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

The Ricci iteration towards cscK metrics

Kewei Zhang School of Mathematical Sciences, Beijing Normal University, Beijing, 100875, People’s Republic of China. [email protected] Dedicated to Gang Tian on the occasion of his 65th65^{th} birthday
Abstract.

Motivated by the problem of finding constant scalar curvature Kähler metrics, we investigate a Ricci iteration sequence of Rubinstein that discretizes the pseudo-Calabi flow. While the long time existence of the flow is still an open question, we show that the iteration sequence does exist for all steps, along which the K-energy decreases. We further show that the iteration sequence, modulo automorphisms, converges smoothly to a constant scalar curvature Kähler metric if there is one, thus confirming a conjecture of Rubinstein from 2007 and extending results of Darvas–Rubinstein to arbitrary Kähler classes.

1. Introduction

A long standing problem in Kähler geometry is to find constant scalar curvature Kähler (cscK) metrics in a given Kähler class. Namely, for a compact Kähler manifold (X,ω)(X,\omega) of dimension nn, we want to search for a Kähler form ω{ω}\omega^{*}\in\{\omega\} that satisfies

trωRic(ω)=R¯,\mathrm{tr}_{\omega^{*}}\operatorname{Ric}(\omega^{*})=\bar{R},

where R¯:=2πnKX{ω}n1{ω}n\bar{R}:=2\pi n\frac{-K_{X}\cdot\{\omega\}^{n-1}}{\{\omega\}^{n}} is the average of the scalar curvature.

Regarding the existence of such metrics, the influential Tian’s properness conjecture (cf. [41, Remark 5.2],[38, Conjecture 7.12]) predicts that the existence of cscK metrics is equivalent to some suitable notion of properness of Mabuchi’s K-energy functional. Tian’s conjecture is central in Kähler geometry and has attracted much work over the past two decades including motivating much work on equivalence between algebro-geometric notions of stability and existence of canonical metrics, as well as on the interface of pluripotential theory and Monge–-Ampère equations. We refer to the surveys [37, 31, 39, 28, 35].

In [20], using the Finsler geometry of the space of Kähler metrics, the authors reduce Tian’s conjecture to a purely PDE regularity problem, which has been recently solved in [13]. Therefore we now have complete solution to Tian’s properness conjecture.

On the other hand, provided the properness of the K-energy, how to produce a cscK metric is also a challenging problem in its own right. In [13] the authors show that certain continuity path provides an approach towards cscK metrics. In this work we show that one can also produce a cscK metric using some dynamical system.

To present our results, we begin by recalling an elementary result in Kähler geometry.

Lemma 1.1.

A closed (1,1)(1,1)-form θ\theta on (X,ω)(X,\omega) satisfies trωθ=Const.\mathrm{tr}_{\omega}\theta=\mathrm{Const.} if and only if θ\theta is harmonic with respect to the Kähler metric ω\omega.

Therefore, ω\omega^{*} is cscK if and only if Ric(ω)\operatorname{Ric}(\omega^{*}) is a harmonic form with respect to ω\omega^{*}. This viewpoint is by no means new, which was explored in early works of Calabi, Futaki, Bando and Mabuchi; see, e.g., [2]. When combined with the framework of geometric flows, this motivates one to consider the following variant of the Kähler Ricc flow:

(1.1) tωt=Ric(ωt)+HRic(ωt),ω0=ω.\partial_{t}\omega_{t}=-\operatorname{Ric}(\omega_{t})+\mathrm{HRic}(\omega_{t}),\ \omega_{0}=\omega.

Here, given any Kähler form α\alpha, HRic(α)\mathrm{HRic}(\alpha) denotes the harmonic part of Ric(α)\operatorname{Ric}(\alpha) with respect to α\alpha. If the flow (1.1) smoothly converges to a limit ω\omega_{\infty}, then one has

Ric(ω)=HRic(ω),\operatorname{Ric}(\omega_{\infty})=\mathrm{HRic}(\omega_{\infty}),

namely, ω\omega_{\infty} is a cscK metric.

Remark 1.2.

If 2πc1(X)=λ{ω}2\pi c_{1}(X)=\lambda\{\omega\}, then the flow (1.1) reduces to

tωt=Ric(ωt)+λωt,\partial_{t}\omega_{t}=-\operatorname{Ric}(\omega_{t})+\lambda\omega_{t},\

which is exactly the classical normalized Kähler Ricci flow in the study of Kähler–Einstein and Kähler Ricci soliton metrics (see e.g. [11, 45, 29, 46, 42, 43, 1, 14]). By [11, 44] we know that such flow has long time existence.

The flow (1.1) first appeared in [21] (see also [36] for a related flow) and was then briefly studied in [34, 33]. Later in [15], this flow was systematically investigated and the authors call it the pseudo-Calabi flow. Note that the flow (1.1) can be viewed as a variant of the Calabi flow [10], and it can be reduced to a coupled system of parabolic equations. Indeed, it is easy to show that (1.1) is equivalent to the following coupled equations:

{ωtn=eFtωn,Ft˙=ΔtFt+R¯trωtRic(ω).\begin{cases}\omega_{t}^{n}=e^{F_{t}}\omega^{n},\\ \dot{F_{t}}=\Delta_{t}F_{t}+\bar{R}-\mathrm{tr}_{\omega_{t}}\operatorname{Ric}(\omega).\end{cases}

So we obtain a parabolic version of the coupled equations for cscK metrics that are studied in [12, 13]. This being said, it is still a highly non-trivial problem to study this flow, with its long time existence and limiting behavior largely open.

In this paper we adopt a somewhat different viewpoint. We consider the discretization of the pseudo-Calabi flow (1.1) that was first proposed by Rubinstein [34, 33]. More precisely, given τ>0\tau>0, we investigate the following Ricci iteration that appeared in [34, Definition 2.1] and [33, (41)]:

(1.2) ωi+1ωiτ=Ric(ωi+1)+HRic(ωi+1),i,ω0=ω.\frac{\omega_{i+1}-\omega_{i}}{\tau}=-\operatorname{Ric}(\omega_{i+1})+\mathrm{HRic}(\omega_{i+1}),\ i\in{\mathbb{N}},\ \omega_{0}=\omega.

Part of the interest in this Ricci iteration is that, clearly, cscK metrics are fixed points. Therefore (1.2) aims to provide a natural theoretical and numerical approach to uniformization in the challenging case of cscK metrics. In [34, Conjecture 2.1], Rubinstein proposed the following.

Conjecture 1.3.

Let XX be a compact Kähler manifold, and assume that there exists a constant scalar curvature Kähler metric in a Kähler class Ω\Omega. Then for any ωΩ\omega\in\Omega the Ricci iteration (1.2) exists for all ii\in{\mathbb{N}} and converges in an appropriate sense to a constant scalar curvature metric.

In addition, the Ricci iteration could be a source of new insights for the study of the flow (1.1), which is known to be a rather difficult problem in the field of geometric flows. For instance, just as in the case of Calabi flow, the long time existence of the flow (1.1) is still unknown (see [15, Conjecture 8.3]). However, after discretization, we can prove the following long time existence result for the sequence (1.2).

Theorem 1.4.

There exists a uniform constant τ0(0,]\tau_{0}\in(0,\infty], depending only on XX and the Kähler class {ω}\{\omega\}, such that for any τ(0,τ0)\tau\in(0,\tau_{0}) the iteration sequence (1.2) exists for all ii\in{\mathbb{N}}, with each ωi\omega_{i} being uniquely determined by ω0\omega_{0}.

This result gives a strong evidence for the long time existence of the flow (1.1). Indeed, sending τ0\tau\to 0, the iteration sequence {ωi}\{\omega_{i}\} is expected to converge to the flow (1.1) (this is interestingly still a conjecture even for the Ricci iteration associated to the Kähler Ricci flow; compare the classical Rothe’s method for parabolic equations). The fact that the sequence {ωi}\{\omega_{i}\} exists for all ii should imply that the flow (1.1) exists for all tt. This is of course a heuristic viewpoint, which hopefully can be made more rigorous in future study.

It is proved in [15, Theorem 3.1] that Mabuchi’s K-energy decreases along the flow (1.1). We show that this is also the case for the Ricci iteration (1.2), which was previously only known in the case where c1(X)=λ{ω}c_{1}(X)=\lambda\{\omega\} (see [33, Proposition 4.2]).

Theorem 1.5.

Along the iteration sequence {ωi}i\{\omega_{i}\}_{i\in{\mathbb{N}}}, the K-energy KωK_{\omega} satisfies

Kω(ωi+1)Kω(ωi) for all i.K_{\omega}(\omega_{i+1})\leq K_{\omega}(\omega_{i})\text{ for all }i\in{\mathbb{N}}.

The equality holds for some ii\in{\mathbb{N}} if and only if ωi=ω0\omega_{i}=\omega_{0} is cscK for all ii\in{\mathbb{N}}.

Since a cscK metric, if exists, minimizes the K-energy. The above result suggests that the iteration sequence (1.2) has the tendency to be attracted by a cscK metric in a suitable sense. We show that this is indeed the case, thus confirming Conjecture 1.3.

Theorem 1.6.

Assume that there exists a cscK metric in {ω}\{\omega\}. Then for any τ>0\tau>0 the iteration sequence ωi\omega_{i} exists, and, up to biholomorphic automorphisms, converges to a cscK metric smoothly.

Our results extend those in the previous works [34, 33, 26, 5, 19, 25], where c1(X)c_{1}(X) is assumed to be proportional to {ω}\{\omega\}. Moreover, Theorem 1.6 also gives strong evidence that the flow (1.1) shall converge to a cscK metric, if there is one in {ω}\{\omega\} (cf. [33, Conjecture 7.4] and [15, Question 8.5]). In view of Tian’s properness conjecture, Theorem 1.6 also shows that the properness of K-energy (modulo group actions, in the sense of Definition 3.13) implies that one can find a cscK metrics using the dynamical system (1.2).

Compared to the recent work of Darvas–Rubinstein [19] in the Fano case, the main difficulty we are faced with is the lack of Ding functional in our general setting. As we shall see, this technicality can be circumvented with the help of the estimates in [13, §3], which are needed for the smooth convergence in Theorem 1.6.

For the direction of Ricci iteration in the real case, we refer the reader to [32, 25, 9]. See also [22] for a Ricci iteration in the local setting.

Organization. After recalling some standard notions and facts in §2 and §3, we prove Theorem 1.4 and Theorem 1.5 in §4. Relying on [13, §3], we will derive some a priori estimates for the Ricci iteration in §5, which allows us to prove Theorem 1.6 in §6. Finally in §7 we point out that our work can be extended to the setting of twisted cscK metrics.

Acknowledgements. It is the author’s great pleasure to dedicate this paper to Prof. Tian, on the occasion of his 65th birthday, for his constant guidance over the years. The author is also grateful to T. Darvas, W. Jian, Y. Rubinstein and Y. Shi for helpful discussions and comments. Part of this work was done during the pleasant and inspiring visit at the Tianyuan Mathematics Research center in Oct. 2023.

The author is supported by NSFC grants 12101052, 12271040, and 12271038.

2. Energy functionals

We recall several standard functionals that will be used throughout this paper.

Let (X,ω)(X,\omega) be a compact Kähler manifold of dimension nn, and set

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

Put V:=Xωn.V:=\int_{X}\omega^{n}. And let

Ric(ωφ):=1¯logdetωφ2πc1(X),R(ωφ):=trωφRic(ωφ),\operatorname{Ric}(\omega_{\varphi}):=-\sqrt{-1}\partial\bar{\partial}\log\det\omega_{\varphi}\in 2\pi c_{1}(X),\ R(\omega_{\varphi}):=\mathrm{tr}_{\omega_{\varphi}}\operatorname{Ric}(\omega_{\varphi}),
R¯:=1VXR(ω)ωn=nVXRic(ω)ωn1=2πnc1(X){ω}n1{ω}n.\bar{R}:=\frac{1}{V}\int_{X}R(\omega)\omega^{n}=\frac{n}{V}\int_{X}\operatorname{Ric}(\omega)\wedge\omega^{n-1}=2\pi n\frac{c_{1}(X)\cdot\{\omega\}^{n-1}}{\{\omega\}^{n}}.

For any u,vωu,v\in\mathcal{H}_{\omega}, define

I(u,v)=I(ωu,ωv):=1VX(vu)(ωunωvn).I(u,v)=I(\omega_{u},\omega_{v}):=\frac{1}{V}\int_{X}(v-u)(\omega^{n}_{u}-\omega^{n}_{v}).
E(u,v):=1(n+1)VX(vu)i=0nωuiωvni.E(u,v):=\frac{1}{(n+1)V}\int_{X}(v-u)\sum_{i=0}^{n}\omega^{i}_{u}\wedge\omega^{n-i}_{v}.
J(u,v)=J(ωu,ωv):=1VX(vu)ωunE(u,v).J(u,v)=J(\omega_{u},\omega_{v}):=\frac{1}{V}\int_{X}(v-u)\omega^{n}_{u}-E(u,v).
Ent(u,v)=Ent(ωu,ωv):=1VXlogωvnωunωvn.Ent(u,v)=Ent(\omega_{u},\omega_{v}):=\frac{1}{V}\int_{X}\log\frac{\omega^{n}_{v}}{\omega^{n}_{u}}\omega^{n}_{v}.

Note that by Jensen’s inequality, it always holds that Ent(u,v)0Ent(u,v)\geq 0.

For any closed (1,1)(1,1) form χ\chi, define

𝒥χ(u,v):=1VX(vu)χi=0n1ωuiωvn1iχ¯E(u,v),\mathcal{J}^{\chi}(u,v):=\frac{1}{V}\int_{X}(v-u)\chi\wedge\sum_{i=0}^{n-1}\omega_{u}^{i}\wedge\omega^{n-1-i}_{v}-\bar{\chi}E(u,v),

where

χ¯:=nVXχωn1=n{χ}{ω}n1{ω}n.\bar{\chi}:=\frac{n}{V}\int_{X}\chi\wedge\omega^{n-1}=n\frac{\{\chi\}\cdot\{\omega\}^{n-1}}{\{\omega\}^{n}}.

The K-energy is defined by

K(u,v)=K(ωu,ωv):=Ent(u,v)+𝒥Ric(ωu)(u,v).K(u,v)=K(\omega_{u},\omega_{v}):=Ent(u,v)+\mathcal{J}^{-\operatorname{Ric}(\omega_{u})}(u,v).

The χ\chi-twisted K-energy is

Kχ(u,v)=Kχ(ωu,ωv):=K(u,v)+𝒥χ(u,v).K^{\chi}(u,v)=K^{\chi}(\omega_{u},\omega_{v}):=K(u,v)+\mathcal{J}^{\chi}(u,v).

If we choose χ:=ωu\chi:=\omega_{u}, then integration by parts gives

(2.1) 𝒥ωu(u,v)=(IJ)(u,v).\mathcal{J}^{\omega_{u}}(u,v)=(I-J)(u,v).

More generally, if χ:=ωw\chi:=\omega_{w} for some wωw\in\mathcal{H}_{\omega}, then

𝒥ωw(u,v)=(IJ)(u,v)+1VX(wu)(ωvnωun),\mathcal{J}^{\omega_{w}}(u,v)=(I-J)(u,v)+\frac{1}{V}\int_{X}(w-u)(\omega^{n}_{v}-\omega^{n}_{u}),

so in particular we have that

(2.2) 𝒥ωw(u,w)=J(u,w).\mathcal{J}^{\omega_{w}}(u,w)=-J(u,w).

One has the following variation formulas (for any u,vωu,v\in\mathcal{H}_{\omega} and fC(X,)f\in C^{\infty}(X,\mathbb{R})):

(2.3) {ddt|t=0E(u,v+tf)=1VXfωvn,ddt|t=0𝒥χ(u,v+tf)=1VXf(trωvχχ¯)ωvn,ddt|t=0Kχ(u,v+tf)=1VXf(R¯χ¯+trωvχR(ωv))ωvn.\begin{cases}\frac{d}{dt}\big{|}_{t=0}E(u,v+tf)=\frac{1}{V}\int_{X}f\omega^{n}_{v},\\ \frac{d}{dt}\big{|}_{t=0}\mathcal{J}^{\chi}(u,v+tf)=\frac{1}{V}\int_{X}f(\mathrm{tr}_{\omega_{v}}\chi-\bar{\chi})\omega^{n}_{v},\\ \frac{d}{dt}\big{|}_{t=0}K^{\chi}(u,v+tf)=\frac{1}{V}\int_{X}f(\bar{R}-\bar{\chi}+\mathrm{tr}_{\omega_{v}}\chi-R(\omega_{v}))\omega^{n}_{v}.\\ \end{cases}

They imply the well known cocycle relations (for u,v,wωu,v,w\in\mathcal{H}_{\omega}):

E(u,v)+E(v,w)=E(u,w).E(u,v)+E(v,w)=E(u,w).
𝒥χ(u,v)+𝒥χ(v,w)=𝒥χ(u,w).\mathcal{J}^{\chi}(u,v)+\mathcal{J}^{\chi}(v,w)=\mathcal{J}^{\chi}(u,w).
Kχ(u,v)+Kχ(v,w)=Kχ(u,w).K^{\chi}(u,v)+K^{\chi}(v,w)=K^{\chi}(u,w).

One can then further deduce the following cocycle relations:

J(u,v)+J(v,w)=J(u,w)+1VX(vw)(ωunωvn).J(u,v)+J(v,w)=J(u,w)+\frac{1}{V}\int_{X}(v-w)(\omega^{n}_{u}-\omega^{n}_{v}).
(IJ)(u,v)+(IJ)(v,w)=(IJ)(u,w)+1VX(vu)(ωwnωvn).(I-J)(u,v)+(I-J)(v,w)=(I-J)(u,w)+\frac{1}{V}\int_{X}(v-u)(\omega^{n}_{w}-\omega^{n}_{v}).

The following result proved in Tian’s work [40] will be used repeatedly.

Lemma 2.1.

For any u,vωu,v\in\mathcal{H}_{\omega}, it holds that

1nJ(u,v)(IJ)(u,v)nJ(u,v).\frac{1}{n}J(u,v)\leq(I-J)(u,v)\leq nJ(u,v).

Moreover, one has

I(u,v)0,J(u,v)0,(IJ)(u,v)0.I(u,v)\geq 0,\ J(u,v)\geq 0,\ (I-J)(u,v)\geq 0.

If one of them takes equality, then they all do, in which case u=vu=v.

Lemma 2.2.

For any u,v,wωu,v,w\in\mathcal{H}_{\omega} one has

𝒥ωw(u,v)𝒥ωw(u,w)=J(v,w).\mathcal{J}^{\omega_{w}}(u,v)-\mathcal{J}^{\omega_{w}}(u,w)=J(v,w).

So in particular, 𝒥ωw(u,v)𝒥ωw(u,w)\mathcal{J}^{\omega_{w}}(u,v)\geq\mathcal{J}^{\omega_{w}}(u,w), and the equality holds if and only if v=wv=w.

Proof.

Using cocycle relation, we can write (recall (2.2))

𝒥ωw(u,v)𝒥ωw(u,w)=𝒥ωw(v,w)=J(v,w)\displaystyle\mathcal{J}^{\omega_{w}}(u,v)-\mathcal{J}^{\omega_{w}}(u,w)=-\mathcal{J}^{\omega_{w}}(v,w)=J(v,w)

to conclude. ∎

Convention. Given an energy functional F{I,J,𝒥χ,K,Kχ}F\in\{I,J,\mathcal{J}^{\chi},K,K^{\chi}\} and uωu\in\mathcal{H}_{\omega}, we also use the notation

Fω(u)=Fω(ωu):=F(0,u) and Eω(u):=E(0,u)F_{\omega}(u)=F_{\omega}(\omega_{u}):=F(0,u)\text{ and }E_{\omega}(u):=E(0,u)

in the circumstances where ω\omega is viewed as a background metric.

Definition 2.3.

The twisted K-energy KωχK^{\chi}_{\omega} is said to be proper if there exist γ>0\gamma>0 and C>0C>0 such that

Kωχ(u)γ(IωJω)(u)C for all uω.K^{\chi}_{\omega}(u)\geq\gamma(I_{\omega}-J_{\omega})(u)-C\text{ for all }u\in\mathcal{H}_{\omega}.

3. The metric completion

We will work with the finite energy space ω1\mathcal{E}_{\omega}^{1} introduced in [23] and use the d1d_{1}-distance on it introduced by Darvas [18]. They provide useful tools for proving our main result concerning convergence of the Ricci iteration. We briefly recall the machinery, referring to [17] and references therein for more details.

Let

PSHω:={φL1(ωn):φ is upper semi-continuous and ωφ=ω+1¯φ0}.\mathrm{PSH}_{\omega}:=\{\varphi\in L^{1}(\omega^{n}):\varphi\text{ is upper semi-continuous and }\omega_{\varphi}=\omega+\sqrt{-1}\partial\bar{\partial}\varphi\geq 0\}.

Following Guedj–Zeriahi [23, Definition 1.1] we define the subset of full mass potentials:

ω:={φPSHω:limj{φj}(ω+1¯max{φ,j})n=0}.\mathcal{E}_{\omega}:=\big{\{}\varphi\in\mathrm{PSH}_{\omega}:\lim_{j\to-\infty}\int_{\{\varphi\leq j\}}(\omega+\sqrt{-1}\partial\bar{\partial}\max\{\varphi,j\})^{n}=0\big{\}}.

For each φω\varphi\in\mathcal{E}_{\omega}, define ωφn:=limj1{φ>j}(ω+1¯max{φ,j})n\omega^{n}_{\varphi}:=\lim_{j\to-\infty}\textbf{1}_{\{\varphi>j\}}(\omega+\sqrt{-1}\partial\bar{\partial}\max\{\varphi,j\})^{n}. The measure (ω+1¯max{φ,j})n(\omega+\sqrt{-1}\partial\bar{\partial}\max\{\varphi,j\})^{n} is defined by the work of Beford–Taylor [3] since max{φ,j}\max\{\varphi,j\} is bounded. Consequently, φω\varphi\in\mathcal{E}_{\omega} if and only if Xωφn=Xωn\int_{X}\omega^{n}_{\varphi}=\int_{X}\omega^{n}, justifying the name of ω\mathcal{E}_{\omega}.

Next, define a further subset, the space of finite 11-energy potentials:

ω1:={φω:X|φ|ωφn<}.\mathcal{E}_{\omega}^{1}:=\big{\{}\varphi\in\mathcal{E}_{\omega}:\int_{X}|\varphi|\omega^{n}_{\varphi}<\infty\big{\}}.

Consider the following weak Finsler metric on ω\mathcal{H}_{\omega} [18]:

ξφ:=1VX|ξ|ωφn,ξTφω=C(X,).||\xi||_{\varphi}:=\frac{1}{V}\int_{X}|\xi|\omega^{n}_{\varphi},\ \xi\in T_{\varphi}\mathcal{H}_{\omega}=C^{\infty}(X,\mathbb{R}).

We denote by d1d_{1} the associated pseudo-metric and recall the following result characterizing the d1d_{1}-metric completion of ω\mathcal{H}_{\omega} [18, Theorem 3.5]:

Theorem 3.1.

(ω,d1)(\mathcal{H}_{\omega},d_{1}) is a metric space whose completion can be identified with (ω1)(\mathcal{E}_{\omega}^{1}), where

d1(u0,u1):=limkd1(u0k,u1k),d_{1}(u_{0},u_{1}):=\lim_{k\to\infty}d_{1}(u_{0}^{k},u_{1}^{k}),

for any smooth decreasing sequences {uik}kω\{u_{i}^{k}\}_{k\in{\mathbb{N}}}\subset\mathcal{H}_{\omega} converging pointwise to uiω1u_{i}\in\mathcal{E}^{1}_{\omega}, i=0,1i=0,1.

Let us now recall several properties of the d1d_{1} metric that will be used in this paper.

Lemma 3.2.

([18, Theorem 3]) For any u,vω1u,v\in\mathcal{E}^{1}_{\omega}, one has

1Cd1(u,v)X|uv|(ωun+ωvn)Cd1(u,v)\frac{1}{C}d_{1}(u,v)\leq\int_{X}|u-v|(\omega^{n}_{u}+\omega^{n}_{v})\leq Cd_{1}(u,v)

for some dimensional constant C>1C>1.

Lemma 3.3.

([20, Proposition 5.5]) There exists a constant C>1C>1 depending only on (X,ω)(X,\omega) such that

1CJω(φ)Cd1(0,φ)CJω(φ)+C for any φ0.\frac{1}{C}J_{\omega}(\varphi)-C\leq d_{1}(0,\varphi)\leq CJ_{\omega}(\varphi)+C\text{ for any }\varphi\in\mathcal{H}_{0}.

One can extend the functionals Entω,Iω,Jω,EωEnt_{\omega},I_{\omega},J_{\omega},E_{\omega} and 𝒥ωχ\mathcal{J}^{\chi}_{\omega} to the space ω1\mathcal{E}^{1}_{\omega}.

Lemma 3.4.

([7]) All the functionals Iω,Jω,Eω,𝒥ωχI_{\omega},J_{\omega},E_{\omega},\mathcal{J}_{\omega}^{\chi} are d1d_{1}-continuous (lsc). The entropy EntωEnt_{\omega} is d1d_{1}-lower semi-continuous. Moreover, for any uω1u\in\mathcal{E}^{1}_{\omega}, there exists ωuid1u\mathcal{H}_{\omega}\ni u_{i}\xrightarrow{d_{1}}u such that Entω(ui)Entω(u)Ent_{\omega}(u_{i})\to Ent_{\omega}(u).

We need the following compactness result going back to [5] (see [20, Theorem 5.6] for a convenient formulation for our context).

Lemma 3.5.

For any A>0A>0, the set

{φω1|d1(0,φ)A and Entω(φ)A}\{\varphi\in\mathcal{E}^{1}_{\omega}|d_{1}(0,\varphi)\leq A\text{ and }Ent_{\omega}(\varphi)\leq A\}

is compact in ω1\mathcal{E}^{1}_{\omega} with respect to the d1d_{1}-topology.

For any u,vω1u,v\in\mathcal{E}^{1}_{\omega}, one can also define

I(u,v):=1VX(vu)(ωunωvn),I(u,v):=\frac{1}{V}\int_{X}(v-u)(\omega^{n}_{u}-\omega^{n}_{v}),
E(u,v):=Eω(v)Eω(u),E(u,v):=E_{\omega}(v)-E_{\omega}(u),

and

J(u,v):=1VX(vu)ωunE(u,v).J(u,v):=\frac{1}{V}\int_{X}(v-u)\omega^{n}_{u}-E(u,v).

Recall that (see [17, Proposition 3.40])

(3.1) |E(u,v)|d1(u,v).|E(u,v)|\leq d_{1}(u,v).
Lemma 3.6.

Given ui,u,vi,vω1u_{i},u,v_{i},v\in\mathcal{E}^{1}_{\omega} such that uid1uu_{i}\xrightarrow{d_{1}}u and vid1vv_{i}\xrightarrow{d_{1}}v, then

limiI(ui,vi)=I(u,v),limiJ(ui,vi)=J(u,v).\lim_{i\to\infty}I(u_{i},v_{i})=I(u,v),\ \lim_{i\to\infty}J(u_{i},v_{i})=J(u,v).
Proof.

We only deal with the JJ-functional, since the proof for the II-functional is similar. One has

|J(ui,vi)J(u,v)|\displaystyle|J(u_{i},v_{i})-J(u,v)| |1VX(viui)ωuin1VX(vu)ωun|+|E(u,ui)|+|E(v,vi)|\displaystyle\leq|\frac{1}{V}\int_{X}(v_{i}-u_{i})\omega^{n}_{u_{i}}-\frac{1}{V}\int_{X}(v-u)\omega^{n}_{u}|+|E(u,u_{i})|+|E(v,v_{i})|

By (3.1) it suffices to estimate |X(viui)ωuinX(vu)ωun||\int_{X}(v_{i}-u_{i})\omega^{n}_{u_{i}}-\int_{X}(v-u)\omega^{n}_{u}|, which can be bounded from above by

|X(viv)\displaystyle|\int_{X}(v_{i}-v) (ωuinωun)|+|X(uiu)(ωuinωun)|+X(|viv|+|uiu|)ωun\displaystyle(\omega^{n}_{u_{i}}-\omega^{n}_{u})|+|\int_{X}(u_{i}-u)(\omega^{n}_{u_{i}}-\omega^{n}_{u})|+\int_{X}(|v_{i}-v|+|u_{i}-u|)\omega^{n}_{u}
+|Xv(ωuinωun)|+|Xu(ωuinωun)|.\displaystyle+|\int_{X}v(\omega^{n}_{u_{i}}-\omega^{n}_{u})|+|\int_{X}u(\omega^{n}_{u_{i}}-\omega^{n}_{u})|.

All of these terms go to zero, thanks to [17, Proposition 3.48 and Corollary 3.51] (see also [6, Lemma 3.13, Lemma 5.8]). ∎

The next quasi-triangle inquality is proved in [5, Theorem 1.8].

Lemma 3.7.

There exists a dimensional constant Cn>0C_{n}>0 such that for any u,v,wω1u,v,w\in\mathcal{E}^{1}_{\omega}

I(u,v)Cn(I(u,w)+I(w,v)).I(u,v)\leq C_{n}(I(u,w)+I(w,v)).

We will need the following convergence criterion, which is a simple consequence of [5, Proposition 2.3].

Lemma 3.8.

Assume that ui,uω1u_{i},u\in\mathcal{E}^{1}_{\omega} such that Eω(ui)=Eω(u)=0E_{\omega}(u_{i})=E_{\omega}(u)=0 and d1(0,ui)Ad_{1}(0,u_{i})\leq A for some A>0A>0 independent of ii. Then

uid1uI(ui,u)0.u_{i}\xrightarrow{d_{1}}u\Leftrightarrow I(u_{i},u)\to 0.
Proof.

That uid1uu_{i}\xrightarrow{d_{1}}u implies I(ui,u)0I(u_{i},u)\to 0 follows from Lemma 3.2.

Now assume that I(ui,u)0I(u_{i},u)\to 0. By Lemma 3.2 the bound on d1(0,ui)d_{1}(0,u_{i}) implies the L1L^{1} bound for ui,u_{i}, which in turn implies a bound on |supXui||\sup_{X}u_{i}|. Put ui:=uisupXuiu_{i}^{\prime}:=u_{i}-\sup_{X}u_{i} and u:=usupXuu^{\prime}:=u-\sup_{X}u. Then I(ui,u)0I(u_{i}^{\prime},u^{\prime})\to 0 as well. By [5, Proposition 2.3] we know that uid1uu_{i}^{\prime}\xrightarrow{d_{1}}u^{\prime} and hence Eω(ui)=supXuiEω(u)=supXuE_{\omega}(u_{i}^{\prime})=-\sup_{X}u_{i}\to E_{\omega}(u^{\prime})=-\sup_{X}u. Then from supXuisupXu\sup_{X}u_{i}\to\sup_{X}u and uid1uu_{i}^{\prime}\xrightarrow{d_{1}}u^{\prime} we deduce that uid1uu_{i}\xrightarrow{d_{1}}u (using Lemma 3.2). So we conclude. ∎

Corollary 3.9.

Given two sequences ui,viω1u_{i},v_{i}\in\mathcal{E}^{1}_{\omega} with Eω(ui)=Eω(vi)=0E_{\omega}(u_{i})=E_{\omega}(v_{i})=0. Assume that uid1uu_{i}\xrightarrow{d_{1}}u and I(ui,vi)0I(u_{i},v_{i})\to 0. Then vid1uv_{i}\xrightarrow{d_{1}}u as well.

Proof.

First, that uid1uu_{i}\xrightarrow{d_{1}}u implies that I(0,ui)CI(0,u_{i})\leq C by Lemma 3.2. So one has

J(0,vi)I(0,vi)Cn(I(0,ui)+I(ui,vi))C.J(0,v_{i})\leq I(0,v_{i})\leq C_{n}(I(0,u_{i})+I(u_{i},v_{i}))\leq C^{\prime}.

This implies that d1(0,vi)C′′d_{1}(0,v_{i})\leq C^{\prime\prime} by Lemma 3.3. Moreover, Lemma 3.7 and 3.2 imply that

I(vi,u)Cn(I(vi,ui)+I(ui,u))0.I(v_{i},u)\leq C_{n}(I(v_{i},u_{i})+I(u_{i},u))\to 0.

So we conclude from the previous lemma that vid1uv_{i}\xrightarrow{d_{1}}u. ∎

We will frequently use the space

0:={φω|Eω(φ)=0}.\mathcal{H}_{0}:=\{\varphi\in\mathcal{H}_{\omega}|E_{\omega}(\varphi)=0\}.

Recall that,

G:=Aut0(X),G:=Aut_{0}(X),

the connected component of complex Lie group of holomorphic automorphisms of XX, naturally acts on 0\mathcal{H}_{0}.

Lemma 3.10.

([20, Lemma 5.8]) For any φ0\varphi\in\mathcal{H}_{0} and fGf\in G, let f.φ0f.\varphi\in\mathcal{H}_{0} be the unique potential such that fωφ=ωf.φf^{*}\omega_{\varphi}=\omega_{f.\varphi}. Then

f.φ=f.0+fφ.f.\varphi=f.0+f^{*}\varphi.
Lemma 3.11.

For any u,v0u,v\in\mathcal{H}_{0} and fGf\in G, one has

I(u,v)=I(f.u,f.v),J(u,v)=J(f.u,f.v).I(u,v)=I(f.u,f.v),\ J(u,v)=J(f.u,f.v).
Proof.

One has

I(u,v)\displaystyle I(u,v) =1VX(vu)(ωunωvn)\displaystyle=\frac{1}{V}\int_{X}(v-u)(\omega^{n}_{u}-\omega^{n}_{v})
=1VX(fvfu)(fωunfωvn)\displaystyle=\frac{1}{V}\int_{X}(f^{*}v-f^{*}u)(f^{*}\omega^{n}_{u}-f^{*}\omega^{n}_{v})
=I(fωu,fωv)=I(ωf.u,ωf.v)=I(f.u,f.v).\displaystyle=I(f^{*}\omega_{u},f^{*}\omega_{v})=I(\omega_{f.u},\omega_{f.v})=I(f.u,f.v).

For JJ-functional, we can write

J(u,v)\displaystyle J(u,v) =1VX(vu)ωun=1VX(fvfu)fωun\displaystyle=\frac{1}{V}\int_{X}(v-u)\omega^{n}_{u}=\frac{1}{V}\int_{X}(f^{*}v-f^{*}u)f^{*}\omega^{n}_{u}
=1V(fv+f.0fuf.0)ωf.un\displaystyle=\frac{1}{V}(f^{*}v+f.0-f^{*}u-f.0)\omega^{n}_{f.u}
=1VX(f.vf.u)ωf.un=J(f.u,f.v).\displaystyle=\frac{1}{V}\int_{X}(f.v-f.u)\omega^{n}_{f.u}=J(f.u,f.v).

Finally, we recall that GG acts on 0\mathcal{H}_{0} isometrically.

Lemma 3.12.

([20, Lemma 5.9]) For any u,v0u,v\in\mathcal{H}_{0} and fGf\in G one has

d1(u,v)=d1(f.u,f.v).d_{1}(u,v)=d_{1}(f.u,f.v).

Then define (as in [20])

d1,G(u,v):=inffGd1(u,f.v).d_{1,G}(u,v):=\inf_{f\in G}d_{1}(u,f.v).
Definition 3.13.

The K-energy KωK_{\omega} is said to be proper modulo GG if there exist γ>0\gamma>0 and C>0C>0 such that

Kω(u)γd1,G(0,u)C for all u0.K_{\omega}(u)\geq\gamma d_{1,G}(0,u)-C\text{ for all }u\in\mathcal{H}_{0}.

4. Basic properties of the Ricci iteration

In this part, we prove Theorem 1.4 and Theorem 1.5.

We begin by introducing the following analytic threshold.

(4.1) γ(X,{ω}):=sup{γ|infφω(Kω(φ)γ(IωJω)(φ))>},\gamma(X,\{\omega\}):=\sup\bigg{\{}\gamma\in\mathbb{R}\bigg{|}\inf_{\varphi\in\mathcal{H}_{\omega}}(K_{\omega}(\varphi)-\gamma(I_{\omega}-J_{\omega})(\varphi))>-\infty\bigg{\}},
Lemma 4.1.

The threshold γ(X,{ω})\gamma(X,\{\omega\}) is finite and is independent of the choice of ω\omega in its cohomology class, hence the notation.

Proof.

The finiteness of γ(X,{ω})\gamma(X,\{\omega\}) follows from

Kω(φ)𝒥ωRic(ω)(φ)Cd1(0,φ),φ0,K_{\omega}(\varphi)\geq\mathcal{J}_{\omega}^{-\operatorname{Ric}(\omega)}(\varphi)\geq-Cd_{1}(0,\varphi),\ \varphi\in\mathcal{H}_{0},

where we used [13, (4.2)]. So, by Lemma 3.3 and Lemma 2.1, one can find C0C\gg 0 such that Kω(φ)+C(IωJω)(φ)CK_{\omega}(\varphi)+C(I_{\omega}-J_{\omega})(\varphi)\geq-C for all φω\varphi\in\mathcal{H}_{\omega}.

To show that γ(X,{ω})\gamma(X,\{\omega\}) does not depend on the choice of ω\omega, we use the cocycle relations recalled in §2. It suffices to note the estimate

|(IJ)(u,w)(IJ)(v,w)||(IJ)(u,v)|+1VX|uv|(ωwn+ωvn)CuvC0,|(I-J)(u,w)-(I-J)(v,w)|\leq|(I-J)(u,v)|+\frac{1}{V}\int_{X}|u-v|(\omega^{n}_{w}+\omega^{n}_{v})\leq C||u-v||_{C^{0}},

for any u,v,wωu,v,w\in\mathcal{H}_{\omega}, where C>0C>0 is a dimensional constant. ∎

Proof of Theorem 1.4.

Taking trace of (1.2) we see that

(4.2) R(ωi+1)=R¯+trωi+1ωinτ,R(\omega_{i+1})=\bar{R}+\frac{\mathrm{tr}_{\omega_{i+1}}\omega_{i}-n}{\tau},

which is a twisted cscK equation. As we now argue, given any ωi{ω}\omega_{i}\in\{\omega\}, the existence of ωi+1{ω}\omega_{i+1}\in\{\omega\} solving (4.2) is guaranteed by the main result in [13] (see also [24]), once τ\tau is chosen to be small enough.

Indeed, for any γ<γ(X,{ω})\gamma<\gamma(X,\{\omega\}) and any Kähler form α{ω}\alpha\in\{\omega\}, the twisted K-energy Kαγ(IαJα)K_{\alpha}-\gamma(I_{\alpha}-J_{\alpha}) is proper by Lemma 4.1. Now choosing τ0(0,]\tau_{0}\in(0,\infty] so that 1/τ0γ(X,{ω})-1/\tau_{0}\leq\gamma(X,\{\omega\}), we see that for any τ(0,τ0)\tau\in(0,\tau_{0}) and ωi{ω}\omega_{i}\in\{\omega\}, the ωiτ\frac{\omega_{i}}{\tau}-twisted K-energy

Kωiωiτ=Kωi+𝒥ωiωiτ=Kωi+1τ(IωiJωi)K^{\frac{\omega_{i}}{\tau}}_{\omega_{i}}=K_{\omega_{i}}+\mathcal{J}^{\frac{\omega_{i}}{\tau}}_{\omega_{i}}=K_{\omega_{i}}+\frac{1}{\tau}(I_{\omega_{i}}-J_{\omega_{i}})

is proper (here we used (2.1)), which implies the solvability of (4.2) by [13, Theorem 4.1]. Moreover, such ωi+1\omega_{i+1} is uniquely determined, by [7, Theorem 4.13]. This completes the proof of Theorem 1.4. It is also clear that one can take τ0=\tau_{0}=\infty once γ(X,{ω})0\gamma(X,\{\omega\})\geq 0.

Proof of Theorem 1.5.

Notice that ωi+1\omega_{i+1} minimizes the twisted K-energy KωiωiτK_{\omega_{i}}^{\frac{\omega_{i}}{\tau}} (see [13, Corollary 4.5]), so that

Kωi(ωi+1)+1τ(IωiJωi)(ωi+1)=Kωiωiτ(ωi+1)Kωiωiτ(ωi)=0.K_{\omega_{i}}(\omega_{i+1})+\frac{1}{\tau}(I_{\omega_{i}}-J_{\omega_{i}})(\omega_{i+1})=K^{\frac{\omega_{i}}{\tau}}_{\omega_{i}}(\omega_{i+1})\leq K^{\frac{\omega_{i}}{\tau}}_{\omega_{i}}(\omega_{i})=0.

This implies that

Kω(ωi+1)Kω(ωi)=Kωi(ωi+1)1τ(IωiJωi)(ωi+1)0,K_{\omega}(\omega_{i+1})-K_{\omega}(\omega_{i})=K_{\omega_{i}}(\omega_{i+1})\leq-\frac{1}{\tau}(I_{\omega_{i}}-J_{\omega_{i}})(\omega_{i+1})\leq 0,

thanks to the cocycle property of the K-energy.

When the equality holds for some ii, one has (IωiJωi)(ωi+1)=0(I_{\omega_{i}}-J_{\omega_{i}})(\omega_{i+1})=0, which means that ωi=ωi+1\omega_{i}=\omega_{i+1}. Then (4.2) shows that ωi=ωi+1\omega_{i}=\omega_{i+1} are both cscK metrics. Moreover, from

R(ωi)=R¯+trωiωi1nτR(\omega_{i})=\bar{R}+\frac{\mathrm{tr}_{\omega_{i}}\omega_{i-1}-n}{\tau}

we get trωiωi1=n\mathrm{tr}_{\omega_{i}}\omega_{i-1}=n, and hence ωi1\omega_{i-1} is harmonic with respect to ωi\omega_{i}. This forces that ωi1=ωi\omega_{i-1}=\omega_{i}, by the uniqueness of harmonic forms. So we eventually see that ω0=ωi\omega_{0}=\omega_{i} for all ii, which is a fixed cscK metric. Thus we conclude Theorem 1.5. ∎

5. A priori estimates of the Ricci iteration

In this part we derive some a priori estimates for the iteration sequence (1.2). By Theorem 1.4 we can find some τ>0\tau>0 so that the iteration carries on forever. Up to scaling the Kähler class, we assume without loss of generality that τ=1\tau=1. Taking trace of (1.2) we then have

R(ωi+1)=R¯n+trωi+1ωi,ω0=ω.R(\omega_{i+1})=\bar{R}-n+\mathrm{tr}_{\omega_{i+1}}\omega_{i},\ \omega_{0}=\omega.

Write

ωi=ω+1¯ui,ui0.\omega_{i}=\omega+\sqrt{-1}\partial\bar{\partial}u_{i},u_{i}\in\mathcal{H}_{0}.

Also, let FiC(X,)F_{i}\in C^{\infty}(X,\mathbb{R}) be such that

(5.1) (ω+1¯ui)n=eFiωn.(\omega+\sqrt{-1}\partial\bar{\partial}u_{i})^{n}=e^{F_{i}}\omega^{n}.

Then

(5.2) ΔωiFi=trωi(Ric(ω)ωi1)+nR¯.\Delta_{\omega_{i}}F_{i}=\mathrm{tr}_{\omega_{i}}(\operatorname{Ric}(\omega)-\omega_{i-1})+n-\bar{R}.

In other words,

Δωi(Fi+ui1)=trωi(Ric(ω)ω)+nR¯.\Delta_{\omega_{i}}(F_{i}+u_{i-1})=\mathrm{tr}_{\omega_{i}}(\operatorname{Ric}(\omega)-\omega)+n-\bar{R}.

Therefore, we are in the situation considered in [13, §3].

We first derive the C0C^{0} bound for uiu_{i} and FiF_{i}.

Proposition 5.1.

Assume that there is some constant A>0A>0 such that

Entω(ui)+d1(0,ui)A for all i.Ent_{\omega}(u_{i})+d_{1}(0,u_{i})\leq A\text{ for all }i\in{\mathbb{N}}.

Then there exists some constant B1B_{1} depending only on X,ωX,\omega and AA such that

|Fi|+|ui|B1 for all i.|F_{i}|+|u_{i}|\leq B_{1}\text{ for all }i\in{\mathbb{N}}.
Proof.

First, using Lemma 3.2, the bound d1(0,ui)Ad_{1}(0,u_{i})\leq A implies that the uiu_{i} has uniform L1L^{1} bound, which in turn gives that

|supXui|C1 for all i,|\sup_{X}u_{i}|\leq C_{1}\text{ for all }i\in{\mathbb{N}},

where C1=C1(X,ω,A)>0C_{1}=C_{1}(X,\omega,A)>0. Moreover, for any p>0p>0, Zeriahi’s version of the Skoda–Tian type estimate [47] (see [17, Corollary 4.16] for a formulation that fits our context) implies that there exists C2=C2(X,ω,A,p)>0C_{2}=C_{2}(X,\omega,A,p)>0 such that

(5.3) XepuiωnC2 for all i.\int_{X}e^{-pu_{i}}\omega^{n}\leq C_{2}\text{ for all }i\in{\mathbb{N}}.

Then one can apply [13, Corollary 3.2] to find a constant C3=C3(X,ω,A)>0C_{3}=C_{3}(X,\omega,A)>0 such that

Fi+ui1C3,|ui|C3 for all i.F_{i}+u_{i-1}\leq C_{3},\ |u_{i}|\leq C_{3}\text{ for all }i\in{\mathbb{N}}.

Since u0=0u_{0}=0, we conclude by induction that there exists some C4=C4(X,ω,A)>0C_{4}=C_{4}(X,\omega,A)>0 such that

FiC4,|ui|C4 for all i.F_{i}\leq C_{4},\ |u_{i}|\leq C_{4}\text{ for all }i\in{\mathbb{N}}.

Then [13, Lemma 3.3] further implies that there exists some B1=B1(X,ω,A)>0B_{1}=B_{1}(X,\omega,A)>0 such that

|Fi|+|ui|B1 for all i.|F_{i}|+|u_{i}|\leq B_{1}\text{ for all }i\in{\mathbb{N}}.

Corollary 5.2.

Assume that there is some constant A>0A>0 such that

Entω(ui)+d1(0,ui)A for all i.Ent_{\omega}(u_{i})+d_{1}(0,u_{i})\leq A\text{ for all }i\in{\mathbb{N}}.

Then for any q>1q>1 there exists some constant B21B_{2}\geq 1 depending only on X,ω,qX,\omega,q and AA such that

n+Δωui1B2 for all i.n+\Delta_{\omega}u_{i}\geq\frac{1}{B_{2}}\text{ for all }i\in{\mathbb{N}}.

and

(5.4) X(n+Δωui)qωnB2 for all i.\int_{X}(n+\Delta_{\omega}u_{i})^{q}\omega^{n}\leq B_{2}\text{ for all }i\in{\mathbb{N}}.
Proof.

The first inequality follows from

n+ΔωuineF/nneB1/n.n+\Delta_{\omega}u_{i}\geq ne^{F/n}\geq ne^{-B_{1}/n}.

The second inequality follows from (5.3), Proposition 5.1 and [13, Corollary 3.4]. ∎

Proposition 5.3.

Assume that there is some constant A>0A>0 such that

Entω(ui)+d1(0,ui)A for all i.Ent_{\omega}(u_{i})+d_{1}(0,u_{i})\leq A\text{ for all }i\in{\mathbb{N}}.

Then there exists some constant B4B_{4} depending only on X,ωX,\omega and AA such that

maxX|ωi(Fi+ui1)|ωi2+maxX(n+Δωui)B4 for all i.\max_{X}|\nabla_{\omega_{i}}(F_{i}+u_{i-1})|^{2}_{\omega_{i}}+\max_{X}(n+\Delta_{\omega}u_{i})\leq B_{4}\text{ for all }i\in{\mathbb{N}}.
Proof.

The proof follows closely the one of [13, Theorem 3.2]. The basic idea is to estimate

Δωi(e12(Fi+ui1)|ωi(Fi+ui1))|ωi2+(n+Δωui))\Delta_{\omega_{i}}(e^{\frac{1}{2}(F_{i}+u_{i-1})}|\nabla_{\omega_{i}}(F_{i}+u_{i-1}))|^{2}_{\omega_{i}}+(n+\Delta_{\omega}u_{i}))

and then apply Nash–Moser iteration.

To simplify the notation, we put Δ:=Δω\Delta:=\Delta_{\omega}, and use the subscript ii to denote the operators associated with the metric ωi\omega_{i}, e.g., tri:=trωi\mathrm{tr}_{i}:=\mathrm{tr}_{\omega_{i}}, Δi:=Δωi\Delta_{i}:=\Delta_{\omega_{i}}. Also, put

wi:=Fi+ui1.w_{i}:=F_{i}+u_{i-1}.

So one has

Δiwi=tri(Ric(ω)ω)+nR¯.\Delta_{i}w_{i}=\mathrm{tr}_{i}(\operatorname{Ric}(\omega)-\omega)+n-\bar{R}.

In what follows, the constants C>0C>0 will change from line to line, which are all uniform (may depend on X,ω,AX,\omega,A, but are independent of ii).

Now we compute Δi(n+Δui)\Delta_{i}(n+\Delta u_{i}). As in [13, (3.25)], we have

Δi(n+Δui)\displaystyle\Delta_{i}(n+\Delta u_{i}) Ctriω(n+Δui)+ΔFiR(ω)\displaystyle\geq-C\mathrm{tr}_{i}\omega(n+\Delta u_{i})+\Delta F_{i}-R(\omega)
Ctriω(n+Δui)+Δwi(n+Δui1)C.\displaystyle\geq-C\mathrm{tr}_{i}\omega(n+\Delta u_{i})+\Delta w_{i}-(n+\Delta u_{i-1})-C.

Using (5.1) and Proposition 5.1 one can estimate

triωneFi(n+Δui)n1C(n+Δui)n1.\mathrm{tr}_{i}\omega\leq ne^{-F_{i}}(n+\Delta u_{i})^{n-1}\leq C(n+\Delta u_{i})^{n-1}.

Also, one can estimate Δwi\Delta w_{i}:

|Δwi|12C|(wi)kk¯|2(1+(ui)kk¯)2+C2(1+(ui)kk¯)212C|(wi)kk¯|2(1+(ui)kk¯)2+C2(n+Δui)2.|\Delta w_{i}|\leq\frac{1}{2C}\frac{|(w_{i})_{k\bar{k}}|^{2}}{(1+(u_{i})_{k\bar{k}})^{2}}+\frac{C}{2}(1+(u_{i})_{k\bar{k}})^{2}\leq\frac{1}{2C}\frac{|(w_{i})_{k\bar{k}}|^{2}}{(1+(u_{i})_{k\bar{k}})^{2}}+\frac{C}{2}(n+\Delta u_{i})^{2}.

So we get

(5.5) Δi(n+Δui)C(n+Δui)n12C|(wi)kk¯|2(1+(ui)kk¯)2C2(n+Δui)2(n+Δui1)C.\Delta_{i}(n+\Delta u_{i})\geq-C(n+\Delta u_{i})^{n}-\frac{1}{2C}\frac{|(w_{i})_{k\bar{k}}|^{2}}{(1+(u_{i})_{k\bar{k}})^{2}}-\frac{C}{2}(n+\Delta u_{i})^{2}-(n+\Delta u_{i-1})-C.

Next, we compute Δi(e12wi|iwi|i2)\Delta_{i}(e^{\frac{1}{2}w_{i}}|\nabla_{i}w_{i}|^{2}_{i}). As in [13, (3.43) and (3.49)], we have

Δi(e12wi|iwi|i2)\displaystyle\Delta_{i}(e^{\frac{1}{2}w_{i}}|\nabla_{i}w_{i}|^{2}_{i}) 2e12wiiwiiiΔiwi+e12wi|(wi)kl¯|2(1+(ui)kk¯)(1+(ui)ll¯)\displaystyle\geq 2e^{\frac{1}{2}w_{i}}\nabla_{i}w_{i}\cdot_{i}\nabla_{i}\Delta_{i}w_{i}+e^{\frac{1}{2}w_{i}}\frac{|(w_{i})_{k\bar{l}}|^{2}}{(1+(u_{i})_{k\bar{k}})(1+(u_{i})_{l\bar{l}})}
Ce12wi|iwi|2((n+Δui)2n1+(n+Δui)+1).\displaystyle-Ce^{\frac{1}{2}w_{i}}|\nabla_{i}w_{i}|^{2}\big{(}(n+\Delta u_{i})^{2n-1}+(n+\Delta u_{i})+1\big{)}.
2e12wiiwiiiΔiwi+1C|(wi)kl¯|2(1+(ui)kk¯)(1+(ui)ll¯)\displaystyle\geq 2e^{\frac{1}{2}w_{i}}\nabla_{i}w_{i}\cdot_{i}\nabla_{i}\Delta_{i}w_{i}+\frac{1}{C}\frac{|(w_{i})_{k\bar{l}}|^{2}}{(1+(u_{i})_{k\bar{k}})(1+(u_{i})_{l\bar{l}})}
Ce12wi|iwi|2((n+Δui)2n1+(n+Δui)+1).\displaystyle-Ce^{\frac{1}{2}w_{i}}|\nabla_{i}w_{i}|^{2}\big{(}(n+\Delta u_{i})^{2n-1}+(n+\Delta u_{i})+1\big{)}.

We used Proposition 5.1 in the last inequality. Note that, in the above estimate, the term involving |(wi)kl¯|2|(w_{i})_{k\bar{l}}|^{2} is dropped in [13, (3.49)] since it plays no role in loc. cit., but for our purpose we do need it to dominate the bad term |(wi)kk¯|2|(w_{i})_{k\bar{k}}|^{2} in (5.5).

Putting these estimates together, we then arrive at

Δi(e12wi|iwi|i2+(n+\displaystyle\Delta_{i}(e^{\frac{1}{2}w_{i}}|\nabla_{i}w_{i}|^{2}_{i}+(n+ Δui))2e12wiiwiiiΔiwiC((n+Δui)n+(n+Δui)2+1)\displaystyle\Delta u_{i}))\geq 2e^{\frac{1}{2}w_{i}}\nabla_{i}w_{i}\cdot_{i}\nabla_{i}\Delta_{i}w_{i}-C\big{(}(n+\Delta u_{i})^{n}+(n+\Delta u_{i})^{2}+1\big{)}
Ce12wi|iwi|2((n+Δui)2n1+(n+Δui)+1)(n+Δui1).\displaystyle-Ce^{\frac{1}{2}w_{i}}|\nabla_{i}w_{i}|^{2}\big{(}(n+\Delta u_{i})^{2n-1}+(n+\Delta u_{i})+1\big{)}-(n+\Delta u_{i-1}).

Putting

Ui:=e12wi|iwi|i2+(n+Δui)+1U_{i}:=e^{\frac{1}{2}w_{i}}|\nabla_{i}w_{i}|^{2}_{i}+(n+\Delta u_{i})+1

and using n+ΔuiB21n+\Delta u_{i}\geq B_{2}^{-1} we can further simplify to get

ΔiUi\displaystyle\Delta_{i}U_{i} 2e12wiiwiiiΔiwiCUi((n+Δui)2n1+1)(n+Δui1)\displaystyle\geq 2e^{\frac{1}{2}w_{i}}\nabla_{i}w_{i}\cdot_{i}\nabla_{i}\Delta_{i}w_{i}-CU_{i}\big{(}(n+\Delta u_{i})^{2n-1}+1\big{)}-(n+\Delta u_{i-1})
2e12wiiwiiiΔiwiCUi((n+Δui)2n1+(n+Δui1)+1),\displaystyle\geq 2e^{\frac{1}{2}w_{i}}\nabla_{i}w_{i}\cdot_{i}\nabla_{i}\Delta_{i}w_{i}-CU_{i}\big{(}(n+\Delta u_{i})^{2n-1}+(n+\Delta u_{i-1})+1\big{)},

where we used that Ui1U_{i}\geq 1. Put

G~i:=C((n+Δui)2n1+(n+Δui1)+1),\tilde{G}_{i}:=C\big{(}(n+\Delta u_{i})^{2n-1}+(n+\Delta u_{i-1})+1\big{)},

then we have the following key estimate:

ΔiUi2e12wiiwiiiΔiwiUiG~i.\Delta_{i}U_{i}\geq 2e^{\frac{1}{2}w_{i}}\nabla_{i}w_{i}\cdot_{i}\nabla_{i}\Delta_{i}w_{i}-U_{i}\tilde{G}_{i}.

Then for any p>1p>1, we obtain

X(p1)\displaystyle\int_{X}(p-1) Uip2|iUi|i2ωn=XUip1(ΔiUi)ωin\displaystyle U_{i}^{p-2}|\nabla_{i}U_{i}|_{i}^{2}\omega^{n}=\int_{X}U_{i}^{p-1}(-\Delta_{i}U_{i})\omega^{n}_{i}
XUipG~iωinX2Uip1e12wiiwiiiΔiwiωin.\displaystyle\leq\int_{X}U^{p}_{i}\tilde{G}_{i}\omega^{n}_{i}-\int_{X}2U^{p-1}_{i}e^{\frac{1}{2}w_{i}}\nabla_{i}w_{i}\cdot_{i}\nabla_{i}\Delta_{i}w_{i}\omega^{n}_{i}.

One can deal with the bad term e12wiiwiiiΔiwie^{\frac{1}{2}w_{i}}\nabla_{i}w_{i}\cdot_{i}\nabla_{i}\Delta_{i}w_{i} using integration by parts:

X2Uip1\displaystyle-\int_{X}2U^{p-1}_{i} e12wiiwiiiΔiwiωin=X2Uip1e12wi(Δiwi)2ωin\displaystyle e^{\frac{1}{2}w_{i}}\nabla_{i}w_{i}\cdot_{i}\nabla_{i}\Delta_{i}w_{i}\omega^{n}_{i}=\int_{X}2U_{i}^{p-1}e^{\frac{1}{2}w_{i}}(\Delta_{i}w_{i})^{2}\omega^{n}_{i}
+XUip1e12wi|iwi|i2Δiwiωin+X2(p1)Uip2e12wiiUiiiwiΔiwiωin.\displaystyle+\int_{X}U^{p-1}_{i}e^{\frac{1}{2}w_{i}}|\nabla_{i}w_{i}|^{2}_{i}\Delta_{i}w_{i}\omega^{n}_{i}+\int_{X}2(p-1)U_{i}^{p-2}e^{\frac{1}{2}w_{i}}\nabla_{i}U_{i}\cdot_{i}\nabla_{i}w_{i}\Delta_{i}w_{i}\omega^{n}_{i}.

Using the simple fact that e12wi|iwi|i2Uie^{\frac{1}{2}w_{i}}|\nabla_{i}w_{i}|^{2}_{i}\leq U_{i}, we can estimate

XUip1e12wi|iwi|i2ΔiwiωinXUip|Δiwi|ωinXUip((Δiwi)2+1)ωin\int_{X}U^{p-1}_{i}e^{\frac{1}{2}w_{i}}|\nabla_{i}w_{i}|^{2}_{i}\Delta_{i}w_{i}\omega^{n}_{i}\leq\int_{X}U_{i}^{p}|\Delta_{i}w_{i}|\omega^{n}_{i}\leq\int_{X}U_{i}^{p}((\Delta_{i}w_{i})^{2}+1)\omega^{n}_{i}

and

X2Uip2e12wiiUiiiwiΔiwiωinX12Uip2|iUi|i2ωin+X2Uip1e12wi(Δiwi)2ωin.\int_{X}2U_{i}^{p-2}e^{\frac{1}{2}w_{i}}\nabla_{i}U_{i}\cdot_{i}\nabla_{i}w_{i}\Delta_{i}w_{i}\omega^{n}_{i}\leq\int_{X}\frac{1}{2}U^{p-2}_{i}|\nabla_{i}U_{i}|^{2}_{i}\omega^{n}_{i}+\int_{X}2U_{i}^{p-1}e^{\frac{1}{2}w_{i}}(\Delta_{i}w_{i})^{2}\omega^{n}_{i}.

Putting these together and using Ui1U_{i}\geq 1, one can derive that (as in [13, (3.54)])

Xp12Uip2|iUi|i2ωinXpUipGieFiωn,\int_{X}\frac{p-1}{2}U_{i}^{p-2}|\nabla_{i}U_{i}|_{i}^{2}\omega_{i}^{n}\leq\int_{X}pU_{i}^{p}G_{i}e^{F_{i}}\omega^{n},

where

Gi:=G~i+(Δiwi)2+2e12wi(Δiw)2+1.G_{i}:=\tilde{G}_{i}+(\Delta_{i}w_{i})^{2}+2e^{\frac{1}{2}w_{i}}(\Delta_{i}w)^{2}+1.

The rest of the proof uses Nash–Moser iteration, which goes through in exactly the same way as in [13, p.960-962]. The only difference lies in the L4nL^{4n}-bound for GiG_{i} (compare [13, (3.63)]), since in our definition of GiG_{i} there is an additional term (n+Δui1)(n+\Delta u_{i-1}). But thanks to (5.4), this term has uniform LqL^{q}-bound (which only depends on X,ω,A,qX,\omega,A,q and is independent of ii). So the estimates in [13] goes through for us as well.

Thus we finally arrive at UiLC=C(X,ω,A)||U_{i}||_{L^{\infty}}\leq C=C(X,\omega,A), finishing the proof.

Corollary 5.4.

Assume that there is some constant A>0A>0 such that

Entω(ui)+d1(0,ui)A for all i.Ent_{\omega}(u_{i})+d_{1}(0,u_{i})\leq A\text{ for all }i\in{\mathbb{N}}.

Then there exists some constant B5>1B_{5}>1 depending only on X,ωX,\omega and AA such that

B51ωωiB5ω for all i.B_{5}^{-1}\omega\leq\omega_{i}\leq B_{5}\omega\text{ for all }i\in{\mathbb{N}}.
Proof.

This follows immediately from ωineB1ωn\omega_{i}^{n}\geq e^{-B_{1}}\omega^{n} and trωωiB4\mathrm{tr}_{\omega}\omega_{i}\leq B_{4}. ∎

By classical elliptic estimates and bootstrapping, we then have the following uniform estimates.

Corollary 5.5.

Assume that there is some constant A>0A>0 such that

Entω(ui)+d1(0,ui)A for all i.Ent_{\omega}(u_{i})+d_{1}(0,u_{i})\leq A\text{ for all }i\in{\mathbb{N}}.

Then for any α(0,1)\alpha\in(0,1) and k1k\geq 1, there exists some constant Bk,α>1B_{k,\alpha}>1 depending only on X,ω,α,kX,\omega,\alpha,k and AA such that

uiCk,αBk,α for all i.||u_{i}||_{C^{k,\alpha}}\leq B_{k,\alpha}\text{ for all }i\in{\mathbb{N}}.
Proof.

First, the estimate B51ωωiB5ωB_{5}^{-1}\omega\leq\omega_{i}\leq B_{5}\omega implies that (5.2) is uniformly elliptic with bounded right hand side. Then arguing as in the proof of [12, Proposition 4.2], one has uiC3,αu_{i}\in C^{3,\alpha} and FiC1,αF_{i}\in C^{1,\alpha} for all ii\in{\mathbb{N}}. This implies that the equation (5.2) has C1,αC^{1,\alpha}-coefficients and right hand (since we already know that ui1u_{i-1} has C3,αC^{3,\alpha} bound). This gives C3,αC^{3,\alpha} bound for FiF_{i}. Differentiating the equation (5.1) twice one then gets a linear elliptic equation for the second derivatives of uiu_{i} with CαC^{\alpha} coefficients and right hand side. So we get the C4,αC^{4,\alpha}-bound for uiu_{i}. Continuing in this way we get all the Ck,αC^{k,\alpha} bounds for uiu_{i}. ∎

Remark 5.6.

In the above discussion we have set τ=1\tau=1 to simplify the exposition. In general, the equation we are dealing with is

{(ω+1¯ui)n=eFiωn,Δωi(Fi+ui1/τ)=trωi(Ric(ω)ω/τ)+n/τR¯.\begin{cases}(\omega+\sqrt{-1}\partial\bar{\partial}u_{i})^{n}=e^{F_{i}}\omega^{n},\\ \Delta_{\omega_{i}}(F_{i}+u_{i-1}/\tau)=\mathrm{tr}_{\omega_{i}}(\operatorname{Ric}(\omega)-\omega/\tau)+n/\tau-\bar{R}.\\ \end{cases}

In this case the estimates we get will depend on τ\tau as well, and unfortunately they blowup as τ0\tau\searrow 0. If one can get uniform estimates independent of τ,\tau, then the iteration should converge to the flow (1.1).

On the other hand, if we are allowed to take τ0\tau\gg 0 (for instance when the K-energy is bounded from below), then the boundedness assumption on d1(0,ui)d_{1}(0,u_{i}) can be removed, since it is merely used to get the estimate (5.3):

Xepui/τωnC,\int_{X}e^{-pu_{i}/\tau}\omega^{n}\leq C,

which now holds for free when τ0\tau\gg 0 by using Tian’s α\alpha-invariant [40]; see [13, Lemma 4.20] for a similar situation.

6. Smooth convergence of the Ricci iteration

Assume that (X,ω)(X,\omega) admits a cscK metric ω\omega^{*} in {ω}\{\omega\}. Then by [8, Theorem 1.5] we know that the K-energy is proper modulo G:=Aut0(X)G:=Aut_{0}(X). Hence one can choose τ0=\tau_{0}=\infty in Theorem 1.4. Then for any τ>0\tau>0, we wish to show that the iteration sequence {ωi}i\{\omega_{i}\}_{i\in{\mathbb{N}}} defined by (1.2) converges in a suitable sense to a cscK metric. Up to scaling the Kähler class, we will assume without loss of generality that τ=1\tau=1. To make further simplification, we will first deal with the case where the cscK metric is unique, in which case the K-energy is proper (by [20, 4, 8]), i.e., γ(X,{ω})>0\gamma(X,\{\omega\})>0 (recall (4.1)).

Therefore, we have that

  • There exists γ>0\gamma>0 and C0>0C_{0}>0 such that

    Kω(φ)γ(IωJω)(φ)C0 for any φω.K_{\omega}(\varphi)\geq\gamma(I_{\omega}-J_{\omega})(\varphi)-C_{0}\text{ for any }\varphi\in\mathcal{H}_{\omega}.
  • There exists a sequence {ωi}i\{\omega_{i}\}_{i\in{\mathbb{N}}} satisfying

    ωi+1ωi=Ric(ωi+1)+HRic(ωi+1),ω0=ω.\omega_{i+1}-\omega_{i}=-\operatorname{Ric}(\omega_{i+1})+\mathrm{HRic}(\omega_{i+1}),\ \omega_{0}=\omega.

    Equivalently, one has

    (6.1) R(ωi+1)=R¯n+trωi+1ωi,ω0=ω.R(\omega_{i+1})=\bar{R}-n+\mathrm{tr}_{\omega_{i+1}}\omega_{i},\ \omega_{0}=\omega.
  • Write ω=ω+1¯u\omega^{*}=\omega+\sqrt{-1}\partial\bar{\partial}{u^{*}} and ωi=ω+1¯ui\omega_{i}=\omega+\sqrt{-1}\partial\bar{\partial}u_{i} for u,ui0u^{*},u_{i}\in\mathcal{H}_{0}, where

    0:={φω|Eω(φ)=0}.\mathcal{H}_{0}:=\{\varphi\in\mathcal{H}_{\omega}|E_{\omega}(\varphi)=0\}.

We wish to show the following.

Theorem 6.1.

Assume that there exists a unique cscK potential u0u^{*}\in\mathcal{H}_{0}, then the sequence {ui}i\{u_{i}\}_{i\in{\mathbb{N}}} converges smoothly to uu^{*}.

To prove this, we need some preparations.

Lemma 6.2.

One has

  1. (1)

    ωi\omega_{i} minimizes 𝒥ωωi\mathcal{J}_{\omega}^{\omega_{i}} over ω\mathcal{H}_{\omega}.

  2. (2)

    ωi+1\omega_{i+1} minimizes Kω+𝒥ωωiK_{\omega}+\mathcal{J}^{\omega_{i}}_{\omega} over ω\mathcal{H}_{\omega}.

Proof.

The first assertion follows from Lemmas 2.2. See e.g. [13, Corollary 4.5] for a proof of the second one. ∎

Lemma 6.3.

One can find A>0A>0 such that for all ii\in{\mathbb{N}}

(6.2) Entω(ui)+d1(0,ui)A.Ent_{\omega}(u_{i})+d_{1}(0,u_{i})\leq A.
Proof.

By Theorem 1.5 we have that

0=Kω(ω0)Kω(ωi)=Entω(ui)+𝒥ω(ui)γ(IωJω)(ui)C0.0=K_{\omega}(\omega_{0})\geq K_{\omega}(\omega_{i})=Ent_{\omega}(u_{i})+\mathcal{J}_{\omega}(u_{i})\geq\gamma(I_{\omega}-J_{\omega})(u_{i})-C_{0}.

This implies that (using Lemma 2.1)

Jω(ui)n(IωJω)(ui)nC0γ.J_{\omega}(u_{i})\leq n(I_{\omega}-J_{\omega})(u_{i})\leq\frac{nC_{0}}{\gamma}.

So Lemma 3.3 gives that

d1(0,ui)C1.d_{1}(0,u_{i})\leq C_{1}.

On the other hand, by [13, Lemma 4.4], one has

0Kω(ωi)=Entω(ui)+𝒥ω(ui)Entω(ui)C2d1(0,ui).0\geq K_{\omega}(\omega_{i})=Ent_{\omega}(u_{i})+\mathcal{J}_{\omega}(u_{i})\geq Ent_{\omega}(u_{i})-C_{2}d_{1}(0,u_{i}).

So we obtain that

Entω(ui)C3,Ent_{\omega}(u_{i})\leq C_{3},

finishing the proof. ∎

Lemma 6.4.

One has for all ii\in{\mathbb{N}}

J(ui+1,ui)Kω(ui)Kω(ui+1).J(u_{i+1},u_{i})\leq K_{\omega}(u_{i})-K_{\omega}(u_{i+1}).

So one has I(ui+1,ui)0I(u_{i+1},u_{i})\to 0.

Proof.

Using that ωi+1\omega_{i+1} minimizes Kω+𝒥ωωiK_{\omega}+\mathcal{J}^{\omega_{i}}_{\omega}, we have

Kω(ui+1)+𝒥ωωi(ui+1)Kω(ui)+𝒥ωωi(ui).K_{\omega}(u_{i+1})+\mathcal{J}^{\omega_{i}}_{\omega}(u_{i+1})\leq K_{\omega}(u_{i})+\mathcal{J}^{\omega_{i}}_{\omega}(u_{i}).

Then using 𝒥ωωi(ui+1)𝒥ωi(ui)=J(ui+1,ui)\mathcal{J}^{\omega_{i}}_{\omega}(u_{i+1})-\mathcal{J}^{\omega_{i}}(u_{i})=J(u_{i+1},u_{i}) (recall Lemma 2.2) we conclude the first assertion.

For the second statement, note that the KK-energy is bounded from below in our setting, so Theorem 1.5 implies that {Kω(ui)}i\{K_{\omega}(u_{i})\}_{i\in{\mathbb{N}}} is a convergent sequence. So we conclude from Lemma 2.1. ∎

Corollary 6.5.

If {uik}k\{u_{i_{k}}\}_{k\in{\mathbb{N}}} is a d1d_{1}-convergent subsequence, say uikd1uu_{i_{k}}\xrightarrow{d_{1}}u, then uik1d1uu_{i_{k}-1}\xrightarrow{d_{1}}u as well.

Proof.

This follows from the previous lemma and Corollary 3.9. ∎

Now we are ready to prove Theorem 6.1

Proof of Theorem 6.1.

We first argue that any convergent subsequence {uik}k\{u_{i_{k}}\}_{k\in{\mathbb{N}}} has to converge to uu^{*} in the d1d_{1}-topology. By (6.2) and Lemma 3.5 this will imply that uid1uu_{i}\xrightarrow{d_{1}}u^{*}.

So assume that there exists a subsequence {uik}k\{u_{i_{k}}\}_{k\in{\mathbb{N}}}, converging in d1d_{1} to a limit uω1u_{\infty}\in\mathcal{E}^{1}_{\omega}. Then for any uωu\in\mathcal{H}_{\omega} we deduce that

Kω(u)\displaystyle K_{\omega}(u_{\infty}) limkKω(uik)\displaystyle\leq\lim_{k\to\infty}K_{\omega}(u_{i_{k}})
=limk(Kω(uik)+𝒥ωωik1(uik)𝒥ωωik1(uik))\displaystyle=\lim_{k\to\infty}(K_{\omega}(u_{i_{k}})+\mathcal{J}_{\omega}^{\omega_{i_{k}-1}}(u_{i_{k}})-\mathcal{J}_{\omega}^{\omega_{i_{k}-1}}(u_{i_{k}}))
limk(Kω(u)+𝒥ωωik1(u)𝒥ωωik1(uik1))\displaystyle\leq\lim_{k\to\infty}(K_{\omega}(u)+\mathcal{J}_{\omega}^{\omega_{i_{k}-1}}(u)-\mathcal{J}_{\omega}^{\omega_{i_{k}-1}}(u_{i_{k}-1}))
=limk(Kω(u)+J(u,uik1))=Kω(u)+J(u,u).\displaystyle=\lim_{k\to\infty}(K_{\omega}(u)+J(u,u_{i_{k}-1}))=K_{\omega}(u)+J(u,u_{\infty}).

Here we used that KωK_{\omega} is d1d_{1}-lsc, uiku_{i_{k}} minimizes Kω+𝒥ωωik1K_{\omega}+\mathcal{J}_{\omega}^{\omega_{i_{k}-1}}, uik1u_{i_{k}-1} minimizes 𝒥ωωik1\mathcal{J}_{\omega}^{\omega_{i_{k}-1}}, Lemma 2.2, Corollary 6.5 and Lemma 3.6. Thus we obtain that

Kω(u)Kω(u)+J(u,u)for any uω.K_{\omega}(u_{\infty})\leq K_{\omega}(u)+J(u,u_{\infty})\ \text{for any }u\in\mathcal{H}_{\omega}.

By Lemma 3.4 we then see that uu_{\infty} is a minimizer of the functional

F(u):=Kω(u)+J(u,u),uω1.F_{\infty}(u):=K_{\omega}(u)+J(u,u_{\infty}),u\in\mathcal{E}^{1}_{\omega}.

We now argue that uu_{\infty} must be a cscK potential and hence u=uu_{\infty}=u^{*}.

By Lemma 6.3, Corollary 5.5 and Arzelà–Ascoli, we know that uωu_{\infty}\in\mathcal{H}_{\omega}. So by Lemma 2.2 we can write

F(u)=Kω(u)+𝒥ωωu(u)𝒥ωωu(u)=Kωωu(u)𝒥ωωu(u).F_{\infty}(u)=K_{\omega}(u)+\mathcal{J}_{\omega}^{\omega_{u_{\infty}}}(u)-\mathcal{J}^{\omega_{u_{\infty}}}_{\omega}(u_{\infty})=K_{\omega}^{\omega_{u_{\infty}}}(u)-\mathcal{J}^{\omega_{u_{\infty}}}_{\omega}(u_{\infty}).

So uu_{\infty} minimizes the twisted K-energy KωωuK_{\omega}^{\omega_{u_{\infty}}}. The variation formula (2.3) of KωωuK_{\omega}^{\omega_{u_{\infty}}} then implies that

R(ωu)=R¯n+trωuωu=R¯.R(\omega_{u_{\infty}})=\bar{R}-n+\mathrm{tr}_{\omega_{u_{\infty}}}\omega_{u_{\infty}}=\bar{R}.

Thus ωu\omega_{u_{\infty}} is a cscK metric. By uniqueness assumption we have that u=uu_{\infty}=u^{*}.

Therefore, we have shown that uikd1uu_{i_{k}}\xrightarrow{d_{1}}u^{*} for any convergent subsequence. So uid1uu_{i}\xrightarrow{d_{1}}u^{*} follows. By Corollary 5.5 and Arzelà–Ascoli we then know that uiuu_{i}\to u^{*} smoothly. ∎

If we do not assume the uniqueness of the cscK metric ω\omega^{*}, then the K-energy is proper modulo the action of biholomorpihc automorphisms of XX (see [8, Theorem 1.5]). Modifying our previous proofs and incorporating the ideas from [19], one can actually prove the following result, which improves Theorem 6.1 and extends Darvas–Rubinstein’s work [19, Theorem 1.6] to arbitrary Kähler classes.

Theorem 6.6.

(=Theorem 1.6) Let (X,ω)(X,\omega) be a compact Kähler manifold admitting a cscK metric in {ω}\{\omega\}. Then for any τ>0\tau>0 the iteration sequence (1.2) sequence exists and there exist holomorphic diffeomorphsims gig_{i} such that giωig_{i}^{*}\omega_{i} converges smoothly to a cscK metric.

Proof.

We give the necessary details for the reader’s convenience. As above, we assume without loss of generality that τ=1\tau=1.

First, using that the K-energy decreases along uiu_{i} and is proper modulo G=Aut0(X)G=Aut_{0}(X), we have that

d1,G(0,ui)A0 for all id_{1,G}(0,u_{i})\leq A_{0}\text{ for all }i\in{\mathbb{N}}

Fix a cscK metric ω{ω}\omega^{*}\in\{\omega\} with ω=ω+1¯u\omega^{*}=\omega+\sqrt{-1}\partial\bar{\partial}u^{*} and u0u^{*}\in\mathcal{H}_{0}. Then pick giGg_{i}\in G such that

d1(u,gi.ui)d1,G(u,ui)+1id1(0,u)+d1,G(0,ui)A1.d_{1}(u^{*},g_{i}.u_{i})\leq d_{1,G}(u^{*},u_{i})+\frac{1}{i}\leq d_{1}(0,u^{*})+d_{1,G}(0,u_{i})\leq A_{1}.

Thus we deduce that

d1(0,gi.ui)A1+d1(0,u).d_{1}(0,g_{i}.u_{i})\leq A_{1}+d_{1}(0,u^{*}).

Then using that the K-energy is GG-invariant (see e.g. [13, Lemma 4.11]), one can argue as in the proof of Lemma 6.3 to show that

Entω(gi.ui)A2 for all i.Ent_{\omega}(g_{i}.u_{i})\leq A_{2}\text{ for all }i\in{\mathbb{N}}.

So the sequence {gi.ui}i\{g_{i}.u_{i}\}_{i\in{\mathbb{N}}} is d1d_{1}-precompact. We wish to show that it converges to uu^{*} smoothly. To this end, we need some uniform estimates for the sequence.

By Lemmas 3.11, 2.1, 6.4 and Theorem 1.5, we know that

I(gi.ui,gi.ui1)\displaystyle I(g_{i}.u_{i},g_{i}.u_{i-1}) =I(ui,ui1)(n+1)J(ui,ui1)\displaystyle=I(u_{i},u_{i-1})\leq(n+1)J(u_{i},u_{i-1})
(n+1)(Kω(ui1)Kω(ui))0.\displaystyle\leq(n+1)(K_{\omega}(u_{i-1})-K_{\omega}(u_{i}))\to 0.

And also, one has (by Lemma 3.7)

J(0,gi.ui1)I(0,gi.ui1)Cn(I(0,gi.ui)+I(gi.ui,gi.ui1)).J(0,g_{i}.u_{i-1})\leq I(0,g_{i}.u_{i-1})\leq C_{n}(I(0,g_{i}.u_{i})+I(g_{i}.u_{i},g_{i}.u_{i-1})).

So we derive that (using Lemma 3.3)

d1(0,gi.ui1)A3.d_{1}(0,g_{i}.u_{i-1})\leq A_{3}.

The upshot is that, there exists some A>0A>0 such that

Entω(gi.ui)+d1(0,gi.ui1)A for all i1.Ent_{\omega}(g_{i}.u_{i})+d_{1}(0,g_{i}.u_{i-1})\leq A\text{ for all }i\geq 1.

For simplicity let us put

vi:=gi.ui and hi1:=gi.ui1.v_{i}:=g_{i}.u_{i}\text{ and }h_{i-1}:=g_{i}.u_{i-1}.

Then from (6.1) we deduce that

R(ωvi)=R¯n+trωvi(ω+1¯hi1).R(\omega_{v_{i}})=\bar{R}-n+\mathrm{tr}_{\omega_{v_{i}}}(\omega+\sqrt{-1}\partial\bar{\partial}h_{i-1}).

This is equivalent to (cf. [13, Lemma 4.19])

(6.3) {(ω+1¯vi)n=eFiωn.Δωvi(Fi+hi1)=trωvi(Ric(ω)ω)+nR¯.\begin{cases}(\omega+\sqrt{-1}\partial\bar{\partial}v_{i})^{n}=e^{F_{i}}\omega^{n}.\\ \Delta_{\omega_{v_{i}}}(F_{i}+h_{i-1})=\mathrm{tr}_{\omega_{v_{i}}}(\operatorname{Ric}(\omega)-\omega)+n-\bar{R}.\\ \end{cases}

And we have that

Entω(vi)+d1(0,hi1)A for all i1.Ent_{\omega}(v_{i})+d_{1}(0,h_{i-1})\leq A\text{ for all }i\geq 1.

Then as in Proposition 5.1 we can obtain the C0C^{0} estimate:

|vi|B1 for all i1.|v_{i}|\leq B_{1}\text{ for all }i\geq 1.

This implies that (by Lemma 3.2)

d1(0,vi)=d1(0,gi.ui)=d1(gi1.0,ui)=d1(gi+1.(gi1.0),hi+1)B2.d_{1}(0,v_{i})=d_{1}(0,g_{i}.u_{i})=d_{1}(g_{i}^{-1}.0,u_{i})=d_{1}(g_{i+1}.(g_{i}^{-1}.0),h_{i+1})\leq B_{2}.

Put

fi:=gi1gi+1,f_{i}:=g^{-1}_{i}\circ g_{i+1},

Then we have

d1(fi.0,0)d1(fi.0,hi+1)+d1(0,hi+1)B2+A3 for all i1.d_{1}(f_{i}.0,0)\leq d_{1}(f_{i}.0,h_{i+1})+d_{1}(0,h_{i+1})\leq B_{2}+A_{3}\text{ for all }i\geq 1.

By the proof of [20, Proposition 6.8], {fi}i1\{f_{i}\}_{i\geq 1} is contained in a bounded set of G.G. In particular, all derivatives of fif_{i} up to order mm, say, are bounded by some CmC_{m} independently of ii. Since one has

hi=gi+1.ui=fi.vi,h_{i}=g_{i+1}.u_{i}=f_{i}.v_{i},

then viv_{i} and hih_{i} enjoy the same a priori estimates. Now the same arguments as in §5 apply to the system of equations (6.3) as well. We conclude that there are uniform Ck,αC^{k,\alpha} estimates (independent of ii) for viv_{i} and hih_{i}.

Now we are ready to show that viuv_{i}\to u^{*} smoothly.

By Arzelà–Ascoli it suffices to argue that vid1uv_{i}\xrightarrow{d_{1}}u^{*}. We prove by contradiction. Assume that there exists a subsequence such that vikd1vv_{i_{k}}\xrightarrow{d_{1}}v_{\infty} for some vω1v_{\infty}\in\mathcal{E}^{1}_{\omega} with d1(u,v)>ε>0d_{1}(u^{*},v_{\infty})>\varepsilon>0. By our uniform estimates for viv_{i} and Arzelà–Ascoli we know that v0.v_{\infty}\in\mathcal{H}_{0}.

For any u0u\in\mathcal{H}_{0}, one has (as in the proof of Theorem 6.1)

Kω(v)\displaystyle K_{\omega}(v_{\infty}) limkKω(vik)=limkKω(uik)\displaystyle\leq\lim_{k\to\infty}K_{\omega}(v_{i_{k}})=\lim_{k\to\infty}K_{\omega}(u_{i_{k}})
=limk(Kω(uik)+𝒥ωuik1(uik)𝒥ωuik1(uik))\displaystyle=\lim_{k\to\infty}(K_{\omega}(u_{i_{k}})+\mathcal{J}^{\omega_{u_{i_{k}-1}}}(u_{i_{k}})-\mathcal{J}^{\omega_{u_{i_{k}-1}}}(u_{i_{k}}))
limk(Kω(gik1.u)+𝒥ωuik1(gik1.u)𝒥ωuik1(uik1))\displaystyle\leq\lim_{k\to\infty}(K_{\omega}(g_{i_{k}}^{-1}.u)+\mathcal{J}^{\omega_{u_{i_{k}-1}}}(g_{i_{k}}^{-1}.u)-\mathcal{J}^{\omega_{u_{i_{k}-1}}}(u_{i_{k}-1}))
=limk(Kω(u)+J(gik1.u,uik1))\displaystyle=\lim_{k\to\infty}(K_{\omega}(u)+J(g^{-1}_{i_{k}}.u,u_{i_{k}-1}))
=limk(Kω(u)+J(u,hik1))=Kω(u)+J(u,v).\displaystyle=\lim_{k\to\infty}(K_{\omega}(u)+J(u,h_{i_{k}-1}))=K_{\omega}(u)+J(u,v_{\infty}).

Here we used that KωK_{\omega} and JJ are GG-invariant (recall Lemma 3.11). Moreover, in the last equality we used that hik1d1vh_{i_{k}-1}\xrightarrow{d_{1}}v_{\infty}. Indeed, Lemma 3.11 and 6.4 imply that I(vi,hi1)=I(ui,ui1)0I(v_{i},h_{i-1})=I(u_{i},u_{i-1})\to 0. So Corollary 3.9 implies that limkhik1=limkvik=v\lim_{k}h_{i_{k}-1}=\lim_{k}v_{i_{k}}=v_{\infty}, as claimed.

From above we observe that vv_{\infty} is a minimizer for the functional

F(u):=Kω(u)+J(u,v),u0.F_{\infty}(u):=K_{\omega}(u)+J(u,v_{\infty}),\ u\in\mathcal{H}_{0}.

This further implies that vv_{\infty} is a minimizer of FF_{\infty} over ω\mathcal{H}_{\omega}. Then as in the proof of Theorem 6.1, we conclude that vv_{\infty} is a cscK potential.

By [4, Theorem 1.3] there exists fGf\in G such that v=f.uv_{\infty}=f.u^{*}. So we obtain that

d1(vik,u)1ikd1,G(vik,u)d1(f1.vik,u)=d1(vik,v).d_{1}(v_{i_{k}},u^{*})-\frac{1}{i_{k}}\leq d_{1,G}(v_{i_{k}},u^{*})\leq d_{1}(f^{-1}.v_{i_{k}},u^{*})=d_{1}(v_{i_{k}},v_{\infty}).

By choice the right hand goes to zero, while the left hand side is strictly bigger than ε2>0\frac{\varepsilon}{2}>0 for any k1k\gg 1. This is a contradiction. So we finish the proof. ∎

As in [19], one expects that the appearance of gig_{i} in the above theorem is actually redundant, which might require substantial new ideas; compare also [13, Proposition 4.17] in the setting of continuity method. In the case of Kähler Ricci flow, the analogous problem is studied in [45, 30, 46, 42, 16].

7. The case of twisted cscK metrics

In the study of cscK metrics, it is often beneficial to allow for some twisted terms. More precisely, given a closed smooth real (1,1)(1,1) form, one can study the following χ\chi-twisted cscK equation:

(7.1) R(ωu)=R¯χ¯+trωuχ.R(\omega_{u})=\bar{R}-\bar{\chi}+\mathrm{tr}_{\omega_{u}}\chi.

This is equivalent to saying that Ric(ωu)χ\operatorname{Ric}(\omega_{u})-\chi is harmonic with respect to ωu\omega_{u}. Therefore, to search for χ\chi-twisted cscK metrics, we are led to the following twisted flow:

(7.2) tωt=Ric(ωt)+Hωt(Ric(ωt)χ)+χ,ω0=ω.\partial_{t}\omega_{t}=-\operatorname{Ric}(\omega_{t})+\mathrm{H}_{\omega_{t}}(\operatorname{Ric}(\omega_{t})-\chi)+\chi,\ \omega_{0}=\omega.

Here HωtH_{\omega_{t}} denotes the harmonic projection operator of the metric ωt\omega_{t}. When 2πc1(X)=λ{ω}+{χ}2\pi c_{1}(X)=\lambda\{\omega\}+\{\chi\}, this flow becomes

tωt=Ric(ωt)+λωt+χ,ω0=ω,\partial_{t}\omega_{t}=-\operatorname{Ric}(\omega_{t})+\lambda\omega_{t}+\chi,\ \omega_{0}=\omega,

which is the twisted Kähler Ricci flow studied in [27, 16].

Discretizing the flow (7.2), we get (for some given τ>0\tau>0)

(7.3) ωi+1ωiτ=Ric(ωi+1)+Hωi+1(Ric(ωi+1)χ)+χ,i,ω0=ω.\frac{\omega_{i+1}-\omega_{i}}{\tau}=-\operatorname{Ric}(\omega_{i+1})+\mathrm{H}_{\omega_{i+1}}(\operatorname{Ric}(\omega_{i+1})-\chi)+\chi,\ i\in{\mathbb{N}},\ \omega_{0}=\omega.

The next result can be proved following exactly the same strategy as we did for the untwisted case. Hence we omit the details.

Theorem 7.1.

Assume that χ0\chi\geq 0. There exists a constant τ0(0,]\tau_{0}\in(0,\infty] depending only on X,{ω}X,\{\omega\} and {χ}\{\chi\} such that for any τ(0,τ0)\tau\in(0,\tau_{0}), the iteration sequence (7.3) exists for all ii\in{\mathbb{N}}, with each ωi\omega_{i} being uniquely determined by ω0\omega_{0}, along which the χ\chi-twisted K-energy KωχK_{\omega}^{\chi} decreases. Moreover, if there exists a unique χ\chi-twisted cscK metric ω{ω}\omega^{*}\in\{\omega\}, then for any τ>0\tau>0 the sequence ωi\omega_{i} converges to ω\omega^{*} smoothly.

Note that, if χ>0\chi>0, the uniqueness of χ\chi-twisted cscK metric is automatic by [7, Theorem 4.13]. This might be useful, since one can study the flow (1.1) or the iteration (1.2) by adding a small amount of χ\chi (cf. the perturbation trick in [4, §4]).

One can try to extend our work further to the case of conical cscK metrics, extremal metrics and other canonical metrics. We leave this to the interested readers.

\AtNextBibliography

References

  • [1] Richard Bamler “Convergence of Ricci flows with bounded scalar curvature” In Ann. of Math. (2) 188.3, 2018, pp. 753–831 DOI: 10.4007/annals.2018.188.3.2
  • [2] Shigetoshi Bando “An obstruction for Chern class forms to be harmonic” In Kodai Math. J. 29.3, 2006, pp. 337–345 DOI: 10.2996/kmj/1162478766
  • [3] Eric Bedford and B.. Taylor “A new capacity for plurisubharmonic functions” In Acta Math. 149.1-2, 1982, pp. 1–40 DOI: 10.1007/BF02392348
  • [4] Robert J. Berman and Bo Berndtsson “Convexity of the KK-energy on the space of Kähler metrics and uniqueness of extremal metrics” In J. Amer. Math. Soc. 30.4, 2017, pp. 1165–1196 DOI: 10.1090/jams/880
  • [5] Robert J. Berman et al. “Kähler-Einstein metrics and the Kähler-Ricci flow on log Fano varieties” In J. Reine Angew. Math. 751, 2019, pp. 27–89 DOI: 10.1515/crelle-2016-0033
  • [6] Robert J. Berman, Sébastien Boucksom, Vincent Guedj and Ahmed Zeriahi “A variational approach to complex Monge-Ampère equations” In Publ. Math. Inst. Hautes Études Sci. 117, 2013, pp. 179–245 DOI: 10.1007/s10240-012-0046-6
  • [7] Robert J. Berman, Tamás Darvas and Chinh H. Lu “Convexity of the extended K-energy and the large time behavior of the weak Calabi flow” In Geom. Topol. 21.5, 2017, pp. 2945–2988 DOI: 10.2140/gt.2017.21.2945
  • [8] Robert J. Berman, Tamás Darvas and Chinh H. Lu “Regularity of weak minimizers of the K-energy and applications to properness and K-stability” In Ann. Sci. Éc. Norm. Supér. (4) 53.2, 2020, pp. 267–289 DOI: 10.24033/asens.2422
  • [9] T. Buttsworth, A. Pulemotov, Y.. Rubinstein and W. Ziller “On the Ricci iteration for homogeneous metrics on spheres and projective spaces” In Transform. Groups 26.1, 2021, pp. 145–164 DOI: 10.1007/s00031-020-09602-3
  • [10] Eugenio Calabi “Extremal Kähler metrics” In Seminar on Differential Geometry No. 102, Ann. of Math. Stud. Princeton Univ. Press, Princeton, NJ, 1982, pp. 259–290
  • [11] Huai Dong Cao “Deformation of Kähler metrics to Kähler-Einstein metrics on compact Kähler manifolds” In Invent. Math. 81.2, 1985, pp. 359–372 DOI: 10.1007/BF01389058
  • [12] Xiuxiong Chen and Jingrui Cheng “On the constant scalar curvature Kähler metrics (I)—A priori estimates” In J. Amer. Math. Soc. 34.4, 2021, pp. 909–936 DOI: 10.1090/jams/967
  • [13] Xiuxiong Chen and Jingrui Cheng “On the constant scalar curvature Kähler metrics (II)—Existence results” In J. Amer. Math. Soc. 34.4, 2021, pp. 937–1009 DOI: 10.1090/jams/966
  • [14] Xiuxiong Chen and Bing Wang “Space of Ricci flows (II)—Part B: Weak compactness of the flows” In J. Differential Geom. 116.1, 2020, pp. 1–123 DOI: 10.4310/jdg/1599271253
  • [15] Xiuxiong Chen and Kai Zheng “The pseudo-Calabi flow” In J. Reine Angew. Math. 674, 2013, pp. 195–251 DOI: 10.1515/crelle.2012.033
  • [16] Tristan C. Collins and Gábor Székelyhidi “The twisted Kähler-Ricci flow” In J. Reine Angew. Math. 716, 2016, pp. 179–205 DOI: 10.1515/crelle-2014-0010
  • [17] Tamás Darvas “Geometric pluripotential theory on Kähler manifolds” In Advances in complex geometry 735, Contemp. Math. Amer. Math. Soc., Providence, RI, 2019, pp. 1–104 DOI: 10.1090/conm/735/14822
  • [18] Tamás Darvas “The Mabuchi geometry of finite energy classes” In Adv. Math. 285, 2015, pp. 182–219 DOI: 10.1016/j.aim.2015.08.005
  • [19] Tamás Darvas and Yanir A. Rubinstein “Convergence of the Kähler-Ricci iteration” In Anal. PDE 12.3, 2019, pp. 721–735 DOI: 10.2140/apde.2019.12.721
  • [20] Tamás Darvas and Yanir A. Rubinstein “Tian’s properness conjectures and Finsler geometry of the space of Kähler metrics” In J. Amer. Math. Soc. 30.2, 2017, pp. 347–387 DOI: 10.1090/jams/873
  • [21] Daniel Guan “Extremal solitons and exponential CC^{\infty} convergence of the modified Calabi flow on certain P1\mathbb{C}{\rm P}^{1} bundles” In Pacific J. Math. 233.1, 2007, pp. 91–124 DOI: 10.2140/pjm.2007.233.91
  • [22] Vincent Guedj, Boris Kolev and Nader Yeganefar “Kähler-Einstein fillings” In J. Lond. Math. Soc., II. Ser. 88.3, 2013, pp. 737–760 DOI: 10.1112/jlms/jdt031
  • [23] Vincent Guedj and Ahmed Zeriahi “The weighted Monge-Ampère energy of quasiplurisubharmonic functions” In J. Funct. Anal. 250.2, 2007, pp. 442–482 DOI: 10.1016/j.jfa.2007.04.018
  • [24] Yoshinori Hashimoto “Existence of twisted constant scalar curvature Kähler metrics with a large twist” In Math. Z. 292.3-4, 2019, pp. 791–803 DOI: 10.1007/s00209-018-2133-y
  • [25] Ryan Hunter “Monge-Ampère iteration” Id/No 73 In Sel. Math., New Ser. 25.5, 2019, pp. 70 DOI: 10.1007/s00029-019-0519-2
  • [26] Julien Keller “Ricci iterations on Kähler classes” In J. Inst. Math. Jussieu 8.4, 2009, pp. 743–768 DOI: 10.1017/S1474748009000103
  • [27] Jiawei Liu “The generalized Kähler Ricci flow” In J. Math. Anal. Appl. 408.2, 2013, pp. 751–761 DOI: 10.1016/j.jmaa.2013.06.047
  • [28] D.. Phong, Jian Song and Jacob Sturm “Complex Monge-Ampère equations” In Surveys in differential geometry. Vol. XVII 17, Surv. Differ. Geom. Int. Press, Boston, MA, 2012, pp. 327–410 DOI: 10.4310/SDG.2012.v17.n1.a8
  • [29] D.. Phong, Jian Song, Jacob Sturm and Ben Weinkove “The Kähler-Ricci flow and the ¯\overline{\partial} operator on vector fields” In J. Differential Geom. 81.3, 2009, pp. 631–647 URL: http://projecteuclid.org/euclid.jdg/1236604346
  • [30] D.. Phong, Jian Song, Jacob Sturm and Ben Weinkove “The Moser-Trudinger inequality on Kähler-Einstein manifolds” In Amer. J. Math. 130.4, 2008, pp. 1067–1085 DOI: 10.1353/ajm.0.0013
  • [31] D.. Phong and Jacob Sturm “Lectures on stability and constant scalar curvature” In Current developments in mathematics, 2007 Int. Press, Somerville, MA, 2009, pp. 101–176
  • [32] Artem Pulemotov and Yanir A. Rubinstein “Ricci iteration on homogeneous spaces” In Trans. Am. Math. Soc. 371.9, 2019, pp. 6257–6287 DOI: 10.1090/tran/7498
  • [33] Yanir A. Rubinstein “Some discretizations of geometric evolution equations and the Ricci iteration on the space of Kähler metrics” In Adv. Math. 218.5, 2008, pp. 1526–1565 DOI: 10.1016/j.aim.2008.03.017
  • [34] Yanir A. Rubinstein “The Ricci iteration and its applications” In C. R., Math., Acad. Sci. Paris 345.8, 2007, pp. 445–448 DOI: 10.1016/j.crma.2007.09.020
  • [35] Yanir A. Rubinstein “Tian’s properness conjectures: an introduction to Kähler geometry” In Geometric analysis—in honor of Gang Tian’s 60th birthday 333, Progr. Math. Birkhäuser/Springer, Cham, 2020, pp. 381–443 DOI: 10.1007/978-3-030-34953-0“˙16
  • [36] Santiago R. Simanca “Heat flows for extremal Kähler metrics” In Ann. Sc. Norm. Super. Pisa Cl. Sci. (5) 4.2, 2005, pp. 187–217
  • [37] R.. Thomas “Notes on GIT and symplectic reduction for bundles and varieties” In Surveys in differential geometry. Vol. X 10, Surv. Differ. Geom. Int. Press, Somerville, MA, 2006, pp. 221–273 DOI: 10.4310/SDG.2005.v10.n1.a7
  • [38] Gang Tian “Canonical metrics in Kähler geometry” Notes taken by Meike Akveld, Lectures in Mathematics ETH Zürich Birkhäuser Verlag, Basel, 2000, pp. vi+101 DOI: 10.1007/978-3-0348-8389-4
  • [39] Gang Tian “Existence of Einstein metrics on Fano manifolds” In Metric and differential geometry 297, Progr. Math. Birkhäuser/Springer, Basel, 2012, pp. 119–159 DOI: 10.1007/978-3-0348-0257-4“˙5
  • [40] Gang Tian “On Kähler-Einstein metrics on certain Kähler manifolds with C1(M)>0C_{1}(M)>0 In Invent. Math. 89.2, 1987, pp. 225–246 DOI: 10.1007/BF01389077
  • [41] Gang Tian “The KK-energy on hypersurfaces and stability” In Comm. Anal. Geom. 2.2, 1994, pp. 239–265 DOI: 10.4310/CAG.1994.v2.n2.a4
  • [42] Gang Tian, Shijin Zhang, Zhenlei Zhang and Xiaohua Zhu “Perelman’s entropy and Kähler-Ricci flow on a Fano manifold” In Trans. Amer. Math. Soc. 365.12, 2013, pp. 6669–6695 DOI: 10.1090/S0002-9947-2013-06027-8
  • [43] Gang Tian and Zhenlei Zhang “Regularity of Kähler-Ricci flows on Fano manifolds” In Acta Math. 216.1, 2016, pp. 127–176 DOI: 10.1007/s11511-016-0137-1
  • [44] Gang Tian and Zhou Zhang “On the Kähler-Ricci flow on projective manifolds of general type” In Chinese Ann. Math. Ser. B 27.2, 2006, pp. 179–192 DOI: 10.1007/s11401-005-0533-x
  • [45] Gang Tian and Xiaohua Zhu “Convergence of Kähler-Ricci flow” In J. Am. Math. Soc. 20.3, 2007, pp. 675–699 DOI: 10.1090/S0894-0347-06-00552-2
  • [46] Gang Tian and Xiaohua Zhu “Convergence of the Kähler-Ricci flow on Fano manifolds” In J. Reine Angew. Math. 678, 2013, pp. 223–245 DOI: 10.1515/crelle.2012.021
  • [47] Ahmed Zeriahi “Volume and capacity of sublevel sets of a Lelong class of plurisubharmonic functions” In Indiana Univ. Math. J. 50.1, 2001, pp. 671–703 DOI: 10.1512/iumj.2001.50.2062