

Given (x,t)(x,t), dνx,t;s(y):=K(x,t;y,s)dgs(y)=(4πτ)n2e(f+ϕ)dgs\mathrm{d}\nu_{x,t;s}(y):=K(x,t;y,s)\mathrm{d}g_{s}(y)=(4\pi\tau)^{-\frac{n}{2}}e^{-(f+\phi)}\mathrm{d}g_{s}, the pointed Nash entropy is defined by

𝒩x,tϕ(τ):=𝒩ϕ[gtτ,ftτ,τ],𝒩sϕ(x,t):=𝒩x,tϕ(ts),\mathcal{N}^{\phi}_{x,t}(\tau):=\mathcal{N}^{\phi}[g_{t-\tau},f_{t-\tau},\tau],\ \ \mathcal{N}^{\phi*}_{s}(x,t):=\mathcal{N}^{\phi}_{x,t}(t-s),

the pointed Perelman entropy is defined by

𝒲x,tϕ(τ):=𝒲ϕ[gtτ,ftτ,τ].\mathcal{W}^{\phi}_{x,t}(\tau):=\mathcal{W}^{\phi}[g_{t-\tau},f_{t-\tau},\tau].

The pointed entropy is not monotone due to the existence of the torsion HH, we need to modify the definition of entropy to get the monotonicity.

Definition 2.11.

Define

Ψx,t(τ):=tτtM|H|2dνx,t;sds,Px,t(τ):=0τs1Ψx,t(s)ds.\Psi_{x,t}(\tau):=\int_{t-\tau}^{t}\int_{M}|H|^{2}\mathrm{d}\nu_{x,t;s}\mathrm{d}s,\ \ P_{x,t}(\tau):=\int_{0}^{\tau}s^{-1}\Psi_{x,t}(s)\mathrm{d}s.

Then we define the generalized pointed Nash entropy by

𝒩x,tH(τ):=𝒩x,tϕ(τ)13Px,t(τ).\mathcal{N}^{H}_{x,t}(\tau):=\mathcal{N}^{\phi}_{x,t}(\tau)-\frac{1}{3}P_{x,t}(\tau).

And define the generalized pointed Perelman entropy by

𝒲x,tH(τ):=𝒲x,tϕ(τ)13(Ψx,t(τ)+Px,t(τ)).\mathcal{W}^{H}_{x,t}(\tau):=\mathcal{W}^{\phi}_{x,t}(\tau)-\frac{1}{3}\left(\Psi_{x,t}(\tau)+P_{x,t}(\tau)\right).
Remark 2.12.

Note that Ψx,t(0)=0\Psi_{x,t}(0)=0, Ψx,t(0)=|H|2(x,t)\Psi^{\prime}_{x,t}(0)=|H|^{2}(x,t), so the function Px,tP_{x,t} is well-defined.

Proposition 2.13.

For the generalized pointed entropy, we have

ddτ(τ𝒩x,tH(τ))=𝒲x,tH(τ),ddτ𝒲x,tH(τ)=2τM|Ricf+ϕ12τg|2dν13τΨx,t(τ)0.\frac{\mathrm{d}}{\mathrm{d}\tau}(\tau\mathcal{N}^{H}_{x,t}(\tau))=\mathcal{W}^{H}_{x,t}(\tau),\ \ \frac{d}{d\tau}\mathcal{W}^{H}_{x,t}(\tau)=-2\tau\int_{M}|\mathrm{Ric}^{f+\phi}-\frac{1}{2\tau}g|^{2}\mathrm{d}\nu-\frac{1}{3\tau}\Psi_{x,t}(\tau)\leqslant 0.
Proof.
ddτ(τ𝒩x,tH(τ))=ddτ(τ𝒩x,tϕ(τ))13ddτ(τPx,t(τ))=𝒲x,tϕ(τ)13(Px,t(τ)+Ψx,t(τ)).\frac{\mathrm{d}}{\mathrm{d}\tau}(\tau\mathcal{N}^{H}_{x,t}(\tau))=\frac{\mathrm{d}}{\mathrm{d}\tau}(\tau\mathcal{N}^{\phi}_{x,t}(\tau))-\frac{1}{3}\frac{\mathrm{d}}{\mathrm{d}\tau}(\tau P_{x,t}(\tau))=\mathcal{W}^{\phi}_{x,t}(\tau)-\frac{1}{3}\left(P_{x,t}(\tau)+\Psi_{x,t}(\tau)\right).
ddτ𝒲x,tH(τ)\displaystyle\frac{\mathrm{d}}{\mathrm{d}\tau}\mathcal{W}^{H}_{x,t}(\tau) =ddτ𝒲x,tϕ(τ)13ddτ(Px,t(τ)+Ψx,t(τ))\displaystyle=\frac{\mathrm{d}}{\mathrm{d}\tau}\mathcal{W}^{\phi}_{x,t}(\tau)-\frac{1}{3}\frac{\mathrm{d}}{\mathrm{d}\tau}\left(P_{x,t}(\tau)+\Psi_{x,t}(\tau)\right)
=2τM|Ricf+ϕ12τg|2dνx,t;tτ+13M|H|2dν13(M|H|2dν+1τΨx,t(τ))\displaystyle=-2\tau\int_{M}|\mathrm{Ric}^{f+\phi}-\frac{1}{2\tau}g|^{2}\mathrm{d}\nu_{x,t;t-\tau}+\frac{1}{3}\int_{M}|H|^{2}\mathrm{d}\nu-\frac{1}{3}\left(\int_{M}|H|^{2}\mathrm{d}\nu+\frac{1}{\tau}\Psi_{x,t}(\tau)\right)
=2τM|Ricf+ϕ12τg|2dνx,t;tτ13τΨx,t(τ)0.\displaystyle=-2\tau\int_{M}|\mathrm{Ric}^{f+\phi}-\frac{1}{2\tau}g|^{2}\mathrm{d}\nu_{x,t;t-\tau}-\frac{1}{3\tau}\Psi_{x,t}(\tau)\leqslant 0.

Remark 2.14.

Perelman entropy in Ricci flow becomes constant implies gg is a gradient shrinking soliton. Our generalized pointed Perelman entropy becomes constant implies the flow is actually Ricci flow and gg is also gradient shrinking soliton. It’s natural since by [GFS21, Proposition 4.28], the generalized shrinking soliton is just the shrinking soliton with H0H\equiv 0.

Corollary 2.15.

We have

𝒩x,tH(τ)=1τ0τ𝒲x,tH(s)ds𝒲x,tH(τ),ddτ𝒩x,tH(τ)0,\mathcal{N}^{H}_{x,t}(\tau)=\frac{1}{\tau}\int^{\tau}_{0}\mathcal{W}^{H}_{x,t}(s)\mathrm{d}s\geqslant\mathcal{W}^{H}_{x,t}(\tau),\ \ \frac{\mathrm{d}}{\mathrm{d}\tau}\mathcal{N}^{H}_{x,t}(\tau)\leqslant 0,
Mτ(Rϕ+|f|2)dνtτn2+13Ψx,t(τ).\int_{M}\tau(R^{\phi}+|\nabla f|^{2})\mathrm{d}\nu_{t-\tau}\leqslant\frac{n}{2}+\frac{1}{3}\Psi_{x,t}(\tau).
Proof.

The first follows from τddτ𝒩H=𝒲H𝒩H0\tau\frac{\mathrm{d}}{\mathrm{d}\tau}\mathcal{N}^{H}=\mathcal{W}^{H}-\mathcal{N}^{H}\leqslant 0. The second conclusion follows from 𝒲H=τ(Rϕ+|f|2)dν+𝒩Hn213Ψx,t(τ)\mathcal{W}^{H}=\tau\int(R^{\phi}+|\nabla f|^{2})\mathrm{d}\nu+\mathcal{N}^{H}-\frac{n}{2}-\frac{1}{3}\Psi_{x,t}(\tau). ∎

We need an oscillation estimation of Nash entropy analogous to [Bam20a, Corollary 5.11], which can be divided by two parts to estimate. The first term, weighted entropy, can be dealt as [Bam20a, Theorem 5.9].

Proposition 2.16.

Suppose Rϕ(,s)RminϕR^{\phi}(\cdot,s)\geqslant R^{\phi}_{\min}, then

|𝒩sϕ|(n2(ts)Rminϕ+13(ts)Ψx,t(ts))12,n2(ts)𝒩sϕ0.|\nabla\mathcal{N}^{\phi*}_{s}|\leqslant\left(\frac{n}{2(t-s)}-R^{\phi}_{\min}+\frac{1}{3(t-s)}\Psi_{x,t}(t-s)\right)^{\frac{1}{2}},\ \ \ -\frac{n}{2(t-s)}\leqslant\square\mathcal{N}^{\phi*}_{s}\leqslant 0.
Proof.

The proof follows the argument in [Bam20a, Proof of Theorem 5.9] with a modification. We assume s=0s=0 after application of a time-shift.

𝒩0ϕ(x,t)=K(x,t;y,0)(logK(x,t;y,0)+ϕ(y,0))dg0(y)n2log(4πt)n2.\mathcal{N}^{\phi*}_{0}(x,t)=-\int K(x,t;y,0)(\log K(x,t;y,0)+\phi(y,0))\mathrm{d}g_{0}(y)-\frac{n}{2}\log(4\pi t)-\frac{n}{2}.

For any vector vTxMv\in T_{x}M with |v|t=1|v|_{t}=1,

v𝒩0ϕ(x,t)\displaystyle\partial_{v}\mathcal{N}^{\phi*}_{0}(x,t) =vK(x,t;y,0)(logK(x,t;y,0)+1+ϕ(y,0))dg0(y)\displaystyle=-\int\partial_{v}K(x,t;y,0)(\log K(x,t;y,0)+1+\phi(y,0))\mathrm{d}g_{0}(y)
=vK(x,t;y,0)(fn2𝒩0ϕ(x,t))dg0(y)\displaystyle=\int\partial_{v}K(x,t;y,0)\left(f-\frac{n}{2}-\mathcal{N}^{\phi*}_{0}(x,t)\right)\mathrm{d}g_{0}(y)
=vK(x,t;y,0)K(x,t;y,0)(fn2𝒩0ϕ(x,t))dνx,t;0(y).\displaystyle=\int\frac{\partial_{v}K(x,t;y,0)}{K(x,t;y,0)}\left(f-\frac{n}{2}-\mathcal{N}^{\phi*}_{0}(x,t)\right)\mathrm{d}\nu_{x,t;0}(y).
(fn2𝒩0ϕ(x,t))2dν(y)2t|f|2dνn2tRminϕ+23Ψx,t(t).\int\left(f-\frac{n}{2}-\mathcal{N}^{\phi*}_{0}(x,t)\right)^{2}\mathrm{d}\nu(y)\leqslant 2t\int|\nabla f|^{2}\mathrm{d}\nu\leqslant n-2tR^{\phi}_{\min}+\frac{2}{3}\Psi_{x,t}(t).
|v𝒩0ϕ|2(x,t)12t(n2tRminϕ+23Ψx,t(t)).|\partial_{v}\mathcal{N}^{\phi*}_{0}|^{2}(x,t)\leqslant\frac{1}{2t}\left(n-2tR^{\phi}_{\min}+\frac{2}{3}\Psi_{x,t}(t)\right).

For the last two inequality, we use the L2L^{2}-Poincaré inequality in [HN14, Theorem 1.10] and the integral bounds on the gradient of the heat kernel in [Bam20a, Proposition 4.2]. Both of them hold for super Ricci flow.

The proof of the second result is same as that in [Bam20a].

𝒩0ϕ(x,t)\displaystyle\square\mathcal{N}^{\phi*}_{0}(x,t) =Mx,t(K(x,t;y,0)(logK(x,t;y,0)+ϕ(y,0)))dg0(y)n2t\displaystyle=-\int_{M}\square_{x,t}\left(K(x,t;y,0)(\log K(x,t;y,0)+\phi(y,0))\right)\mathrm{d}g_{0}(y)-\frac{n}{2t}
=M(x,tK(logK+ϕ(y,0)+1)|xK(x,t;y,0)|2K(x,t;y,0))dg0(y)n2t\displaystyle=-\int_{M}\left(\square_{x,t}K(\log K+\phi(y,0)+1)-\frac{|\nabla_{x}K(x,t;y,0)|^{2}}{K(x,t;y,0)}\right)\mathrm{d}g_{0}(y)-\frac{n}{2t}
=M|xK(x,t;y,0)|2K(x,t;y,0)dg0(y)n2t.\displaystyle=\int_{M}\frac{|\nabla_{x}K(x,t;y,0)|^{2}}{K(x,t;y,0)}\mathrm{d}g_{0}(y)-\frac{n}{2t}.

Then the second results also follow from the integral bounds of the heat kernel, see [Bam20a, Proposition 4.2]. ∎

Definition 2.17.

Given two probability measure μ1,μ2\mu_{1},\mu_{2} on (M,g)(M,g), define the W1W_{1}-Wasserstein distance by

dW1g(μ1,μ2):=sup{Mfd(μ1μ2):fC(M),|f|1}.d_{W_{1}}^{g}(\mu_{1},\mu_{2}):=\sup\left\{\int_{M}f\mathrm{d}(\mu_{1}-\mu_{2}):f\in C^{\infty}(M),|\nabla f|\leqslant 1\right\}.
Lemma 2.18.

Let (M,g(t),H(t))(M,g(t),H(t)) be a generalized Ricci flow, x1,x2Mx_{1},x_{2}\in M, then

tdW1g(t)(νx1,t1(t),νx2,t2(t))\displaystyle t\mapsto d_{W_{1}}^{g(t)}(\nu_{x_{1},t_{1}}(t),\nu_{x_{2},t_{2}}(t))

is non-decreasing for tt1,t2t\leqslant t_{1},t_{2}.

Proof.

Note that the generalized Ricci flow is a super Ricci flow, then it follows from [Bam20a, Lemma 2.7]. ∎

Corollary 2.19.

Suppose Rϕ(,s)RminϕR^{\phi}(\cdot,s)\geqslant R^{\phi}_{\min}, |H|2A|H|^{2}\leqslant A on t[s,t]t\in[s,t^{*}], s<tt1,t2s<t^{*}\leqslant t_{1},t_{2}, then for x1,x2Mx_{1},x_{2}\in M

𝒩sϕ(x1,t1)𝒩sϕ(x2,t2)(n2(ts)Rminϕ+13A)12dW1t(νx1,t1(t),νx2,t2(t))+n2log(t2sts).\mathcal{N}^{\phi*}_{s}(x_{1},t_{1})-\mathcal{N}^{\phi*}_{s}(x_{2},t_{2})\leqslant\left(\frac{n}{2(t^{*}-s)}-R^{\phi}_{\min}+\frac{1}{3}A\right)^{\frac{1}{2}}d_{W_{1}}^{t^{*}}\left(\nu_{x_{1},t_{1}}(t^{*}),\nu_{x_{2},t_{2}}(t^{*})\right)+\frac{n}{2}\log\left(\frac{t_{2}-s}{t^{*}-s}\right).
Proof.

It follows from Proposition 2.16 with same argument as [Bam20a, Proof of Corollary 5.11]. ∎

We can get a gradient bound on PsP^{*}_{s} by direct computation.

Proposition 2.20.

Suppose |H|2A|H|^{2}\leqslant A, Ps(x,t):=Px,t(ts)P^{*}_{s}(x,t):=P_{x,t}(t-s), we have |Ps(x,t)|CnAts|\nabla P^{*}_{s}(x,t)|\leqslant C_{n}A\sqrt{t-s}.

Proof.

By changing the order of integration, we have

Ptτ(x,t)\displaystyle P^{*}_{t-\tau}(x,t) =0τs1Ψx,t(s)ds=0τ0sMs1|H|2dνtrdrds\displaystyle=\int_{0}^{\tau}s^{-1}\Psi_{x,t}(s)\mathrm{d}s=\int_{0}^{\tau}\int_{0}^{s}\int_{M}s^{-1}|H|^{2}\mathrm{d}\nu_{t-r}\mathrm{d}r\mathrm{d}s
=0τrτMs1|H|2dνtrdsdr=0τlog(τs)M|H|2dνtsds.\displaystyle=\int_{0}^{\tau}\int_{r}^{\tau}\int_{M}s^{-1}|H|^{2}\mathrm{d}\nu_{t-r}\mathrm{d}s\mathrm{d}r=\int_{0}^{\tau}\log\left(\frac{\tau}{s}\right)\int_{M}|H|^{2}\mathrm{d}\nu_{t-s}\mathrm{d}s.
|Ptτ|\displaystyle\left|\nabla P^{*}_{t-\tau}\right| 0τlog(τs)M|H|2|K(x,t;y,ts)|K(x,t;y,ts)dνts(y)ds\displaystyle\leqslant\int_{0}^{\tau}\log\left(\frac{\tau}{s}\right)\int_{M}|H|^{2}\frac{\left|\nabla K(x,t;y,t-s)\right|}{K(x,t;y,t-s)}\mathrm{d}\nu_{t-s}(y)\mathrm{d}s
CnA0τs12log(τs)ds=CnAτ.\displaystyle\leqslant C_{n}A\int_{0}^{\tau}s^{-\frac{1}{2}}\log\left(\frac{\tau}{s}\right)\mathrm{d}s=C_{n}A\sqrt{\tau}.

where we use integral bounds on the gradient of heat kernel along super Ricci flow in [Bam20a, Proposition 4.2]. ∎

The oscillation bound of PsP^{*}_{s} follows from a straightforward computation:

Proposition 2.21.

Suppose 0<t<t1<t20<t^{*}<t_{1}<t_{2}, |H|2+|H|2C|H|^{2}+|\nabla H|^{2}\leqslant C, dW1t(νx1,t1(t),νx2,t2(t))δd_{W_{1}}^{t^{*}}(\nu_{x_{1},t_{1}}(t^{*}),\nu_{x_{2},t_{2}}(t^{*}))\leqslant\delta, then

P0(x1,t1)P0(x2,t2)+C(t1(1δ)0tlog(t2t2s)ds).P^{*}_{0}(x_{1},t_{1})\leqslant P^{*}_{0}(x_{2},t_{2})+C\left(t_{1}-(1-\delta)\int_{0}^{t^{*}}\log\left(\frac{t_{2}}{t_{2}-s}\right)\mathrm{d}s\right).

In particular, assume t2t<ε<101t_{2}-t^{*}<\varepsilon<10^{-1}, 0<T1<t2<T0<T^{-1}<t_{2}<T, then

P0(x1,t1)P0(x2,t2)+C(εTεlogε+2δT).P^{*}_{0}(x_{1},t_{1})\leqslant P^{*}_{0}(x_{2},t_{2})+C(\varepsilon T-\varepsilon\log\varepsilon+2\delta T).
Proof.

As the computation in Proof of Proposition 2.20, we have

P0(xi,ti)=0tilog(titis)M|H|2dνxi,ti(s)ds=0t+tti,\displaystyle P^{*}_{0}(x_{i},t_{i})=\int_{0}^{t_{i}}\log\left(\frac{t_{i}}{t_{i}-s}\right)\int_{M}|H|^{2}\mathrm{d}\nu_{x_{i},t_{i}}(s)\mathrm{d}s=\int_{0}^{t^{*}}+\int_{t^{*}}^{t_{i}},

where i=1,2i=1,2. For the second term, we have

0ttilog(titis)M|H|2dνxi,ti(s)dsCttilog(titis)ds,0\leqslant\int_{t^{*}}^{t_{i}}\log\left(\frac{t_{i}}{t_{i}-s}\right)\int_{M}|H|^{2}\mathrm{d}\nu_{x_{i},t_{i}}(s)\mathrm{d}s\leqslant C\int_{t^{*}}^{t_{i}}\log\left(\frac{t_{i}}{t_{i}-s}\right)\mathrm{d}s,

where we use |H|2C|H|^{2}\leqslant C. Then we have

P0(x1,t1)P0(x2,t2)\displaystyle P^{*}_{0}(x_{1},t_{1})-P^{*}_{0}(x_{2},t_{2})\leqslant 0tlog(t1t1s)M|H|2dνx1,t1(s)ds+Ctt1log(t1t1s)ds\displaystyle\int_{0}^{t^{*}}\log\left(\frac{t_{1}}{t_{1}-s}\right)\int_{M}|H|^{2}\mathrm{d}\nu_{x_{1},t_{1}}(s)\mathrm{d}s+C\int_{t^{*}}^{t_{1}}\log\left(\frac{t_{1}}{t_{1}-s}\right)\mathrm{d}s
0tlog(t2t2s)M|H|2dνx2,t2(s)ds.\displaystyle-\int_{0}^{t^{*}}\log\left(\frac{t_{2}}{t_{2}-s}\right)\int_{M}|H|^{2}\mathrm{d}\nu_{x_{2},t_{2}}(s)\mathrm{d}s.

By lemma 2.18, we have dW1t(νx1,t1(t),νx2,t2(t))δd_{W_{1}}^{t}(\nu_{x_{1},t_{1}}(t),\nu_{x_{2},t_{2}}(t))\leqslant\delta for any ttt\leqslant t^{*}.

0tlog(t1t1s)M|H|2dνx1,t1(s)ds0tlog(t2t2s)M|H|2dνx2,t2(s)ds\displaystyle\int_{0}^{t^{*}}\log\left(\frac{t_{1}}{t_{1}-s}\right)\int_{M}|H|^{2}\mathrm{d}\nu_{x_{1},t_{1}}(s)\mathrm{d}s-\int_{0}^{t^{*}}\log\left(\frac{t_{2}}{t_{2}-s}\right)\int_{M}|H|^{2}\mathrm{d}\nu_{x_{2},t_{2}}(s)\mathrm{d}s
=\displaystyle= I1+I2,\displaystyle\ I_{1}+I_{2},

where

I1\displaystyle I_{1} :=0t[log(t1t1s)log(t2t2s)]M|H|2dνx1,t1(s)ds\displaystyle:=\int_{0}^{t^{*}}\left[\log\left(\frac{t_{1}}{t_{1}-s}\right)-\log\left(\frac{t_{2}}{t_{2}-s}\right)\right]\int_{M}|H|^{2}\mathrm{d}\nu_{x_{1},t_{1}}(s)\mathrm{d}s
C0t[log(t1t1s)log(t2t2s)]ds,\displaystyle\leqslant C\int_{0}^{t^{*}}\left[\log\left(\frac{t_{1}}{t_{1}-s}\right)-\log\left(\frac{t_{2}}{t_{2}-s}\right)\right]\mathrm{d}s,
I2\displaystyle I_{2} :=0tlog(t2t2s)M|H|2d(νx1,t1(s)νx2,t2(s))ds\displaystyle:=\int_{0}^{t^{*}}\log\left(\frac{t_{2}}{t_{2}-s}\right)\int_{M}|H|^{2}\mathrm{d}\left(\nu_{x_{1},t_{1}}(s)-\nu_{x_{2},t_{2}}(s)\right)\mathrm{d}s
Cδ0tlog(t2t2s)ds.\displaystyle\leqslant C\delta\int_{0}^{t^{*}}\log\left(\frac{t_{2}}{t_{2}-s}\right)\mathrm{d}s.

The inequality on I1I_{1} is due to |H|2C|H|^{2}\leqslant C and the fact that log(t1t1s)log(t2t2s)0\log\left(\frac{t_{1}}{t_{1}-s}\right)-\log\left(\frac{t_{2}}{t_{2}-s}\right)\geqslant 0. The inequality on I2I_{2} is due to the definition of dW1d_{W_{1}} and ||H|2|2|H||H|C\left|\nabla|H|^{2}\right|\leqslant 2|H||\nabla H|\leqslant C.

Combining above, we have

P0(x1,t1)P0(x2,t2)\displaystyle P^{*}_{0}(x_{1},t_{1})-P^{*}_{0}(x_{2},t_{2})\leqslant C{0t[log(t1t1s)log(t2t2s)]ds+δ0tlog(t2t2s)ds\displaystyle C\left\{\int_{0}^{t^{*}}\left[\log\left(\frac{t_{1}}{t_{1}-s}\right)-\log\left(\frac{t_{2}}{t_{2}-s}\right)\right]\mathrm{d}s+\delta\int_{0}^{t^{*}}\log\left(\frac{t_{2}}{t_{2}-s}\right)\mathrm{d}s\right.
+tt1log(t1t1s)ds}\displaystyle\left.+\int_{t^{*}}^{t_{1}}\log\left(\frac{t_{1}}{t_{1}-s}\right)\mathrm{d}s\right\}
\displaystyle\leqslant C{0t1log(t1t1s)ds0tlog(t2t2s)ds+δ0tlog(t2t2s)ds}\displaystyle C\left\{\int_{0}^{t_{1}}\log\left(\frac{t_{1}}{t_{1}-s}\right)\mathrm{d}s-\int_{0}^{t^{*}}\log\left(\frac{t_{2}}{t_{2}-s}\right)\mathrm{d}s+\delta\int_{0}^{t^{*}}\log\left(\frac{t_{2}}{t_{2}-s}\right)\mathrm{d}s\right\}
\displaystyle\leqslant C(t1(1δ)0tlog(t2t2s)ds).\displaystyle C\left(t_{1}-(1-\delta)\int_{0}^{t^{*}}\log\left(\frac{t_{2}}{t_{2}-s}\right)\mathrm{d}s\right).

If t2t<ε<101t_{2}-t^{*}<\varepsilon<10^{-1}, 0<T1<t2<T0<T^{-1}<t_{2}<T, then

0tlog(t2t2s)ds=tlogt2t2logt2+t+(t2t)log(t2t)T+logT2T,\displaystyle\int_{0}^{t^{*}}\log\left(\frac{t_{2}}{t_{2}-s}\right)\mathrm{d}s=t^{*}\log t_{2}-t_{2}\log t_{2}+t^{*}+(t_{2}-t^{*})\log(t_{2}-t^{*})\leqslant T+\log T\leqslant 2T,
t10tlog(t2t2s)ds=\displaystyle t_{1}-\int_{0}^{t^{*}}\log\left(\frac{t_{2}}{t_{2}-s}\right)\mathrm{d}s= (t1t)+(t2t)logt2(t2t)log(t2t)\displaystyle(t_{1}-t^{*})+(t_{2}-t^{*})\log t_{2}-(t_{2}-t^{*})\log(t_{2}-t^{*})
\displaystyle\leqslant ε+εlogTεlogε\displaystyle\varepsilon+\varepsilon\log T-\varepsilon\log\varepsilon
\displaystyle\leqslant εTεlogε.\displaystyle\varepsilon T-\varepsilon\log\varepsilon.

So

P0(x1,t1)P0(x2,t2)C(εTεlogε+2δT).\displaystyle P^{*}_{0}(x_{1},t_{1})-P^{*}_{0}(x_{2},t_{2})\leqslant C(\varepsilon T-\varepsilon\log\varepsilon+2\delta T).

2.2. Weighted noncollapsed

Definition 2.22.

Define the ϕ\phi-weighted volume by

|B(x,t,r)|ϕ:=B(x,t,r)eϕ(y,t)dgt(y).|B(x,t,r)|_{\phi}:=\int_{B(x,t,r)}e^{-\phi(y,t)}\mathrm{d}g_{t}(y).
Remark 2.23.

Since ϕ=16|H|20\square\phi=\frac{1}{6}|H|^{2}\geqslant 0, if we assume ϕ\phi has nonnegative initial data, then ϕ(x,t)\phi(x,t) will always be nonnegative by maximum principle. Then we have |B(x,t,r)|ϕ|B(x,t,r)|gt|B(x,t,r)|_{\phi}\leqslant|B(x,t,r)|_{g_{t}}. Thus weighted noncollapsed condition is stronger than the conventional condition.

Theorem 2.24.

Suppose [t0r2,t0]I[t_{0}-r^{2},t_{0}]\subset I, |Rϕ|+|H|2Cr2|R^{\phi}|+|H|^{2}\leqslant Cr^{-2} on B(x0,t0,r)×[t0r2,t0]B(x_{0},t_{0},r)\times[t_{0}-r^{2},t_{0}], ϕ0\phi\geqslant 0, then

|B(x0,t0,r)|ϕcexp(𝒩x0,t0H(r2))rn.|B(x_{0},t_{0},r)|_{\phi}\geqslant c\exp(\mathcal{N}^{H}_{x_{0},t_{0}}(r^{2}))r^{n}.
Proof.

The proof follows the argument in [Bam20a, Proof of Theorem 6.1] with a modification. After parabolic scaling, we can assume r=1r=1, t0=0t_{0}=0.

Take h(y):=η(d0(x0,y))Cc(B(x0,0,1))h(y):=\eta(d_{0}(x_{0},y))\in C^{\infty}_{c}(B(x_{0},0,1)), where η\eta is cut-off function, such that 0η10\leqslant\eta\leqslant 1, η|[0,12]1\eta|_{[0,\frac{1}{2}]}\equiv 1, η|[1,)0\eta|_{[1,\infty)}\equiv 0, 0η40\geqslant\eta^{\prime}\geqslant-4. a:=Mh2eϕdg0a:=\int_{M}h^{2}e^{-\phi}\mathrm{d}g_{0}, then |B(x0,0,12)|ϕa|B(x0,0,1)|ϕ|B(x_{0},0,\frac{1}{2})|_{\phi}\leqslant a\leqslant|B(x_{0},0,1)|_{\phi}.

Take vC(M×[1,0])v\in C^{\infty}(M\times[-1,0]), such that ϕv=0\square^{*}_{\phi}v=0 and v(,0)=a1h2v(\cdot,0)=a^{-1}h^{2}, then veϕdg=a1h2eϕdg0=1\int ve^{-\phi}\mathrm{d}g=\int a^{-1}h^{2}e^{-\phi}\mathrm{d}g_{0}=1. vt=(4πτ)n2eftv_{t}=(4\pi\tau)^{-\frac{n}{2}}e^{-f_{t}}, where τ:=1t\tau:=1-t. f1=logv1n2log(8π)f_{-1}=-\log v_{-1}-\frac{n}{2}\log(8\pi), f0=log(a1h2)n2log(4π)f_{0}=-\log(a^{-1}h^{2})-\frac{n}{2}\log(4\pi).

𝒩ϕ[g1,f1,2]=f1v1eϕdg1n2=v1logv1eϕdg1Cn.\mathcal{N}^{\phi}[g_{-1},f_{-1},2]=\int f_{-1}v_{-1}e^{-\phi}\mathrm{d}g_{-1}-\frac{n}{2}=-\int v_{-1}\log v_{-1}e^{-\phi}\mathrm{d}g_{-1}-C_{n}.

Note that v1(z)eϕ(z,1)=K(y,0;z,1)a1h2(y)eϕ(y,0)dg0(y)v_{-1}(z)e^{-\phi(z,-1)}=\int K(y,0;z,-1)a^{-1}h^{2}(y)e^{-\phi(y,0)}\mathrm{d}g_{0}(y), then by Jensen inequality,

v1logv1(z)eϕ(z,1)K(y,0;z,1)(logK(y,0;z,1)+ϕ(z,1))a1h2(y)eϕ(y,0)dg0(y).v_{-1}\log v_{-1}(z)\leqslant\int e^{\phi(z,-1)}K(y,0;z,-1)\left(\log K(y,0;z,-1)+\phi(z,-1)\right)a^{-1}h^{2}(y)e^{-\phi(y,0)}\mathrm{d}g_{0}(y).

Note that 𝒩y,0ϕ(1)=K(y,0;z,1)(logK(y,0;z,1)+ϕ(z,1))dg1(z)+Cn\mathcal{N}^{\phi}_{y,0}(1)=-\int K(y,0;z,-1)\left(\log K(y,0;z,-1)+\phi(z,-1)\right)\mathrm{d}g_{-1}(z)+C_{n}, then

𝒩ϕ[g1,f1,2]𝒩y,0ϕ(1)a1h2(y)eϕ(y,0)dg0(y)Cn𝒩x,0ϕ(1)C𝒩x,0H(1)C.\mathcal{N}^{\phi}[g_{-1},f_{-1},2]\geqslant\int\mathcal{N}^{\phi}_{y,0}(1)a^{-1}h^{2}(y)e^{-\phi(y,0)}\mathrm{d}g_{0}(y)-C_{n}\geqslant\mathcal{N}^{\phi}_{x,0}(1)-C\geqslant\mathcal{N}^{H}_{x,0}(1)-C.
01𝒲ϕ[gs,fs,1+s]ds\displaystyle\int_{0}^{1}\mathcal{W}^{\phi}[g_{-s},f_{-s},1+s]\mathrm{d}s =01dds((1+s)𝒩ϕ[gs,fs,1+s])ds\displaystyle=\int_{0}^{1}\frac{\mathrm{d}}{\mathrm{d}s}\left((1+s)\mathcal{N}^{\phi}[g_{-s},f_{-s},1+s]\right)\mathrm{d}s
=2𝒩ϕ[g1,f1,2]𝒩ϕ[g0,f0,1]\displaystyle=2\mathcal{N}^{\phi}[g_{-1},f_{-1},2]-\mathcal{N}^{\phi}[g_{0},f_{0},1]
2𝒩x,0ϕ(1)𝒩ϕ[g0,f0,1]C,\displaystyle\geqslant 2\mathcal{N}^{\phi}_{x,0}(1)-\mathcal{N}^{\phi}[g_{0},f_{0},1]-C,
01𝒲ϕ[gs,fs,1+s]ds𝒲ϕ[g0,f0,1]+C,\int_{0}^{1}\mathcal{W}^{\phi}[g_{-s},f_{-s},1+s]\mathrm{d}s\leqslant\mathcal{W}^{\phi}[g_{0},f_{0},1]+C,
𝒲ϕ[g0,f0,1]+𝒩ϕ[g0,f0,1]\displaystyle\mathcal{W}^{\phi}[g_{0},f_{0},1]+\mathcal{N}^{\phi}[g_{0},f_{0},1] =(Rϕ+|f0|2+2f0)dν0Cn\displaystyle=\int\left(R^{\phi}+|\nabla f_{0}|^{2}+2f_{0}\right)\mathrm{d}\nu_{0}-C_{n}
a1(4|h|22h2logh2)eϕdg0+2loga+C\displaystyle\leqslant a^{-1}\int(4|\nabla h|^{2}-2h^{2}\log h^{2})e^{-\phi}\mathrm{d}g_{0}+2\log a+C
a1(100+2e1)|B(x0,0,1)|ϕ+2log|B(x0,0,1)|ϕ+C\displaystyle\leqslant a^{-1}(100+2e^{-1})|B(x_{0},0,1)|_{\phi}+2\log|B(x_{0},0,1)|_{\phi}+C
200|B(x0,0,1)|ϕ|B(x0,0,12)|ϕ+2log|B(x0,0,1)|ϕ+C.\displaystyle\leqslant 200\frac{|B(x_{0},0,1)|_{\phi}}{|B(x_{0},0,\frac{1}{2})|_{\phi}}+2\log|B(x_{0},0,1)|_{\phi}+C.

Then by parabolic rescale, we have

rn|B(x0,0,r)|ϕcexp(𝒩x,0H(r2))exp(100|B(x0,0,r)|ϕ|B(x0,0,r2)|ϕ).r^{-n}|B(x_{0},0,r)|_{\phi}\geqslant c\exp(\mathcal{N}^{H}_{x,0}(r^{2}))\exp\left(-100\frac{|B(x_{0},0,r)|_{\phi}}{|B(x_{0},0,\frac{r}{2})|_{\phi}}\right).

Then there exists c0=c0(c,n)c_{0}=c_{0}(c,n), we claim that rn|B(x0,0,r)|ϕc0exp(𝒩x,0H(r2))r^{-n}|B(x_{0},0,r)|_{\phi}\geqslant c_{0}\exp(\mathcal{N}^{H}_{x,0}(r^{2})). If not, there exists r0>0r_{0}>0, r0n|B(x0,0,r0)|ϕ<c0exp(𝒩x,0H(r02))r_{0}^{-n}|B(x_{0},0,r_{0})|_{\phi}<c_{0}\exp(\mathcal{N}^{H}_{x,0}(r_{0}^{2})), then

|B(x0,0,r0)|ϕ|B(x0,0,r0/2)|ϕ>1001log(cc0)>2n.\displaystyle\frac{|B(x_{0},0,r_{0})|_{\phi}}{|B(x_{0},0,r_{0}/2)|_{\phi}}>100^{-1}\log(\frac{c}{c_{0}})>2^{n}.

so r0/2r_{0}/2 is also a counterexample for the claim. Repeating this process, we get for any kk, rk=2kr0r_{k}=2^{-k}r_{0} violates the claim, however it’s impossible if we set c0<ωnc_{0}<\omega_{n} since as r0r\to 0, rn|B(x0,0,r)|ϕωneϕ(x0,0)r^{-n}|B(x_{0},0,r)|_{\phi}\to\omega_{n}e^{-\phi(x_{0},0)}, exp(𝒩x,0H(r02))exp(ϕ(x0,0)13|H|2(x0,0))\exp(\mathcal{N}^{H}_{x,0}(r_{0}^{2}))\to\exp(-\phi(x_{0},0)-\frac{1}{3}|H|^{2}(x_{0},0)). ∎

Remark 2.25.

Note that the key point in the proof is the monotonicity of the generalized entropy 𝒩x,tH\mathcal{N}^{H}_{x,t}. For the weighted entropy 𝒩x,tϕ\mathcal{N}^{\phi}_{x,t}, the last argument of the proof fails.

Bamler introduced the concept of HnH_{n}-center instead of worldline to represent the same point in different time with error.

Definition 2.26 ([Bam20a, Definition 3.10]).

A point (z,t)M×I(z,t)\in M\times I is called an HnH_{n}-center of a point (x0,t0)M×I(x_{0},t_{0})\in M\times I if tt0t\leqslant t_{0} and Vart(δz,νx0,t0;t)Hn(t0t)\operatorname{Var}_{t}(\delta_{z},\nu_{x_{0},t_{0};t})\leqslant H_{n}(t_{0}-t), where the variance Var(μ1,μ2):=M×Md2(x1,x2)dμ1(x1)dμ2(x2)\operatorname{Var}(\mu_{1},\mu_{2}):=\int_{M\times M}d^{2}(x_{1},x_{2})\mathrm{d}\mu_{1}(x_{1})\mathrm{d}\mu_{2}(x_{2}).

In [Bam20a, Proposition 3.12], Bamler shows that along the super Ricci flow, HnH_{n}-center always exists due to the HnH_{n}-concentration of such flow. Considering the HnH_{n}-center instead, we can have a similar result using the weaker entropy 𝒩x,tϕ\mathcal{N}^{\phi}_{x,t}.

Theorem 2.27.

Suppose [t0r2,t0]I[t_{0}-r^{2},t_{0}]\subset I, Rϕ(,t0r2)RminϕR^{\phi}(\cdot,t_{0}-r^{2})\geqslant R^{\phi}_{\min}, |H|2Ar2|H|^{2}\leqslant Ar^{-2}, (z,t0r2)(z,t_{0}-r^{2}) is HnH_{n}-center of (x0,t0)(x_{0},t_{0}), then

|B(z,t0r2,2Hnr)|ϕcexp(𝒩x0,t0ϕ(r2))rn,|B(z,t_{0}-r^{2},\sqrt{2H_{n}}r)|_{\phi}\geqslant c\exp(\mathcal{N}^{\phi}_{x_{0},t_{0}}(r^{2}))r^{n},

where c=cexp(2(n2Rminϕr2+23A)12)c=c\exp(-2(n-2R^{\phi}_{\min}r^{2}+\frac{2}{3}A)^{\frac{1}{2}}).

Proof.

The proof is analogous to the argument in [Bam20a, Proof of Theorem 6.2]. After parabolic scaling, we can assume r=t0=1r=t_{0}=1. Denote by dνt:=dνx0,1;t\mathrm{d}\nu_{t}:=\mathrm{d}\nu_{x_{0},1;t}, B:=B(z,0,2Hn)B:=B(z,0,\sqrt{2H_{n}}). By [Bam20a, Proposition 3.13], we have ν0(B)12\nu_{0}(B)\geqslant\frac{1}{2}.

M|fn2𝒩x0,1ϕ(1)|dν0(M(fn2𝒩x0,1ϕ(1))2dν0)12(n2Rminϕ+23Ψx0,1(1))12.\int_{M}\left|f-\frac{n}{2}-\mathcal{N}^{\phi}_{x_{0},1}(1)\right|\mathrm{d}\nu_{0}\leqslant\left(\int_{M}\left(f-\frac{n}{2}-\mathcal{N}^{\phi}_{x_{0},1}(1)\right)^{2}\mathrm{d}\nu_{0}\right)^{\frac{1}{2}}\leqslant\left(n-2R^{\phi}_{\min}+\frac{2}{3}\Psi_{x_{0},1}(1)\right)^{\frac{1}{2}}.
1ν0(B)Bfdν0\displaystyle\frac{1}{\nu_{0}(B)}\int_{B}f\mathrm{d}\nu_{0} n2+𝒩x0,1ϕ(1)1ν0(B)M|fn2𝒩x0,1ϕ(1)|dν0\displaystyle\geqslant\frac{n}{2}+\mathcal{N}^{\phi}_{x_{0},1}(1)-\frac{1}{\nu_{0}(B)}\int_{M}\left|f-\frac{n}{2}-\mathcal{N}^{\phi}_{x_{0},1}(1)\right|\mathrm{d}\nu_{0}
n2+𝒩x0,1ϕ(1)2(n2Rminϕ+23Ψx0,1(1))12.\displaystyle\geqslant\frac{n}{2}+\mathcal{N}^{\phi}_{x_{0},1}(1)-2\left(n-2R^{\phi}_{\min}+\frac{2}{3}\Psi_{x_{0},1}(1)\right)^{\frac{1}{2}}.

Let u:=ν0(B)1(4π)n2efu:=\nu_{0}(B)^{-1}(4\pi)^{-\frac{n}{2}}e^{-f}, then Bueϕdg0=1\int_{B}ue^{-\phi}\mathrm{d}g_{0}=1.

Bulogueϕdg0\displaystyle\int_{B}u\log ue^{-\phi}\mathrm{d}g_{0} =1ν0(B)fdν0n2log(4π)log(ν0(B))\displaystyle=-\frac{1}{\nu_{0}(B)}\int f\mathrm{d}\nu_{0}-\frac{n}{2}\log(4\pi)-\log(\nu_{0}(B))
𝒩x0,1ϕ(1)n2+2(n2Rminϕ+23Ψx0,1(1))12+log2+n2log(4π).\displaystyle\leqslant-\mathcal{N}^{\phi}_{x_{0},1}(1)-\frac{n}{2}+2\left(n-2R^{\phi}_{\min}+\frac{2}{3}\Psi_{x_{0},1}(1)\right)^{\frac{1}{2}}+\log 2+\frac{n}{2}\log(4\pi).

By Jensen’s inequality,

1|B|ϕBulogueϕdg0(1|B|ϕBueϕdg0)log(1|B|ϕBueϕdg0)=1|B|ϕlog|B|ϕ,\frac{1}{|B|_{\phi}}\int_{B}u\log ue^{-\phi}\mathrm{d}g_{0}\geqslant\left(\frac{1}{|B|_{\phi}}\int_{B}ue^{-\phi}\mathrm{d}g_{0}\right)\log\left(\frac{1}{|B|_{\phi}}\int_{B}ue^{-\phi}\mathrm{d}g_{0}\right)=-\frac{1}{|B|_{\phi}}\log|B|_{\phi},
log|B|ϕ𝒩x0,1ϕ(1)2(n2Rminϕ+23Ψx0,1(1))12Cn.\log|B|_{\phi}\geqslant\mathcal{N}^{\phi}_{x_{0},1}(1)-2\left(n-2R^{\phi}_{\min}+\frac{2}{3}\Psi_{x_{0},1}(1)\right)^{\frac{1}{2}}-C_{n}.

3. Heat kernel Estimate

Theorem 3.1.

Suppose [s,t]I[s,t]\subset I, Rϕ(,s)RminϕR^{\phi}(\cdot,s)\geqslant R^{\phi}_{\min}, |H|2A|H|^{2}\leqslant A, then

K(x,t;y,s)C(ts)n2exp(𝒩sϕ(x,t)),K(x,t;y,s)\leqslant\frac{C}{(t-s)^{\frac{n}{2}}}\exp(-\mathcal{N}^{\phi*}_{s}(x,t)),

where C=C(Rminϕ(ts),A(ts))C=C(R^{\phi}_{\min}(t-s),A(t-s)).

Proof.

The proof follows from the argument in [Bam20a, Proof of Theorem 7.1]. After parabolic scaling, we can assume s=0s=0, t=1t=1. By [CCG+10, Theorem 26.25], we always have K(x,t;y,0)Ztn2exp(𝒩0ϕ(x,t))K(x,t;y,0)\leqslant Zt^{-\frac{n}{2}}\exp(-\mathcal{N}^{\phi*}_{0}(x,t)) for t(0,1]t\in(0,1], for some large ZZ, however, ZZ may depends on the flow g(t)g(t). We will show that there exists Z¯=Z¯(Rminϕ(ts),A(ts))\underline{Z}=\underline{Z}(R^{\phi}_{\min}(t-s),A(t-s)), such that if ZZ¯Z\geqslant\underline{Z}, the bound will still hold after replacing ZZ by Z2\frac{Z}{2}. Using this argument, we have K(x,t;y,0)Z¯tn2exp(𝒩0ϕ(x,t))K(x,t;y,0)\leqslant\underline{Z}t^{-\frac{n}{2}}\exp(-\mathcal{N}^{\phi*}_{0}(x,t)). Note that the bound is parabolic scaling invariant, it suffices to show the bound holds for t=1t=1. Denote by u(x,t):=K(x,t;y,0)u(x,t):=K(x,t;y,0), u=0\square u=0.

ddtMueϕdgt=Mu(ϕ1)eϕdgt=MRϕueϕdgtRminϕMueϕdgt.\frac{\mathrm{d}}{\mathrm{d}t}\int_{M}ue^{-\phi}\mathrm{d}g_{t}=-\int_{M}u(\square^{*}_{\phi}1)e^{-\phi}\mathrm{d}g_{t}=-\int_{M}R^{\phi}ue^{-\phi}\mathrm{d}g_{t}\leqslant-R^{\phi}_{\min}\int_{M}ue^{-\phi}\mathrm{d}g_{t}.

Then we have MueϕdgtCexp(ϕ(y,0))\int_{M}ue^{-\phi}\mathrm{d}g_{t}\leqslant C\exp(-\phi(y,0)). Denote by v:=(t12)|u|2+u2v:=(t-\frac{1}{2})|\nabla u|^{2}+u^{2}.

v2(t12)|2u|22|u|2.\square v\leqslant-2(t-\frac{1}{2})|\nabla^{2}u|^{2}-2|\nabla u|^{2}.

Then v0\square v\leqslant 0 for t[12,1]t\in[\frac{1}{2},1]. We assume t12t\geqslant\frac{1}{2} from now on, we have

v(x,t)\displaystyle v(x,t) MK(x,t;y,12)v(y,12)dg12(y)=MK(x,t;y,12)u2(y,12)dg12(y)\displaystyle\leqslant\int_{M}K(x,t;y,\frac{1}{2})v(y,\frac{1}{2})\mathrm{d}g_{\frac{1}{2}}(y)=\int_{M}K(x,t;y,\frac{1}{2})u^{2}(y,\frac{1}{2})\mathrm{d}g_{\frac{1}{2}}(y)
2nZ2MK(x,t;y,12)exp(2𝒩0ϕ(y,12))dg12(y).\displaystyle\leqslant 2^{n}Z^{2}\int_{M}K(x,t;y,\frac{1}{2})\exp(-2\mathcal{N}^{\phi*}_{0}(y,\frac{1}{2}))\mathrm{d}g_{\frac{1}{2}}(y).

Take (z,12)(z,\frac{1}{2}) be the HnH_{n}-center of (x,t)(x,t), we have dW112(δz,νx,t(12))12Hnd_{W_{1}}^{\frac{1}{2}}(\delta_{z},\nu_{x,t}(\frac{1}{2}))\leqslant\sqrt{\frac{1}{2}H_{n}}. We have 𝒩0ϕ(y,12)𝒩0ϕ(z,12)+Cd12(y,z)𝒩0ϕ(x,t)+Cd12(y,z)+C-\mathcal{N}^{\phi*}_{0}(y,\frac{1}{2})\leqslant-\mathcal{N}^{\phi*}_{0}(z,\frac{1}{2})+Cd_{\frac{1}{2}}(y,z)\leqslant-\mathcal{N}^{\phi*}_{0}(x,t)+Cd_{\frac{1}{2}}(y,z)+C.

v(x,t)\displaystyle v(x,t) 2nZ2MK(x,t;y,12)exp(2𝒩0ϕ(y,12))dg12(y)\displaystyle\leqslant 2^{n}Z^{2}\int_{M}K(x,t;y,\frac{1}{2})\exp(-2\mathcal{N}^{\phi*}_{0}(y,\frac{1}{2}))\mathrm{d}g_{\frac{1}{2}}(y)
CZ2exp(2𝒩0ϕ(x,t))MK(x,t;y,12)exp(Cd12(y,z))dg12(y).\displaystyle\leqslant CZ^{2}\exp(-2\mathcal{N}^{\phi*}_{0}(x,t))\int_{M}K(x,t;y,\frac{1}{2})\exp(Cd_{\frac{1}{2}}(y,z))\mathrm{d}g_{\frac{1}{2}}(y).

By [Bam20a, Proof of Theorem 7.1, Theorem 3.14], we have

MK(x,t;y,12)exp(Cd12(y,z))dg12(y)C.\int_{M}K(x,t;y,\frac{1}{2})\exp(Cd_{\frac{1}{2}}(y,z))\mathrm{d}g_{\frac{1}{2}}(y)\leqslant C.

Note that v(x,t)(t12)|u|2v(x,t)\geqslant(t-\frac{1}{2})|\nabla u|^{2}, we have for t[34,1]t\in[\frac{3}{4},1], |u|(x,t)CZexp(𝒩0ϕ(x,t))|\nabla u|(x,t)\leqslant CZ\exp(-\mathcal{N}^{\phi*}_{0}(x,t)).

Fix (x0,1)(x_{0},1), denote by dν:=K(x0,1;,)dg\mathrm{d}\nu:=K(x_{0},1;\cdot,\cdot)\mathrm{d}g, let ρ(0,12)\rho\in(0,\frac{1}{2}) to be determined, then t1:=1ρ2[34,1]t_{1}:=1-\rho^{2}\in[\frac{3}{4},1]. Let (z1,t1)(z_{1},t_{1}) be the HnH_{n}-center of (x0,1)(x_{0},1).

𝒩0ϕ(z1,t1)𝒩0ϕ(x0,1)+CdW1t1(δz1,νt1)𝒩0ϕ(x0,1)+CHnρ2.-\mathcal{N}^{\phi*}_{0}(z_{1},t_{1})\leqslant-\mathcal{N}^{\phi*}_{0}(x_{0},1)+Cd_{W_{1}}^{t_{1}}(\delta_{z_{1}},\nu_{t_{1}})\leqslant-\mathcal{N}^{\phi*}_{0}(x_{0},1)+C\sqrt{H_{n}\rho^{2}}.

We have for ρρ¯\rho\leqslant\bar{\rho}, 𝒩0ϕ(z1,t1)𝒩0ϕ(x0,1)+log2-\mathcal{N}^{\phi*}_{0}(z_{1},t_{1})\leqslant-\mathcal{N}^{\phi*}_{0}(x_{0},1)+\log 2. Then 𝒩0ϕ(y,t1)𝒩0ϕ(x0,1)+log2+Cdt1(y,z1)-\mathcal{N}^{\phi*}_{0}(y,t_{1})\leqslant-\mathcal{N}^{\phi*}_{0}(x_{0},1)+\log 2+Cd_{t_{1}}(y,z_{1}). Denote by B:=B(z1,t1,100Hnρ)B:=B(z_{1},t_{1},\sqrt{100H_{n}}\rho).

u(x0,1)=Mudνt1=B+MBudνt1.u(x_{0},1)=\int_{M}u\mathrm{d}\nu_{t_{1}}=\int_{B}+\int_{M\setminus B}u\mathrm{d}\nu_{t_{1}}.

By Theorem 2.27, |B|ϕcexp(𝒩0ϕ(x0,1))ρn|B|_{\phi}\geqslant c\exp(\mathcal{N}^{\phi*}_{0}(x_{0},1))\rho^{n}. For any x,x′′Bx^{\prime},x^{\prime\prime}\in B, we have u(x,t1)u(x′′,t1)+CZexp(𝒩0ϕ(x,t))ρu(x^{\prime},t_{1})\leqslant u(x^{\prime\prime},t_{1})+CZ\exp(-\mathcal{N}^{\phi*}_{0}(x,t))\rho. Integrate by |B|ϕ1eϕdgt1|B|_{\phi}^{-1}e^{-\phi}\mathrm{d}g_{t_{1}},

u(x,t1)\displaystyle u(x^{\prime},t_{1}) 1|B|ϕBueϕdgt1+CZexp(𝒩0ϕ(x,t))ρ\displaystyle\leqslant\frac{1}{|B|_{\phi}}\int_{B}ue^{-\phi}\mathrm{d}g_{t_{1}}+CZ\exp(-\mathcal{N}^{\phi*}_{0}(x,t))\rho
Cexp(𝒩0ϕ(x,t))(Zρ+Cρn).\displaystyle\leqslant C\exp(-\mathcal{N}^{\phi*}_{0}(x,t))(Z\rho+C\rho^{-n}).

Then for the first term, we have Budνt1Cexp(𝒩0ϕ(x,t))(Zρ+Cρn)\int_{B}u\mathrm{d}\nu_{t_{1}}\leqslant C\exp(-\mathcal{N}^{\phi*}_{0}(x,t))(Z\rho+C\rho^{-n}). For the second term, νt1(MB)1001\nu_{t_{1}}(M\setminus B)\leqslant 100^{-1}, MBudνt1Zt1n2MBexp(𝒩0ϕ(y,t1))dνt1(y)\int_{M\setminus B}u\mathrm{d}\nu_{t_{1}}\leqslant Zt_{1}^{-\frac{n}{2}}\int_{M\setminus B}\exp(-\mathcal{N}^{\phi*}_{0}(y,t_{1}))\mathrm{d}\nu_{t_{1}}(y). For ρρ¯\rho\leqslant\bar{\rho}, since 𝒩0ϕ(z1,t1)𝒩0ϕ(x0,1)+log2-\mathcal{N}^{\phi*}_{0}(z_{1},t_{1})\leqslant-\mathcal{N}^{\phi*}_{0}(x_{0},1)+\log 2, we have

MBudνt1\displaystyle\int_{M\setminus B}u\mathrm{d}\nu_{t_{1}} 2ZMBexp(𝒩0ϕ(y,t1))dνt1(y)\displaystyle\leqslant 2Z\int_{M\setminus B}\exp(-\mathcal{N}^{\phi*}_{0}(y,t_{1}))\mathrm{d}\nu_{t_{1}}(y)
4Zexp(𝒩0ϕ(x0,1))MBexp(Cdt1(y,z1))dνt1(y).\displaystyle\leqslant 4Z\exp(-\mathcal{N}^{\phi*}_{0}(x_{0},1))\int_{M\setminus B}\exp(Cd_{t_{1}}(y,z_{1}))\mathrm{d}\nu_{t_{1}}(y).

By [Bam20a, (7.18)], we have

MBexp(Cdt1(y,z1))dνt1(y)ν(MB)+Cρ1100+Cρ.\int_{M\setminus B}\exp(Cd_{t_{1}}(y,z_{1}))\mathrm{d}\nu_{t_{1}}(y)\leqslant\nu(M\setminus B)+C\rho\leqslant\frac{1}{100}+C\rho.

Then we have MBudνt1exp(𝒩0ϕ(x,t))(110Z+CZρ)\int_{M\setminus B}u\mathrm{d}\nu_{t_{1}}\leqslant\exp(-\mathcal{N}^{\phi*}_{0}(x,t))(\frac{1}{10}Z+CZ\rho).

u(x0,1)exp(𝒩0ϕ(x,t))(CZρ+Cρn+110Z).u(x_{0},1)\leqslant\exp(-\mathcal{N}^{\phi*}_{0}(x,t))(CZ\rho+C\rho^{-n}+\frac{1}{10}Z).

First fix ρ=ρ0\rho=\rho_{0} such that Cρ0110C\rho_{0}\leqslant\frac{1}{10}, then for ZZ¯=5Cρ0nZ\geqslant\underline{Z}=5C\rho_{0}^{-n}, we have

u(x0,1)12Zexp(𝒩0ϕ(x,t)).u(x_{0},1)\leqslant\frac{1}{2}Z\exp(-\mathcal{N}^{\phi*}_{0}(x,t)).

. ∎

Theorem 3.2.

Suppose [s,t]I[s,t]\subset I, Rϕ(,s)RminϕR^{\phi}(\cdot,s)\geqslant R^{\phi}_{\min}, |H|2A|H|^{2}\leqslant A, then there exist constants C,C0C,C_{0} depending on Rminϕ(ts)R^{\phi}_{\min}(t-s) and A(ts)A(t-s), such that

|xK|K(x,t;y,s)C(ts)12log(C0exp(𝒩sϕ(x,t))(ts)n2K(x,t;y,s)).\frac{|\nabla_{x}K|}{K}(x,t;y,s)\leqslant\frac{C}{(t-s)^{\frac{1}{2}}}\sqrt{\log\left(\frac{C_{0}\exp(-\mathcal{N}^{\phi*}_{s}(x,t))}{(t-s)^{\frac{n}{2}}K(x,t;y,s)}\right)}.
Proof.

By [Bam20a, Proof of Theorem 7.5], Theorem 3.1 and Corollary 2.19 imply this theorem, since other argument holds for super Ricci flow. ∎

4. ε\varepsilon-Regularity

We need some notions to state the theorem.

Definition 4.1.

Backward parabolic curvature radius is defined by

rRm(x,t):=sup{r>0:|Rm|r2on B(x,t,r)×[tr2,t]}.r_{\mathrm{Rm}}(x,t):=\sup\{r>0:|\mathrm{Rm}|\leqslant r^{-2}\text{on\ }B(x,t,r)\times[t-r^{2},t]\}.

Bamler introduce a new notion of parabolic neighborhood instead of conventional neighborhood in [Bam20a], which is an important concept in [Bam20a, Bam20b, Bam20c].

Definition 4.2 ([Bam20a]).

Suppose that (x0,t0)M×I(x_{0},t_{0})\in M\times I, r>0r>0 and t0r2It_{0}-r^{2}\in I. The backward PP^{*}-parabolic rr-balls is defined by

P(x0,t0,r):={(x,t)M×[t0r2,t0]:dW1t0r2(νx0,t0;t0r2,νx,t;t0r2)<r}.P^{*}_{-}(x_{0},t_{0},r):=\{(x,t)\in M\times[t_{0}-r^{2},t_{0}]:d_{W_{1}}^{t_{0}-r^{2}}(\nu_{x_{0},t_{0};t_{0}-r^{2}},\nu_{x,t;t_{0}-r^{2}})<r\}.
Theorem 4.3.

Suppose [tr2,t]I[t-r^{2},t]\subset I, r2|H|2+r4|H|2Cr^{2}|H|^{2}+r^{4}|\nabla H|^{2}\leqslant C, RϕCr2R^{\phi}\geqslant-Cr^{-2}, there exists ε(n,C)>0\varepsilon(n,C)>0, if 𝒩x,tH(r2)ε\mathcal{N}^{H}_{x,t}(r^{2})\geqslant-\varepsilon, then rRm(x,t)εrr_{\mathrm{Rm}}(x,t)\geqslant\varepsilon r.

Proof.

The proof follows from the argument in [Bam20a, Proof of Theorem 10.2]. After a parabolic scaling, we can assume t=r=1t=r=1. Assume there exists a sequence of counterexamples (Mi,(gi,t)t[0,1],(Hi,t),(ϕi,t),(xi,1))(M_{i},(g_{i,t})_{t\in[0,1]},(H_{i,t}),(\phi_{i,t}),(x_{i},1)), ri:=rRm(xi,1)<εi0r_{i}:=r_{\mathrm{Rm}}(x_{i},1)<\varepsilon_{i}\to 0, 𝒩xi,1H(1)εi\mathcal{N}^{H}_{x_{i},1}(1)\geqslant-\varepsilon_{i}.

Choose a sequence AiA_{i}\to\infty with 10Airi010A_{i}r_{i}\to 0. By a point-picking argument in [Bam20a, Claim 10.5], there exists (xi,ti)P(xi,1,10Airi)(x_{i}^{\prime},t_{i}^{\prime})\in P^{*}_{-}(x_{i},1,10A_{i}r_{i}) satisfies ri:=rRm(xi,ti)ri0r_{i}^{\prime}:=r_{\mathrm{Rm}}(x_{i}^{\prime},t_{i}^{\prime})\leqslant r_{i}\to 0 and rRm110rir_{\mathrm{Rm}}\geqslant\frac{1}{10}r_{i}^{\prime} on P(xi,ti,Airi)P^{*}_{-}(x_{i}^{\prime},t_{i}^{\prime},A_{i}r_{i}^{\prime}).

By RϕCR^{\phi}\geqslant-C, Corollary 2.19 and Proposition 2.21, we have

𝒩0ϕ(xi,ti)\displaystyle\mathcal{N}^{\phi*}_{0}(x_{i}^{\prime},t_{i}^{\prime}) 𝒩0ϕ(xi,1)CAirin2log(ti1(10Airi)2),\displaystyle\geqslant\mathcal{N}^{\phi*}_{0}(x_{i},1)-CA_{i}r_{i}-\frac{n}{2}\log\left(\frac{t_{i}^{\prime}}{1-(10A_{i}r_{i})^{2}}\right),
P0(xi,ti)\displaystyle P^{*}_{0}(x_{i}^{\prime},t_{i}^{\prime}) P0(x1,1)C(Airi)2log(10Airi).\displaystyle\leqslant P^{*}_{0}(x_{1},1)-C(A_{i}r_{i})^{2}\log(10A_{i}r_{i}).

Then 𝒩xi,1H(1)0\mathcal{N}^{H}_{x_{i},1}(1)\to 0, Airi0A_{i}r_{i}\to 0 implies 𝒩xi,tiH(ti)0\mathcal{N}^{H}_{x_{i}^{\prime},t_{i}^{\prime}}(t_{i}^{\prime})\to 0. Note that ϕi(y,ti)dν(y)=𝒩xi,tiϕ(0)0\int\phi_{i}(y,t_{i}^{\prime})\mathrm{d}\nu(y)=-\mathcal{N}^{\phi}_{x_{i}^{\prime},t_{i}^{\prime}}(0)\to 0, ddτϕidν=16|Hi|2dν0\frac{\mathrm{d}}{\mathrm{d}\tau}\int\phi_{i}\mathrm{d}\nu=-\frac{1}{6}\int|H_{i}|^{2}\mathrm{d}\nu\leqslant 0. Then for t[0,ti]t\in[0,t_{i}^{\prime}], ϕi(t)dν\int\phi_{i}(t)\mathrm{d}\nu uniformly converges to 0.

By [Bam20a, Corollary 9.6], we can take a sequence AiAiA_{i}^{\prime}\leqslant A_{i}, AiA_{i}^{\prime}\to\infty such that P(xi,ti,Airi)P(xi,ti,Airi)P_{-}(x_{i}^{\prime},t_{i}^{\prime},A_{i}^{\prime}r_{i}^{\prime})\subset P^{*}_{-}(x_{i}^{\prime},t_{i}^{\prime},A_{i}r_{i}^{\prime}), then rRm110rir_{\mathrm{Rm}}\geqslant\frac{1}{10}r_{i}^{\prime} on P(xi,ti,Airi)P_{-}(x_{i}^{\prime},t_{i}^{\prime},A_{i}^{\prime}r_{i}^{\prime}). Define g~i(t):=ri2gi(ri2t+ti),\tilde{g}_{i}(t):=r_{i}^{\prime-2}g_{i}(r_{i}^{\prime 2}t+t_{i}^{\prime}), H~i(t):=ri2Hi(ri2t+ti)\tilde{H}_{i}(t):=r_{i}^{\prime-2}H_{i}(r_{i}^{\prime 2}t+t_{i}^{\prime}), we obtain a sequence of flows (M~i(\tilde{M}_{i},(g~i,t)t[(Ai)2,0](\tilde{g}_{i,t})_{t\in[-(A_{i}^{\prime})^{2},0]}, (H~i,t),(ϕ~i,t)))(\tilde{H}_{i,t}),(\tilde{\phi}_{i,t}))) satisfies that rRm110r_{\mathrm{Rm}}\geqslant\frac{1}{10} on P(xi,0,Ai)P_{-}(x_{i}^{\prime},0,A_{i}^{\prime}), rRm(xi,0)=1r_{\mathrm{Rm}}(x_{i}^{\prime},0)=1, |H~i,t|2Cri20|\tilde{H}_{i,t}|^{2}\leqslant Cr_{i}^{\prime 2}\to 0, and for any fixed T>0T>0, 𝒩~xi,0H(T)=𝒩xi,tiH(ri2T)0\tilde{\mathcal{N}}^{H}_{x_{i}^{\prime},0}(T)=\mathcal{N}^{H}_{x_{i}^{\prime},t_{i}^{\prime}}(r_{i}^{\prime 2}T)\to 0, ϕ~i(T)dν~i0\int\tilde{\phi}_{i}(T)\mathrm{d}\tilde{\nu}_{i}\to 0.

Then the injectivity radius at (xi,0)(x_{i}^{\prime},0) is uniformly bounded from below by Theorem 2.24. Since |Rm||\mathrm{Rm}|, |H||H|, |H||\nabla H| are all uniformly bounded, after passing to a subsequence, these flows converge to a smooth and complete pointed ancient flow (M,(g,t)t0,x,H=0)(M_{\infty},(g_{\infty,t})_{t\leqslant 0},x_{\infty},H_{\infty}=0). Since |H~i|20|\tilde{H}_{i}|^{2}\to 0, the limit flow is actually Ricci flow. By the upper bound of heat kernel and its gradient, K~i(xi,0;,)\tilde{K}_{i}(x_{i}^{\prime},0;\cdot,\cdot) converge locally uniformly to a positive solution vC(M×)v_{\infty}\in C^{\infty}(M_{\infty}\times\mathbb{R}_{-}) of conjugate heat equation. Since ϕ~i(T)dν~i0\int\tilde{\phi}_{i}(T)\mathrm{d}\tilde{\nu}_{i}\to 0, ϕ~i=16|H~i|20\square\tilde{\phi}_{i}=\frac{1}{6}|\tilde{H}_{i}|^{2}\to 0, |H~i|20|\nabla\tilde{H}_{i}|^{2}\to 0, we have ϕ~iϕ0\tilde{\phi}_{i}\to\phi_{\infty}\equiv 0.

Since for any fixed T>0T>0, 𝒩~xi,0H(T)0\tilde{\mathcal{N}}^{H}_{x_{i}^{\prime},0}(T)\to 0, then 𝒲~xi,0H(T)TT+1𝒲~xi,0H(s)ds=(T+1)𝒩~xi,0H(T+1)T𝒩~xi,0H(T)0\tilde{\mathcal{W}}^{H}_{x_{i}^{\prime},0}(T)\geqslant\int_{T}^{T+1}\tilde{\mathcal{W}}^{H}_{x_{i}^{\prime},0}(s)\mathrm{d}s=(T+1)\tilde{\mathcal{N}}^{H}_{x_{i}^{\prime},0}(T+1)-T\tilde{\mathcal{N}}^{H}_{x_{i}^{\prime},0}(T)\to 0, which implies

T0M(2τ|Ricf~i+ϕ~i12τg~i|213τΨ~xi,0)dνdt0.\int_{-T}^{0}\int_{M}\left(-2\tau\left|\mathrm{Ric}^{\tilde{f}_{i}+\tilde{\phi}_{i}}-\frac{1}{2\tau}\tilde{g}_{i}\right|^{2}-\frac{1}{3\tau}\tilde{\Psi}_{x_{i}^{\prime},0}\right)\mathrm{d}\nu\mathrm{d}t\to 0.

Then Ricf+ϕ12τg=0\mathrm{Ric}^{f_{\infty}+\phi_{\infty}}-\frac{1}{2\tau}g_{\infty}=0. ϕ=0\phi_{\infty}=0 implies Ric+2f12τg=0\mathrm{Ric}+\nabla^{2}f_{\infty}-\frac{1}{2\tau}g_{\infty}=0, thus gg_{\infty} is a gradient shrinking soliton. Since |Rm||\mathrm{Rm}| is uniformly bounded, gg_{\infty} must be flat, which contradicts with rRm(x,0)=1<r_{\mathrm{Rm}}(x_{\infty},0)=1<\infty. ∎

5. Some Estimates on HH

In the previous sections, a bound on |H||H| is always needed. Such bounds may not always hold without assumptions, so we will derive some estimates under centain curvature conditions.

5.1. Integral bounds on HH

By the monotonicity of the weighted scalar curvature, (2.6) gives the first integral bounds depends on the scalar curvature RR and the volume Vol(gt)\mathrm{Vol}(g_{t}):

M|H|2dgt12M(RR0ϕ)dgt.\int_{M}|H|^{2}\mathrm{d}g_{t}\leqslant 12\int_{M}\left(R-R^{\phi}_{0}\right)\mathrm{d}g_{t}.
Proposition 5.1.

We have

Vol(gt)e3R0ϕt(Vol(g0)+0tM2e3R0ϕsRdgsds).\mathrm{Vol}(g_{t})\leqslant e^{-3R^{\phi}_{0}t}\left(\mathrm{Vol}(g_{0})+\int_{0}^{t}\int_{M}2e^{3R^{\phi}_{0}s}R\mathrm{d}g_{s}\mathrm{d}s\right).

In particular, if MRdgtC0\int_{M}R\mathrm{d}g_{t}\leqslant C_{0}, then we have

  • If R0ϕ>0R^{\phi}_{0}>0, then Vol(gt)C\mathrm{Vol}(g_{t})\leqslant C, M|H|2dgtC\int_{M}|H|^{2}\mathrm{d}g_{t}\leqslant C.

  • If R0ϕ=0R^{\phi}_{0}=0, then Vol(gt)C(t+1)\mathrm{Vol}(g_{t})\leqslant C(t+1), M|H|2dgtC\int_{M}|H|^{2}\mathrm{d}g_{t}\leqslant C.

  • If R0ϕ<0R^{\phi}_{0}<0, then Vol(gt)Ce3R0ϕt\mathrm{Vol}(g_{t})\leqslant Ce^{-3R^{\phi}_{0}t}, M|H|2dgtCe3R0ϕt\int_{M}|H|^{2}\mathrm{d}g_{t}\leqslant Ce^{-3R^{\phi}_{0}t}.

In particular, if RC0R\leqslant C_{0}, then we have

Vol(gt)Ce(2C03R0ϕ)t,M|H|2dgtCeCt.\mathrm{Vol}(g_{t})\leqslant Ce^{\left(2C_{0}-3R^{\phi}_{0}\right)t},\ \ \int_{M}|H|^{2}\mathrm{d}g_{t}\leqslant Ce^{Ct}.
Proof.
ddtVol(gt)=M(14|H|2R)dgt2MRdgt3R0ϕVol(gt).\frac{\mathrm{d}}{\mathrm{d}t}\mathrm{Vol}(g_{t})=\int_{M}\left(\frac{1}{4}|H|^{2}-R\right)\mathrm{d}g_{t}\leqslant 2\int_{M}R\mathrm{d}g_{t}-3R^{\phi}_{0}\mathrm{Vol}(g_{t}).

Then the results follow from the Gronwall’s inequality. ∎

Proposition 5.2.
ddtM|H|2dgt=M(2|dH|232|H2|2+14|H|4+6Ric,H2R|H|2)dgt.\frac{\mathrm{d}}{\mathrm{d}t}\int_{M}|H|^{2}\mathrm{d}g_{t}=\int_{M}\left(-2|d^{*}H|^{2}-\frac{3}{2}|H^{2}|^{2}+\frac{1}{4}|H|^{4}+6\left<\mathrm{Ric},H^{2}\right>-R|H|^{2}\right)\mathrm{d}g_{t}.
Proof.

Direct calculation. ∎

5.2. Pointwise bounds on HH

In many cases, the integral bound is not enough. In previous sections, we need to estimate the term

Ψx,t(τ)=tτtM|H|2dνx,t;sds.\Psi_{x,t}(\tau)=\int_{t-\tau}^{t}\int_{M}|H|^{2}\mathrm{d}\nu_{x,t;s}\mathrm{d}s.

Note that Ψx,t(0)=M|H|2dνx,t;t=|H|2(x,t)\Psi_{x,t}^{\prime}(0)=\int_{M}|H|^{2}\mathrm{d}\nu_{x,t;t}=|H|^{2}(x,t), then Ψx,t(τ)=|H|2(x,t)τ+O(τ2)\Psi_{x,t}(\tau)=|H|^{2}(x,t)\tau+O(\tau^{2}). Thus to estimate the term Ψx,t\Psi_{x,t}, the pointwise bound on HH is necessary.

Proposition 5.3.
(5.1) |H|2=2|H|232|H2|26RH2|H|232n|H|4+Cn|Rm||H|2,\square|H|^{2}=-2|\nabla H|^{2}-\frac{3}{2}|H^{2}|^{2}-6R_{H}\leqslant-2|\nabla H|^{2}-\frac{3}{2n}|H|^{4}+C_{n}|\mathrm{Rm}||H|^{2},

where RH:=RijklHijrHklsgrsR_{H}:=R_{ijkl}H^{ijr}H^{kls}g_{rs}.

Proof.
t|H|2=32Ric12H2,H2+2ΔdH,H,\frac{\partial}{\partial t}|H|^{2}=3\left<2\mathrm{Ric}-\frac{1}{2}H^{2},H^{2}\right>+2\left<\Delta_{d}H,H\right>,

where Δd=(dd+dd)\Delta_{d}=-(dd^{*}+d^{*}d) is Hodge Laplacian. Denote the connection Laplacian by Δ\Delta, then by the Weitzenböck formula, we have

2ΔdH,H=2ΔHRic(H),H=Δ|H|22|H|22Ric(H),H,2\left<\Delta_{d}H,H\right>=2\left<\Delta H-\mathrm{Ric}(H),H\right>=\Delta|H|^{2}-2|\nabla H|^{2}-2\left<\mathrm{Ric}(H),H\right>,

where Ric(H)\mathrm{Ric}(H) is the Weitzenböck curvature operator on the tensor defined by

Ric(H)(X1,,Xk):=i,j(R(ej,Xi)H)(X1,,ej,,Xk).\mathrm{Ric}(H)(X_{1},\cdots,X_{k}):=\sum_{i,j}(R(e_{j},X_{i})H)(X_{1},\cdots,e_{j},\cdots,X_{k}).

In the normal coordinate,

(Ric(H))ijk=(RsisrHrjk+RsijrHsrk+RsikrHsjr+RsjirHrsk+RsjsrHirk+RsjkrHisr\displaystyle(\mathrm{Ric}(H))_{ijk}=-\left(R_{sisr}H_{rjk}+R_{sijr}H_{srk}+R_{sikr}H_{sjr}+R_{sjir}H_{rsk}+R_{sjsr}H_{irk}+R_{sjkr}H_{isr}\right.
+RskirHrjs+RskjrHirs+RsksrHijr),\displaystyle\left.+R_{skir}H_{rjs}+R_{skjr}H_{irs}+R_{sksr}H_{ijr}\right),

where

RsisrHrjk=RirHrjk,RsjsrHirk=RjrHirk,RsksrHijr=RkrHijr.-R_{sisr}H_{rjk}=R_{ir}H_{rjk},\ -R_{sjsr}H_{irk}=R_{jr}H_{irk},\ -R_{sksr}H_{ijr}=R_{kr}H_{ijr}.

By Bianchi identity,

RsijrHsrkRsjirHrsk=RijsrHsrk+RjsirHsrkRsjirHrsk=RijsrHsrk.-R_{sijr}H_{srk}-R_{sjir}H_{rsk}=R_{ijsr}H_{srk}+R_{jsir}H_{srk}-R_{sjir}H_{rsk}=R_{ijsr}H_{srk}.

Thus

(Ric(H))ijk=RirHrjk+RjrHirk+RkrHijr+RijsrHsrk+RiksrHsjr+RjksrHisr,(\mathrm{Ric}(H))_{ijk}=R_{ir}H_{rjk}+R_{jr}H_{irk}+R_{kr}H_{ijr}+R_{ijsr}H_{srk}+R_{iksr}H_{sjr}+R_{jksr}H_{isr},
Ric(H),H=3Ric,H2+3RijsrHsrkHijk=3Ric,H2+3RH,\left<\mathrm{Ric}(H),H\right>=3\left<\mathrm{Ric},H^{2}\right>+3R_{ijsr}H_{srk}H_{ijk}=3\left<\mathrm{Ric},H^{2}\right>+3R_{H},
t|H|2\displaystyle\frac{\partial}{\partial t}|H|^{2} =6Ric,H232|H2|2+Δ|H|22|H|26Ric,H26RH\displaystyle=6\left<\mathrm{Ric},H^{2}\right>-\frac{3}{2}|H^{2}|^{2}+\Delta|H|^{2}-2|\nabla H|^{2}-6\left<\mathrm{Ric},H^{2}\right>-6R_{H}
=Δ|H|22|H|232|H2|26RH.\displaystyle=\Delta|H|^{2}-2|\nabla H|^{2}-\frac{3}{2}|H^{2}|^{2}-6R_{H}.

The inequality of (5.1) follows from |A|21n(trA)2|A|^{2}\geqslant\frac{1}{n}(\operatorname{tr}A)^{2} and |RH|=|RmHH|Cn|Rm||H|2.|R_{H}|=|\mathrm{Rm}*H*H|\leqslant C_{n}|\mathrm{Rm}||H|^{2}.

Corollary 5.4.

Suppose |Rm|K|\mathrm{Rm}|\leqslant K, then |H|2max(CnK,maxx|H|2(x,0))|H|^{2}\leqslant\max\left(C_{n}K,\max_{x}|H|^{2}(x,0)\right).

Proof.

Apply maximum principle to (5.1). ∎

In general cases, bounds on Riemann curvature tensor seems to be necessary. However, the more important cases of generalized Ricci flow is in the lower dimension. For the dimensional reasons, the equation will be a bit simplified. Since the 3-form in 2-manifolds is trivial, the generalized Ricci flow is exactly the same as the Ricci flow. In dimension n=3n=3, the tensor HH will be deduced to a scalar function:

Proposition 5.5 ([GFS21, Proposition 4.38]).

Suppose n=3n=3, (gt,Ht)(g_{t},H_{t}) solves the generalized Ricci flow, the function htC(M)h_{t}\in C^{\infty}(M) satisfies Ht=htdgtH_{t}=h_{t}\mathrm{d}g_{t}, then

{tg=2Ric+h2g,th=Δh+Rh32h3.\left\{\begin{aligned} \frac{\partial}{\partial t}g&=-2\mathrm{Ric}+h^{2}g,\\ \frac{\partial}{\partial t}h&=\Delta h+Rh-\frac{3}{2}h^{3}.\end{aligned}\right.
Corollary 5.6.

Suppose n=3n=3, |R|C0|R|\leqslant C_{0}, then |H|C|H|\leqslant C.

Proof.

Assume hh attains its maximum at (x0,t0)(x_{0},t_{0}), t0>0t_{0}>0, h(x0,t0)=maxt[0,T]h(x,t)h(x_{0},t_{0})=\max_{t\in[0,T]}h(x,t).

0h(x0,t0)Rh(x0,t0)32h3(x0,t0).0\leqslant\square h(x_{0},t_{0})\leqslant Rh(x_{0},t_{0})-\frac{3}{2}h^{3}(x_{0},t_{0}).

Then h(x0,t0)(23C0)12h(x_{0},t_{0})\leqslant\left(\frac{2}{3}C_{0}\right)^{\frac{1}{2}}. The lower bound is similar. The result follows from |H|2=6h2|H|^{2}=6h^{2}. ∎

In dimension n=4n=4, the bounds on scalar curvature RR seems only to give a time-dependent bound on HL2||H||_{L^{2}}. However, by a careful computation, we will show the Ricci curvature bound implies the pointwise bound on HH.

Proposition 5.7.

Suppose n=4n=4, for a suitable orthonormal basis {ei}\{e_{i}\} of TpMT_{p}M, we have

(H2)ij(p)=13|H|2(1110).(H^{2})_{ij}(p)=\frac{1}{3}|H|^{2}\left(\begin{matrix}1&&&\\ &1&&\\ &&1&\\ &&&0\end{matrix}\right).

Thus there exists a vector field XHX_{H} as the eigenvector of 0. In fact, we can take XH=(H)Γ(TM)X_{H}=\left(\star H\right)^{\sharp}\in\Gamma^{\infty}(TM).

Proof.

It’s obvious if H(p)=0H(p)=0. So we assume H(p)0H(p)\neq 0. Take (H)=λe4\left(\star H\right)^{\sharp}=\lambda e_{4} and extend e4e_{4} to an orthonormal basis {ei}\{e_{i}\}. Then

H=(λe4)=λe1e2e3.\displaystyle H=-\star(\lambda e^{4})=\lambda e^{1}\wedge e^{2}\wedge e^{3}.

So λ2=16|H|2\lambda^{2}=\frac{1}{6}|H|^{2}. (H2)4k=(H2)k4=ie4H,iekH=0(H^{2})_{4k}=(H^{2})_{k4}=\left<i_{e_{4}}H,i_{e_{k}}H\right>=0 for k{1,2,3,4}k\in\{1,2,3,4\}. For i,j{1,2,3}i,j\in\{1,2,3\}

(H2)ij=k,l{1,2,3}HiklHjkl=δijk,l{1,2,3}Hikl2=2λ2δij=13|H|2δij.\displaystyle(H^{2})_{ij}=\sum_{k,l\in\{1,2,3\}}H_{ikl}H_{jkl}=\delta_{ij}\sum_{k,l\in\{1,2,3\}}H^{2}_{ikl}=2\lambda^{2}\delta_{ij}=\frac{1}{3}|H|^{2}\delta_{ij}.

Proposition 5.8.

Suppose n=4n=4, we have

RH=13R|H|22Ric,H2=13R|H|2+23Ric(XH)|H|2.R_{H}=\frac{1}{3}R|H|^{2}-2\left<\mathrm{Ric},H^{2}\right>=-\frac{1}{3}R|H|^{2}+\frac{2}{3}\mathrm{Ric}(X_{H})|H|^{2}.

Combine with (5.1), we have

|H|2=2|H|212|H|4+2R|H|24Ric(XH)|H|2.\square|H|^{2}=-2|\nabla H|^{2}-\frac{1}{2}|H|^{4}+2R|H|^{2}-4\mathrm{Ric}(X_{H})|H|^{2}.

Thus RicK\mathrm{Ric}\geqslant-K, RKR\leqslant K implies |H|2max(12K,maxx|H|2(x,0))|H|^{2}\leqslant\max(12K,\max_{x}|H|^{2}(x,0)).

Proof.

Take the normal coordinate and let Ric=(Rij)\mathrm{Ric}=(R_{ij}) to be diagonal.

RH=i,j,k,l,rRijklHijrHklr={i,j}={k,l}+{i,j}{k,l},R_{H}=\sum_{i,j,k,l,r}R_{ijkl}H_{ijr}H_{klr}=\sum_{\{i,j\}=\{k,l\}}+\sum_{\{i,j\}\neq\{k,l\}},
{i,j}={k,l}=4[K12(H1232+H1242)+K13(H1232+H1342)+K14(H1242+H1342)\displaystyle\sum_{\{i,j\}=\{k,l\}}=-4\left[K_{12}(H^{2}_{123}+H^{2}_{124})+K_{13}(H^{2}_{123}+H^{2}_{134})+K_{14}(H^{2}_{124}+H^{2}_{134})\right.
+K23(H1232+H2342)+K24(H1242+H2342)+K34(H1342+H2342)],\displaystyle\left.+K_{23}(H^{2}_{123}+H^{2}_{234})+K_{24}(H^{2}_{124}+H^{2}_{234})+K_{34}(H^{2}_{134}+H^{2}_{234})\right],

where Kij=RijjiK_{ij}=R_{ijji} is the sectional curvature. Note that

(K12+K13+K23)H1232=(12RR44)H1232.(K_{12}+K_{13}+K_{23})H^{2}_{123}=\left(\frac{1}{2}R-R_{44}\right)H^{2}_{123}.

Thus we have

{i,j}={k,l}=2R(H1232+H1242+H1342+H2342)+4(R11H2342+R22H1342+R33H1242+R44H1232).\sum_{\{i,j\}=\{k,l\}}=-2R\left(H^{2}_{123}+H^{2}_{124}+H^{2}_{134}+H^{2}_{234}\right)+4\left(R_{11}H^{2}_{234}+R_{22}H^{2}_{134}+R_{33}H^{2}_{124}+R_{44}H^{2}_{123}\right).

Using

(H2)11=2(H1232+H1242+H1342)=2(16|H|2H2342),(H^{2})_{11}=2(H^{2}_{123}+H^{2}_{124}+H^{2}_{134})=2\left(\frac{1}{6}|H|^{2}-H^{2}_{234}\right),

we have

(5.2) {i,j}={k,l}\displaystyle\sum_{\{i,j\}=\{k,l\}} =13R|H|2+23|H|2iRii2iRii(H2)ii\displaystyle=-\frac{1}{3}R|H|^{2}+\frac{2}{3}|H|^{2}\sum_{i}R_{ii}-2\sum_{i}R_{ii}(H^{2})_{ii}
=13R|H|22Ric,H2.\displaystyle=\frac{1}{3}R|H|^{2}-2\left<\mathrm{Ric},H^{2}\right>.
(5.3) {i,j}{k,l}=8\displaystyle\sum_{\{i,j\}\neq\{k,l\}}=8 [(R1213+R2434)H124H134+(R1214+R2334)H123H143+(R1223+R1434)H124H234\displaystyle\left[(R_{1213}+R_{2434})H_{124}H_{134}+(R_{1214}+R_{2334})H_{123}H_{143}+(R_{1223}+R_{1434})H_{124}H_{234}\right.
(R1224+R1334)H123H243+(R1314+R2324)H132H142+(R1323+R1424)H134H234]\displaystyle\left.(R_{1224}+R_{1334})H_{123}H_{243}+(R_{1314}+R_{2324})H_{132}H_{142}+(R_{1323}+R_{1424})H_{134}H_{234}\right]
=8\displaystyle=8 (R23H124H134R24H123H143+R13H124H234+R14H123H243\displaystyle\left(-R_{23}H_{124}H_{134}-R_{24}H_{123}H_{143}+R_{13}H_{124}H_{234}+R_{14}H_{123}H_{243}\right.
R34H132H142R12H134H234)\displaystyle\ \ \left.-R_{34}H_{132}H_{142}-R_{12}H_{134}H_{234}\right)
=0\displaystyle=0 .

Combine (5.2) and (5.3), we have

RH=13R|H|22Ric,H2.R_{H}=\frac{1}{3}R|H|^{2}-2\left<\mathrm{Ric},H^{2}\right>.

For the suitable orthonormal basis in Proposition 5.7, XH=e4X_{H}=e_{4}.

Ric,H2=iRii(H2)ii=13|H|2(R11+R22+R33)=13R|H|213R44|H|2,\left<\mathrm{Ric},H^{2}\right>=\sum_{i}R_{ii}(H^{2})_{ii}=\frac{1}{3}|H|^{2}\left(R_{11}+R_{22}+R_{33}\right)=\frac{1}{3}R|H|^{2}-\frac{1}{3}R_{44}|H|^{2},
RH=13R|H|2+23Ric(XH)|H|2.R_{H}=-\frac{1}{3}R|H|^{2}+\frac{2}{3}\mathrm{Ric}(X_{H})|H|^{2}.

Recently, an interesting bound of torsion is found in the case of the pluriclosed flow on complex surface with special initial data, see [Ye23].

Theorem 5.9 ([Ye23, Theorem 1.3]).

For a compact complex surface (M4,J,g(t))(M^{4},J,g(t)) that admits a pluriclosed flow with Hermitian-symplectic initial data on [0,τ)[0,\tau), if the Chern scalar curvature satisfies

RC(x,t)K,xM,t[0,τ),\displaystyle R^{C}(x,t)\leqslant K,\ x\in M,t\in[0,\tau),

then we have

|H(x,t)|g(x,t)2max{2K,maxx|H(x,0)|g(x,0)2},xM,t[0,τ).\displaystyle|H(x,t)|^{2}_{g(x,t)}\leqslant\max\left\{2K,\max_{x}|H(x,0)|^{2}_{g(x,0)}\right\},\ x\in M,t\in[0,\tau).

References