1 Introduction
1.1 Background and motivation
Let
$(X,\omega _0)$
be a compact Kähler manifold, and let
$\omega ^{\bullet }(t),t\in [0,T)$
be a family of Kähler metrics on X which solve the Kähler-Ricci flow

for some
$0<T\leqslant +\infty $
. In this paper, we are interested in the case when
$\omega ^{\bullet }(t)$
is an immortal solution (i.e., when
$T=+\infty $
). Thanks to a result of Tian-Zhang [Reference Tian and Zhang37] (see also [Reference Tsuji45]), we know that the solution
$\omega ^{\bullet }(t)$
is immortal if and only if the canonical bundle
$K_X$
is nef, which means that
$c_1(K_X)$
lies in the closure of the cone of Kähler classes in
$H^{1,1}(X,\mathbb {R})$
. This condition does not depend on
$\omega _0$
, and manifolds with
$K_X$
nef are also known as smooth minimal models.
The Abundance Conjecture in birational geometry, and its natural extension to Kähler manifolds, predicts that if the canonical bundle of a compact Kähler manifold is nef, then it must be semiample, which means that
$K_X^p$
is base-point-free for some
$p\geqslant 1$
. This conjecture is known when
$\dim X\leqslant 3$
by [Reference Campana, Höring and Peternell1, Reference Das and Ou6, Reference Das and Ou7].
Throughout the rest of the paper, we will assume that
$K_X$
is semiample. It is then known (see, for example, [Reference Lazarsfeld25, Theorem 2.1.27]) that global sections of
$K_X^p$
for
$p\geqslant 1$
sufficiently divisible define a surjective holomorphic map
$f:X\to B\subset \mathbb {CP}^N$
(the Iitaka fibration of X) with connected fibers onto a normal projective variety B (known as the canonical model of X), of dimension m equal to the Kodaira dimension of X (in particular, we have
$0\leqslant m\leqslant \dim X$
). The smooth fibers of f are then Calabi-Yau manifolds, of dimension
$n:=\dim X-m$
, which are pairwise diffeomorphic but in general are not pairwise biholomorphic.
In the two extreme cases when
$m=0$
or
$m=\dim X$
, the behavior of the flow is completely understood thanks to the work of many people (see, for example, the recent survey [Reference Tosatti42] and references therein), so we will furthermore assume from now on that
$0<m<\dim X$
, which is known as ‘intermediate Kodaira dimension’. Thus, we have
$\dim X=m+n$
, and both the fibers and the base of f are positive-dimensional.
The simplest examples of this setup arise when
$m=n=1$
, where X is a minimal properly elliptic surface, B is a compact Riemann surface, and
$f:X\to B$
is an elliptic fibration. In this case, the behavior of the Kähler-Ricci flow (1.1) was first studied by Song-Tian [Reference Song and Tian31], who shortly afterwards also considered the case of general
$m,n$
in [Reference Song and Tian32]. A major difficulty in this setting is that the total volume of
$(X,\omega ^{\bullet }(t))$
is easily seen to converge to zero as
$t\to +\infty $
, and this ‘collapsing’ behavior makes it extremely hard to analyze the flow. As we will now explain, in [Reference Song and Tian31, Reference Song and Tian32], it was shown that the metrics
$\omega ^{\bullet }(t)$
collapse in the weak topology to the pullback of a canonical metric on B, and our main goal is to obtain higher order regularity and a uniform Ricci curvature bound for
$\omega ^{\bullet }(t)$
(away from the singular fibers of f) and thus prove two conjectures of Song-Tian.
When X is projective, the condition that
$K_X$
be nef means that X is a smooth minimal model. The connection between the Minimal Model Program in birational geometry and the behavior of the Kähler-Ricci flow was first discovered independently by Cascini-La Nave [Reference Cascini and La Nave3] and Song-Tian [Reference Song and Tian31], and remains an area of current research. These works outlined a conjectural picture for the behavior of the Kähler-Ricci flow on any projective (or more generally compact Kähler) manifold. When
$K_X$
is not nef, singularities must develop in finite time, and the flow should implement the corresponding birational contractions or collapse the fibers of a Mori fiber space. The case when
$K_X$
is nef (so the manifold is a smooth minimal model) is the topic of our paper.
1.2 Setup
We now describe our setup in more detail. As mentioned above, we have a compact Kähler manifold
$(X^{m+n},\omega _0)$
with semiample canonical bundle and intermediate Kodaira dimension m (so
$m,n>0$
), and
$\omega ^{\bullet }(t)$
denotes the immortal solution of the Kähler-Ricci flow (1.1). Let
$f:X\to B$
be the Iitaka fibration of X, and let
$S\subset X$
be the preimage of the union of the set of singular values of f and the singular set of B. Thus, by construction,
$f:X\backslash S\to B\backslash f(S)$
is a proper holomorphic submersion with n-dimensional connected Calabi-Yau fibers
$X_{z}=f^{-1}(z), z\in B\backslash f(S)$
. By Ehresmann’s Lemma (and the connectedness of
$B\backslash f(S)$
), the fibers
$X_z$
are pairwise diffeomorphic, but, in general, their complex structure varies with z, and this variation can be encoded in a smooth semipositive Weil-Petersson form
$\omega _{\mathrm {WP}}\geqslant 0$
on
$B\backslash f(S)$
, defined in [Reference Song and Tian31] (see also [Reference Tosatti39, §5.6]).
By [Reference Song and Tian32], there exists a smooth Kähler metric
$\omega _{\mathrm {can}}$
on
$B\backslash f(S)$
satisfying the twisted Kähler-Einstein equation

The pullback of
$\omega _{\mathrm {can}}$
to
$X\backslash S$
will also be denoted by the same symbol, for convenience. The metric
$\omega _{\mathrm {can}}$
extends to a closed positive current on B, and in [Reference Song and Tian31, Reference Song and Tian32] it is shown that as
$t\to +\infty $
, we have

weakly as currents on X as well as in the
$C^{0}_{\mathrm {loc}}(X\backslash S)$
topology of Kähler potentials. Motivated by this, in [Reference Song and Tian31, p.612], [Reference Song and Tian32, p.306], [Reference Tian35, Conjecture 4.5.7], [Reference Tian36, p.258], Song-Tian posed the following:
Conjecture 1.1. Let
$(X,\omega _0)$
be a compact Kähler manifold with
$K_X$
semiample and intermediate Kodaira dimension
$0<m<\dim X$
, and let
$\omega ^{\bullet }(t)$
solve (1.1). Then the convergence (1.3) happens in the locally smooth topology as tensors on
$X\backslash S$
.
Explicitly, Conjecture 1.1 asks to show that given any
$K\Subset X\backslash S$
and
$k\in \mathbb {N}$
, we have

There have been a number of partial results towards Conjecture 1.1, often using techniques that were first developed for a family of elliptic PDEs that describe the collapsing of families of Ricci-flat Kähler metrics on a Calabi-Yau manifold with a fibration structure, and which share some of the features of (1.1) (see, for example, the survey [Reference Tosatti40]). Indeed, Fong-Zhang [Reference Fong and Zhang12] adapted work of the third-named author [Reference Tosatti38] to prove that (1.3) holds in the
$C^{1,\alpha }_{\mathrm { loc}}(X\backslash S)$
topology of Kähler potentials (
$\alpha <1$
), and the works [Reference Fong and Zhang12, Reference Hein and Tosatti19, Reference Tosatti and Zhang44] proved Conjecture 1.1 when the smooth fibers of f are tori or finite quotients of tori (see also [Reference Gill13] and [Reference Tosatti39, §5.14]), using and improving a method of Gross-Tosatti-Zhang [Reference Gross, Tosatti and Zhang14]. Later, Tosatti-Weinkove-Yang proved that (1.3) holds in
$C^0_{\mathrm {loc}}(X\backslash S)$
, and this was improved to
$C^{\alpha }_{\mathrm {loc}}(X\backslash S),\alpha <1$
by Chu-Lee [Reference Chu and Lee4] adapting the techniques of Hein-Tosatti [Reference Hein and Tosatti20], which also allowed Fong-Lee [Reference Fong and Lee11] to prove Conjecture 1.1 when all smooth fibers are pairwise biholomorphic.
In a later work [Reference Song and Tian33], Song-Tian proved that the scalar curvature of
$\omega ^{\bullet }(t)$
remains uniformly bounded on X, independent of
$t\geqslant 0$
. They then conjectured a similar statement for the Ricci curvature, away from the singular fibers of f (see [Reference Tian36, Conjecture 4.7]):
Conjecture 1.2. Let
$(X,\omega _0)$
be a compact Kähler manifold with
$K_X$
semiample and intermediate Kodaira dimension
$0<m<\dim X$
, and let
$\omega ^{\bullet }(t)$
solve (1.1). Then the Ricci curvature of
$\omega ^{\bullet }(t)$
remains uniformly bounded on compact subsets of
$X\backslash S$
, independent of t.
This is only known when the smooth fibers of f are tori, or finite quotients of tori [Reference Fong and Lee11] (hence, in particular, it holds on minimal properly elliptic surfaces), or when the smooth fibers are pairwise biholomorphic [Reference Chu and Lee4]. It is known that, in general, the conjectural Ricci bound cannot be improved to a full Riemann curvature bound (on compact subsets of
$X\backslash S$
): by [Reference Tosatti and Zhang44], this holds if and only if the smooth fibers are tori or finite quotients.
It is well known that the Kähler-Ricci flow (1.1) reduces to a scalar PDE, of parabolic complex Monge-Ampère type, for a family of evolving Kähler potentials. Following [Reference Song and Tian32], we construct a closed real
$(1,1)$
-form
$\omega _F$
on
$X\backslash S$
, which is of the form
$\omega _F=\omega _0+{i\partial \bar \partial }\rho $
, such that for every
$z\in B\backslash f(S)$
, we have that
$\omega _F|_{X_z}$
is the unique Ricci-flat Kähler metric on
$X_z$
cohomologous to
$\omega _0|_{X_z}$
. While
$\omega _F$
is not semipositive definite in general (see [Reference Cao, Guenancia and Păun2] for a counterexample), given any compact set
$K\Subset X\backslash S$
, we can find
$t_0$
such that for all
$t\geqslant t_0$
,

is a Kähler metric on K, with fibers of size
$\approx e^{-t/2}$
and base of size
$\approx 1$
. On
$X\backslash S$
, we can then write
$\omega ^{\bullet }(t)=\omega ^\natural (t)+{i\partial \bar \partial }\varphi (t)$
, where the potentials
$\varphi (t)$
satisfy

for
$t\geqslant 0$
(see, for example, [Reference Tosatti39, §5.7] and [Reference Tosatti, Weinkove and Yang43, §3.1]. Then, since we know the weak convergence in (1.3), Conjecture 1.1 is equivalent to the a priori estimates

for all
$k\in \mathbb {N}$
and all
$t\geqslant 0$
. Furthermore, since
$\varphi (t)$
is uniformly bounded in
$L^\infty (X)$
by [Reference Song and Tian32] (which uses [Reference Demailly and Pali8, Reference Eyssidieux, Guedj and Zeriahi9] – see also [Reference Guo, Phong and Tong17] for a new proof), these estimates are also equivalent to

for all
$k\in \mathbb {N}$
and all
$t\geqslant 0$
.
1.3 Main result
The main result of this paper gives a complete solution of Conjectures 1.1 and 1.2:
In fact, in both conjectures, we prove much more precise statements. The higher order estimates for
$\omega ^{\bullet }(t)$
are derived as consequences of a very detailed asymptotic expansion for
$\omega ^{\bullet }(t)$
, which is in the same spirit as the expansion recently obtained in [Reference Hein and Tosatti21] by Hein-Tosatti for collapsing Ricci-flat metrics on Calabi-Yau manifolds. As for the Ricci curvature bound, we show that on
$X\backslash S$
we have

where on any fixed compact subset of
$X\backslash S$
we have
$|\mathrm {Err}|_{g^{\bullet }(t)}\to 0$
, as
$t\to +\infty $
. Thus, in a strong sense, the Ricci curvature of the evolving metrics
$\omega ^{\bullet }(t)$
is asymptotic to
$-\omega _{\mathrm {can}}$
. Furthermore, our bound on the Ricci curvature (and on all of the pieces of the asymptotic expansion of the metric) is an a priori bound: it only depends on the uniform constants in Lemma 4.1, which are due to [Reference Fong and Zhang12, Reference Song and Tian33, Reference Tosatti, Weinkove and Yang43].
The starting point of our analysis, which was proved in [Reference Fong and Zhang12] by adapting [Reference Tosatti38] in the elliptic setting, is the following estimate: given
$K\Subset X\backslash S$
, there is
$C>0$
such that on K we have

for all
$t\geqslant t_0$
. In other words,
$\omega ^{\bullet }(t)$
is shrinking in the fiber directions and remains of bounded size in the base directions. Since the linearized operator of (1.6) is the time-dependent heat operator of
$\omega ^{\bullet }(t)$
, we see from (1.10) that the ellipticity is degenerating in the fiber directions as
$t\to +\infty $
, and so there is no clear way to approach the a priori estimates (1.8). Indeed, the local analog of such estimates are false; see the discussion in [Reference Hein and Tosatti20] in the elliptic case.
However, it turns out that we can work locally on the base (but using crucially that the fibers are compact without boundary), and since f is differentiably a locally trivial fiber bundle over
$B\backslash f(S)$
, we may without loss assume that our base B is now simply the Euclidean unit ball in
$\mathbb {C}^m$
, and
$f:B\times Y\to B$
is just the projection onto the first factor, where Y is a closed manifold and
$B\times Y$
is equipped with a complex structure J (not necessarily a product) such that f is
$(J,J_{\mathbb {C}^m})$
holomorphic. The fibers
$\{z\}\times Y, z\in B$
are then compact n-dimensional Calabi-Yau manifolds diffeomorphic to Y. Under this trivialization, the Ricci-flat Kähler metric
$\omega _F|_{X_z}$
defines a Riemannian metric
$g_{Y,z}$
on
$\{z\}\times Y$
, which we extend trivially to
$B\times Y$
, and we use these to define a family of shrinking Riemannian product metrics

on
$B\times Y$
, which are uniformly equivalent to
$\omega ^\natural (t)$
and hence to
$\omega ^{\bullet }(t)$
. We will also denote by
$g(t):=g_0(t)$
the shrinking product metrics with z equal to the origin in B.
1.4 Overview of the proof
As in [Reference Hein and Tosatti20, Reference Hein and Tosatti21], the first attempt to overcome the issue of degenerating ellipticity is to try to prove much more – namely, try to prove uniform bounds for
$\varphi (t)$
or
$\omega ^{\bullet }(t)$
in the shrinking norms
$C^k(K,g(t))$
, since
$g^{\bullet }(t)$
is uniformly equivalent to
$g(t)$
. This, however, cannot be proved in general since we know from [Reference Tosatti, Weinkove and Yang43] that
$e^t\omega ^{\bullet }(t)|_{X_z}$
converge smoothly to
$\omega _F|_{X_z}$
, and since
$g_{Y,z}$
and
$g_{Y,z'}$
are not in general parallel with respect to each other, the shrinking
$C^k$
norms of
$g_{z}(t)$
and
$g_{z'}(t)$
are not uniformly equivalent as
$t\to +\infty $
. To address this issue, the first and third-named authors defined in [Reference Hein and Tosatti21] a connection
${\mathbb {D}}$
on
$B\times Y$
which on each fiber
$\{z\}\times Y$
acts like the Levi-Civita connection of
$g_z(t)$
, and using its parallel transport operator, they defined new shrinking
$C^{k,\alpha }$
norms,
$0<\alpha <1$
. We will consider the natural parabolic extension of these norms to space-time derivatives in Section 2 below. Since parabolic Hölder seminorms behave differently according to the parity of k, we will only work with
$k=2j$
even (cf. Remark 2.5).
The hope would then be to show that
$\omega ^{\bullet }(t)-\omega ^\natural (t)={i\partial \bar \partial }\varphi (t)$
is uniformly bounded in these shrinking
$C^{2j+\alpha ,j+{\alpha }/2}$
norms. This turns out to be true when
$j=0$
, but false starting from
$j=1$
. This phenomenon, which was discovered in [Reference Hein and Tosatti21] in the elliptic setting, manifests itself only when the complex structure J is not a product and the fibers are not tori or quotients. In a nutshell, the variation of complex structures, and the non-flatness of
$g_{z}(t)$
, destroy these desired shrinking norm bounds. However, with much work, we are able to construct a collection of ‘obstruction functions’ on
$B\times Y$
(up to shrinking B) and decompose the solution
${i\partial \bar \partial }\varphi (t)$
into a sum of finitely many terms
$\gamma _1(t),\dots ,\gamma _j(t)$
(constructed roughly speaking using the fiberwise
$L^2$
projections of
$\Delta ^{g^\natural (t)}\varphi (t)$
onto the space of obstructions), and a remainder
$\eta _j(t)$
. We then show via a contradiction and blowup argument that the remainder
$\eta _j(t)$
is bounded in the shrinking
$C^{2j+\alpha ,j+{\alpha }/2}$
norm, while the terms
$\gamma _1(t),\dots ,\gamma _j(t)$
are not, but they satisfy strong enough estimates which guarantee that they are bounded in the
$C^{2j+\alpha ,j+{\alpha }/2}$
norm of a fixed metric
$\omega _X$
on X. As mentioned earlier, the higher order estimates on all these pieces depend only on the constant in the
$C^0$
estimate (1.10), and on the other constants that appear in Lemma 4.1 (including the uniform bound on the scalar curvature of
$\omega ^{\bullet }(t)$
from [Reference Song and Tian33]), and thus ultimately, they depend only on the geometry of X and on the initial metric
$\omega _0$
.
This procedure is iterated by replacing j with
$j+1$
, and new obstruction functions are constructed by measuring the failure of the remainder
$\eta _j(t)$
to be bounded in the shrinking
$C^{2(j+1)+\alpha ,j+1+{\alpha }/2}$
norm. This way, we can split
$\eta _j(t)=\gamma _{j+1}(t)+\eta _{j+1}(t)$
and obtain the next order in the expansion. As in [Reference Hein and Tosatti21], there is an extra technical difficulty, which arises from the fact that the terms
$\gamma _j(t)$
are constructed by plugging in
$\eta _{j-1}(t)$
and the obstruction functions into an approximate elliptic Green operator, which has an extra parameter
$k\in \mathbb {N}$
that measures the quality of the approximation. Thus, all the terms in the expansion also end up depending on k, which is large and chosen a priori, and the procedure works for
$j\leqslant k$
.
The resulting asymptotic expansion of
$\omega ^{\bullet }(t)$
is described in detail in Theorem 4.2 below, which is the main technical result of the paper. It is the parabolic analog of [Reference Hein and Tosatti21, Theorem 4.1], and its proof follows the same overall method via blowup and contradiction, but there are some new key difficulties. First, as mentioned earlier, the (shrinking) parabolic Hölder norms that we use are better behaved when the order of derivatives is even, which compels us to use
$C^{2j+\alpha ,j+{\alpha }/2}$
norms instead of
$C^{j+\alpha ,(j+{\alpha })/2}$
(see, for example, Lemma 2.4 and Remark 2.5). More importantly, since the approximate Green operator that we use in this paper is the same as in [Reference Hein and Tosatti21], it provides an approximate parametrix for the Laplacian of
$\omega ^\natural (t)$
(in a rough sense) but not for the heat operator (it seems far from clear that a similar strategy could be implemented with an approximate heat kernel construction). Because of this, to obtain a contradiction at the end of the blowup argument (which is divided into 3 cases, with the last case itself divided into 3 subcases A, B and C), we now have to deal with new terms that come from taking time derivatives of the solution, which are not taken care of by construction, unlike [Reference Hein and Tosatti21]. To make matters worse, in the blowup argument, the evolving Kähler potential has
$L^\infty $
norm that is blowing up, so it cannot be passed to a limit to obtain a contradiction. Dealing with these issues requires substantial work.
Another new difficulty, compared to [Reference Hein and Tosatti21], is that the case
$j=0$
(i.e., where we prove
$C^{{\alpha },{\alpha }/2}$
estimates) does not behave in the same way as the cases
$j\geqslant 1$
because the parabolic complex Monge-Ampère equation also involves
$\varphi (t)$
without derivatives landing on it, unlike the elliptic complex Monge-Ampère equation where only
${i\partial \bar \partial }\varphi $
enters. To deal with this issue, we employ a different blowup quantity for
$j=0$
, which is closer in spirit to our earlier works [Reference Hein and Tosatti20, Reference Fong and Lee11]. As a result, different ideas will be required to close the blowup argument, according to whether
$j=0$
or
$j\geqslant 1$
. Furthermore, when
$j\geqslant 1$
, we are forced to add one new term to the main blowup quantity (when compared to [Reference Hein and Tosatti21]), to gain better control on the fiber average of the Kähler potential and its time derivative, and we later have to show that this new term can be dealt with in the blowup argument. Next, in subcase A, dealing with these terms forces us to refine the Selection Theorem 3.1 where the obstruction functions are chosen, and when
$j=0$
, we need a whole new argument. In subcase B, we employ an energy argument inspired by [Reference Fong and Lee11, Claim 3.2], and in subcase C a different energy argument has to be applied fiber by fiber.
Once the asymptotic expansion is established, the smooth convergence of Conjecture 1.1 follows easily. However, proving the Ricci curvature bound for
$\omega ^{\bullet }(t)$
in Conjecture 1.2 requires substantial work by plugging in the expansion with
$j=1$
into the formula for the Ricci curvature (as two derivatives of the logarithm of the volume form) and using our explicit a priori estimates for the terms of the expansion to deduce boundedness of Ricci. Here again, we encounter a new difficulty compared to [Reference Hein and Tosatti21], which arises from the fact that one of the estimates in Theorem 4.2 is weaker than the corresponding one in the elliptic setting, because of the fact that we can only work with even order norms. During the course of the proof of the Ricci bound, we also prove a fact of independent interest in Proposition 5.1, by showing that
$\varphi +\dot {\varphi }$
minus its fiberwise average decays to zero (away from the singular fibers) faster than
$e^{-t}$
(see (5.33)). This improves on earlier work of Fong-Zhang [Reference Fong and Zhang12, p.110] (see also [Reference Tosatti39, Lemma 5.13]) and Tosatti-Weinkove-Yang [Reference Tosatti, Weinkove and Yang43, Lemma 3.1 (iv)].
Remark 1.4. We conjecture that the Ricci curvature of
$\omega ^{\bullet }(t)$
remains uniformly bounded also near the singular fibers of f. One could imagine settling this for some minimal elliptic surfaces by developing a parabolic version of the Gross-Wilson gluing result in [Reference Gross and Wilson15] (thanks to J. Lott for this suggestion), and for some Lefschetz fibered
$3$
-folds by developing a parabolic version of Li’s gluing result in [Reference Li27].
Remark 1.5. It is natural to ask whether we really need to assume that our compact Kähler manifold X with
$K_X$
nef satisfies the Abundance Conjecture (thanks to S. Karigiannis and J. Cheng for raising this point). The reader can verify that the results in [Reference Fong and Zhang12, Reference Song and Tian32, Reference Song and Tian33, Reference Tosatti, Weinkove and Yang43] on which we rely, as well as our main theorems, are also valid under the a priori weaker assumption that
$c_1(K_X)$
is a semiample
$(1,1)$
-class [Reference Tosatti41, Def.3.4]: there is a surjective holomorphic map
$f:X\to B$
with connected fibers onto a normal compact Kähler analytic space B such that
$c_1(K_X)=f^{*}[\omega ]$
for some Kähler class
$[\omega ]$
on B. However, a very recent result of Das-Hacon [Reference Das and Hacon5, Theorem 4.4], which was prompted by our questions to C. Hacon as well as the related [Reference Tosatti41, Question 3.5], shows that under this hypothesis
$K_X$
is already semiample, and it is elementary to deduce from this that f is the Iitaka fibration of
$K_X$
. We thank also M. Păun for discussions about this point.
1.5 Organization of the paper
In Section 2 we introduce our parabolic shrinking norms and seminorms and prove an interpolation inequality, the crucial Proposition 2.6 and a Schauder estimate. Section 3 contains the proof of the Selection Theorem 3.1 where the obstruction functions are selected. Section 4 is the main part of the paper and is where the asymptotic expansion is proved in Theorem 4.2. Lastly, in Section 5, we give the proof of our main Theorem 1.3.
2 Parabolic Hölder norms and interpolation
The setup where we are working in was described in the Introduction.
2.1
${\mathbb {D}}$
-derivatives
Recall that our main goal is to establish higher order estimates for the metrics
$\omega ^{\bullet }(t)$
on
$B\times Y$
which evolve by the normalized Kähler-Ricci flow (1.1). We know from Lemma 4.1 (i) below that
$\omega ^{\bullet }(t)$
is uniformly equivalent to
$\omega ^\natural (t)=(1-e^{-t})\omega _{\mathrm {can}}+e^{-t}\omega _F$
, which is shrinking in the fiber directions as
$t\to +\infty $
. As mentioned above in the overview of proof, the fiberwise Ricci-flat metrics
$g_{Y,z}$
are in general quite different from each other as
$z\in B$
varies, and this forces us to define a new connection
${\mathbb {D}}$
which along each fiber
$\{z\}\times Y$
acts like the Levi-Civita connection of
$g_z(t)=g_{\mathbb {C}^m}+e^{-t}g_{Y,z}$
. This is what was achieved by the first and third-named authors in [Reference Hein and Tosatti21, §2.1], and we now recall their construction.
Definition 2.1. For
$z\in B\subset \mathbb {C}^m$
, we let
$\nabla ^z$
be the Levi-Civita connection of the product metric
$g_{z}(t)=g_{\mathbb {C}^m}+e^{-t} g_{Y,z}$
on
$B\times Y$
, which is independent of
$t\geqslant 0$
. Let
${\mathbb {D}}$
be the connection on the tangent bundle of
$B\times Y$
and on all of its tensor bundles defined by

for all tensors
$\eta $
on
$B\times Y$
and
$x\in B\times Y$
.
For the detailed discussion of the properties of
${\mathbb {D}}$
, we refer readers to [Reference Hein and Tosatti21, §2.1]. Given a curve
$\gamma $
in
$B\times Y$
which contains the points
$a,b$
, we let
$\mathbb {P}^\gamma _{ab}$
denote the
${\mathbb {D}}$
-parallel transport from a to b along the
$\gamma $
. A curve
$\gamma $
is called a
$\mathbb {P}$
-geodesic if
$\dot {\gamma }$
is
${\mathbb {D}}$
-parallel along
$\gamma $
. Two examples of
$\mathbb {P}$
-geodesics are horizontal paths
$(z(t),y_0)$
where
$z(t)$
is an affine segment in
$\mathbb {C}^m$
, and vertical paths
$(z_0,y(t))$
where
$y(t)$
is a
$g_{Y,z_0}$
-geodesic in
$\{z_0\}\times Y$
. These are the only
$\mathbb {P}$
-geodesics that we will use in the paper, as every two points in
$B\times Y$
can be connected by concatenating two of these
$\mathbb {P}$
-geodesics, where the vertical one is minimal. We may also write
$\mathbb {P}_{ab}$
instead of
$\mathbb {P}^\gamma _{ab}$
if the
$\mathbb {P}$
-geodesic
$\gamma $
joining a and b is not emphasized.
2.2
$\mathfrak {D}$
-derivatives
${\mathbb {D}}$
-derivatives that we just defined are spatial derivatives. It will be very convenient to use a similar shorthand notation when we also allow time derivatives. Thus, given a time-dependent contravariant tensor
$\eta $
and
$k\in \mathbb {N}$
, we define

which is a sum of tensors of different types. We will also use the notation

when we only take spatial base derivatives, as well as time derivatives. Observe also that if g is any Riemannian product metric on
$B\times Y$
, then we have the pointwise equality

which we will use implicitly many times.
In our setting,
$\{g_{Y,z}\}_{z\in B},$
is a smooth family of Riemannian metrics on Y, so (up to shrinking B slightly) we can find
$\Lambda>1$
so that

In particular, the norm measured with respect to
$g_{Y,0}$
is uniformly comparable to that of
$g_{Y,z}$
for
$z\in B$
.
2.3 Hölder seminorms
We now use the connection
${\mathbb {D}}$
to define a parabolic Hölder norm on
$B\times Y\times [0,+\infty )$
. For
$p=(z,y)\in B\times Y,t\geqslant 0, 0<R\leqslant \sqrt {t}$
and (shrinking) product metrics
$g_\zeta (\tau )=g_{\mathbb {C}^m}+e^{-\tau } g_{Y,\zeta }$
, we define the parabolic domain

The parabolic domain with respect to any other product metric is defined analogously. We will very often simply take
$\zeta =0\in B$
.
Definition 2.2. For any
$0<{\alpha }<1$
,
$R>0$
,
$p\in B\times Y$
,
$t\geqslant 0$
and smooth tensor field
$\eta $
on
$B\times Y\times [t-R^2,t]$
, given a product metric g (such as
$g=g_z(\tau )$
for some
$z\in B$
and
$\tau \geqslant 0$
), we define

where the supremum is taken among all
$(x,s)$
and
$(x',s')$
in
$Q_{g,R}(p,t)$
in which x and
$x'$
are either horizontally or vertically joined by a
$\mathbb {P}$
-geodesic.
In the case when we use
$g=g_z(\tau )$
and
$\tau $
is allowed to go to
$+\infty $
, we will refer to these as shrinking parabolic Hölder seminorms. Nevertheless, for each fixed
$R>0$
, we will have

for all
$\tau>\tau _0(R,Y)$
. This will be the setting where the parabolic Hölder seminorm are applied in the whole paper. In this case, we will simply denote it by

when the metric g and the shrinking rate
$\tau $
play no role.
Lastly, as in [Reference Hein and Tosatti21, (4.101)], it will also be useful to consider (shrinking) parabolic Hölder seminorms
$[\eta ]_{{\alpha },{\alpha }/2,\mathrm {base}, Q_{g,R}(p,t),g}$
which are defined as in (2.7) but where the supremum is taken only among
$(x,s)$
and
$(x',s')$
in
$Q_{g,R}(p,t)$
such that x and
$x'$
are horizontally joined by a
$\mathbb {P}$
-geodesic.
2.4 Parabolic interpolation
We need an interpolation inequality between the highest order (i.e.,
$C^{k+{\alpha },(k+{\alpha })/2}$
) and the lowest order (i.e.,
$L^\infty $
) norms of a tensor. In the parabolic framework, it will be more convenient to interpolate with the top even order (cf. Remark 2.5). This can be viewed as a parabolic version of [Reference Hein and Tosatti21, Proposition 2.8], and as in there it is crucial that the constants in the interpolation inequality are independent of the shrinking size parameter
$\tau \geqslant 0$
.
Proposition 2.3. For any
$k\in \mathbb {N}_{>0}$
and
${\alpha }\in (0,1)$
, there exists
$C_k=C_k({\alpha },\Lambda )>0$
(where
$\Lambda $
is given in (2.5)) such that the following holds. Let
$\eta $
be a smooth contravariant p-tensor on
$B\times Y$
. Then for all
$(x_0, t_0)\in B\times Y\times \mathbb {R}$
,
$ 0<\rho <R$
and
$\tau \geqslant 0$
such that
$ Q_{g_0(\tau ),R}(x_0,t_0)\Subset B\times Y\times \mathbb {R}$
, we have

Moreover, for any
$j\in \mathbb {N}$
and
${\beta }\in (0,1)$
with
$j+{\beta }<2k+{\alpha }$
, we have

Proof. We first show (2.10). Fix a pair of
$(p,q)$
such that
$0<j=p+2q\leqslant 2k$
, and assume first that
$p>0$
. Since
$d^{g_0(\tau )}(x,x_0)<\rho $
for
$(x,t)\in Q_{g_0(\tau ),\rho }(x_0,t_0)$
, we can treat
$\partial _t^q \eta |_{(x,t)}$
as a smooth tensor on
$B_{\mathbb {C}^m}(z_0,\rho )\times Y$
by freezing t so that [Reference Hein and Tosatti21, Proposition 2.8] applies to conclude

Thus, it remains to show the interpolation on time derivatives (i.e., we assume in the rest that
$p=0$
, so
$q>0$
). For each
$(x,t)\in Q_{g_0(\tau ),\rho }(x_0,t_0)$
, fix
$s=(R-\rho )^2>0$
so that
$(x,t-s)\in Q_{g_0(\tau ),R}(x_0,t_0)$
. Then there exists
$t_s\in [t-s,t]$
so that

which allows us to estimate

If
$q=k$
, we arrive at

Otherwise,
$q<k$
, and we have

Applying this dichotomy inductively, with suitable replacements of
$\rho $
and R at each step, we conclude that there exists
$C>0$
so that for each
$1\leqslant q\leqslant k$
,

By combining this with (2.12), we see that (2.10) follows.
It remains to prove (2.11). Fix
$(x,t),(x',s)\in Q_{g_0(\tau ),\rho }(x_0,t_0)$
such that x and
$x'$
are joined either horizontally or vertically by a
$\mathbb {P}$
-geodesic. Denote
$d=d^{g_0(\tau )}(x,x')+|t-s|^{\frac {1}{2}}$
. We want to estimate
$|\mathfrak {D}^j \eta (x,t)-\mathbb {P}_{x'x} \mathfrak {D}^j\eta (x',s)|_{g_0(\tau )}$
. Fix a pair of
$(p,q)$
such that
$0<p+2q=j\leqslant 2k$
and
${\beta }\in (0,1)$
with
$j+\beta <2k+\alpha $
.
If
$d\geqslant \frac 1{4\Lambda } (R-\rho )$
where
$\Lambda $
is the constant in (2.5), then using the triangle inequality and the boundedness of the operator norm of
$\mathbb {P}$
from [Reference Hein and Tosatti21, §2.1.1], we deduce that

so that the conclusion follows from (2.10).
If
$d<\frac 1{4\Lambda } (R-\rho )$
,
$j=2k$
and
${\beta }<{\alpha }$
, then

which is acceptable. It remains to consider the case when
$d<\frac 1{4\Lambda } (R-\rho )$
and
$j=p+2q<2k$
. Here, using again the boundedness of
$\mathbb {P}$
, we can estimate

where the first term is the spatial Hölder seminorm of the tensor
$\partial _t^q\eta (t)$
with t frozen, and
$x=(z,y)\in \mathbb {C}^m\times Y$
. Applying [Reference Hein and Tosatti21, Lemma 2.5] to the first term in the last line of (2.18) gives

As for the second term in the last line of (2.18), assume first that
$j+2=p+2(q+1)\leqslant 2k$
. In this case, we can argue similarly by estimating the difference in term of time derivatives

Hence, under the assumption that
$j+2=p+2(q+1)\leqslant 2k$
, we can combine (2.18), (2.19) and (2.20) to get

The conclusion then follows by combining with (2.10) since
$d\leqslant \frac 1{4\Lambda }(R-\rho )$
.
It remains to consider the case when
$p+2q<2k<p+2q+2$
(i.e.,
$p+2q=2k-1$
). This implies that p is odd, and hence,
$p\geqslant 1$
. Let v run over a
$g_0(\tau )$
-orthonormal basis of tangent vectors which are either horizontal or vertical. Let
$\gamma (u)$
be the unique
$\mathbb {P}$
-geodesic with
$\gamma (0)=x$
and
$\dot \gamma (0)=v$
with
$u\in (0,R-\rho )$
. Denote
$\sigma (u,\cdot )=\mathbb {P}^{-1}_{\gamma (0),\gamma (u)}{\mathbb {D}}^{p-1}\partial _t^{q} \eta (\gamma (u),\cdot )$
so that
${\mathbb {D}}_v {\mathbb {D}}^{p-1}\partial _t^{q}\eta (x,\cdot )=\partial _u|_{u=0}\sigma (u,\cdot )$
. By the mean value theorem, there exists
$\theta \in [0,1]$
such that

where
$\sigma '$
denotes the u-derivative. Since v is arbitrary, we conclude that

The result then follows using (2.10) again. This completes the proof.
We end this subsection by showing that a function on
$\mathbb {R}^n\times (-\infty ,0]$
with bounded
$(2k+{\alpha })$
parabolic Hölder seminorm and vanishing parabolic
$2k$
jet at
$(x,t)=(0,0)$
will be bounded in
$C^{2k+{\alpha },k+{\alpha }/2}_{\mathrm {loc}}$
.
Lemma 2.4. Let u be a smooth function on
$\mathbb {R}^n\times (-\infty ,0]$
such that

for some
$R,\Lambda _0>0,k\in \mathbb {N}$
and
$\mathfrak {D}^{\ell }u|_{(0,0)}=0$
for all
$0\leqslant \ell \leqslant 2k$
. Then for all
$0<r\leqslant R$
and
$0\leqslant m\leqslant 2k$
, there exists
$C_0(n,m)>0$
such that

Moreover, for all
${\beta }\in (0,{\alpha })$
, there exists
$C_1(n,\Lambda _0,{\beta })>0$
such that

Remark 2.5. This Lemma is false as stated if we replace
$2k$
with an odd integer, and this is the main reason why in our main Theorem 4.2, we will restrict to even order derivatives. The simplest counterexample is the function
$u(x,t)=t$
in
$\mathbb {R}\times \mathbb {R}$
, which satisfies
$u(0,0)=0$
,
$\mathfrak {D} u|_{(0,0)}=0$
and
$[\mathfrak {D} u]_{{\alpha },{\alpha }/2,\mathbb {R}\times \mathbb {R}}=0$
but (2.25) fails for
$m=0$
.
To fix this, one has to redefine the parabolic Hölder seminorms of odd order by adding an additional term; see [Reference Lieberman29, p.46]. If one were to do this, then the statement of Lemma 2.4 would also hold when
$2k$
is replaced by an odd integer. However, the additional term that one would need to add would not be compatible with our blowup arguments in section 4, especially with the ‘non-escaping property’ in Section 4.5.
Proof. Write
$Q_r=B_{\mathbb {R}^n}(r)\times [-r^2,0]$
for notational convenience. We only prove the bound for
$\|\mathfrak {D}^m u\|_{\infty }$
in (2.25) since the bound for Hölder seminorm in (2.26) is similar.
By considering
$\tilde u(x,t)=\Lambda _0 ^{-1}R^{-2k-{\alpha }}u(Rx,R^2t)$
for
$(x,t)\in B_{\mathbb {R}^n}(1)\times [-1,0]$
, we can assume
$\Lambda _0=1=R=1$
and
$0<r\leqslant 1$
. We prove the result by induction on k. In case
$k=0$
, the jet assumption is equivalent to
$u(0,0)=0$
, and hence for all
$0<r\leqslant 1$
,

so that the conclusion holds.
Next, we consider the induction step, so we assume that the conclusion holds for all
$0\leqslant \ell \leqslant k$
and prove it for
$k+1\geqslant 1$
. Given a smooth function u with
$[\mathfrak {D}^{2k+2}u]_{{\alpha },{\alpha }/2,Q_1}\leqslant 1$
, given any
$0\leqslant m\leqslant 2k$
, every derivative
$\mathfrak {D}^{m+2}u$
can be written as
$\mathfrak {D}^m v$
where
$v=\partial _t u$
or
$v={\mathbb {D}}^2 u$
(evaluated at some pair of tangent vectors). The function v satisfies
$[\mathfrak {D}^{2k}v]_{{\alpha },{\alpha }/2,Q_1}\leqslant 1$
and
$\mathfrak {D}^{\ell }v|_{(0,0)}=0$
for all
$0\leqslant \ell \leqslant 2k$
. The induction hypothesis then implies
$\|\mathfrak {D}^{m+2} u\|_{\infty ,Q_r}\leqslant C_{k} r^{2k+{\alpha }-m}$
for all
$0\leqslant m\leqslant 2k$
and
$0<r\leqslant 1$
. It remains to extend it to
$m=-1,-2$
(i.e., to bound u and
${\mathbb {D}} u$
). Let
$(x,t)\in Q_r$
and fix a unit vector
$e_1$
, and estimate

and we bound each of the numbered terms as follows. By the mean value theorem, there exists
$\theta \in [0,1]$
such that

and using [Reference Hein and Tosatti21, Lemma 2.5], we can bound this by
$r\|\mathfrak {D}^2 u\|_{\infty ,Q_r}\leqslant C r^{2k+{\alpha }+1}.$
For the second term, the mean value theorem again shows that there is
$\theta '\in [0,1]$
so that

For the third term, using the mean value theorem and [Reference Hein and Tosatti21, Lemma 2.5], we can find
$\theta "\in [0,1]$
so that

and for the fourth term, we again use [Reference Hein and Tosatti21, Lemma 2.5] to bound

and putting these all together proves that
$|{\mathbb {D}}_1 u(x,t)|\leqslant Cr^{2k+{\alpha }+1}$
, and hence,
$\|\mathfrak {D} u\|_{\infty ,Q_r}\leqslant C_{k} r^{2k+{\alpha }+1}$
since
$e_1$
is arbitrary. The upper bound for
$|u|$
is now straightforward using the bounds on
${\mathbb {D}} u$
and
$\partial _t u$
. This completes the proof of the inductive step.
2.5 Bounds on Hölder seminorms imply decay
In this section, we establish a generalization of [Reference Hein and Tosatti21, Theorem 2.11] to our setting. Recall that at each point
$x=(z,y)\in B\times Y$
,
$\omega _{F,z}$
is the unique Kähler-Ricci flat metric on each fiber
$X_z$
which is cohomologous to
$\omega _0|_{X_z}$
. We can assume that
$\int _{\{z\}\times Y} \omega _{F,z}^n=1$
for all
$z\in B$
. For any function f in space-time
$B\times Y\times \mathbb {R}$
, we will use
$\underline{f}(z,t)$
to denote its fiberwise average:

The following result will be crucial for us:
Proposition 2.6. Suppose
$g=g_{\mathbb {C}^m}+\delta ^2 g_{Y,0}$
is a metric on
$B\times Y$
, where
$0<\delta \leqslant 1$
is arbitrary. For any
$k\in \mathbb {N}, {\alpha }\in (0,1)$
and
$0<\rho <R<1$
with
$\rho \geqslant \Lambda \delta $
, there exists
$C(k,{\alpha },\rho ,R)>0$
such that for all smooth function
$\varphi $
on
$B\times Y\times \mathbb {R}$
with
$\underline{\varphi }=0$
,
$x_0=(0,y_0), t_0\in \mathbb {R}$
and for all
$0\leqslant j\leqslant 2k$
, we have

Moreover, for all
${\beta }\in (0,1)$
such that
$j+{\beta }<2k+{\alpha }$
,

Moreover, the same estimates hold if
$\varphi $
is replaced by
$\eta ={i\partial \bar \partial }\varphi $
where
$\underline{\varphi }=0$
.
Remark 2.7. We require
$\rho \geqslant \Lambda \delta $
here so as to ensure that

which is needed in order to apply [Reference Hein and Tosatti21, Theorem 2.11].
Proof. Suppose we can show that

Then (2.34) and (2.35) would follow from this and the interpolation Proposition 2.3, as in [Reference Hein and Tosatti21, (2.61)–(2.62)]. To prove (2.37), given
$(x,t)\in Q_{g,\rho }(x_0,t_0)$
, write as usual
$x=(z,y)$
and freeze the time variable in
$\varphi (\cdot ,t)$
. Assuming first that
$k\geqslant 1$
, similarly to [Reference Hein and Tosatti21, (2.81)], we can use
$\underline{\varphi }(t)=0$
to estimate

where in the second inequality, we used [Reference Hein and Tosatti21, Lemma 2.10], and (2.37) follows when
$k\geqslant 1$
. Lastly, when
$k=0$
, the argument is straightforward: using again that
$\underline{\varphi }(t)=0$
, we bound

which completes the proof of (2.34) and (2.35) for
$\varphi $
. Lastly, the analogous estimates for
$\eta ={i\partial \bar \partial }\varphi $
follow in a similar fashion, by first using interpolation Proposition 2.3 to reduce ourselves to proving

and then proving (2.40) by freezing the time variable t and applying [Reference Hein and Tosatti21, (2.61)] to
$\eta (\cdot , t)$
and getting

which concludes the proof.
2.6 Parabolic Schauder estimates
In the course of the proof of our main Theorem, we also need two parabolic Schauder estimates on cylinders, which will be used when linearizing the Kähler-Ricci flow equation, and which are analogs of [Reference Hein and Tosatti21, Proposition 2.15]. Let
$(z_{\ell },y_{\ell })\to (z_\infty ,y_\infty )$
be a convergent family of points in
$B\times Y$
. For
$\ell \geqslant 1$
, consider the diffeomorphism
$\Lambda _{\ell }: (\check z,\check y)\mapsto (z_{\ell }+e^{-t_{\ell }/2}\check z,\check y)$
, and let
$\check J_{\ell }$
be the pullback of the complex structure J via
$\Lambda _{\ell }$
, which converges to
$\check J_\infty =J_{\mathbb {C}^m}+J_{Y,z_\infty }$
locally smoothly. Similarly, we let
$\check {{\mathbb {D}}}_{\ell }$
denote the pullback of the connection
${\mathbb {D}}$
so that
$\check {{\mathbb {D}}}_{\ell }\to \check {{\mathbb {D}}}_\infty =\nabla ^{\mathbb {C}^m}+\nabla ^{g_{Y,z_\infty }}$
locally smoothly in spacetime. By the translation, we may assume that our new base point is
$\check p_{\ell }=(\check z_{\ell },\check y_{\ell })=(0,\check y_{\ell })\to (0,\check y_\infty )$
. We rescale the geometric quantities in a parabolic way centered at
$t_{\ell }$
, such as, for example,
$\check g_{\ell }(\check t)=e^{t_{\ell }}\Lambda _{\ell }^{*} g(t_{\ell }+e^{-t_{\ell }}\check t),$
where recall that we denote by
$g(t)=g_{\mathbb {C}^m}+e^{-t}g_{Y,0}$
, so that
$\check {g}_{\ell }(0)=g_{\mathbb {C}^m}+g_{Y,0}$
.
The first Schauder estimate is for scalar functions:
Proposition 2.8. Let
$U\subset \mathbb {C}^m\times Y$
be an open set containing
$(0,y_\infty )$
. Let
$\check g_{\ell }, \check {\omega }_{\ell }^\sharp $
be Riemannian resp.
$\check J_{\ell }$
-Kähler metrics on U that converge locally smoothly to a Riemannian resp.
$\check J_\infty $
-Kähler metric
$\check {g}_\infty ,\check {\omega }_\infty ^\sharp $
on U. Then for all
$a\in \mathbb {N}_{>0}$
,
${\alpha }\in (0,1)$
and
$R>0$
, there exists
$\ell _0>0$
and
$C>0$
such that for all
$0<\rho <R$
,
$\ell \geqslant \ell _0$
and all smooth function u defined on
$U\times \mathbb {R}$
, we have that

whenever
$\check Q_{\check g_{\ell }(0),R}(\check p_{\ell },0 )\subset U\times \mathbb {R}$
.
Proof. We let
$\sigma $
be small enough so that with respect to
$\check g_\infty =g_{\mathbb {C}^m}+g_{Y,z_\infty }$
, the
$\check g_\infty $
-geodesic ball of size
$\sigma $
is geodesically convex and admits a normal coordinate chart centered at any
$(z,y)\in \mathbb {C}^m\times Y$
. This is possible since Y is compact and
$\mathbb {C}^m$
is flat with respect to
$\check g_\infty $
. Since
$\check g_{\ell }(t)\to g_{\mathbb {C}^m}+g_{Y,z_\infty }$
as
$\ell \to +\infty $
and
$\check {\mathbb {D}}_{\ell }\to \check {\mathbb {D}}_\infty $
locally smoothly, for any large
$R>0$
, there exists
$\ell _0$
such that for all
$\ell>\ell _0$
and
$(\check p,\check t)\in \check Q_{\check g_{\ell }(0),2R}(\check p_{\ell },0)$
,
$B_{\check g_{\ell }(0)}(\check p,\sigma )$
is geodesically convex and is compactly contained in a Euclidean ball of radius
$2\sigma $
. Moreover, we can assume that the Kähler structure
$\check {\omega }^\sharp _{\ell }, \check J_{\ell }$
is smoothly close to the product structure
$\check {\omega }^\sharp _\infty , \check {J}_\infty $
in the parabolic domain
$\check Q_{\check g_{\ell }(0),2R}(\check p_{\ell },0 )$
.
The product metric
$\check g_{\ell }(0)$
is uniformly comparable to the Euclidean metric on
$\check Q_{\check g_{\ell }(0),\sigma }(\check p,\check t)$
for all
$(\check p,\check t)\in \check Q_{\check g_{\ell }(0),R}(\check p_{\ell },0 )$
. By the standard Euclidean parabolic Schauder estimates (see, for example, [Reference Lieberman29, Theorem 4.9 and p.84]), there exists
$C>0$
such that for all
$0<\rho _1<\rho _2<\sigma $
, and functions u on
$\check Q_{\check g_{\ell }(0),\sigma }(\check p,\check t)$
where
$(\check p,\check t)\in \check Q_{\check g_{\ell }(0),R}(\check p_{\ell },0 )$
, we have

where the derivatives, Hölder norms are computed using the standard Euclidean metric and
$\Delta _{\check \omega _{\ell }^\sharp }$
denotes the Laplacian of the metric
$\check \omega _{\ell }^\sharp $
.
We first use interpolation to eliminate the terms in the last line of (2.43). For
$0<\rho _1<\rho _2<\sigma $
, we let
$\rho _2^{\prime }=\frac 12 (\rho _1+\rho _2)$
so that (2.43) holds with
$\rho _2$
replaced by
$\rho _2^{\prime }$
and C replaced by a slightly larger C. Hence, the standard Euclidean interpolation (or Proposition 2.3) yields

where we have used
$\rho _2^{\prime }-\rho _2=\frac 12 (\rho _2-\rho _1)$
.
We now want to estimate the
$L^\infty $
norm of
$\left(\frac {\partial }{\partial t}-\Delta _{\check \omega _{\ell }^\sharp }\right)u$
in terms of the
$L^\infty $
norm of u and of the Hölder seminorm of top order derivatives of u. To do this, we interpolate again. Let
$\varepsilon \in (0,1)$
be a small constant to be determined, and given
$0<\rho _1<\rho _2<\sigma $
, denote
$\rho _2^{\prime \prime }=\rho _1+(1-\varepsilon ) (\rho _2-\rho _1)$
. We consider (2.44) with
$\rho _2$
replaced by
$\rho _2^{\prime \prime }$
. Since
$\check \omega _{\ell }^\sharp $
is uniformly bounded in
$C^\infty $
, the standard Euclidean interpolation (or Proposition 2.3) yields

Therefore, inserting (2.44) and (2.45) into (2.43), we can choose
$\varepsilon $
sufficiently small so that

for all
$0<\rho _1<\rho _2<\sigma $
. We can then apply the iteration lemma in [Reference Hein and Tosatti21, Lemma 2.9] to obtain

We now claim that one can interchange the Euclidean derivatives
$\partial $
and Euclidean parallel transport P in the definition of Hölder norms with
$\check {\mathbb {D}}_{\ell }$
and
$\check {\mathbb {P}}_{\ell }$
. Although in our definition of Hölder norms we only consider horizontal and vertical
$\mathbb {P}$
-geodesics while the standard Euclidean Hölder norms consider all possible segments, it is immediate to see from [Reference Hein and Tosatti21, (2.31)] that the difference is harmless. To compare
$\partial $
with
$\check {\mathbb {D}}_{\ell }$
, we note that on each
$\check Q_{\check g_{\ell }(0),\sigma }(\check p,\check t)$
, we can write
$\partial =\check {\mathbb {D}}_{\ell }+\check \Gamma _{\ell }$
(indeed independent of
$\check {t}$
) so that
$\check \Gamma _{\ell }\to \check \Gamma _\infty $
where
$\check \Gamma _\infty (\check p)=0$
by our choice of normal coordinate centered at
$\check p$
. In particular, the local smooth convergence implies that
$\check \Gamma _{\ell }\to 0$
in
$C^k$
for fixed k uniformly as
$\ell \to +\infty $
and
$\sigma \to 0$
, and then ODE estimates show that
$P-\check {\mathbb {P}}_{\ell }\to 0$
in
$C^k$
. In particular, switching from
$\partial ^p$
to
$\check {\mathbb {D}}_{\ell }^p$
will generate an error of the form (ignoring combinatorial constants):

Since
$0<\rho _1<\rho _2\leqslant \sigma <1$
, and since all terms in (2.48) with
$\check {\Gamma }_{\ell }$
are
$o(1)$
, the right-hand side in (2.48) can be absorbed in the second term in the left-hand side of (2.43). Thus, going back to (2.47), we can change the
$\partial ,P$
with
$\check {D}_{\ell },\check {\mathbb {P}}_{\ell }$
, and using interpolation again, we obtain

where the coefficient
$\frac 12$
in the last term is achieved by choosing
$\sigma $
sufficiently small thanks to the local smooth convergence of
$\check \Gamma _{\ell }$
. We fix
$\sigma $
from now on. It now follows from [Reference Hein and Tosatti21, Lemma 2.9] again that we have

for all
$0<\rho _1<\rho _2<\sigma $
. This in particular shows the desired conclusion for all small
$0<\rho _1<\rho _2<\sigma $
with arbitrary center
$(\check p,\check t)$
in the compact set
$ \check Q_{\check g_{\ell }(0),R}(\check p,0 )$
.
Now we prove the Hölder control on
$ \check Q_{\check g_{\ell }(0),\rho }(\check p_{\ell },0 )$
for
$0<\rho <R$
. We can assume
$R\geqslant \sigma $
. If
$\rho <\frac 12 \sigma \leqslant \frac 12 R$
so that
$\sigma -\rho \geqslant \frac 12\sigma \geqslant C^{-1}(R-\rho )$
, then (2.50) implies

Hence, it remains to consider the case
$\rho \geqslant \frac 12\sigma $
. Given any two points
$(\check p,\check t)$
and
$(\check q,\check s)$
in
$ \check Q_{\check g_{\ell }(0),\rho }(\check p_{\ell },0 )$
with
$r=d^{\check g_{\ell }(0)}(\check p,\check q)+|\check t-\check s|^{\frac {1}{2}}$
, we choose a sequence of points
$\{(\check p_i,\check t_i) \}_{i=1}^N$
inside
$\check Q_{\check g_{\ell }(0),\rho }(\check p_{\ell },0 )$
such that
$(\check p_1,\check t_1)=(\check p,\check t)$
and
$(\check p_N,\check t_N)=(\check q,\check s)$
. We can choose them in a way so that
$r_i=d^{\check g_{\ell }(0)}(\check p_i,\check p_{i+1})+|\check t_i-\check t_{i+1}|^{\frac {1}{2}}\leqslant \frac 14 \sigma $
is uniformly comparable to r and
$N\leqslant C \sigma ^{-2}$
. Here, the square comes from the time direction. Moreover, we can assume
$ \check Q_{\check g_{\ell }(0),\frac 14\sigma }(\check p_i,\check t_i)\subset \check Q_{\check g_{\ell }(0),\rho }(\check p_{\ell },0 ) $
for all
$i=2,\dots ,N-1$
(i.e., except
$(\check p,\check t)$
and
$(\check q,\check s)$
). For each
$i=1,\dots ,N-1$
, we can apply (2.50) again to obtain

for all
$i=1,\dots ,N-1$
. Here, we have used the fact that
$\sigma \geqslant C^{-1}R\geqslant C^{-1}(R-\rho )$
. Since

using (2.52) completes the proof after taking supremum over all
$(\check p,\check t)$
and
$(\check q,\check s)$
in
$\check Q_{\check g_{\ell }(0),\rho }(\check p_{\ell },0 )$
.
The second Schauder estimate is for real
$(1,1)$
-forms:
Proposition 2.9. Let
$U\subset \mathbb {C}^m\times Y$
be an open set containing
$(0,y_\infty )$
. Let
$\check g_{\ell }, \check {\omega }_{\ell }^\sharp $
be Riemannian resp.
$\check J_{\ell }$
-Kähler metrics on U that converges locally smoothly to a Riemannian resp.
$\check J_\infty $
-Kähler metric
$\check {g}_\infty ,\check {\omega }_\infty ^\sharp $
on U. Then for all
$a\in \mathbb {N}_{>0}$
,
${\alpha }\in (0,1)$
and
$R>0$
, there exists
$\ell _0>0$
and
$C>0$
such that for all
$0<\rho <R$
,
$\ell \geqslant \ell _0$
and all real
$\check {J}_{\ell }$
-
$(1,1)$
form
$\eta $
defined on
$U\times \mathbb {R}$
, we have that

whenever
$\check Q_{\check g_{\ell }(0),R}(\check p_{\ell },0 )\subset U\times \mathbb {R}$
, and where
$\Delta _{\check \omega _{\ell }^\sharp }\eta $
denotes the Hodge Laplacian of
$\check \omega _{\ell }^\sharp $
acting on differential forms.
Proof. The proof shares some similarities with the proof of Proposition 2.8. After the same preliminary remarks as there, we first work in the Euclidean setting, with the Hodge Laplacian
$\Delta _{\check \omega _{\ell }^\sharp }\eta $
. We can then apply standard parabolic Schauder estimate to the uniformly parabolic system given by
$(\partial _t-\Delta _{\check \omega _{\ell }^\sharp })$
acting on real
$(1,1)$
-forms [Reference Ladyženskaja, Solonnikov and Ural’ceva24, Chapter 7], which shows that there exists
$C>0$
such that for all
$0<\rho _1<\rho _2<\sigma $
, and
$\eta $
lives in
$\check Q_{\check g_{\ell }(0),\sigma }(\check p,\check t)$
where
$(\check p,\check t)\in \check Q_{\check g_{\ell }(0),R}(\check p_{\ell },0 )$
, we have

where the derivatives and Hölder norms are computed using standard Euclidean metric.
We first eliminate the terms in the last line of (2.55) by interpolation: given
$0<\rho _1<\rho _2<\sigma $
, we let
$\rho _2^{\prime }=\frac {1}{2}(\rho _1+\rho _2)$
, so that (2.55) holds with
$\rho _2$
replaced by
$\rho _2^{\prime }$
, and so standard interpolation gives

using that
$\rho ^{\prime }_2-\rho _2=\frac {1}{2}(\rho _2-\rho _1)$
. We now want to estimate the
$L^\infty $
norm of
$\left(\partial _t-\Delta _{\check \omega _{\ell }^\sharp } \right) \eta $
. Let
$\varepsilon \in (0,1)$
be a constant to be determined, and denote by
$\rho _2^{\prime \prime }=\rho _2+\varepsilon (\rho _2-\rho _1)$
. By interpolation and the
$C^\infty $
boundedness of
$\check g_{\ell }^\sharp $
, we have

Since
$a\geqslant 1$
, we can choose
$\varepsilon $
small enough so that
$C\varepsilon ^{2a+{\alpha }-2}\leqslant \frac 12$
. Inserting this in (2.56) (replacing
$\rho _2^{\prime \prime }$
by
$\rho _2$
), and plugging into (2.55) gives

and the iteration lemma in [Reference Hein and Tosatti21, Lemma 2.9] then gives

This is the direct analog of (2.47). After this, the rest of the proof of (2.54) proceeds exactly as in the proof of Proposition 2.8.
3 The Selection Theorem
In our main theorem, we will need the analog of the Selection Theorem [Reference Hein and Tosatti21, Theorem 3.11], adapted to our parabolic setting, and to the specific structure of the parabolic complex Monge-Ampère equation that we are dealing with. As in [Reference Hein and Tosatti21], to state this, we will need some preparatory notation. First, we fix two natural numbers
$0\leqslant j\leqslant k$
and a Euclidean ball B centered at the origin, and as usual, we have fixed the fiberwise Calabi-Yau volume forms on the fibers
$\{z\}\times Y$
. Given
$t,k$
and two smooth functions
$A\in C^\infty (B,\mathbb {R})$
and
$G\in C^\infty (B\times Y,\mathbb {R})$
with fiberwise average zero and fiberwise
$L^2$
norm
$1$
, we constructed in [Reference Hein and Tosatti21, §3.2] a function
$\mathfrak {G}_{t,k}(A,G)\in C^\infty (B\times Y,\mathbb {R})$
also with fiberwise average zero, which is in some appropriate sense an approximate right inverse of
$\Delta ^{\omega ^\natural _t}$
applied to
$AG$
; see [Reference Hein and Tosatti21, Lemma 3.7] for a precise statement. We will need the following quasi-explicit formula from [Reference Hein and Tosatti21, Lemma 3.8]: given any
$A,G$
as above, there are t-independent smooth functions
$\Phi _{\iota , r}(G)$
on
$B\times Y$
such that for all
$t\geqslant 0$
, we have

where here and in the rest,
$\circledast $
denotes some tensorial contraction, and we have

For all
$i<j$
, we suppose that we have smooth functions
$G_{i,p,k}\in C^\infty (B\times Y,\mathbb {R}), 1\leqslant p\leqslant N_{i,k}$
, which have fiberwise average zero and are fiberwise
$L^2$
orthonormal, with
$G_{i,p,k}=0$
when
$i=0$
. The main goal is to find smooth functions
$G_{j,p,k}$
which satisfy a certain property that we now describe.
We are also given sequences of real numbers
$t_{\ell }\to +\infty $
and
$\delta _{\ell }>0$
with
$\delta _{\ell }\to 0$
and
$\lambda _{\ell }:=\delta _{\ell } e^{\frac {t_{\ell }}{2}}\to +\infty $
. Consider the diffeomorphisms

where
$B_R:=B_{\mathbb {C}^m}(0,R)$
, and for any function u on
$B\times Y\times [0,t_{\ell }]$
, we will write
$\check {u}_{\ell }=\Sigma _{\ell }^{*}u$
, and for a time-dependent
$2$
-form
$\alpha $
(with
$t\in [0,t_{\ell }]$
) we will write
$\check {\alpha }_{\ell }=e^{t_{\ell }}\Sigma _{\ell }^{*}\alpha $
. In particular, note that
$\check {\omega }_{\ell , \mathrm {can}}=e^{t_{\ell }}\Sigma _{\ell }^{*}\omega _{\mathrm {can}}$
is a (time-independent) Kähler metric on
$B_{e^{\frac {t_{\ell }}{2}}}$
uniformly equivalent to Euclidean (independent of
$\ell $
).
We will also need to factor
$\Sigma _{\ell } = \Psi _{\ell }\circ \Xi _{\ell }$
, where


and given a function u on
$B\times Y\times [0,t_{\ell }]$
, we will write
$\hat {u}=\Psi _{\ell }^{*}u$
, and given a time-dependent
$2$
-form
$\alpha $
, we will write
$\hat {\alpha }=\lambda _{\ell }^2\Psi _{\ell }^{*}\alpha $
. We will also use the notation

so that for example
$\Xi _{\ell }(\check {Q}_R)=\hat {Q}_{R\delta _{\ell }^{-1}}$
.
Let also
$\eta ^\ddagger _{\ell }$
be an arbitrary sequence of
$(1,1)$
-forms on
$B\times [0, t_{\ell }]$
with coefficients (spacetime) polynomials of degree at most
$2j$
which satisfy
$\hat {\eta }^\ddagger _{\ell }\to 0$
locally smoothly in spacetime (which implies that
$\check {\eta }^\ddagger _{\ell }\to 0$
locally smoothly as well), let
$B^\sharp _{\ell }$
be an arbitrary sequence of smooth functions on
$B\times [0,t_{\ell }]$
such that
$\hat {B}^\sharp _{\ell }\to 0$
locally smoothly, and for
$1\leqslant i\leqslant j$
, let
$A^\sharp _{i,p,k}$
be arbitrary (spacetime) polynomials of degree at most
$2j$
on B such that
$\hat {A}^\sharp _{\ell ,i,p,k}=\lambda _{\ell }^2\Psi _{\ell }^{*}A^\sharp _{i,p,k}$
satisfy that there is some
$0<\alpha _0<1$
such that given any
$R>0$
, there is
$C>0$
with

for all
$0\leqslant \iota \leqslant 2j$
, or equivalently that
$\check {A}^\sharp _{\ell ,i,p,k}=e^{t_{\ell }}\Sigma _{\ell }^{*}A^\sharp _{i,p,k}=\delta _{\ell }^2\Xi _{\ell }^{*}\hat {A}^\sharp _{\ell ,i,p,k}$
satisfy

With these, we define for
$1\leqslant i\leqslant j$

so
$\gamma ^{\sharp }_{t,j,k}$
depends on how we choose the functions
$G_{j,p,k}$
. It is proved by the argument in (4.200) below that our assumption (3.7) on
$\hat {A}^\sharp _{\ell ,i,p,k}$
implies that for any
$R>0$
, there is
$C>0$
with

or equivalently

for all
$\iota \geqslant 0$
and all
$1\leqslant i\leqslant j$
(these can be just taken as assumptions for now). Observe that the constants C in (3.10) and (3.11) depend on the choice of the functions
$G_{j,p,k}$
, but the exponent
$\alpha _0$
does not. We also define

and

which has the property that
$\hat {\omega }^\sharp _{\ell }$
is a Kähler metric on
$\hat {Q}_R$
for all
$R>0$
, and
$\ell $
sufficiently large, using that
$\hat {\eta }^\ddagger _{\ell }$
is pulled back from B and goes to zero locally uniformly, and the estimate (3.10) with
$\iota =0$
for
$\hat {\eta }^\dagger _{\ell }$
. Passing to the check picture, we obtain

The key quantity we are interested in is then

which can be compared to the corresponding quantity [Reference Hein and Tosatti21, (3.47)] in the elliptic setting. Observe that by definition, we have
$\Sigma _{\ell }^{*}B^\sharp _{\ell }=\Xi _{\ell }^{*}\hat {B}^\sharp _{\ell }$
.
To clarify, when we will apply the Selection Theorem later, the functions
$B^\sharp _{\ell }$
will be defined by

where
$\partial _{\hat {t}}\underline{{\hat {\chi }}_{\ell }^\sharp }$
is a spacetime polynomial of degree at most
$2j$
. The fact that
$\hat {B}^\sharp _{\ell }\to 0$
locally smoothly will follow from (4.158). We will also later define

Hence, we will have

Given these preliminaries, the following is then the key result:
Theorem 3.1 (Selection Theorem)
Suppose we are given
$0\leqslant j\leqslant k$
and when
$j>1$
, we are also given smooth function
$G_{i,p,k},1\leqslant i\leqslant j-1,1\leqslant p\leqslant N_{i,k}$
, on
$B\times Y$
which are fiberwise
$L^2$
orthonormal and have fiberwise average zero. Then there are a concentric ball
$B'=B_{\mathbb {C}^m}(0,r)\subset B$
and smooth functions
$G_{j,p,k},1\leqslant p\leqslant N_{j,k}$
on
$B'\times Y$
(identically zero if
$j=0$
), with fiberwise average zero so that
$G_{i,p,k},1\leqslant p\leqslant N_{i,k},1\leqslant i\leqslant j$
are all fiberwise
$L^2$
orthonormal with the following property: if
$\delta _{\ell },t_{\ell }>0$
are any sequences with
$t_{\ell }\to +\infty $
,
$\delta _{\ell }\to 0$
and
$\delta _{\ell } e^{\frac {t_{\ell }}{2}}\to +\infty ,$
and if
$A_{i,p,k}^\sharp ,B^\sharp _{\ell },\eta _{\ell }^\dagger ,\eta _{\ell }^\ddagger ,\omega _{\ell }^\sharp $
are as above, and if (3.15) converges locally uniformly on
$\mathbb {C}^m\times Y\times (-\infty ,0]$
to some limiting function
$\mathcal {F}$
as
$\ell \to \infty $
, then on
$\Sigma _{\ell }^{-1}(B'\times Y\times [-r^2,0])$
, we can write (3.15) as

where
$f_{\ell ,0}, f_{\ell ,i,p}$
are functions pulled back from
$B'\times (-r^2,0]$
such that
$\hat {f}_{\ell , 0}=\Psi _{\ell }^{*}f_{\ell ,0},\hat {f}_{\ell ,i,p}=\Psi _{\ell }^{*}f_{\ell ,i,p}$
converge locally smoothly to zero, and
$o(1)$
is a term that converges locally smoothly to zero. Lastly, (3.15) converges to
$\mathcal {F}$
locally smoothly.
Remark 3.2. The argument follows closely the proof of the Selection Theorem 3.11 in [Reference Hein and Tosatti21], but apart from the obvious change from space to space-time, there are some other differences that we now briefly discuss, which arise from the different structure of the parabolic complex Monge-Ampère equation that we have compared to its elliptic counterpart in [Reference Hein and Tosatti21]. The first term in (3.15) is reminescent of [Reference Hein and Tosatti21, (3.47)], but it now has a logarithm. This change will be quite immaterial, since
$\log (x)\approx x-1$
for
$x\approx 1$
. Next, compared to [Reference Hein and Tosatti21, (3.47)], the quantity in (3.15) also contains two more pieces. The last term with
$B^\sharp _{\ell }$
is trivially acceptable since it can be absorbed into
$f_{\ell ,0}$
, while the term involving
$\check {\mathfrak {G}}_{\check {t},k}$
will have to be dealt with, and it will turn out to be ‘lower order’ compared to the first term in (3.15). Putting these all together will allow us to follow the proof of [Reference Hein and Tosatti21, Theorem 3.11] very closely.
Proof. The proof is by induction on j. First, we treat the base case
$j=0$
. In this case, by definition, there are no obstruction functions and the quantity in (3.15) reduces to

As in [Reference Hein and Tosatti21], we introduce the notation

where the functions
$\check \omega _{F,\mathbf {bb}},\check \omega _{F,\mathbf {bf}},\check \omega _{F,\mathbf {ff}}$
so defined are uniformly bounded on
$B_{e^{\frac {t_{\ell }}{2}}}\times Y\times [-e^{t_{\ell }}t_{\ell },0]$
. Following [Reference Hein and Tosatti21, (3.50)], we then compute

where the
$O(e^{-t_{\ell }})$
is in the locally smooth topology, and so

where
$f_{\ell }$
is some sequence of smooth functions pulled back from B. Passing to the hat picture (i.e., letting
$\hat {f}_{\ell }=\Psi _{\ell }^{*}f_{\ell }$
), our assumption that
$\hat {\eta }^\ddagger _{\ell }\to 0$
locally smoothly implies that
$\hat {f}_{\ell }\to 0$
locally smoothly. It thus follows that

and recalling that
$\hat {f}_{\ell }\to 0$
and
$\hat {B}^\sharp _{\ell }\to 0$
locally smoothly, as well as
$\delta _{\ell }^{-\alpha }e^{-{t_{\ell }}}=o(1)$
, we see that (3.19) holds. Lastly, since by assumption (3.20) converges locally uniformly to
$\mathcal {F}$
, the same is true for
$\delta _{\ell }^{-\alpha }\Sigma _{\ell }^{*}(B^\sharp _{\ell }-f_{\ell })$
, and the same argument as in [Reference Hein and Tosatti21, (3.52)–(3.54)] shows that this convergence is locally smooth. This concludes the proof of Theorem 3.1 in the case
$j=0$
.
We then treat the inductive step, and assume
$j\geqslant 1$
. By assumption, the obstruction functions
$G_{i,p,k}$
with
$1\leqslant i<j$
have already been selected on
$B'\times Y$
(recall that
$B'=B_r$
), and we need to select the
$G_{j,p,k}$
’s. As in [Reference Hein and Tosatti21], this will be done via an iterative procedure, with iteration parameter
$\kappa $
, initially set at
$\kappa =0$
, and at each step assuming we have already selected some obstruction functions
$G^{[q]}_{j,p,k}, 1\leqslant q\leqslant \kappa $
(this being the empty list when
$\kappa =0$
), and with the iterative step consisting of selecting some new obstruction functions
$G^{[\kappa +1]}_{j,p,k}$
to add to these. After this will be achieved, we will then show that if we perform this iterative step
$\overline {\kappa }$
times (for some uniform
$\overline {\kappa }$
) and define the obstruction function
$G_{j,p,k}$
by putting together all the
$G^{[q]}_{j,p,k}$
’s obtained at all iterations
$1\leqslant q\leqslant \overline {\kappa }+1$
, then the desired conclusion (3.19) holds.
To start the proof, we give a couple of definitions following [Reference Hein and Tosatti21]. We will say that a sequence of functions on
$\Sigma _{\ell }^{-1}(B_r\times Y\times (-r^2,0])$
satisfies condition
$(\star )$
if it equals

for some
$N\in \mathbb {N}$
, where the functions
$f_{\ell ,0},f_{\ell ,i}$
are smooth and pulled back from
$B_r\times (-r^2,0]$
and
$\hat {f}_{\ell ,0}=\Psi _{\ell }^{*}f_{\ell ,0}, \hat {f}_{\ell ,i}=\Psi _{\ell }^{*}f_{\ell ,i}$
converge locally smoothly to zero, the time-independent functions
$h_i$
are smooth on
$B_r\times Y$
with fiberwise average zero, and the
$o(1)$
is a term that converges locally smoothly to zero. This definition is tailored to our desired conclusion in (3.19). As in [Reference Hein and Tosatti21, Remark 3.13], we see that if a sequence of functions satisfies
$(\star )$
and converges locally uniformly to some limit, then this convergence is actually smooth.
For
$\kappa \geqslant 0$
, given also arbitrary spacetime polynomials
$A^{\sharp ,[q]}_{j,p,k},1\leqslant q\leqslant \kappa ,$
of degree at most
$2j$
such that
$\check {A}^{\sharp ,[q]}_{\ell ,j,p,k}=e^{t_{\ell }}\Sigma _{\ell }^{*}A^{\sharp ,[q]}_{j,p,k}$
satisfy (3.8), we construct as in (3.9)

for
$1\leqslant q\leqslant \kappa $
(setting
$\gamma ^{\sharp ,[0]}_{t,j,k}=0$
) and let

This is a Kähler metric on
$B_r\times Y\times (-r^2,0]$
, and we can then consider the function

The following is the analog of [Reference Hein and Tosatti21, Lemma 3.14]:
Lemma 3.3. Suppose either
$\kappa =0$
or
$\kappa \geqslant 1$
and we have selected the functions
$G_{j,p,k}^{[q]}$
as above for
$1\leqslant q\leqslant \kappa .$
Then the sequence of functions
$\mathcal {B}_{\ell }^{[\kappa ]}$
satisfies
$(\star )$
. Furthermore, we have

where the term
$O(e^{-\alpha _0\frac {t_{\ell }}{2}})$
is in
$L^\infty _{\mathrm {loc}}$
, while the last term is a function from the base which goes to zero locally smoothly.
Proof. For ease of notation, define

which have the property that
$\hat {\omega }^\square _{\ell }$
are Kähler metrics on
$\hat {Q}_R$
for all
$R>0$
and
$\ell $
large, and which in the check picture become

We first consider (in the original undecorated picture)

and so

Note that the functions

with
$q=0$
are
$o(1)$
and pulled back from B, while when
$q>0$
they are not pulled back from B, but they are visibly of the form
$f_{t,0}+\sum _{i=1}^N f_{t,i}h_i$
with the same notation as above, where the
$\hat {f}_{t,0},\hat {f}_{t,i}$
converge smoothly to zero at least as
$O(e^{-qt})$
, and the functions
$h_i$
have fiberwise average zero and do not depend on the choice of
$\eta ^\ddagger _t$
. An analogous statement holds for

and if we take the logarithm and use the Taylor expansion of
$\log (1+x)$
, this shows that

for some possibly different functions
$f_{t,0}, f_{t,i}$
, and where the term
$O(e^{-(j+1)t})$
is in
$C^p_{\mathrm {loc}}$
for all
$p\geqslant 0$
. Passing to the check picture, we have

and so

and
$\delta _{\ell }^{-2j-\alpha }e^{-(j+1)t_{\ell }}=o(1)$
by assumption, namely the LHS of (3.38) satisfies
$(\star )$
, as well as

The next step is to show that the quantity

also satisfies
$(\star )$
. We have just discussed the first piece, and the second piece equals

and for the terms with
$\check {\gamma }^\sharp _{\ell ,1,k}+\cdots +\check {\gamma }^\sharp _{\ell ,j-1,k}+\check {\gamma }^{\sharp ,[1]}_{\ell ,j,k}+\cdots +\check {\gamma }^{\sharp ,[\kappa ]}_{\ell ,j,k}$
, we recall from (3.9) and (3.26) that we have

with the bounds (3.11), where the approximate Green operator
$\check {\mathfrak {G}}_{\check {t},k}$
is given schematically by

by (3.1). Plugging this into (3.41) and arguing as we did above reveals that the quantity in (3.40) satisfies
$(\star )$
and that, furthermore,

Lastly, to prove that
$\mathcal {B}_{\ell }^{[\kappa ]}$
satisfies
$(\star )$
, it remains to consider the piece

The fact that this term satisfies
$(\star )$
again follows immediately from (3.43) together with (3.8), which also give that

and this completes the proof that
$\mathcal {B}_{\ell }^{[\kappa ]}$
satisfies
$(\star )$
and that (3.29) holds.
Now that Lemma 3.3 is established, we can start the first step of the iteration, when
$\kappa =0$
and we need to select the obstruction functions
$G^{[1]}_{j,p,k}$
. To do this, we consider
$\mathcal {B}_{\ell }^{[0]}$
, which by Lemma 3.3 satisfies
$(\star )$
, and let
$\{h_i\}$
be the corresponding functions in its expansion (3.25). Applying the approximate fiberwise Gram-Schmidt [Reference Hein and Tosatti21, Proposition 3.1] to the functions
$h_i$
together with the
$G_{i,p,k}, 1\leqslant i<j$
, produces our desired list
$G^{[1]}_{j,p,k}$
(on
$B_r\times Y$
, up to shrinking r), so that we may assume that the functions
$h_i$
in (3.25) lie in the fiberwise linear span of the
$G_{i,p,k}, 1\leqslant i<j$
together with the
$G^{[1]}_{j,p,k}$
. This completes the first step (
$\kappa =0$
).
Next, we consider a subsequent step
$\kappa \geqslant 1$
of the iteration, so we assume we are given the lists
$G_{i,p,k}, 1\leqslant i<j$
and
$G^{[q]}_{j,p,k}, 1\leqslant q\leqslant \kappa $
; hence, we have the function
$\mathcal {B}_{\ell }^{[\kappa ]}$
in (3.28), and we want to construct the obstruction functions
$G^{[\kappa +1]}_{j,p,k}$
. In order to do this, we must compare
$\mathcal {B}_{\ell }^{[\kappa ]}$
and
$\mathcal {B}_{\ell }^{[\kappa -1]}$
. We have

As in [Reference Hein and Tosatti21, (3.75)–(3.76)], for
$1\leqslant i\leqslant m+n$
, we have

where the
$O(\cdot ), o(\cdot )$
are in the locally smooth topology. Using (3.43), we can write (for any
$1\leqslant q\leqslant \kappa $
)

As in [Reference Hein and Tosatti21, (3.80)], we decompose (3.49) schematically as the sum of
$6$
pieces

(which depend on
$\ell ,j,k$
, but we omit this from the notation for simplicity). Observe here that for all
$(\check {z},\check {y},\check {t})\in \check {Q}_{R\delta _{\ell }^{-1}}$
, we have

where, by assumption,
$e^{-t_{\ell }}\delta _{\ell }^{-2}\to 0$
, so the term
$e^{-re^{-t_{\ell }}\check {t}}$
in (3.50) is
$1+o(1)$
. Now, as in [Reference Hein and Tosatti21, (3.85)–(3.91)], we see that
$\mathrm {I}^{[q]}_{0,0}$
is the dominant term, in the sense that

whenever
$\bullet \neq \mathrm {I}$
or
$(\iota ,r)\neq (0,0)$
. This together with (3.11) implies that

where given any
$R>0$
, there is a
$C>0$
such that for all
$\ell $
,

Relation (3.2) shows that

while from the definition of
$I^{[q]}_{0,0}$
and (3.8), we have

while the argument in [Reference Hein and Tosatti21, (3.83)] gives the reverse bound

and (3.53), (3.54), (3.55), (3.56) imply in particular that

From (3.53), (3.54), (3.55), (3.58) and the Taylor expansion of
$\log (1+x)$
, we see that

where
$F^{[q]}_{\ell }$
satisfies (3.54).
This deals with the term on the second line of (3.47). As for the last line, we can use (3.43) to expand

and applying the obvious parabolic extension of [Reference Hein and Tosatti21, Lemma 3.10] to balls of radius
$R\delta _{\ell }^{-1}$
gives for
$\iota \geqslant 0$
,

and so

Inserting (3.62) into (3.60), and using also (3.57) and the fact that
$e^{-re^{-t_{\ell }}\check {t}}=1+o(1)$
from (3.51), we see that

and combining this with (3.47) and (3.59), we finally obtain

where
$E^{[\kappa ]}_{\ell }$
satisfies

Now, from Lemma 3.3, we know that both
$\mathcal {B}_{\ell }^{[\kappa ]}$
and
$\mathcal {B}_{\ell }^{[\kappa -1]}$
satisfy
$(\star )$
, and from (3.64), we see that so does
$\delta _{\ell }^{-2j-\alpha }E^{[\kappa ]}_{\ell }$
, and so it has an expansion of the form (3.25). As in [Reference Hein and Tosatti21], we apply the approximate fiberwise Gram-Schmidt [Reference Hein and Tosatti21, Proposition 3.1] to the functions
$h_i$
together with the
$G_{i,p,k}, 1\leqslant i<j$
and
$G^{[q]}_{j,p,k}, 1\leqslant q\leqslant \kappa $
; this produces our desired list
$G^{[\kappa +1]}_{j,p,k}$
(on
$B_r\times Y$
, up to shrinking r), so that we may assume that the functions
$h_i$
in (3.25) lie in the fiberwise linear span of the
$G_{i,p,k}, 1\leqslant i<j$
together with the
$G^{[q]}_{j,p,k}, 1\leqslant q\leqslant \kappa +1$
. This completes the step from
$\kappa $
to
$\kappa +1$
in our iterative procedure.
Iterating (3.64) shows that for every
$\kappa \geqslant 1$
, we have

with

for
$1\leqslant q\leqslant \kappa $
, and also

which follows immediately from (3.29).
We can now repeat the iterative step
$\overline {\kappa }:=\lceil \frac {2j+\alpha }{\alpha _0}\rceil $
and then we stop, so the last set of functions which are added to the list are the
$G_{j,p,k}^{[\overline {\kappa }+1]}$
. Our choice of
$\overline {\kappa }$
is made so that

The resulting
$G_{j,p,k}^{[q]}$
with
$1\leqslant q\leqslant \overline {\kappa }+1$
are then renamed simply
$G_{j,p,k}$
. These, together with the
$G_{i,p,k}, 1\leqslant i<j$
, are the obstruction functions that we seek. It remains to show that the statement of the Selection Theorem 3.1 holds with this choice of obstruction functions. By definition, the quantity in (3.15) equals
$\mathcal {B}_{\ell }^{[\overline {\kappa }+1]}$
(up to the term with
$\Sigma _{\ell }^{*}B^\sharp _{\ell }$
, which we can ignore since it can be absorbed into
$f_{\ell ,0}$
in (3.19)). We know that
$\mathcal {B}_{\ell }^{[\overline {\kappa }+1]}$
satisfies
$(\star )$
thanks to Lemma 3.3. As mentioned earlier, because of this we know that if it converges locally uniformly, then it converges locally smoothly, which is the last claim in the Selection Theorem 3.1. The last thing to prove is that if
$\mathcal {B}_{\ell }^{[\overline {\kappa }+1]}$
converges locally uniformly, then (3.19) holds, and thanks to (3.66) (with
$\kappa =\overline {\kappa }+1$
) and to our choice of obstruction functions, it suffices to show that
$\delta _{\ell }^{-2j-\alpha }E^{[\overline {\kappa }+1]}_{\ell }$
is
$o(1)$
in the locally smooth topology. Since this term satisfies
$(\star )$
(as mentioned earlier), it suffices to show that it is
$o(1)$
in the
$L^\infty _{\mathrm {loc}}$
topology, and this follows from (3.66), (3.67), (3.68), (3.69) and our main assumption that
$\mathcal {B}_{\ell }^{[\overline {\kappa }+1]}=O(1)$
in
$L^\infty _{\mathrm {loc}}$
, by using the same iteration argument as [Reference Hein and Tosatti21, (3.101)–(3.112)]. This completes the proof of Theorem 3.1.
4 Asymptotic expansion
In this section, we will prove our main technical result, Theorem 4.2, which gives an asymptotic expansion for the metrics
$\omega ^{\bullet }(t)$
which evolve under the Kähler-Ricci flow (1.1). Recall that
$\omega ^{\bullet }(t)=\omega ^\natural (t)+{i\partial \bar \partial }\varphi (t)$
, where the potentials
$\varphi (t)$
solve the parabolic complex Monge-Ampère equation (1.6), which we can write as

4.1 Known estimates
First, let us recall a few of the known estimates for the Kähler-Ricci flow (1.1) and its equivalent formulation (1.6). There are many other facts that are known about this flow (see, for example, [Reference Tosatti39, §5] or [Reference Tosatti42, §7] for overviews), but the following are the only ones that we will need:
Lemma 4.1. Assume the setup in Section 1.2. Then there exists
$C>0$
such that on
$B\times Y\times [0,+\infty )$
, we have
-
(i)
$C^{-1}\omega ^\natural (t)\leqslant \omega ^{\bullet }(t)\leqslant C\omega ^\natural (t)$ ,
-
(ii)
$|\varphi (t)|+|\dot \varphi (t)|\to 0$ as
$t\to +\infty $ ,
-
(iii)
$|\omega ^{\bullet }(t)- \omega _{\mathrm {can}}|_{g^{\bullet }(t)}\to 0$ as
$t\to +\infty ,$
-
(iv)
$|R({g^{\bullet }(t)})|\leqslant C$ ,
-
(v)
$|\dot {\varphi }(t)+\ddot {\varphi }(t)|\leqslant C$ ,
-
(vi)
$|\nabla (\varphi (t)+\dot {\varphi }(t))|_{g^{\bullet }(t)}\leqslant C.$
Proof. Item (i) is proved in [Reference Fong and Zhang12] (and is an adaptation of [Reference Tosatti38]; see also [Reference Song and Tian31] for the case
$m=n=1$
). Item (ii) is proved in [Reference Tosatti, Weinkove and Yang43, Lemma 3.1], and item (iii) in [Reference Tosatti, Weinkove and Yang43, Theorem 1.2] (see especially the very end of its proof on p.685). Item (iv) is the main theorem of [Reference Song and Tian33], and this implies (v) thanks to the relation [Reference Tosatti39, p.345]

To prove (vi), we use [Reference Song and Tian33, Proposition 3.1] which gives

where
$\Omega $
is a smooth positive volume form on X such that
$-{i\partial \bar \partial }\log \Omega =-\omega _B$
is the pullback of a Kähler form on B. However, by [Reference Tosatti39, Proposition 5.9], we have
$-{i\partial \bar \partial }\log (\omega _{\mathrm { can}}^m\wedge \omega _F^n)=-\omega _{\mathrm {can}},$
so on
$B\times Y$
, we have

which is a
$(1,1)$
-form pulled back from B; hence, the logarithm of the ratio of these two volume forms restricted to the fibers
$\{z\}\times Y$
is pluriharmonic, hence constant. Thus,

for some smooth function G on B, and so using (4.1) and (4.3), we get

as desired.
4.2 Statement of the asymptotic expansion
Given any
$j\in \mathbb {N}, 0\leqslant 2j\leqslant k$
and
$z\in B$
, during the course of the proof of Theorem 4.2 below, we will work by induction on j. By applying repeatedly the Selection Theorem 3.1, and consequently shrinking our ball at each step, we will obtain in particular a collection
$G_{i,p,k}, 1\leqslant i\leqslant j, 1\leqslant p\leqslant N_{i,k}$
of smooth function on
$B\times Y$
with fiberwise average zero, which are fiberwise
$L^2$
orthonormal. For each such
$G_{i,p,k}$
and
$t>0$
, as in [Reference Hein and Tosatti21, (3.6)], we define
$P_{t,i,p,k}=P_{t,G_{i,p,k}}$
, where

for any
$(1,1)$
form
${\alpha }$
on
$B\times Y$
and H with
$\int _{\{z\}\times Y}H\omega _F^n=0$
for all
$z\in \mathbb {C}^m$
. Throughout the proof, we will fix a reference shrinking product metric
$g(t)=g_{\mathbb {C}^m}+e^{-t} g_{Y,z_0}$
. It will only be used to measure the norms and distance but not the connection, and thus the exact choice of
$z_0$
is unimportant thanks to (2.5) (we will usually take
$z_0=0$
). We will also need the t-dependent approximate Green operator
$\mathfrak {G}_{t,k}$
defined in [Reference Hein and Tosatti21, §3.2], to which we refer for its basic properties.
Theorem 4.2. For all
$j,k\in \mathbb {N},0\leqslant 2j\leqslant k$
,
$z\in B$
, there exists
$B'=B_{\mathbb {C}^m}(z,R)\Subset B$
and functions
$G_{i,p,k}, 1\leqslant i\leqslant j, 1\leqslant p\leqslant N_{i,k}$
as above, such that on
$B'\times Y$
, we have a decomposition

with the following properties. For all
${\alpha }\in (0,1)$
and
$r<R$
, there is
$C>0$
such that for all
$t\geqslant 0$
,


where
$Q_r(z,t)=\left(B_{\mathbb {C}^m}(z,r)\times Y\right)\times [t-r^2,t]$
. Furthermore, we have

for
$1\leqslant i\leqslant j$
where
$A_{i,p,k}(t)=P_{t,i,p,k}(\eta _{i-1,k}(t))$
are functions from the base, and we have







Remark 4.3. Key differences between Theorem 4.2 and [Reference Hein and Tosatti21, Theorem 4.1] are the estimates in (4.14), (4.15). These will be crucial for us in the proof, to deal with the term
$\partial _t\varphi +\varphi $
in the complex Monge-Ampère equation (4.1). Another difference is that the bounds in (4.16) are worse than those in [Reference Hein and Tosatti21, (4.12)], due to the fact that in this paper, we can only consider even order Hölder norms.
4.3 Setup of induction scheme
We start with a given
$z\in B$
. For any given
$k\in \mathbb {N}$
, we prove the Theorem by induction on j. We treat both the base case and induction case together, although they will have to be considered separately at certain steps of the proof. Given k with
$0\leqslant 2j\leqslant k$
, if
$j>0$
, we assume Theorem 4.2 holds at the
$(j-1)$
-th step, so there exists
$B_{\mathbb {C}^m}(z,r)\subset B$
such that we already have the decomposition of
$\omega ^{\bullet }(t)$
at the
$(j-1)$
-th step satisfying the desired estimates on
$B_{\mathbb {C}^m}(z,r)\times Y\times [0,+\infty )$
. We aim to refine the decomposition at the j-th step as well as define it for
$j=0$
.
As mentioned in the Introduction, we can write
$\omega ^{\bullet }=\omega ^\natural +{i\partial \bar \partial } \varphi $
. When
$j=0$
, we take
$\gamma _0={i\partial \bar \partial }\underline{\varphi }$
and
$\eta _{0,k}={i\partial \bar \partial } (\varphi -\underline{\varphi })$
so that
$\omega ^{\bullet }=\omega ^\natural +\gamma _0+\eta _{0,k}$
. If
$j\geqslant 1$
, suppose we already have the decomposition

on
$\left(B_{\mathbb {C}^m}(z,r)\times Y\right)\times [0,+\infty )$
. We further decompose
$\eta _{j-1,k}$
into
$\gamma _{j,k}+\eta _{j,k}$
as follows. When
$j>1$
, up to shrinking
$r>0$
, we can assume that we already have selected smooth functions
$G_{i,p,k},1\leqslant i\leqslant j-1, 1\leqslant p\leqslant N_{i,k}$
on
$B_{\mathbb {C}^m}(z,r)\times Y$
, which are fiberwise
$L^2$
orthonormal and have fiberwise average zero. When
$j\geqslant 1$
, we then apply the Selection Theorem 3.1 which up to shrinking r further, gives us a list of functions
$G_{j,p,k}, 1\leqslant p\leqslant N_{j,k}$
on
$B_{\mathbb {C}^m}(z,r)\times Y$
, which are fiberwise
$L^2$
orthonormal and have fiberwise average zero, so that the conclusion of the Selection Theorem 3.1 holds for the collection
$G_{i,p,k}, 1\leqslant p\leqslant N_{i,k},1\leqslant i\leqslant j$
. With this collection of function, we define

where P is given by (4.7) and

where
$\mathfrak {G}_{t,k}$
is defined in [Reference Hein and Tosatti21, §3.2]. Finally, we define
$\eta _{j,k}:=\eta _{j-1,k}-\gamma _{j,k}$
so that

on
$B_{\mathbb {C}^m}(z,r)\times Y\times [0,+\infty )$
. For ease of notation, by scaling and translation of our coordinates, we may assume without loss that we have this decomposition
$B_{\mathbb {C}^m}(z,r)=B_{\mathbb {C}^m}(1)=B$
.
4.3.1 The base case of the induction
$j=0$
The base case of the induction, where
$j=0$
, needs to be treated separately, and although the overall scheme of proof is the same as when
$j\geqslant 1$
, there will be some crucial differences.
First of all, let us examine the estimates that we need to prove in order to establish Theorem 4.2 for
$j=0$
. The estimates (4.16), (4.17), (4.18) are vacuous by definition. By Lemma 4.1 (iii), we have that
$\|{i\partial \bar \partial }\varphi \|_{\infty , B\times Y\times [t-1,t],g(t)}=o(1)$
as
$t\to +\infty $
, and the fiber integration argument in [Reference Tosatti38, p. 436] then gives
$\|\gamma _0\|_{\infty , B\times [t-1,t],g_{\mathbb {C}^m}}=o(1)$
as well, which implies (4.12). Similarly, Lemma 4.1 (ii) implies that
$\|\partial _t\varphi +\varphi \|_{\infty , B\times Y\times [t-1,t]}=o(1)$
, and taking fiberwise average, this easily implies that
$\|\partial _t\underline{\varphi }+\underline{\varphi }\|_{\infty , B\times [t-1,t]}=o(1)$
too, which implies (4.14).
Next, using the bounds
$\|\partial _{t}^2\varphi +\partial _{t}\varphi \|_{\infty , B\times Y\times [t-1,t]}\leqslant C$
and
$\|\nabla (\partial _{t}\varphi +\varphi )\|_{\infty , B\times Y\times [t-1,t], g(t)}\leqslant C$
from Lemma 4.1 (v), (vi), which together with the
$L^\infty $
bound for
$\partial _t\varphi +\varphi $
imply the same bounds for the fiber average

we can bound for any
$x,x'\in B$
and
$t\geqslant 0$
and
$s\in [t-1,t]$
,

which gives

which implies (4.15).
Thus, when
$j=0$
, it suffices to establish (4.9), (4.10) and (4.13). The final claim we will need is that if we suppose we have proved that for all
$t\geqslant 0$
we have

where
$Q_r(z,t)$
is as in the statement of Theorem 4.2, then the estimates (4.9), (4.10) and (4.13) will all hold. To prove this claim, we use the following ‘non-cancellation’ inequality

which is straightforward to prove using [Reference Hein and Tosatti21, (4.215)], except that here there is no stretching involved. Plugging (4.26) and Lemma 4.1 (iii) into (4.27) gives

which is exactly (4.13), and recalling that
$\eta _{0,k}={i\partial \bar \partial }\varphi -{i\partial \bar \partial }\underline{\varphi }$
, we can use (4.26), (4.28), the triangle inequality and the boundedness of
$\mathbb {P}$
to estimate

which proves (4.10). Lastly, (4.9) follows from this and Proposition 2.6, using that the potential
$\varphi -\underline{\varphi }$
of
$\eta _{0,k}$
has fiberwise average zero. This completes the proof of the claim and shows that in order to establish Theorem 4.2 for
$j=0$
, it suffices to prove the single estimate (4.26).
4.3.2 Estimates from induction hypothesis
Suppose
$j\geqslant 1$
and the conclusion holds at the
$(j-1)$
-th step. We first observe that by [Reference Hein and Tosatti21, (4.16)], the operator
$P_{t,j,p,k}$
satisfies

for any
$(1,1)$
form
${\alpha }$
on the total space and
$t\geqslant 0$
. In particular, we can put
${\alpha }=\eta _{i-1,k}(t)$
for
$t\geqslant 0$
and
$1\leqslant i\leqslant j$
and use also (4.9) to see that

for all
${\beta }\in (0,1)$
,
$1\leqslant i\leqslant j$
,
$1\leqslant p\leqslant N_{i,k}$
and
$t\geqslant 0$
.
4.3.3 Reduction to estimating the Hölder seminorms, when
$j\geqslant 1$
Suppose again that
$j\geqslant 1$
and fix a real number
${\alpha }\in (0,1)$
. We first show that (4.9), (4.12), (4.14), (4.16) and (4.17) on
$B_{\mathbb {C}^m}(\rho )\times Y\times [0,+\infty )$
(for some
$\rho <1$
) would follow immediately once we establish the Hölder seminorm bounds (4.10), (4.13), (4.15) and (4.18) on a slightly larger domain.
We first address (4.9) and (4.12). Since the potential of
$\eta _{j,k}$
has fiberwise average zero, (4.9) follows directly from Proposition 2.6 and (4.10). Next, as in Section 4.3.1, we observe that the estimate in Lemma 4.1 (iii) implies that
$\|\gamma _0\|_{\infty , B\times [t-1,t],g_{\mathbb {C}^m}}=o(1)$
. Then, (4.12) follows by interpolating between this and (4.13) using Proposition 2.3. Similarly, the estimate in Lemma 4.1 (ii) implies that
$\|\partial _t\underline{\varphi }+\underline{\varphi }\|_{\infty , B\times [t-1,t]}=o(1)$
, and interpolating between this and (4.15) via Proposition 2.3, we obtain (4.14). The remaining task is to show (4.16) and (4.17). By (4.18), (4.31), we can interpolate from
$Q_\rho (0,t)$
to
$Q_R(0,t)$
(
$\rho <R<1)$
using Proposition 2.3 and get

By choosing

(which is small), we see that for each
$1\leqslant \iota \leqslant 2j+2$
,

which is (4.16). Finally for (4.17), by interpolating (4.16) (with
$\iota =2j+2$
) with (4.18) using Proposition 2.3, we obtain

By choosing

we arrive at

This shows (4.17). Thus, to prove Theorem 4.2, it suffices to prove (4.26) when
$j=0$
, and to prove (4.10), (4.13), (4.15) and (4.18) when
$j\geqslant 1$
.
4.4 Setup of primary blowup quantity
To this end, we denote

which by definition satisfies
$\eta _{j,k}={i\partial \bar \partial }\psi _{j,k}$
. Of course, when
$j=0$
, we have by definition
$\psi _{0,k}=\varphi -\underline{\varphi }$
. For
$x=(z,y),x'=(z',y')\in B\times Y$
which are either horizontally or vertically joined and
$0<t'<t$
, we consider the quantities

and for
$j\geqslant 1$
,

as well as

For each
$x=(z,y)\in B\times Y$
and
$t\geqslant 0$
, we define the blowup quantity

where the sup is taken over all
$x'=(z',y')\in B\times Y$
with
$|z'-z|<\frac 14 \left| |z|-1 \right|$
and
$x'$
is either horizontally or vertically joined with x. We want to show that there is
$C>0$
such that for all
$t\geqslant 0$
,

Since
$g(t)$
is uniformly comparable to
$g(s)$
if
$|t-s|<1$
, a bound on
$\mu _j$
implies (4.26) when
$j=0$
, and implies (4.10), (4.13), (4.15) and (4.18) when
$j\geqslant 1$
, and would thus conclude the proof of Theorem 4.2.
Observe that the quantity
$\mathcal {H}_0$
is closer in spirit to the one used in our earlier works [Reference Hein and Tosatti20, (5.7)] and [Reference Chu and Lee4, (3.10)] (which dealt only with the case
$j=0$
), rather than the one used in [Reference Hein and Tosatti21, (4.29)] (which dealt with all
$j\geqslant 0$
at once).
We now setup the contradiction argument, so suppose that (4.43) fails. We can then find a sequence
$t_{\ell }>0$
such that
$\sup _{B\times Y\times [0,t_{\ell }]} \mu _j(x,t)\to +\infty $
as
$\ell \to +\infty $
. Since the solution of the flow is smooth on any compact time interval, we must have
$t_{\ell }\to +\infty $
. Moreover, there exists
$s_{\ell }\in [0,t_{\ell }]$
such that
$\sup _{B\times Y}\mu _{j}(x,s_{\ell })=\sup _{B\times Y\times [0,t_{\ell }]} \mu _j(x,t)$
. Without loss of generality, we can assume
$s_{\ell }=t_{\ell }\to +\infty $
and

For each
$\ell $
, we choose
$x_{\ell }=(z_{\ell },y_{\ell })\in B\times Y$
such that
$\mu _j(x_{\ell },t_{\ell })=\sup _{B\times Y} \mu _j(x,t_{\ell })$
. We also define
$\lambda _{\ell }$
by

so that

and hence,
$\lambda _{\ell } \to +\infty $
. Let
$x_{\ell }^{\prime }\in B\times Y$
be the point realizing
$\sup _{x'} \mathcal {D}_j(x_{\ell },x',t_{\ell })$
and
$t_{\ell }^{\prime }\in [t_{\ell }-1,t_{\ell }]$
realizing
$\sup _{t'}\mathcal {H}_j(x_{\ell },x_{\ell }^{\prime },t_{\ell },t')$
. Without loss of generality, we can also assume
$x_{\ell }\to x_\infty \in \overline {B\times Y}$
.
Consider the diffeomorphisms

Let
$\hat {x}_{\ell }:=(\hat {z}_{\ell },\hat {y}_{\ell })$
, where

so that
$\hat {t}_{\ell }=0, \hat {t}^{\prime }_{\ell }=\lambda _{\ell }^2(t_{\ell }^{\prime }-t_{\ell })$
, and
$\hat {t}=\lambda _{\ell }^2(t-t_{\ell })\in [-\lambda _{\ell }^2 t_{\ell },0]$
. Given a (time-dependent) contravariant
$2$
-tensor
$\alpha $
(such as
$\omega ^{\bullet }(t), g(t)$
, etc.), we define
$\hat {\alpha }_{\ell }:=\lambda _{\ell }^2\Psi _{\ell }^{*}\alpha $
. Thus, for example,
$\hat \omega _{\ell }^{\bullet }(\hat t)=\lambda _{\ell }^2\Psi _{\ell }^{*} \omega ^{\bullet }(t_{\ell }+\lambda _{\ell }^{-2}\hat t).$
The pullback complex structure will be denoted by
$\hat {J}_{\ell }$
. Given a (time-dependent) scalar function F, we will also denote by
$\hat {F}_{\ell }:=\Psi _{\ell }^{*}F$
, so that, for example,
$\hat {G}_{\ell ,i,p,k}=\Psi _{\ell }^{*}G_{i,p,k}$
. However, for the two functions
$A_{i,p,k}$
and
$\varphi $
, we will define instead

where
$\hat {t}\in [-t_{\ell } \lambda _{\ell }^2,0]$
. We define also

Observe that from (4.31), we have that

for all
$1\leqslant i\leqslant j$
and
$1\leqslant p\leqslant N_{i,k}$
and
$-t_{\ell } \lambda _{\ell }^2\leqslant \hat t\leqslant 0$
. For notational convenience, we will still use
$\mathfrak {D}$
and
$\mathbb {P}$
to denote their pullbacks via
$\Psi _{\ell }$
. In particular,
$\hat \omega _{\ell }^{\bullet }=\hat \omega _{\ell }^\natural +{i\partial \bar \partial } \hat \varphi _{\ell }$
satisfies the following Kähler-Ricci flow

and we can equivalently write the complex Monge-Ampère equation (4.1) as

where (following the above convention)
$\hat \omega _{\ell ,\mathrm {can}}=\lambda _{\ell }^2\Psi _{\ell }^{*} \omega _{\mathrm {can}}$
. It is then straightforward to see that for all
$\ell \geqslant 0$
, we have for
$j=0,$

and for
$j\geqslant 1,$

and
$\hat {x}^{\prime }_{\ell }$
was chosen to maximize the difference quotients in (4.54) and (4.55) (which we can call
$\hat {\mathcal {D}}_j(\hat {x}_{\ell },\hat {x}',\hat {t}_{\ell })$
) among all points
$\hat {x}'=(\hat {z}',\hat {y}')\in B_{\mathbb {C}^m}(\lambda _{\ell })\times Y$
, with
$|\hat {z}'-\hat {z}_{\ell }|<\frac 14 \left| |\hat {z}_{\ell }|-\lambda _{\ell } \right|$
which are horizontally or vertically joined to
$\hat {x}_{\ell }$
. Moreover, the points
$\hat {x}_{\ell }$
and
$\hat {t}_{\ell }$
themselves maximize the quantity

among all
$\hat {x}=(\hat z,\hat y)\in B_{\mathbb {C}^m}(\lambda _{\ell })\times Y$
and
$\hat t\in [-t_{\ell }\lambda _{\ell }^2,0]$
; hence, for all such
$\hat {x},\hat {x}',\hat {t}$
, we have

Using
$\left| |\hat z_{\ell }|-\lambda _{\ell }\right|=\lambda _{\ell } \left| |z_{\ell }|-1\right|\to +\infty $
(hence the pointed limit centered at
$\hat x_{\ell }$
will be complete) and

together with the triangle inequality, we see from (4.57) that there exists
$C>0$
such that for any fixed
$R>0$
, there exists
$\ell _R\in \mathbb {N}$
such that for all
$\ell \geqslant \ell _R$
and
$\hat {z}^{\bullet }_{\ell }=\hat {z}_{\ell }$
or
$\hat {z}^{\prime }_{\ell }$
, we have

where here and in the following, the hat decoration over
$B_{\mathbb {C}^m}$
is just to remind the reader that we are in the hat picture. This in particular implies there exists
$C>0$
so that for all fixed R,
$\hat t\in [-t_{\ell }\lambda _{\ell }^2,0]$
and all sufficiently large
$\ell $
, we have for
$j=0,$

and for
$j\geqslant 1,$

where
$\hat Q_R(\hat z,\hat t)=\left(B_{\mathbb {C}^m}(\hat z,R)\times Y\right)\times [-R^2+\hat t,\hat t]$
with
$\hat z$
being either
$\hat z_{\ell }$
or
$\hat z_{\ell }^{\prime }$
.
When
$j=0$
we will need the following ‘non-cancellation’ inequality

which is again straightforward to prove using [Reference Hein and Tosatti21, (4.215)], and plugging in the Hölder bound in (4.60) and the
$L^\infty $
bound
$\|{i\partial \bar \partial }\hat {\varphi }_{\ell }\|_{\infty , \hat Q_R(\hat z,\hat t),\hat {g}_{\ell }(\hat {t})}=o(1)$
which comes from Lemma 4.1 (iii), we see that

and combining this with (4.60), using that
$\hat {\eta }_{0,k}={i\partial \bar \partial }\hat {\psi }_{\ell ,0,k}={i\partial \bar \partial }\hat {\varphi }_{\ell }-{i\partial \bar \partial }\underline{\hat {\varphi }_{\ell }}$
, and using the triangle inequality and the boundedness of
$\mathbb {P}$
, we get

After passing to subsequence, we will split the rest of the proof into three cases, according to the behavior of
$\delta _{\ell }=\lambda _{\ell } e^{-t_{\ell }/2}$
: Case 1:
$\delta _{\ell }\to +\infty $
, Case 2:
$\delta _{\ell }\to \delta _\infty>0$
and Case 3:
$\delta _{\ell }\to 0$
.
4.5 Non-escaping property
In this subsection, we will show that in Cases 2 and 3, the distance between the two blowup points
$\hat {x}_{\ell }$
and
$\hat x_{\ell }^{\prime }$
will not go to infinity. This proof does not apply to Case 1 (at least when
$j\geqslant 1$
), but we will nevertheless establish the same result in that case in (4.97) below.
Proposition 4.4. Suppose
$\delta _{\ell }\leqslant C$
for some
$C>0$
. Then there exists
$C'>0$
such that for all
$\ell>0$
,

Proof. First, we can easily deal with the case
$j=0$
. By Lemma 4.1 (iii), we know that

which implies

as
$\ell \to \infty $
, for fixed
$\hat {z},\hat {t},\rho $
, and applying this to
$\hat {z}=\hat {z}_{\ell }$
or
$\hat {z}^{\prime }_{\ell }$
and
$\hat {t}=0$
or
$\hat {t}^{\prime }_{\ell }$
, we see that the numerator on the RHS of (4.54) is going to zero, which gives us the even stronger statement that

and for
$j=0$
, we do not even need the assumption that
$\delta _{\ell }\leqslant C$
.
Next, we assume
$j\geqslant 1$
. The argument is a modification of [Reference Hein and Tosatti21, Proposition 4.5], and the goal is to estimate each of the terms in the blowup quantity in (4.55). In the following, we denote
$\hat Q_R(\hat z,\hat t)=\left(B_{\mathbb {C}^m}(\hat z,R)\times Y\right)\times [-R^2+\hat t,\hat t]$
with
$\hat z$
being either
$\hat z_{\ell }$
or
$\hat z_{\ell }^{\prime }$
and
$\hat t$
is either
$\hat t_{\ell }=0$
or
$\hat t_{\ell }^{\prime }$
.
We first handle the terms involving
$A_{\ell ,i,p,k}$
. Recall from (4.61) that for all
$-2\leqslant \iota \leqslant 2k$
and each given
$R>0$
,

while from (4.51),

since
$\hat {t}=0$
or
$\hat {t}_{\ell }^{\prime }$
satisfies
$\hat {t}\geqslant -\lambda _{\ell }^2$
. By Proposition 2.3, for all
$1\leqslant r\leqslant 2j$
,

We choose

which is small for any given R. We get for all
$0\leqslant r\leqslant 2j$
that

where
$\hat t$
is either
$\hat t_{\ell }=0$
or
$\hat t_{\ell }^{\prime }$
.
If we let
$r=\iota +2>0$
, then for
$0<r\leqslant 2k+2$
, we can interpolate again, using (4.69), and get

By taking

which is small thanks to our assumption that
$\delta _{\ell }\leqslant C$
, we conclude that for all
$0<r\leqslant 2k+2$
,

Applying (4.73) and (4.76) to balls centered at
$(\hat z_{\ell },\hat t_{\ell })$
and
$(\hat z_{\ell }^{\prime },\hat t_{\ell }^{\prime })$
(of any radius – for example,
$1$
), together with the boundedness of operator norm of
$\mathbb {P}$
from [Reference Hein and Tosatti21, §2.1.1], this gives

We now treat
$\hat \psi _{\ell ,j,k}$
which has fiberwise average zero. By Proposition 2.6 (in case
$\delta _{\ell }$
does not converge to
$0$
, we choose
$\rho $
to be sufficiently large) and (4.61), we have

Applying this to balls centered at
$\hat z_{\ell }$
and
$\hat z_{\ell }^{\prime }$
and invoking [Reference Hein and Tosatti21, §2.1.1] again, we have

It remains to consider the fiberwise average of the potential (i.e.,
$\underline{\hat \varphi _{\ell }}$
). Recalling (4.66) and taking fiber average (using the fiber integration argument in [Reference Tosatti38, p.436]) gives in particular

as
$t\to +\infty $
, which implies

as
$\ell \to \infty $
, for fixed
$\hat {z},\hat {t},\rho $
, and interpolating between this and (4.61) gives

for
$\hat {z}=\hat {z}_{\ell }$
or
$\hat {z}^{\prime }_{\ell }$
and
$\hat {t}=0$
or
$\hat {t}^{\prime }_{\ell }$
. Using again the boundedness of the operator norm of
$\mathbb {P}$
, this implies that

Similarly, from Lemma 4.1 (ii) and taking fiber average, we know that

as
$t\to +\infty $
, which implies

as
$\ell \to \infty $
, for fixed
$\hat {z},\hat {t},\rho $
, and again interpolating between this and (4.61) gives

for
$\hat {z}=\hat {z}_{\ell }$
or
$\hat {z}^{\prime }_{\ell }$
and
$\hat {t}=0$
or
$\hat {t}^{\prime }_{\ell }$
, and hence,

Combining (4.55) with (4.77), (4.79), (4.83) and (4.87), we obtain the desired bound (4.65).
We are now in position to study the flows obtained as complete pointed limits of

as
$\ell \to +\infty $
, where as usual
$\hat {t}\in [-\lambda _{\ell }^2t_{\ell },0]$
. By translation, we can assume
$\hat x_{\ell }=(\hat z_{\ell },\hat y_{\ell })=(0,\hat y_{\ell })\in \mathbb {C}^m\times Y$
and
$\hat y_{\ell }\to \hat y_\infty \in Y$
by compactness of Y after passing to a subsequence.
4.6 Blowup analysis in Case 1:
$\delta _{\ell }\to +\infty $
In this case, the metrics
$\hat {g_{\ell }}(0)$
are blowing up in the fiber directions, so that their pointed blowup limit (modulo local diffeomorphisms that stretch the fibers) would be
$\mathbb {C}^{m+n}$
. While this is the approach taken in our earlier works [Reference Chu and Lee4, Reference Fong and Lee11, Reference Hein and Tosatti20], it turns out that we need a different approach instead (at least when
$j\geqslant 1$
). So, following [Reference Hein and Tosatti21], we consider the diffeomorphisms

pull back time-dependent
$2$
-tensors via
$\Xi _{\ell }$
, rescale them by
$\delta _{\ell }^{-2}$
and denote the new tensors with a check, so, for example,
$\check \omega ^{\bullet }_{\ell }(\check t)=\delta _{\ell }^{-2} \Xi _{\ell }^{*}\hat \omega _{\ell }^{\bullet }(\delta _{\ell }^2 \check t)$
. We also apply the same pullback and rescaling procedure to the scalar functions
$\hat {A}_{\ell ,i,p,k}$
and
$\hat {\varphi }_{\ell }$
.
In this case,
$\check g_{\ell }(\check {t})$
is locally uniformly Euclidean in space-time and
$\check \omega _{\ell }^{\bullet } $
satisfies the Kähler-Ricci flow equation

and the Monge-Ampère equation (4.53) becomes

Thanks to Lemma 4.1 (i), we know that
$\check \omega _{\ell }^{\bullet } (\check t)$
is uniformly equivalent to

which in turn is locally uniformly equivalent to the Euclidean metric. The pullback of the complex structure also converges locally uniformly smoothly to the Euclidean product complex structure due to the stretching. We want to apply the local higher order estimates in [Reference Chu and Lee4, Proposition 2.1] on
$\check Q_1(\check z_{\ell },0)$
and
$\check Q_1(\check z_{\ell }^{\prime },\check t_{\ell }^{\prime })$
, but we do not know whether
$\check {B}_1(\check {z}_{\ell }),\check {B}_1(\check {z}^{\prime }_{\ell })$
are contained in
$B_{e^{\frac {t_{\ell }}{2}}}$
, as we do not have any relation between
$\delta _{\ell }$
and
$\|\hat {z}_{\ell }|-\lambda _{\ell }|$
. However, these are compactly contained in the slightly larger ball
$B_{(1+\sigma )e^{\frac {t_{\ell }}{2}}}$
for any fixed
$\sigma>0$
and all
$\ell $
sufficiently large, and we may assume without loss that
$\check \omega _{\ell }^{\bullet } (\check t)$
is uniformly equivalent to Euclidean on
$B_{(1+\sigma )e^{\frac {t_{\ell }}{2}}}\times Y\times [-(1+\sigma )^2e^{t_{\ell }}t_{\ell },0]$
. Thus, the local higher order estimates give us uniform
$C^\infty $
estimates for
$\check \omega _{\ell }^{\bullet }$
on
$\check Q_1(\check z_{\ell },0)$
and
$\check Q_1(\check z_{\ell }^{\prime },\check t_{\ell }^{\prime })$
. Thus, on these sets, we have uniform
$C^\infty $
bounds for
${i\partial \bar \partial }\check {\varphi }$
, hence on
${i\partial \bar \partial }\underline{\check {\varphi }}$
(by fiber averaging), hence on
$\check {A}_{\ell ,1,p,k}$
(from its definition), hence on
$\check {\gamma }_{\ell , 1,k}$
(also from its definition), hence on
$\check {\eta }_{\ell ,1,k}$
(from its definition), and so forth until
$\check {A}_{\ell ,j,p,k},\check {\gamma }_{\ell ,j,k},\check {\eta }_{\ell ,j,k}$
. From the PDE (4.91), we also get uniform
$C^\infty $
bounds for
$\partial _{\check t} \check \varphi _{\ell }+e^{-t_{\ell }} \check \varphi _{\ell }$
, and so by fiber averaging also on
$\partial _{\check t} \underline{\check \varphi _{\ell }}+e^{-t_{\ell }} \underline{\check \varphi _{\ell }}$
. Also, since
$\check {\eta }_{\ell ,j,k}={i\partial \bar \partial }\psi _{\ell ,j,k}$
is locally smoothly bounded, and
$\psi _{\ell ,j,k}$
has fiberwise average zero, then fiberwise Moser iteration gives us a uniform
$L^\infty $
bound for
$\psi _{\ell ,j,k}$
, and elliptic estimates show that
$\psi _{\ell ,j,k}$
has uniform
$C^\infty $
bounds. Putting these all together, we get in particular


where
$\check Q_{1}$
is either
$\check Q_{1}(\check z_{\ell },0)$
or
$\check Q_{1}(\check z_{\ell }^{\prime },\check t_{\ell }^{\prime })$
. Transferring these back to the hat picture gives


where
$\hat Q_{\delta _{\ell }}$
is either
$\hat Q_{\delta _{\ell }}(\hat z_{\ell },0)$
or
$\hat Q_{\delta _{\ell }}(\hat z_{\ell }^{\prime },\hat t_{\ell }^{\prime })$
. Using (4.95) and the triangle inequality (and the usual bound on the operator norm of
$\mathbb {P}$
), we obtain a uniform upper bound for the numerators of (4.54) and (4.55), and so for
$j\geqslant 1$
, we conclude that

so that the two points
$(\hat z_{\ell },\hat t_{\ell })$
and
$(\hat z_{\ell }^{\prime },\hat t_{\ell }^{\prime })$
are approaching each other (we already know this when
$j=0$
thanks to (4.68)). Thus,
$(\hat x_{\ell }^{\prime },\hat t_{\ell }^{\prime })\in \hat Q_{\delta _{\ell }}(\hat z_{\ell },0)$
for all
$\ell $
large, and so applying (4.96) shows that the quantities in (4.54) and (4.55), which both equal
$1$
, are also bounded above by
$C\delta _{\ell }^{-2j-{\alpha }}\to 0$
, which is a contradiction.
4.7 Blowup analysis in Case 2:
$\delta _{\ell }\to \delta _\infty \in (0,+\infty )$
Without loss of generality, we can assume
$\delta _\infty =1$
. In this case, the blowup model is
$\mathbb {C}^m\times Y$
and

as
$\ell \to +\infty $
in
$C^\infty _{\mathrm {loc}}(\mathbb {C}^m\times Y\times (-\infty ,0])$
. Moreover, the complex structure also converges to a product.
As in Case 1, Lemma 4.1 (i) implies that
$\hat \omega _{\ell }^{\bullet }(\hat t)$
is locally uniformly equivalent to product metric on
$\mathbb {C}^m\times Y$
. Moreover, since
$\hat \omega _{\ell }^{\bullet }$
satisfies the Kähler-Ricci flow equation

we can again apply [Reference Chu and Lee4, Proposition 2.1] on
$\hat Q_R(\hat z_{\ell },0)$
for R sufficiently large to obtain
$C^\infty _{\mathrm {loc}}$
regularity of
$\hat \omega _{\ell }^{\bullet }$
. Arguing as in Case 1, we obtain
$C^\infty $
bounds for all the pieces in the decomposition, and using Proposition 4.4, we can assume
$(\hat x_{\ell }^{\prime },\hat t_{\ell }^{\prime })\to (\hat x_\infty ',\hat t_\infty ')\in (\mathbb {C}^m\times Y)\times (-\infty ,0]$
. Estimating the
$C^{\alpha }$
difference quotients in (4.55) by
$C^\beta $
ones for any
$\beta>\alpha $
, we see that
$d^{\hat g_{\ell }(0)}(\hat x_{\ell },\hat x_{\ell }^{\prime })+|\hat t_{\ell }^{\prime }|^{\frac {1}{2}}\geqslant C^{-1}$
for all
$\ell $
, which when
$j=0$
is already a contradiction to (4.68).
Assuming then that
$j\geqslant 1$
, we see that the limit
$(\hat x_\infty ,0)$
of
$(\hat x_{\ell },0)$
is different from
$(\hat x_\infty ',\hat t_\infty ')$
. By the local uniform higher order regularity, the geometric quantities smoothly subconverge as
$\ell \to +\infty $
. In particular, the limit
$\hat {\omega }^{\bullet }_\infty (\hat {t})$
is an ancient solution on
$\mathbb {C}^m\times Y \times (-\infty ,0]$
of the Kähler-Ricci flow

Smooth convergence also implies that (4.55) still holds in the limit.
By (4.51), we have that
$\hat A_{\ell ,i,p,k}\to 0$
locally uniformly, hence locally smoothly, so its contributions to (4.55) vanish in the limit as
$\ell \to +\infty $
. However, Lemma 4.1 (ii),(iii) also implies that the limits of
${i\partial \bar \partial }\underline{\hat \varphi _{\ell }}$
and
$\partial _{\hat t}{\hat \varphi _{\ell }}+\lambda _{\ell }^{-2}{\hat \varphi _{\ell }}$
(and hence also of
$\partial _{\hat t}\underline{\hat \varphi _{\ell }}+\lambda _{\ell }^{-2}\underline{\hat \varphi _{\ell }}$
) vanish so that their contributions in (4.55) also vanish in the limit.
We are left with killing the limit of the contribution of
$\hat \psi _{\ell ,j,k}$
. For this, recall that from Lemma 4.1 (iv) we have
$|R(g^{\bullet }(t))|\leqslant C$
, and so
$|R(\hat {g}^{\bullet }(\hat {t}))|\leqslant C\lambda _{\ell }^{-2} \to 0$
, thus the limiting metrics
$\hat {\omega }^{\bullet }_\infty (\hat {t})$
are scalar-flat, and hence Ricci-flat and static (using the well-known evolution equation of the scalar curvature under the Kähler-Ricci flow (4.100)). Also, since
$\hat A_{\ell ,i,p,k}\to 0$
locally smoothly, this implies that
$\hat {\gamma }_{\ell ,i,k}\to 0, 1\leqslant i\leqslant j,$
and so
$\hat \omega _\infty ^{\bullet }=\omega _P+\hat \eta _{\infty ,j,k}$
, and in particular,
$\hat \eta _{\infty ,j,k}$
is also static. The Liouville Theorem from [Reference Hein18] shows that
$\nabla ^{g_P}\hat \eta _{\infty ,j,k}=\nabla ^{g_P}\hat \omega _\infty ^{\bullet }\equiv 0$
. Thus
$\hat {\eta }_{\infty ,j,k}={i\partial \bar \partial }\hat {\psi }_{\infty ,j,k}$
is parallel, with bounded
$g_P$
norm (from Lemma 4.1 (i)), so by [Reference Hein and Tosatti20, Proposition 3.12], we have
${i\partial \bar \partial }\hat {\psi }_{\infty ,j,k}={i\partial \bar \partial } p$
for some quadratic polynomial p on
$\mathbb {C}^m$
. This means that
$\hat {\psi }_{\infty ,j,k}-p$
is pluriharmonic on
$\mathbb {C}^m\times Y$
, and hence it is also pulled back from
$\mathbb {C}^m$
(since Y is compact). This clearly implies that
$\hat {\psi }_{\infty ,j,k}$
is pulled back from
$\mathbb {C}^m$
, and since it also has fiberwise average zero, it must vanish identically. Thus, the contribution of
$\hat \psi _{\ell ,j,k}$
to (4.55) also vanishes in the limit, and this gives a contradiction to (4.55).
4.8 Blowup analysis in Case 3:
$\delta _{\ell }\to 0$
In this case, the blowup model is
$\mathbb {C}^m$
which is still collapsed. This is the most difficult case and will occupy most of the rest of the paper.
By Proposition 4.4, we know that
$(\hat x_{\ell }^{\prime },\hat t_{\ell }^{\prime })$
remains at bounded distance from
$(\hat x_{\ell },0)$
. The key claim is the following non-colliding estimate, whose proof will take substantial work:
Claim 4.5. There exists
$\varepsilon _0>0$
such that for all
$\ell \geqslant 1$
, we have

First, we show how to quickly complete the proof of Theorem 4.2 assuming Claim 4.5 holds. When
$j=0$
, it is clear that (4.101) is incompatible with (4.68), while if
$j\geqslant 1$
then (4.55) implies

while the right-hand side is of
$o(1)$
thanks to (4.77), (4.79), (4.83) and (4.87) as
$\ell \to +\infty $
, since
$\delta _{\ell }\to 0$
. This is a contradiction. Therefore, to complete the proof of Theorem 4.2, it remains to prove Claim 4.5.
4.9 Setup of secondary blowup in proving Claim 4.5
For each
$\ell \geqslant 1$
, let

If Claim 4.5 fails to be true, then we may assume that
$d_{\ell }\to 0$
as
$\ell \to +\infty $
. Define a new parameter

and consider the diffeomorphism

As usual, we pull back time-dependent
$2$
-tensors via
$\Theta _{\ell }$
, rescale them by
$d_{\ell }^{-2}$
and denote the new tensors with a tilde, so, for example,
$\tilde {\omega }^{\bullet }_{\ell }(\tilde {t})=d_{\ell }^{-2} \Theta _{\ell }^{*}\hat \omega _{\ell }^{\bullet }(d_{\ell }^2 \tilde {t})$
. We also apply the same pullback and rescaling procedure to the scalar functions
$\hat {A}_{\ell ,i,p,k}$
and
$\hat {\varphi }_{\ell }$
, so
$\tilde A_{\ell ,i,p,k}(\tilde t)=d_{\ell }^{-2}\Theta _{\ell }^{*} \hat A_{\ell ,i,p,k}(d_{\ell }^2\tilde t), \tilde {\varphi }_{\ell }(\tilde {t})=d_{\ell }^{-2}\Theta _{\ell }^{*}\hat {\varphi }_{\ell }(d_{\ell }^2\tilde {t})$
.
The decomposition (4.22) becomes

and the parabolic complex Monge-Ampère equation (4.53) becomes

From (4.64) (when
$j=0$
) and (4.61) (when
$j\geqslant 1$
), we immediately see that for any fixed
$R>0$
and
$\tilde t\in [-\lambda _{\ell }^2d_{\ell }^{-2}t_{\ell },0]$
, we have

if
$j=0$
, and

when
$j\geqslant 1$
, where
$\tilde Q_R(\tilde t)=B_{\mathbb {C}^m}(\tilde z_{\ell },R)\times Y\times [-R^2+\tilde t,\tilde t]$
. Moreover, (4.54) and (4.55) become, respectively,

for
$j=0$
and

for
$j\geqslant 1$
, and by definition, we also have

for all
$\ell>0$
.
For convenience, from now on,
$\tilde Q_r$
will always denote
$B_{\mathbb {C}^m}(0,r)\times Y\times [-r^2,0]$
, where recall that we have translated the first blowup point
$(\tilde {x}_{\ell },\tilde {t}_{\ell })$
in the
$\mathbb {C}^m$
directions so that
$(\tilde z_{\ell },\tilde {t}_{\ell })=(0,0)\in \mathbb {C}^m\times \mathbb {R}$
. As long as
$r>1$
, the cylinder
$\tilde {Q}_r$
always contains the other blowup point
$(\tilde {x}^{\prime }_{\ell },\tilde {t}^{\prime }_{\ell })$
because of (4.112). Following [Reference Hein and Tosatti21, §4.8], our goal is to obtain a contradiction by passing to the limit as
$\ell \to +\infty $
the various pieces of the decomposition, after scaling them by
$d_{\ell }^{-2j-{\alpha }}$
. To do this, we need to perform a jet subtraction centered at
$(0,0)\in \mathbb {C}^m\times (-\infty ,0]$
for functions pulled back from the base. Given
$r\in \mathbb {N}$
, the parabolic r-jet at
$(0,0)$
of a function u in
$\mathbb {C}^m\times (-\infty ,0]$
is given by

using standard multiindex notation, where here we treat
$\tilde {z}$
as real variables. The parabolic degree of such a polynomial is defined by letting the
$\tilde {z}$
variables have degree
$1$
and the
${\tilde {t}}$
variable degree
$2$
. Thus, the degree of
$u^\sharp $
is at most r. We will also write
$u^{*}:=u-u^\sharp $
, whose r-jet at
$(0,0)$
thus vanishes.
With this notation, we define
$\tilde A_{\ell ,i,p,k}^\sharp $
as the parabolic
$2j$
-jet of
$\tilde A_{\ell ,i,p,k}$
at
$(0,0)$
, and define
$\tilde A_{\ell ,i,p,k}^{*}=\tilde A_{\ell ,i,p,k}-\tilde A_{\ell ,i,p,k}^\sharp $
so that
${\mathbb {D}}^{p} \partial _{\tilde {t}}^q \tilde A_{\ell ,i,p,k}^{*}|_{(0,0)}=0$
for any
$p+2q\leqslant 2j$
. As for the potential
$\underline{\tilde {\varphi }_{\ell }}$
, since the PDE (4.107) and the blowup quantity in (4.109) (for
$j\geqslant 1$
) both contain the term
$\partial _{\tilde {t}}{\tilde {\varphi }}_{\ell }+d_{\ell }^2\lambda _{\ell }^{-2}\tilde {\varphi }_{\ell }$
, it will be more convenient to consider instead

so that

We claim that the fiber average
$\underline{\tilde {\chi }_{\ell }}$
satisfies estimates similar to those satisfied by
$ \underline{\tilde {\varphi }_{\ell }}$
in (4.108) and (4.109):
Claim 4.6. For any fixed
$R>0$
, we have


as
$\ell \to +\infty $
.
Proof of Claim 4.6
We first treat
$\partial _{\tilde t}\underline{\tilde {\chi }_{\ell }}$
, and since this will only be used when
$j\geqslant 1$
, we only prove it here in this case (but see (4.137) and (4.147) below for a stronger statement when
$j=0$
). Observe first that for any
$r\in \mathbb {N}$
, we have

Using this, we bound

Thanks to (4.109), the third line from the bottom is
$o(d_{\ell }^{2j+{\alpha }})$
. As for the last two lines, we interpolate between (4.109) and the estimate
$\| \partial _{\tilde t}\underline{\tilde {\varphi }_{\ell }}+d_{\ell }^2\lambda _{\ell }^{-2}\underline{\tilde {\varphi }_{\ell }} \|_{\infty ,\tilde Q_{Rd_{\ell }^{-1}}}=o(1)$
which comes from Lemma 4.1 (ii), and we get for each
$1\leqslant \iota \leqslant 2j$
,

By choosing
$R-\rho \approx 1$
(which is clearly allowed) and replacing R by a slightly smaller one, we see that for all
$1\leqslant \iota \leqslant 2j$
,

We can then estimate


Putting these together proves (4.116).
Next, we treat
${i\partial \bar \partial }\underline{ \tilde {\chi }_{\ell }}$
in a similar fashion (allowing now also
$j=0$
),

and then we continue the argument exactly as above, using the bound (4.108) when
$j=0$
. This proves (4.117).
With this modification in mind, we let
$\underline{ \tilde \chi _{\ell }^\sharp }$
be the (parabolic) jet of
$\underline{ \tilde \chi _{\ell }}=e^{d_{\ell }^2\lambda _{\ell }^{-2}\tilde t}\underline{\tilde {\varphi }_{\ell }}$
at
$(\tilde z,\tilde t)=(0,0)$
of order
$2j+2$
, and define
$\underline{\tilde {\chi }_{\ell }^{*}}:=\underline{ \tilde {\chi }_{\ell }}-\underline{ \tilde {\chi }_{\ell }^\sharp }$
. Define also

so that by definition, we have

Recall that for
$1\leqslant i\leqslant j$
, we had defined

(see also (4.347) below for an explicit formula). We shall further denote

so that

and we will write
$\tilde {\omega }_{\ell }^\sharp :=\tilde \omega _{\ell }^\natural +\tilde {\eta }_{\ell }^\dagger +\tilde {\eta }_{\ell }^\ddagger $
.
Thanks to these jet subtractions, and to Claim 4.6, from (4.108), (4.109), (4.111) and (4.112), we obtain

when
$j=0$
, and

when
$j\geqslant 1$
, as well as

while on the other hand (4.110), remains the same for
$j=0$
.
4.10 Estimates on each component
Our next task is then to obtain precise estimates on all the pieces of the decomposition (4.129), which will allow us to later expand and linearize the Monge-Ampère equation (4.107). In the following, the radii R and S will be any fixed radii, unless otherwise specified. Some of the estimates are analogous to those in [Reference Hein and Tosatti21, §4.9], replacing balls
$\tilde {B}_r$
by parabolic cylinders
$\tilde {Q}_r=B_{\mathbb {C}^m}(0,r)\times Y\times [-r^2,0]$
. We follow closely the arguments there.
4.10.1 Estimates for
$\tilde {\psi }_{\ell ,j,k}$
First, we assume that
$j=0$
. Then from the Hölder seminorm bound for
${i\partial \bar \partial }\hat {\psi }_{\ell ,0,k}$
in (4.64), together with Proposition 2.6, we obtain

and since
$\hat {\psi }_{\ell ,0,k}$
has fiberwise average zero, we can apply fiberwise Moser iteration to this, and get

Using the bounds
$\|\partial _{t}^2\varphi +\partial _{t}\varphi \|_{\infty , B\times Y\times [t-1,t]}\leqslant C$
and
$\|{\mathbb {D}}(\partial _{t}\varphi +\varphi )\|_{\infty , B\times Y\times [t-1,t],g(t)}\leqslant C$
from Lemma 4.1 (v), (vi), we can bound for any
$x,x'\in B_{R\lambda _{\ell }^{-1}}\times Y$
and
$t,s\in (t_{\ell }-\lambda _{\ell }^{-2}R^2,t_{\ell }]$
,

which gives

Repeating the argument with
$\underline{\varphi }$
gives

and combining these, we see that

so from this, the bound
$[{i\partial \bar \partial }\hat {\psi }_{\ell ,0,k}]_{{\alpha },{\alpha }/2,\hat {Q}_R,\hat {g}_{\ell }(0)}\leqslant C$
from (4.64), and the bounds (4.133) and (4.209) below (on
$\hat {\omega }^\sharp _{\ell }$
), we see that

We wish to use the Schauder estimates in Proposition 2.8, and for this, we need to pass to the check picture via the diffeomorphism
$\Xi _{\ell }$
in (3.4), pulling back all geometric quantities and scaling
$2$
-forms (as well as
$\hat {\psi }_{\ell ,0,k}$
) by
$\delta _{\ell }^{-2}$
. We can then apply Proposition 2.8 to
$\check \psi _{\ell ,0,k}$
and then transfer the result back to the hat picture. This shows that for every
$\rho <R$
(where R is fixed), we have

and employing the interpolation inequality in (2.11), and (4.134), (4.139), we can bound the RHS of (4.140) by

and after combining (4.140) and (4.141), we can apply the iteration lemma in [Reference Hein and Tosatti21, Lemma 2.9] and deduce that for every
$\rho <R$
, we have

so in particular, for any fixed R, we deduce that

and finally applying Proposition 2.6 and translating to the tilde picture, we get

for any
$0\leqslant \iota \leqslant 2$
.
Next, when
$j\geqslant 1$
, from the Hölder seminorm bound for
$\mathfrak {D}^{2j+2}\hat {\psi }_{\ell , j,k}$
in (4.61), together with Proposition 2.6, we obtain bounds for the lower-order derivatives of
$\hat {\psi }_{\ell ,j,k}$
, which in the tilde picture translate to

for any
$0\leqslant \iota \leqslant 2j+2$
. Observe that these are formally the same as (4.144) when
$j=0$
.
4.10.2 Estimates for
$\underline{\tilde {\chi }_{\ell }}$
By Lemma 4.1 (ii),(iii),

When
$j=0$
, we have

where we used
$d_{\ell }^{-{\alpha }}[\partial _{\tilde t} \underline{\tilde {\varphi }_{\ell }}+d_{\ell }^2\lambda _{\ell }^{-2} \underline{\tilde {\varphi }_{\ell }}]_{{\alpha },{\alpha }/2, \tilde Q_{Rd_{\ell }^{-1}},\tilde g_{\ell }(0)}=o(1)$
, which follows from (4.137). Similarly,

where we used
$d_{\ell }^{-{\alpha }}[{i\partial \bar \partial }\underline{\tilde {\varphi }_{\ell }}]_{{\alpha },{\alpha }/2, \tilde Q_{Rd_{\ell }^{-1}},\tilde g_{\ell }(0)}\leqslant C$
, which follows from (4.108). Thus, when
$j=0$
, we have

When
$j\geqslant 1$
, from (4.109), (4.116) and (4.117), we obtain the analogous seminorm bounds

and we can interpolate between these and (4.146) to conclude that

We now treat
$\tilde {\chi }_{\ell }^{*}$
. From (4.149), (4.150) and the definition of
$\tilde {\chi }_{\ell }^{*}$
we see that for all
$j\geqslant 0$
,

Using these and Lemma 2.4, we get

for all
$0\leqslant \iota \leqslant 2j$
and
$0<S\leqslant Rd_{\ell }^{-1}$
.
We will also need a bound for the
$L^\infty $
norm of derivatives of
$d_{\ell }^{-2j-{\alpha }}\underline{\tilde {\chi }_{\ell }^{*}}$
of order up to
$2j+2$
, which in general may blow up, but which will nevertheless be useful later. To start, from Lemma 4.1 (ii), we have
$\|\underline{\hat {\chi }_{\ell }}\|_{\infty , \hat {Q}_R}=o(1)\lambda _{\ell }^2,$
while from (4.152), we see in particular that

Standard Euclidean Schauder estimates then imply that

and since we may assume without loss that
$o(1)\lambda _{\ell }^2\geqslant C$
, passing to the tilde picture, we get

and we can then apply Lemma 2.4 to
$\tilde {Q}_R$
and get

for
$0\leqslant \iota \leqslant 2j+2$
.
Since
$\underline{\tilde {\chi }^\sharp _{\ell }}$
is a jet of
$\underline{\tilde {\chi }_{\ell }}$
, it inherits from (4.146) and (4.151) the bounds

for all
$0\leqslant \iota \leqslant 2j$
and R fixed. Moreover, since
$\underline{\tilde {\chi }_{\ell }^\sharp }$
is a polynomial of degree at most
$2j+2$
, it vanishes when differentiated more than
$2j+2$
times.
Furthermore, recalling the definitions of
$\tilde {\eta }^\ddagger _{\ell }$
and
$\tilde {\eta }^\diamondsuit _{\ell }$
in (4.125), from (4.153) and (4.158), we quickly deduce

for
$0\leqslant \iota \leqslant 2j, 0<S\leqslant Rd_{\ell }^{-1}$
, and

also for
$0\leqslant \iota \leqslant 2j$
, R fixed (and derivatives of
$\tilde {\eta }^\ddagger _{\ell }$
of order higher than
$2j$
vanish, of course).
4.10.3 Estimates for
$\tilde A_{\ell ,i,p,k}^{*}$
From (4.131), we have

and hence, Lemma 2.4 implies

for all
$0\leqslant \iota \leqslant 2j$
and fixed
$R>0$
. Taking
$R=S d_{\ell }$
and transferring to the tilde picture, we have

for all
$0\leqslant \iota \leqslant 2j$
and
$S\leqslant R d_{\ell }^{-1}$
.
For higher order derivatives of
$\tilde A_{\ell ,i,p,k}^{*}$
, we use interpolation. Since
$\tilde A^\sharp _{\ell ,i,p,k}$
is a parabolic polynomial of degree at most
$2j$
, (4.131) implies

for all
$0\leqslant \iota \leqslant 2k+2$
. For any
$1\leqslant \iota \leqslant 2k+2$
, Proposition 2.3 in the hat picture gives for
$0<\rho _1<\rho _2\leqslant R$
,

We can follow closely the choice of
$\rho _i$
in the interpolation in [Reference Hein and Tosatti21, §4.9.3] (with j replaced by
$2j$
here) to conclude that

This will play an important role in case
$\varepsilon _{\ell }\geqslant C^{-1}$
. For later purposes, we will need to estimate the dependence of
$C_S$
on S as
$S\to +\infty $
. In fact,
$C_S$
in [Reference Hein and Tosatti21, (4.127)] is given by
$C(AS)^{\frac {{\alpha }^2}{\iota +{\alpha }}}$
where
$A>1$
is given as a function of
$S>1$
by solving the equation
$(A-1)^{-1}A^{\frac {{\alpha }}{\iota +{\alpha }}}= S^{\frac {\iota }{\iota +{\alpha }}}$
. Since
$\frac {{\alpha }}{\iota +{\alpha }}<1$
for
$\iota \geqslant 1$
, A stays bounded as
$S\to +\infty $
. Hence, we can estimate
$C_S$
from above by

for
$S>1$
. Therefore, in the tilde picture we obtain

where
$1\leqslant \iota \leqslant 2k+2$
and
$S>1$
fixed (which is the analog of [Reference Hein and Tosatti21, (4.129)]).
Similarly, in the case when
$\varepsilon _{\ell }\to 0$
, following the derivation of [Reference Hein and Tosatti21, (4.132)], we obtain

which in the tilde picture becomes

for
$1\leqslant \iota \leqslant 2k+2$
and
$S>1$
fixed.
4.10.4 Estimates for
$\tilde A_{\ell ,i,p,k}^\sharp $
By (4.73), we have for all
$0\leqslant \iota \leqslant 2j$
that

for all given
$R>0$
(with C independent of
$R>0$
). Since
$\hat A_{\ell ,i,p,k}^\sharp $
is the
$2j$
-jet of
$\hat A_{\ell ,i,p,k}$
at
$(0,0)$
, we see that all the coefficients of the polynomial
$\hat A_{\ell ,i,p,k}^\sharp $
are bounded by
$C\delta _{\ell }^2 e^{\frac {-2i+2-{\alpha }}{2}\cdot \frac {{\alpha } }{2j+{\alpha }}t_{\ell }}$
, and so

for all
$S>0$
,
$0\leqslant \iota \leqslant 2j$
and
$0<{\beta }<1$
. Transferring to the tilde picture yields

for all
$0<S\leqslant Rd_{\ell }^{-1}$
,
$0\leqslant \iota \leqslant 2j$
and
$0<{\beta }<1$
, c.f. [Reference Hein and Tosatti21, (4.135)–(4.136)].
4.10.5 Estimates for
$\tilde {\eta }_{\ell }^\circ $
and its potential
First, we recall that

Here,
$\hat {\mathfrak {G}}_{\hat {t},k}$
and
$\hat A_{\ell ,i,p,k}$
are
$\hat t$
-dependent while
$\hat G_{\ell ,i,p,k}$
is independent of time.
By applying (3.1) for each fixed
$\hat t$
, we can write

where
$\hat \Phi _{\iota ,q}(\hat G_{\ell ,i,p,k})$
is independent of time. Note also that
$\hat J_{\ell }$
is independent of time and hence [Reference Hein and Tosatti21, (4.141), (4.144)] can be directly carried over, so that

for each
$\iota \geqslant 0,\lceil \frac {\iota }{2}\rceil \leqslant q\leqslant 2k$
and fixed
$R>0$
.
Schematically, we have

To estimate this, we will need (4.162), (4.164) and (4.168):

for all
$0\leqslant \iota \leqslant 2k+2+2j$
and
$S>1$
fixed. Now using (4.176) and (4.178) in (4.177), and recalling that
$\delta _{\ell }=\lambda _{\ell } e^{-\frac {t_{\ell }}{2}},$
yields

for all
$0\leqslant r\leqslant 2j+2$
. Similarly,

for all
$0\leqslant r\leqslant 2j+2$
. Transferring to the tilde picture gives

for all
$0\leqslant r\leqslant 2j+2$
and
$S>1$
fixed.
Next, we consider the case when
$\varepsilon _{\ell }\leqslant C$
, and we take only derivatives and difference quotients in the base and time directions. The argument is similar to [Reference Hein and Tosatti21, (4.159)]. We start by noting that if a certain derivative
$\mathfrak {D}^\iota $
contains precisely u fiber derivatives (
$0\leqslant u\leqslant \iota $
), then we will denote it schematically by
. We then have the easy bounds from [Reference Hein and Tosatti21, (4.153)–(4.154)]

We use (4.162), (4.164) and (4.169) to obtain

for all
$0\leqslant \iota \leqslant 2j$
,
$1\leqslant \iota '\leqslant 2k+2$
and
$S>1$
fixed. Then (4.177), (4.182) and (4.183) imply

for
$0\leqslant r\leqslant 2j+2$
, and fixed
$S>1$
. Arguing similarly for the Hölder seminorms, and transferring to the tilde picture yields

Observe that when
$r=2j+2$
, the leading term arises when
$d=i_1=q=\iota =0$
and
$i_2=2j+2$
, which in the tilde picture is given by

We now bound
$\tilde \eta _{\ell }^\circ $
and its derivatives. Using [Reference Hein and Tosatti21, (4.139)], paying extra attention to the time derivatives, we have

where

Using (4.176) with (4.178), we can estimate
$\mathfrak {D}^r \hat \eta _{\ell }^\circ $
by

where we have used the fact that
$\delta _{\ell }= \lambda _{\ell } e^{-t_{\ell }/2}$
. Similarly,

In particular in the tilde picture,

for all
$0\leqslant r\leqslant 2j$
(which is the analog of [Reference Hein and Tosatti21, (4.151)]).
These estimates are only useful when
$\varepsilon _{\ell }\geqslant C^{-1}$
. In the case when
$\varepsilon _{\ell }\to 0$
, we shall only take derivatives and difference quotients in the base and time directions. We can follow the argument to derive [Reference Hein and Tosatti21, (4.159)], using (4.162), (4.164),(4.166) and (4.167) instead of [Reference Hein and Tosatti21, (4.123), (4.125), (4.130)], and using
$\mathfrak {D}$
instead of
${\mathbb {D}}$
, and taking also time derivatives of
$e^{-q(t_{\ell }+\lambda _{\ell }^{-2}\hat t)}$
in (4.187), we obtain

for all
$0\leqslant r\leqslant 2j$
and S fixed. Also, it is important to note that in the
$L^\infty $
bound in (4.192) with
$r=2j$
, which nominally is
$O(d_{\ell }^{\alpha })$
, the only term which is not actually
$o(d_{\ell }^{\alpha })$
comes from the terms in the sum in (4.187) with
$d=0$
. To see this, we follow verbatim the discussion in [Reference Hein and Tosatti21, (4.161)], which gives us that

locally uniformly.
Similarly, following the derivation of [Reference Hein and Tosatti21, (4.162)], in the case when
$\varepsilon _{\ell }\to 0$
, we have

for all
$0\leqslant r\leqslant 2j$
, fixed
$S>0$
and a fixed metric
$g_X$
.
4.10.6 Estimates for
$\tilde {\eta }_{\ell }^\dagger $
and its potential
Recall that by definition we have

where we have applied (3.1). From (4.172), we have

for all
$S>0$
,
$0\leqslant \iota \leqslant 2j$
and
$0<{\beta }<1$
, while the derivatives of (parabolic) order
$>2j$
vanish since
$\hat A_{\ell ,i,p,k}^\sharp $
is a (parabolic) polynomial of degree at most
$2j$
. By applying
$\mathfrak {D}^r$
to
$\hat \eta _{\ell }^\dagger $
, we have

where

so that (4.176) and (4.196) imply

The Hölder seminorm is similar, and hence,

for all
$r\geqslant 0,{\beta }\in (0,1)$
and
$S\leqslant Rd_{\ell }^{-1}$
with R fixed (as in [Reference Hein and Tosatti21, (4.172)]). Likewise, if we take at most 2 fiber derivatives landing on
$\hat \Phi $
and
$\hat J_{\ell }$
, then we use (4.196) and (4.182) to get

for all
$r\geqslant 0,{\beta }\in (0,1)$
and
$S\leqslant Rd_{\ell }^{-1}$
with R fixed (as in [Reference Hein and Tosatti21, (4.173)]). In particular, they converge to zero. If instead we use fixed metric
$g_X$
, then the derivatives of
$\hat \Phi $
and
$\hat J_{\ell }$
are bounded, and thus,

for all
$r\geqslant 0,{\beta }\in (0,1)$
and fixed
$S>0$
(as in [Reference Hein and Tosatti21, (4.174)]). Similarly to the discussion of [Reference Hein and Tosatti21, (4.175)], we also have

for all
$r\geqslant 0,{\beta }\in (0,1)$
and fixed
$S>0$
(as in [Reference Hein and Tosatti21, (4.175)]).
Next, we estimate the potential of
$\tilde {\eta }_{\ell }^\dagger $
when
$\varepsilon _{\ell }\leqslant C$
. Given
$r\geqslant 0$
, by applying
$\mathfrak {D}^r_{\textbf {bt}}$
to
$ \hat {\mathfrak {G}}_{\hat {t},k}(\hat A_{\ell ,i,p,k}^\sharp ,\hat G_{\ell ,i,p,k})$
with (3.1)

which can be estimated using (4.196) and (4.182) to get

The
$L^\infty $
norm is similar, and hence in the tilde picture we have

for all
$r\geqslant 0$
and S fixed.
4.10.7 Estimates for
$\tilde {\omega }_{\ell }^\sharp $
Recall that

We can follow exactly the same discussion in [Reference Hein and Tosatti21, §4.9.7] to conclude the following. Since
$\tilde {\omega }_{\ell ,\mathrm {can}}=d_{\ell }^{-2}\lambda _{\ell }^2\Theta _{\ell }^{*}\Psi _{\ell }^{*} \omega _{\mathrm {can}}$
and
$(\Psi _{\ell } \circ \Theta _{\ell })(\tilde z,\tilde y)=(d_{\ell } \lambda _{\ell }^{-1}z,y)$
where
$d_{\ell } \lambda _{\ell }^{-1}\to 0$
, the spatial stretching implies

for all
$\iota \geqslant 0,{\beta }\in (0,1)$
and
$S\leqslant Rd_{\ell }^{-1}$
with R fixed (which is the analog [Reference Hein and Tosatti21, (4.181)–(4.182)]). Therefore,

for all
$\iota \geqslant 0,{\beta }\in (0,1)$
and
$S\leqslant Rd_{\ell }^{-1}$
with R fixed. Likewise, if we only differentiate in the base and time directions, then we have the following improvement:

for all
$\iota \geqslant 0,{\beta }\in (0,1)$
with
$\iota +{\beta }>0$
and
$S\leqslant C d_{\ell }^{-1}$
(which is the analog of [Reference Hein and Tosatti21, (4.183)]) as well as

for all
$\iota \geqslant 0,{\beta }\in (0,1)$
and
$S\leqslant C d_{\ell }^{-1}$
(which is the analog of [Reference Hein and Tosatti21, (4.184)–(4.185)])
4.11 Expansion of the Monge-Ampère flow
We rewrite the complex Monge-Ampère equation (4.107) as

where the terms
$\mathcal {E}_1$
and
$\mathcal {E}_2$
are given respectively by the third and second line from the bottom in (4.212). Recall also that

so that if we define

then by definition, we have

and we can further rewrite the (4.212) as

where we defined

The next Proposition gives us control on the error terms
$\mathcal {E}_i$
:
Proposition 4.7. For any fixed
$R>0$
and
$i=1,2,3$
, we have

In particular,

Furthermore, if
$\varepsilon _{\ell }\geqslant C^{-1}$
, then for any fixed
$R>1$
and
$0\leqslant a\leqslant 2j$
, we have




Proof. The estimate (4.218) for
$\mathcal {E}_3$
follows easily from (4.158) and (4.206).
As for
$\mathcal {E}_2$
, recall that

The term
$nd_{\ell }^2\lambda _{\ell }^{-2}\tilde t$
is killed by
$[\mathfrak {D}_{\textbf {bt}}^{2j}\cdot ]_{{\alpha },{\alpha }/2,\mathrm {base}}$
if
$j>0$
, while if
$j=0$
, we have

Estimate (4.218) for
$\mathcal {E}_2$
then follows from this together with (4.208), (4.210) and (4.211).
For
$\mathcal {E}_1$
, we write

where we defined

Thanks to (4.145), (4.159), (4.192) and (4.210), for every
$0\leqslant \iota \leqslant 2j$
and fixed
$R>0$
, we have

If we write schematically
$\mathbf {A}=\operatorname {tr}_{\tilde \omega _{\ell }^\sharp }\left(\tilde {\eta }_{\ell }^\circ +\tilde {\eta }_{\ell }^\diamondsuit +\tilde {\eta }_{\ell ,j,k}\right), \mathbf {B}=\mathcal {E}_4$
, which are both
$o(1)$
locally uniformly, then we have
$\mathcal {E}_1=\log (1+\mathbf {A}+\mathbf {B})-\mathbf {A}.$
For any
$a\geqslant 0$
, we can then write schematically

and then (4.218) for
$\mathcal {E}_1$
follows from (4.228).
Now that (4.218) is established, (4.219) follows immediately from this and (4.216).
When
$\varepsilon _{\ell } \geqslant C^{-1}$
, estimates (4.220) and (4.221) follow from (4.145), (4.159), (4.191), (4.209) together with (4.229). Lastly, to prove (4.222) and (4.223), using (4.216), (4.220) and (4.221), it suffices to show that


for fixed
$R>0$
and
$0\leqslant a\leqslant 2j$
, which is a direct consequence of (4.145), (4.153), (4.181), (4.191) together with (4.209) and (4.229).
The goal is then to kill the RHS of (4.110) when
$j=0$
, and to kill the contributions of
$d_{\ell }^{-2j-{\alpha }}\underline{\tilde {\chi }_{\ell }^{*}}$
,
$d_{\ell }^{-2j-{\alpha }}\tilde {\psi }_{\ell ,j,k}$
and
$d_{\ell }^{-2j-{\alpha }}\tilde A_{\ell ,i,p,k}^{*}$
to (4.132) when
$j\geqslant 1$
. We will split the discussion into three cases (without loss of generality): Subcase A:
$\varepsilon _{\ell }\to +\infty $
, Subcase B:
$\varepsilon _{\ell }\to \varepsilon _\infty>0$
, Subcase C:
$\varepsilon _{\ell }\to 0$
as
$\ell \to +\infty $
where
$\varepsilon _{\ell }=d_{\ell }^{-1}\delta _{\ell }=d_{\ell }^{-1}\lambda _{\ell } e^{-t_{\ell }/2}$
.
4.12 Subcase A:
$\varepsilon _{\ell }\to +\infty $
In this subcase, the background geometry is diverging in the fiber directions, and similarly to the analogous case in [Reference Hein and Tosatti21, §4.10], we will kill all contributions to (4.132) using parabolic Schauder estimates for the linear heat equation. The Selection Theorem 3.1 will also be used crucially. The argument is quite long and involved because of the complexity of the quantitative estimates satisfied by all the pieces in the decomposition of the solution
$\tilde {\omega }^{\bullet }_{\ell }$
. We start with the direct analog of the non-cancellation result in [Reference Hein and Tosatti21, Proposition 4.7].
Proposition 4.8. The following inequalities hold for all
$0\leqslant {\alpha }<1$
,
$a\in \mathbb {N}$
,
$1\leqslant i\leqslant j$
,
$1\leqslant p\leqslant N_{i,k}$
and all
$R>0$
:


Proof. The proof is very similar to that of [Reference Hein and Tosatti21, Proposition 4.7], so we only highlight the differences. The starting point of the proof (Claim 1 in [Reference Hein and Tosatti21, Proof of Proposition 4.7]) is to express
${i\partial \bar \partial }\underline{\hat {\varphi }_{\ell }}$
and
$\hat A_{\ell ,i,p,k}$
as pushforwards of quantities on the total space. This step is essentially identical here, with the only difference being that in the formula for
$\hat A_{\ell ,i,p,k}$
, the term
$e^{-(2k+2)\frac {t}{2}}$
in [Reference Hein and Tosatti21, (4.204)] now becomes
$e^{-(2k+2)(t_{\ell }+\lambda _{\ell }^{-2}\hat t)/2}.$
The extra time-dependent constant
$e^{-(2k+2)\lambda _{\ell }^{-2}\frac {\hat t}{2}}$
will then also need to be differentiated in the analog of [Reference Hein and Tosatti21, (4.234)], which gives us extra cross terms in the analog of [Reference Hein and Tosatti21, (4.247)], but which can be estimated in a similar way resulting in the same upper bound as stated.
The next step is to try to commute the derivative
$\mathfrak {D}^a$
with the pushforward, and make the commutation error terms explicit. Recalling that
$\mathfrak {D}^{a}$
is a sum of terms of the form
${\mathbb {D}}^p\partial _{\hat {t}}^q$
, we observe that
$\partial _{\hat {t}}^q$
trivially commutes with pushforwards (with no error terms), while the commutation of
${\mathbb {D}}^p$
gives exactly the same result as in Claim 2 [Reference Hein and Tosatti21, Proof of Proposition 4.7]; cf. [Reference Hein and Tosatti21, (4.223), (4.240)].
The last step is to estimate the Hölder difference quotient of
$\mathfrak {D}^a{i\partial \bar \partial }\underline{\hat {\varphi }_{\ell }}$
and
$\mathfrak {D}^a\hat A_{\ell ,i,p,k}$
. This is now a space-time Hölder difference quotient, which we can split with the triangle inequality into a space-only difference quotient (which is estimated following the method of Claim 3 in [Reference Hein and Tosatti21, Proof of Proposition 4.7] verbatim), and a time-only difference quotient, which again commutes with pushforward and so can be estimated trivially without any further error terms. This completes the outline of the proof.
For notation convenience, we will denote
$\hat \eta _{\ell }={i\partial \bar \partial }\hat \varphi _{\ell }=\hat \eta _{\ell }^\dagger +\hat \eta _{\ell }^\ddagger +\hat \eta _{\ell }^\diamondsuit +\hat \eta _{\ell }^\circ +\hat \eta _{\ell ,j,k}$
.
4.12.1 The case
$j=0$
Unlike [Reference Hein and Tosatti21], the case
$j=0$
requires a separate treatment. This is due to the fact that the Monge-Ampère equation (4.216) is naturally a parabolic PDE for the scalar potential
$\tilde {\rho }_{\ell }$
, which, however, does not have a uniform bound on its
$L^\infty $
norm, which is an issue when applying Schauder estimates. This is remedied in two different ways according to whether
$j=0$
or
$j>0$
. In this subsection, we treat the case
$j=0$
.
The first crucial claim is that for any fixed
$R>0$
, we have

For the term
$\mathcal {E}_1$
, we have already proved an even strong result in (4.221), so we consider
$\mathcal {E}_2+\mathcal {E}_3$
, which when
$j=0$
equals

and since
$\partial _{\tilde {t}}\underline{\tilde {\chi }^\sharp _{\ell }}$
is a constant (in space and time), it is clear that

and we are left with showing that

For this, we pass to the check picture using the diffeomorphisms
$\Pi _{\ell }$
in (4.285) below, scaling geometric quantities by
$\varepsilon _{\ell }^{-2}$
, so that the quantity in (4.237) equals

and using (3.23), together with the facts that
$[\check {\eta }^\ddagger _{\ell }]_{{\alpha },{\alpha }/2,\check {Q}_{R\varepsilon _{\ell }^{-1}},\check {g}_{\ell }(0)}=0,$
and
$\delta _{\ell }^{-\alpha }[\check {\omega }_{\ell ,\mathrm {can}}]_{{\alpha },{\alpha }/2,\check {Q}_{R\varepsilon _{\ell }^{-1}},\check {g}_{\ell }(0)}\leqslant C\lambda _{\ell }^{-\alpha }=o(1),$
we see that (4.237) holds.
The next issue we face is that (4.144) does not provide us with uniform bounds on
$d_{\ell }^{-{\alpha }}\|\mathfrak {D}^{2}\tilde {\psi }_{\ell ,0,k}\|_{\infty ,\tilde {Q}_R,\tilde {g}_{\ell }(0)},$
for any fixed R, so we are unable to pass
$d_{\ell }^{-{\alpha }}\partial _{\tilde {t}}\tilde {\psi }_{\ell ,0,k}$
or
$d_{\ell }^{-{\alpha }}{i\partial \bar \partial }\tilde {\psi }_{\ell ,0,k}$
to a limit. To fix this, we use the method of [Reference Hein and Tosatti20, Subclaim 1.3], by replacing the whole fiber Y with a coordinate chart and performing a jet subtraction to
$d_{\ell }^{-{\alpha }}\tilde {\psi }_{\ell ,0,k}$
so that the remainder is locally
$C^2$
convergent.
To fix this, recall that
$\tilde {g}_{\ell }(\tilde {t}) = g_{\mathbb {C}^m} + \varepsilon _{\ell }^2 e^{-d_{\ell }^2\lambda _{\ell }^{-2}\tilde {t}} g_{Y,0}$
. Let
$\mathbf {x}^{2m+1}, \ldots , \mathbf {x}^{2m+2n}$
be normal coordinates for
$g_{Y,0}$
centered at
$y_{\ell }$
. Viewed as a map from Y to
$\mathbb {R}^{2n}$
, these depend on
$\ell $
, but we prefer to instead pull back our setup to
$\mathbb {R}^{2n}$
under the inverse map. In this sense, we may then assume without loss that

This is possible thanks to the compactness of Y. Define
$\tilde {\mathbf {x}}^j = \varepsilon _{\ell } \mathbf {x}^j$
, so that
$\tilde {\mathbf {x}}^{2m+1}, \ldots , \tilde {\mathbf {x}}^{2m+2n}$
are normal coordinates for
$\varepsilon _{\ell }^2 g_{Y,0}$
centered at
$y_{\ell }$
. Formally also write
$\tilde {\mathbf {x}}^1, \ldots , \tilde {\mathbf {x}}^{2m}$
for the standard real coordinates on
$\mathbb {C}^m$
. Then
$\tilde {\mathbf {x}}^1, \ldots , \tilde {\mathbf {x}}^{2m+2n}$
are normal coordinates for
$\tilde {g}_{\ell }(0)$
centered at
$\tilde {x}_{\ell }$
with

We then define a function
$\tilde {\psi }^\sharp _{\ell ,0,k}$
on
$\tilde {Q}_{2\varepsilon _{\ell }}$
to be the parabolic
$2$
nd order Taylor polynomial of
$\tilde {\psi }_{\ell ,0,k}$
at the space-time origin
$(0,0)$
, using the spatial coordinates
$\tilde {\mathbf {x}}^i,1\leqslant i\leqslant 2m+2n$
, and we define also
$\tilde {\psi }^{*}_{\ell ,0,k}:=\tilde {\psi }_{\ell ,0,k}-\tilde {\psi }^\sharp _{\ell ,0,k}.$
Since all Euclidean derivatives of
$\tilde {\psi }^{*}_{\ell ,0,k}$
of order at most
$2$
vanish at
$(0,0)$
, using the formula in [Reference Hein and Tosatti21, Lemma 2.3] relating
${\mathbb {D}}$
-derivatives and ordinary derivatives, we see that
$\mathfrak {D}^\iota \tilde {\psi }^{*}_{\ell ,0,k}\big |_{(0,0)}=0,0\leqslant \iota \leqslant 2$
.
Next, we prove the following bounds for
$d_{\ell }^{-\alpha }\tilde {\psi }^\sharp _{\ell ,0,k}$
: for all
$\iota \geqslant 0, 0<\beta <1$
, and
$0<R\leqslant \varepsilon _{\ell }$
,


To prove this claim, note that for any
$z\in B_1,$
the metric
$g_{Y,z}$
is at bounded distance to
$g_{Y,0}$
in
$C^\infty (Y)$
, and thus,

for all
$\iota \geqslant 1, R\leqslant \varepsilon _{\ell }$
.
Let us also note the following bounds for the Euclidean derivatives
$\tilde {\psi }_{\ell ,0,k}$
,

for all multiindices
$\gamma $
with
$|\gamma |+2q\leqslant 2$
. To see this, we first apply the diffeomorphism

and pull back metrics and
$2$
-forms, as well as
$\tilde {\psi }_{\ell ,0,k}$
, via
$\Lambda _{\ell }$
, multiply them by
$\varepsilon _{\ell }^{-2}$
and denote the resulting objects with a check, so, for example, the metrics
$\check {g}_{\ell }(0)=\varepsilon _{\ell }^{-2}\Lambda _{\ell }^{*}\tilde {g}_{\ell }(0)$
on
$\check {Q}_2$
are smoothly convergent to a fixed metric (smoothly comparable to Euclidean), and the pulled back complex structure is approaching the Euclidean one (without loss). The bounds (4.144) transform to

for
$0\leqslant \iota \leqslant 2$
. Since
$\check {g}_{\ell }(0)$
is approximately Euclidean, [Reference Hein and Tosatti21, Lemma 2.6] gives us that the Euclidean
$C^{\alpha ,{\alpha }/2}$
norm of
$d_{\ell }^{-\alpha }\check {\psi }_{\ell ,0,k}$
is also bounded by
$C\varepsilon _{\ell }^{2+\alpha }$
, and translating these back to the tilde picture proves (4.244).
First, we prove (4.241). Given
$(\tilde {x},\tilde {t}),(\tilde {x}',\tilde {t}')\in \tilde {Q}_R$
, call
$d=d^{\tilde {g}_{\ell }(0)}(\tilde {x},\tilde {x}')+|\tilde {t}-\tilde {t}'|^{\frac {1}{2}},$
and given
$p,q\geqslant 0$
with
$p+2q=\iota $
, we can bound

Since
$\tilde {\psi }^\sharp _{\ell ,0,k}$
is the sum of a polynomial of degree at most
$2$
in the
$\tilde {\mathbf {x}}$
variables (constant in time), and of a polynomial of degree at most
$1$
in the
$\tilde {t}$
variable (constant in space), it follows that
${\mathbb {D}}^{p+1}\partial _{\tilde {t}}^q\tilde {\psi }^\sharp _{\ell ,0,k}\equiv 0$
unless
$q=0$
(hence,
$\iota =p$
), and that
${\mathbb {D}}^{p}\partial _{\tilde {t}}^{q+1}\tilde {\psi }^\sharp _{\ell ,0,k}\equiv 0$
unless
$p=q=0$
(and hence,
$\iota =0$
).
We consider these two cases separately, so we first bound the term with
${\mathbb {D}}^{\iota +1}\tilde {\psi }^\sharp _{\ell ,0,k}$
(which is equal to
${\mathbb {D}}^{\iota +1}$
applied to the spatial Taylor polynomial only), by converting
${\mathbb {D}}^{\iota +1}$
into
$\nabla ^{\iota +1}$
using [Reference Hein and Tosatti21, Lemma 2.3] (which involves a certain tensor
$\mathbf {A}$
), and estimating



and estimating the big
$L^\infty $
norm by
$C\varepsilon _{\ell }^{1+\alpha -\iota }$
, as follows.
(1) We have
$ \partial ^b \Gamma = O( \varepsilon _{\ell }^{-b-1})$
by (4.240) and (4.243).
(2) The
$\mathbf {A}$
-tensor in the tilde picture is bounded by
$O(\varepsilon _{\ell }^{-2})$
, since it is schematically of the same type as
$\partial \Gamma $
. By the same reason,
$\nabla _{z,y,\tilde {z}}^{r}\mathbf {A}$
is
$O(\varepsilon _{\ell }^{-r-2})$
.
(3) Writing
$d_{\ell }^{-\alpha }\tilde {\psi }_{\ell ,0,k}=\psi $
, we can then estimate

where in the sum,
$a_1 + \cdots + a_{\ell } + \ell + b = r$
by counting the total number of
$\partial $
s and
$\Gamma $
s in each term, and
$b\leqslant |\gamma |$
. Now recall that
$R\leqslant \varepsilon _{\ell }$
, so that
$\partial ^b (\tilde {\mathbf {x}}-\tilde {\mathbf {x}}_{\ell })^\gamma =O(\varepsilon _{\ell }^{|\gamma |-b})$
. Since
$(\partial ^{|\gamma |}\psi )(\tilde {\mathbf {x}}_{\ell }) = O(\varepsilon _{\ell }^{2+\alpha -|\gamma |})$
by (4.244) and using Step (1), the quantity in (4.249) can be estimated by
$O(\varepsilon _{\ell }^{2+\alpha -r})$
.
(4) From Steps (2) and (3), we can bound

and so using Step (3) again, we obtain the desired bound of
$C\varepsilon _{\ell }^{1+\alpha -\iota }$
for the big
$L^\infty $
norm. This gives us the desired bound

for the first term in the last line of (4.245). As for the other term in that line, it is only nontrivial when
$\iota =p=q=0$
, and in that case, we want to bound
$d_{\ell }^{-\alpha }\|\partial _{\tilde {t}}\tilde {\psi }^\sharp _{\ell ,0,k}\|_{\infty ,\tilde {Q}_{2R},\tilde {g}_{\ell }(0)}$
. Since we have
$\partial _{\tilde {t}}\tilde {\psi }^\sharp _{\ell ,0,k}=(\partial _{\tilde {t}}\tilde {\psi }_{\ell ,0,k})(\tilde {\mathbf {x}}_{\ell },\tilde {t}_{\ell })$
, we obtain from (4.244)

and combining (4.245) with (4.251) and (4.252) proves (4.241).
To prove (4.242), given
$p,q\geqslant 0$
with
$p+2q=\iota $
, to bound
$d_{\ell }^{-\alpha }{\mathbb {D}}^p\partial _{\tilde {t}}^q\tilde {\psi }^\sharp _{\ell ,0,k}$
, we again need to consider only two cases. The first case (
$q=0,p=\iota $
) is the one with only spatial derivatives that land on the spatial Taylor polynomial (writing again
$\psi =d_{\ell }^{-\alpha }\tilde {\psi }_{\ell ,0,k}$
)

whose
$L^\infty $
norm on
$\tilde {Q}_{2R}$
is bounded by
$C\varepsilon _{\ell }^{2+\alpha -\iota }$
thanks to the estimates in Step (3) (with
$r=\iota $
) and Step (4) (with
$\iota $
there replaced by
$\iota -1$
). The second case only happens when
$\iota =2$
and we have only 1 time derivative that lands on the
$\tilde {t}$
-variable Taylor polynomial, which gives simply
$d_{\ell }^{-\alpha }(\partial _{\tilde {t}}\tilde {\psi }_{\ell ,0,k})(\tilde {\mathbf {x}}_{\ell },\tilde {t}_{\ell })$
, and this is bounded by
$C\varepsilon _{\ell }^{\alpha }$
by (4.244). Putting these observations together proves (4.242).
Combining (4.241) for
$\iota =2$
with the bound
$d_{\ell }^{-\alpha }[\mathfrak {D}^2\tilde {\psi }_{\ell ,0,k}]_{{\alpha },{\alpha }/2,\tilde {Q}_{R},\tilde {g}_{\ell }(0)}\leqslant C$
from (4.144), we see that

and so we can apply Lemma 2.4 and get

However, from (4.153), we also get uniform local parabolic
$C^{{\alpha },{\alpha }/2}$
bounds on
$d_{\ell }^{-\alpha }e^{-d_{\ell }^2\lambda _{\ell }^{-2}\tilde t}\partial _{\tilde {t}}\underline{\tilde \chi _{\ell }^{*}}$
and
$d_{\ell }^{-\alpha }{i\partial \bar \partial }\left(e^{-d_{\ell }^2\lambda _{\ell }^{-2}\tilde t}\underline{\tilde \chi _{\ell }^{*} }\right)$
, and so if we define

then from (4.214), we have
$\tilde {\rho }_{\ell }=\tilde {\rho }^{*}_{\ell }+\tilde {\psi }^\sharp _{\ell ,0,k},$
and

which by these estimates has uniform local parabolic
$C^{{\alpha },{\alpha }/2}$
bounds. The same estimates also give us uniform local parabolic
$C^{{\alpha },{\alpha }/2}$
bounds on
$d_{\ell }^{-\alpha }{i\partial \bar \partial }\tilde {\rho }^{*}_{\ell }$
, so passing to a subsequence, we have that

in
$C^{\gamma ,\gamma /2}_{\mathrm {loc}}$
for all
$0<\gamma <{\alpha }$
, for a function
$u_\infty \in C^{{\alpha },{\alpha }/2}_{\mathrm {loc}}$
and a
$(1,1)$
-form
$\eta _\infty \in C^{{\alpha },{\alpha }/2}_{\mathrm {loc}}$
on
$\mathbb {C}^{m+n}\times (-\infty ,0]$
. Moreover, thanks to (4.157) and (4.255), we have

and hence,
$d_{\ell }^{-\alpha }\partial _{\tilde {t}}\tilde {\rho }^{*}_{\ell }\to u_\infty $
locally uniformly.
Next, we observe that by definition, for any fixed
$R>0$
, we have

thanks to (4.136), and similarly from (4.137),

and since

we see that

and using (4.144), this gives

which combined with (4.241) implies

Also, (4.255) implies that

However, from (4.158), we see that

and which combined with (4.261) gives

Plugging (4.265), (4.266) and (4.268) into (4.257) gives

which implies that
$u_\infty $
is constant in space-time, and since its value at
$(0,0)$
vanishes, we conclude that
$u_\infty \equiv 0$
on
$\mathbb {C}^{m+n}\times (-\infty ,0]$
. Thus,
$d_{\ell }^{-\alpha }\partial _{\tilde {t}}\tilde {\rho }^{*}_{\ell }\to 0$
locally uniformly.
Given any
$(\tilde {x},\tilde {t}),(\tilde {x},\tilde {t}')\in \tilde {Q}_R$
, we have

and since we know that
$\int _{t'}^t\left(d_{\ell }^{-{\alpha }}\partial _{\tilde {t}}\tilde {\rho }^{*}_{\ell }\right)(\tilde {x},\tilde {s})d\tilde {s} \to 0$
locally uniformly (as a function of
$\tilde {x}$
), it follows that the RHS of (4.270) converges to zero weakly as currents. Since the LHS of (4.270) converges in
$C^\gamma _{\mathrm {loc}}$
to
$\eta _\infty (\tilde {x},\tilde {t})-\eta _\infty (\tilde {x},\tilde {t}')$
, we conclude that this Hölder continuous
$(1,1)$
-form (with
$\tilde {x}$
varying) is zero as a current, and hence, it is identically zero. This shows that
$\eta _\infty $
is time-independent.
From (4.216), we have

and from (4.234) and (4.241), we see that applying
$d_{\ell }^{-\alpha }[\cdot ]_{{\alpha },{\alpha }/2,\tilde {Q}_R,\tilde {g}_{\ell }(0)}$
to the RHS of (4.271), we get
$o(1)$
as
$\ell \to +\infty $
. We can then multiply (4.271) by
$d_{\ell }^{-\alpha }$
, and since the LHS converges in
$C^{\gamma ,\gamma /2}_{\mathrm {loc}}$
, we can pass to the limit (recalling that
$u_\infty \equiv 0$
) and get

on
$\mathbb {C}^{m+n}\times (-\infty ,0]$
, for some time-independent constant c. Since the value of the LHS of (4.272) at
$(0,0)$
is zero, this forces
$c=0$
; that is,

Since the
$C^{\gamma }_{\mathrm {loc}}$
form
$\eta _\infty $
is time-independent and weakly closed, it can be written as
$\eta _\infty ={i\partial \bar \partial } v_\infty $
for some time-independent function
$v_\infty \in C^{2+\gamma }_{\mathrm {loc}}(\mathbb {C}^{m+n})$
. From (4.273), we see that
$\Delta _{g_{\mathbb {C}^{m+n}}}v_\infty =0$
, so
$\eta _\infty $
is smooth by elliptic regularity, and passing to the limit (4.153) and (4.254), we see that
$|{i\partial \bar \partial } v_\infty |=O(|z|^{\alpha })$
. Thus, each component
$\partial _{\alpha } \partial _{\bar {\beta }}v_\infty $
of
${i\partial \bar \partial } v_\infty $
satisfies
$\Delta _{g_{\mathbb {C}^{m+n}}} \partial _{\alpha } \partial _{\bar {\beta }}v_\infty =0$
and
$|\partial _{\alpha } \partial _{\bar {\beta }}v_\infty |=O(|z|^{\alpha })$
, so by the standard Liouville Theorem for harmonic functions, we have that
$\eta _\infty $
has constant coefficients, and hence, it vanishes identically since its value at
$(0,0)$
is zero.
This implies that

locally uniformly on
$\mathbb {C}^{m+n}\times (-\infty ,0]$
in the coordinates
$(\tilde {\mathbf {x}},\tilde {t})$
. Recall that, by definition, we have

and that
$\underline{\tilde \chi _{\ell }^\sharp }$
is a polynomial on
$\mathbb {C}^m$
of degree at most
$2$
, while
$\tilde {\psi }^\sharp _{\ell ,0,k}$
is a polynomial on
$\mathbb {C}^{m+n}$
(in the
$(\tilde {\mathbf {x}},\tilde {t})$
coordinates) of degree at most
$2$
. To convert
${i\partial \bar \partial }$
into
${\mathbb {D}}$
-derivatives, schematically we have
${i\partial \bar \partial }=\tilde {J}_{\ell }\circledast {\mathbb {D}}^2+({\mathbb {D}}\tilde {J}_{\ell })\circledast {\mathbb {D}}$
, with the bounds (cf. [Reference Hein and Tosatti21, (4.304)])

Since
$\tilde {J}_{\ell }$
and
${\mathbb {D}}$
are independent of
$\tilde {t}$
, it follows that

so, as in the proof of (4.245), we can bound

using (4.276) for the second inequality and (4.242) for the third one. Also, from the bounds (4.158), it follows easily that

and so using (4.274), (4.275), (4.278) and (4.279), and recalling also (4.112), we see that

a contradiction to (4.110).
4.12.2 The case
$j\geqslant 1$
. Killing the contribution of
$\tilde A_{\ell ,i,p,k}$
In the rest of this section, we will assume that
$j\geqslant 1$
. The goal of this subsection is to prove a precise estimate on
$\tilde A_{\ell ,i,p,k}$
: for all
$a\geqslant 2j$
,
$a\in 2\mathbb {N}$
and
${\alpha }\leqslant {\beta }<1$
, there is
$C>0$
such that

where here and in the rest of this section, we use the notation
$O(\varepsilon _{\ell })$
for a radius R such that
$\Lambda \varepsilon _{\ell } \leqslant R\leqslant \Lambda ^2\varepsilon _{\ell }$
, where
$\Lambda>1$
is the fixed constant from (2.5) (so that the
$\tilde {g}_{\ell }(0)$
-geodesic ball centered at
$\tilde {x}_{\ell }$
with radius R contains a Euclidean ball of radius
$R/2$
times the whole Y fiber). Note that since
$j\geqslant 1$
, we have
$a\geqslant 2$
.
Observe that once (4.281) is established for all even
$a\geqslant 2j$
, the same estimate will also hold for all
$a\geqslant 2j$
by interpolation: indeed, if
$a\geqslant 2j+1$
is odd, then for any
$0<\rho <R$
,

and taking
$\rho ,R=O(\varepsilon _{\ell })$
gives the claim.
Thus, once (4.281) is established, taking
$a=2j+2+\iota $
(with
$-2\leqslant \iota \leqslant 2k$
), and
$\beta>\alpha $
gives an
$o(1)$
bound for the
$C^{\beta ,\beta /2}$
seminorm on the cylinder centered at
$(\tilde {x}_{\ell },\tilde {t}_{\ell })$
of radius
$2$
, which contains the other blowup point
$(\tilde {x}^{\prime }_{\ell },\tilde {t}^{\prime }_{\ell })$
which lies at distance
$1$
, and hence an
$o(1)$
bound for the
$C^{\alpha ,\alpha /2}$
seminorm on the same cylinder, which kills the contribution of
$\mathfrak {D}^{2j+2+\iota }\tilde A^{*}_{\ell ,i,p,k}$
in (4.132).
Apart from the fact that in our parabolic setting we only work with derivatives of even order (as was explained earlier), the overall argument to prove (4.281) will be similar to the one to prove [Reference Hein and Tosatti21, (4.252)], replacing
${\mathbb {D}}$
by
$\mathfrak {D}$
, j by
$2j$
and t by
$t_{\ell }$
. As in [Reference Hein and Tosatti21, (4.254)], we use (4.145), (4.153), (4.158), (4.191), (4.202) and Proposition 4.8 to conclude that for all
$a\geqslant 2j$
,
$1\leqslant i\leqslant j, 1\leqslant p\leqslant N_{i,k}$
,
$a\in 2\mathbb {N}$
and
${\alpha }\leqslant {\beta }<1$
, there is
$C>0$
so that

Our goal is to show that each term on the right-hand side of (4.283) is of
$O(\varepsilon _{\ell }^{{\alpha }-{\beta }})$
which implies (4.281).
We first treat the second term,
$[\mathfrak {D}^a_{\textbf {bt}} \tilde \eta _{\ell }]_{{\beta },{\beta }/2,\mathrm {base},\tilde Q_{O(\varepsilon _{\ell })},\tilde g_{\ell }(0)}$
. We start with noting that we can interchange
$\tilde {\eta }_{\ell }$
with
$\tilde {\eta }_{\ell }^\circ +\tilde {\eta }_{\ell }^\diamondsuit +\tilde {\eta }_{\ell ,j,k}$
thanks to (4.160) and (4.201): for all
$a\geqslant 2j$
,
$a\in 2\mathbb {N}$
,

To bound the RHS of (4.284) using the parabolic Schauder estimates in Proposition 2.9, we need to pass to the check picture via the diffeomorphism

pulling back all geometric quantities and scaling
$2$
-forms by
$\varepsilon _{\ell }^{-2}$
. We can then apply Proposition 2.9 to
$\check \eta _{\ell ,j,k}+\check \eta _{\ell }^\diamondsuit +\check \eta _{\ell }^\circ $
and then transfer the result back to the tilde picture. This shows that given any radius
$R=O(\varepsilon _{\ell })$
and
$0<\rho <R$
, and letting
$\tilde {R}=\rho +\frac {1}{2}(R-\rho )$
, we have

using (4.145), (4.159) and (4.191), where here
$\Delta _{\tilde {\omega }_{\ell }^\sharp }$
denotes the Hodge Laplacian acting on forms. This is the parabolic analog of [Reference Hein and Tosatti21, (4.259)]. Recalling that
$\tilde \eta _{\ell ,j,k}+\tilde \eta _{\ell }^\diamondsuit +\tilde {\eta }_{\ell }^\circ ={i\partial \bar \partial }\tilde {\rho }_{\ell },$
and using that the Hodge Laplacian of a Kähler metric commutes with
${i\partial \bar \partial }$
, we have

and using the PDE (4.216), the triangle inequality and the boundedness of
$\mathbb {P}$
-parallel transport, this can be bounded by

and inserting these into (4.286) and interpolating again with Proposition 2.3, we get

The main claim is then the following:
Claim 4.9. For all
$a\geqslant 2j$
,
$a\in 2\mathbb {N}$
,
$R=O(\varepsilon _{\ell })$
,
$0<\rho <R$
,
${\alpha }\leqslant {\beta }<1$
, if we let
$\tilde {R}=\rho +\frac {1}{2}(R-\rho )$
, then we have

Before giving the proof of Claim 4.9, we establish some of its consequences. Suppose Claim 4.9 has been proved for some
$a\geqslant 2j$
. Then plugging it into (4.289) we get for all
$R=O(\varepsilon _{\ell })$
and
$0<\rho <R$
,

and then the iteration lemma in [Reference Hein and Tosatti21, Lemma 2.9] gives

and choosing now
$\rho =O(\varepsilon _{\ell })$
gives

and this can be inserted back into (4.284) to finally give

which would show that the second term on the RHS of (4.283) is
$O(\varepsilon _{\ell }^{{\alpha }-{\beta }})$
.
Furthermore, we can interpolate between (4.293) and the
$L^\infty $
norm bound for
$\tilde \eta _{\ell ,j,k}+\tilde \eta _{\ell }^\diamondsuit +\tilde {\eta }_{\ell }^\circ $
that comes from (4.145), (4.159) and (4.191), using Proposition 2.3 on cylinders of radius
$O(\varepsilon _{\ell })$
to see that

for all
$0\leqslant b\leqslant a$
, and so following the discussion in [Reference Hein and Tosatti21, (4.280)–(4.283)], this implies that the third term on the RHS of (4.283) is of
$O(\varepsilon _{\ell }^{{\alpha }-{\beta }})$
.
To summarize, we have shown that if Claim 4.9 holds, then for all
$a\geqslant 2j$
,
$a\in 2\mathbb {N}$
,
${\alpha }\leqslant {\beta }<1$
,
$1\leqslant i\leqslant j$
and
$1\leqslant p\leqslant N_{i,k}$
, there is
$C>0$
such that

which is the parabolic analog of [Reference Hein and Tosatti21, (4.283)]. The remaining argument follows closely that of [Reference Hein and Tosatti21, §4.10.5]. We use induction on
$1\leqslant i\leqslant j$
to show that for all
$a\geqslant 2j$
and
$a\in 2\mathbb {N}$
, we have

The base case
$i=1$
follows directly from (4.296). Suppose (4.297) holds up to
$i_0-1$
for some
$i_0\geqslant 2$
. Then the first term inside the summation on (4.296) can be estimated by
$C\varepsilon _{\ell }^{{\alpha }-{\beta }}$
since
$a+2k+2\in 2\mathbb {N}$
. We now treat the second term inside the summation – namely, the last line of (4.296). First, we treat the terms with
$0\leqslant b\leqslant 2j$
, by using (4.73) and transferring it to the tilde picture which gives

and using these, we can bound these terms by
$o(\varepsilon _{\ell }^{{\alpha }-{\beta }})$
exactly as in [Reference Hein and Tosatti21, (4.286)]. As for the terms with
$2j<b\leqslant a+2k+2$
, we apply interpolation (i.e., Proposition 2.3) using (4.298) and (4.297) for
$i\leqslant i_0-1$
from the induction hypothesis to show that for
$2j<b\leqslant a+2k+2$
and
$i\leqslant i_0-1$
,

so that the terms in the last line of (4.296) with
$2j<b\leqslant a+2k+2$
are also of
$o(\varepsilon _{\ell }^{{\alpha }-{\beta }})$
by the same argument as [Reference Hein and Tosatti21, (4.290)]. This completes the inductive proof of (4.297), and hence of (4.281), modulo the proof of Claim 4.9, which we now turn to.
4.12.3 Proof of Claim 4.9
The proof of Claim 4.9 goes along similar lines to [Reference Hein and Tosatti21, (4.261)], but with some differences. We will prove the claim by induction on
$a\geqslant 2j$
, following the discussion in [Reference Hein and Tosatti21, §4.10.3–4.10.4]. Recall that the terms
$\mathcal {E}_i, i=1,2,3,$
are defined in (4.212) and (4.217). First, we consider the term
$\mathcal {E}_2+\mathcal {E}_3$
, which by definition equals

and we claim that for all
$b\geqslant 0$
and
${\alpha }\leqslant \beta <1$
, and all fixed
$R>1$
, there is
$C>0$
such that

Observe that (4.301) for
$a\leqslant 2j$
and
${\beta }={\alpha }$
is exactly given by (4.222) and (4.223). To prove it for
$a\geqslant 2j$
and
${\beta }\geqslant \alpha $
, we apply the diffeomorphism
$\Pi _{\ell }$
in (4.285) so that
$\Pi _{\ell }^{*}(\mathcal {E}_2+\mathcal {E}_3)$
equals

and for any
$b\geqslant 0$
, we have


So thanks to (4.301), we see that
$\delta _{\ell }^{-2j-\alpha }\Pi _{\ell }^{*}(\mathcal {E}_2+\mathcal {E}_3)$
is locally uniformly bounded in
$C^{2j+\alpha ,j+{\alpha }/2}$
with respect to the (essentially fixed) metric
$\check {g}_{\ell }(0)$
, so by Ascoli-Arzelà, up to passing to a subsequence, it converges locally uniformly on
$\mathbb {C}^m\times Y\times (-\infty ,0]$
to some limiting function
$\mathcal {F}$
. Since the quantity
$-\delta _{\ell }^{-2j-\alpha }\Pi _{\ell }^{*}(\mathcal {E}_2+\mathcal {E}_3)$
is exactly (3.15), we would like to apply the Selection Theorem 3.1, so we check that its hypotheses are satisfied. The functions
$\hat {A}_{\ell ,i,p,k}$
satisfy (3.7) with
$\alpha _0=\frac {\alpha ^2}{2j+\alpha }$
thanks to (4.172), while the function
$-e^{-\lambda _{\ell }^{-2}\hat t}\partial _{\hat t}\underline{\hat \chi _{\ell }^\sharp }-n\lambda _{\ell }^{-2}\hat {t}$
converges to
$0$
locally smoothly thanks to (4.158). We can thus apply the Selection Theorem and conclude that
$\delta _{\ell }^{-2j-\alpha }\Pi _{\ell }^{*}(\mathcal {E}_2+\mathcal {E}_3)$
converges to
$\mathcal {F}$
locally smoothly, and hence, its derivatives of all orders are uniformly bounded on
$\check Q_{R}$
. Thanks to (4.303), (4.304), this proves that (4.301) holds for all
$b\geqslant 0$
and
${\alpha }\leqslant \beta <1$
. Recalling then that
${i\partial \bar \partial } =\tilde J_{\ell }\circledast {\mathbb {D}}^2+{\mathbb {D}} \tilde J_{\ell }\circledast {\mathbb {D}}$
, and using the bounds in (4.176) for
$\tilde {J}_{\ell }$
and its derivatives, these imply directly that for all
$a\geqslant 2, {\alpha }\leqslant \beta <1,$

Next, we consider the term
$\mathcal {E}_1$
. Recall that

where
$\mathcal {E}_4$
was defined in (4.227). Taking
$\mathfrak {D}^{a}$
derivatives, for
$a\geqslant 2j\geqslant 2,a\in 2\mathbb {N}$
, we again expand it schematically as in (4.229). The first step is to prove estimates for
$\mathcal {E}_4$
, and for this, we observe that
$\varepsilon _{\ell }^{a-2j}d_{\ell }^{-2j-{\alpha }}\mathcal {E}_4$
is identical to the term
$\mathcal {N}$
in [Reference Hein and Tosatti21, (4.268)], and we will bound it following the discussion there. To do this, we need some basic estimates first. Using (4.209), we have for all
$\iota \geqslant 0$
that

Next, we observe that for all
${\alpha }\leqslant {\beta }<1$
and
$0\leqslant \iota <\max (a-1,2j)$
, we have

Indeed, these estimates are already known to hold for
$\iota \leqslant 2j$
and
$\beta ={\alpha }$
thanks to (4.145), (4.153) and (4.181); hence, using interpolation, they also hold for
$\iota <2j$
and
$\alpha \leqslant \beta <1$
, while if
$\iota <a-1, {\alpha }\leqslant {\beta }<1$
, these also hold thanks to the estimates (4.295) which hold by the induction hypothesis. Recalling that in this section we have
$j\geqslant 1$
, it follows that
$\max (a-1,2j)\geqslant 2$
.
From (4.308) and the definition of
$\mathcal {E}_4$
in (4.227), it follows immediately that for
$0\leqslant \iota <\max (a-1,2j)$
and
${\alpha }\leqslant {\beta }<1$
, we have

However, we do not have the estimates (4.308) when
$a-1\leqslant \iota \leqslant a$
and
$\iota \geqslant 2j$
, since our induction argument is only on even values of
$a\geqslant 2j$
, and this is different from the discussion in [Reference Hein and Tosatti21, §4.10.3]. To accommodate for the missing term with derivatives of order between
$a-1$
and a, we apply Proposition 2.3 with (4.308) so that



Given these, we can use the same method as in [Reference Hein and Tosatti21, (4.272)–(4.275)] and see that for every
$\iota $
with
$\max (a-1,2j)\leqslant \iota \leqslant a$
and
$\alpha \leqslant \beta <1$
, we have

and for
$\max (a-1,2j)\leqslant \iota <a, \alpha \leqslant \beta <1$
,

while for
$\iota =a$
,

One important observation that we used here is that whenever we need to use (4.310), (4.311) or (4.312) for some term in
$\mathcal {E}_4$
, the remaining part of this summand in
$\mathcal {E}_4$
is hit by at most
$1+\beta $
derivatives, and since
$1+\beta <2\leqslant \max (a-1,2j)$
, for these other terms, we are allowed to apply (4.308).
Now that we have our estimates for
$\mathcal {E}_4$
, we need estimates on derivatives of
$\operatorname {tr}_{\tilde {\omega }_{\ell }^\sharp }(\tilde {\eta }_{\ell }^\circ +\tilde {\eta }_{\ell }^\diamondsuit +\tilde {\eta }_{\ell ,j,k})$
. For this, from (4.307) and (4.308), we see that for
$0\leqslant \iota <\max (a-1,2j), {\alpha }\leqslant {\beta }<1$
,


while for derivatives of order
$a-1\leqslant \iota \leqslant a$
and
$\iota \geqslant 2j$
, we can argue as above and estimate crudely


Equipped with (4.309), (4.313), (4.314), (4.315), (4.316), (4.317), (4.318) and (4.319), we proceed to estimate derivatives of
$\mathcal {E}_1$
. We first consider the first line of (4.229), which we can write as

and we take
$\varepsilon _{\ell }^{a-2j}d_{\ell }^{-2j-\alpha }[\cdot ]_{{\beta },{\beta }/2,\tilde Q_{\tilde R},\tilde g_{\ell }(0)}$
of this. Since
$|\operatorname {tr}_{\tilde {\omega }_{\ell }^\sharp }(\tilde {\eta }_{\ell }^\circ +\tilde {\eta }_{\ell }^\diamondsuit +\tilde {\eta }_{\ell ,j,k})+\mathcal {E}_4|=o(1)$
, when the difference quotient lands on
$\mathfrak {D}^a \mathrm {tr}$
, we can estimate this by
$o(1)$
times (4.319) (with
$\iota =a$
). Similarly,
$[\operatorname {tr}_{\tilde {\omega }_{\ell }^\sharp }(\tilde {\eta }_{\ell }^\circ +\tilde {\eta }_{\ell }^\diamondsuit +\tilde {\eta }_{\ell ,j,k})+\mathcal {E}_4]_{C^\beta }=o(\varepsilon _{\ell }^{-\beta })$
, so when the difference quotient lands on
$\frac {\mathrm {tr}+\mathcal {E}_4}{1+\mathrm {tr}+\mathcal {E}_4}$
, we can estimate this by
$o(\varepsilon _{\ell }^{-\beta })$
times (4.318) (with
$\iota =a$
). And when the difference quotient lands on
$\frac {\mathfrak {D}^a\mathcal {E}_4}{1+\mathrm {tr}+\mathcal {E}_4}$
, we argue similarly with (4.313) and (4.315). So all together when we apply
$\varepsilon _{\ell }^{a-2j}d_{\ell }^{-2j-\alpha }[\cdot ]_{{\beta },{\beta }/2,\tilde Q_{\tilde R},\tilde g_{\ell }(0)}$
to (4.320), we can bound it by

Lastly, we need to consider what happens when we take
$\varepsilon _{\ell }^{a-2j}d_{\ell }^{-2j-\alpha }[\cdot ]_{{\beta },{\beta }/2,\tilde Q_{\tilde R},\tilde g_{\ell }(0)}$
of the large sum in the second line of (4.229). If all the derivatives that appear there are of order
$<\max (a-1,2j)$
, then this is bounded by
$o(\varepsilon _{\ell }^{\alpha -\beta })$
, while if there is at least one derivative of order at least
$\max (a-1,2j)$
(to which we apply (4.313)–(4.315), (4.318)–(4.319)), then all other derivatives in total are of order at most
$1$
(and to these we can instead apply (4.309), (4.316)–(4.317)). Putting all these together proves that

and combining this with (4.305) completes the proof of Claim 4.9.
For later use, observe also that the same argument gives an analogous bound for the
$L^\infty $
norm of derivatives of
$\mathcal {E}_1$
– namely,

4.12.4 Killing the contributions from
$\underline{\tilde {\chi }^{*}_{\ell }}$
The starting point is (4.232) with
$a=2j$
and radius
$C\delta _{\ell }$
. For
$0\leqslant b\leqslant 2j$
, we first bound the term

using (4.145), (4.159), (4.160), (4.189), (4.190) and (4.202). Using this, we transfer (4.232) to the tilde picture and multiply it by
$d_{\ell }^{\beta -\alpha }$
and get

where we used (4.284) and (4.293). Taking
$\beta>\alpha $
gives us an
$o(1)$
bound for the parabolic
$C^{\beta ,\beta /2}$
seminorm of
$d_{\ell }^{-2j-\alpha }\mathfrak {D}^{2j}\tilde {\eta }^\diamondsuit _{\ell }$
on the cylinder of radius
$2$
centered at
$\tilde {x}_{\ell }$
(which contains the other blowup point
$\tilde {x}^{\prime }_{\ell }$
), and hence an
$o(1)$
bound for the parabolic
$C^{\alpha ,\alpha /2}$
seminorm on this same cylinder. Thanks to the bounds (4.159), the same conclusion holds for the parabolic
$C^{\alpha ,\alpha /2}$
seminorm of

on the same cylinder, which kills one contribution of
$\underline{\tilde {\chi }^{*}_{\ell }}$
to (4.132).
Next, using (4.295) and the bounds (4.307) for
$\tilde {\omega }^\sharp _{\ell }$
, we see that for any
$a\geqslant 0$
, there is
$C>0$
such that


while from (4.301), (4.322), (4.323) and (4.293), we have


and so using these in the PDE (4.216), we get


At this point, we want to deduce from this bounds for the fiber average of
$\partial _{\tilde t} \tilde {\rho }_{\ell }+d_{\ell }^2\lambda _{\ell }^{-2}\tilde {\rho }_{\ell }$
, and this can be done using the following ‘non-cancellation’ estimate, stated in the hat picture, for a smooth time-dependent function f on
$\hat {Q}_R$
(where the fiber average is
$\underline{f}=(\mathrm {pr}_B)_*(f\Psi _{\ell }^{*}\omega _F^n)$
) which states

and which is proved exactly as [Reference Hein and Tosatti21, (4.199)], using Claims 2 and 3 there. We apply this in the tilde picture with
$a=2j$
to the function
$\partial _{\tilde t} \tilde {\rho }_{\ell }+d_{\ell }^2\lambda _{\ell }^{-2}\tilde {\rho }_{\ell }$
, whose fiber average is by definition
$e^{-d_{\ell }^2\lambda _{\ell }^{-2}\tilde {t}}\partial _{\tilde {t}}\underline{\tilde {\chi }^{*}_{\ell }}$
, and we get

using (4.331) and (4.332). Thus, if we take
${\beta }>{\alpha }$
, then the LHS of (4.334) is
$o(1)$
, and so we get an
$o(1)$
bound for the same
$C^{\beta ,\beta /2}$
seminorm on
$\tilde {Q}_R$
(
$R>1$
fixed), and hence,

which thanks to the bounds in (4.153) implies that

which kills the other contribution of
$\underline{\tilde {\chi }^{*}_{\ell }}$
to (4.132).
4.12.5 Killing the contribution of
$\tilde \psi _{\ell ,j,k}$
It remains to kill the contribution from
$\tilde {\psi }_{\ell ,j,k}$
to (4.132). In contrast with [Reference Hein and Tosatti21, §4.10.7] where they had to kill the contribution from
$d_{\ell }^{-2j-\alpha }\mathfrak {D}^{2j}{i\partial \bar \partial }\tilde {\psi }_{\ell ,j,k}$
, here we need to kill
$d_{\ell }^{-2j-\alpha }\mathfrak {D}^{2j+2}\tilde {\psi }_{\ell ,j,k}$
.
From the definition of
$\tilde {\rho }_{\ell }$
in (4.214), we have

and so using (4.332) and (4.334), together with the triangle inequality and the boundedness of
$\mathbb {P}$
-parallel transport, gives

Now thanks to (4.145) and (4.181), we can bound

and so (4.338) implies

Next, from the bound (4.325), as well as the analogous bounds for lower derivatives of
$\tilde {\eta }^\diamondsuit _{\ell }$
which come from (4.159), and the bounds (4.307) for
$\tilde {\omega }^\sharp _{\ell }$
, we see that

and this together with (4.328) gives

which together with (4.340) gives

This estimate can be inserted in the Schauder estimates in Proposition 2.8 (as usual by first going to the check picture, and then changing the result back to the tilde picture), with radii both
$O(\varepsilon _{\ell })$
,

where we used the bounds (4.145) and (4.181) for the
$L^\infty $
norm. As usual, taking
$\beta>\alpha $
, this implies an
$o(1)$
bound for the
$C^{{\beta },{\beta }/2}$
seminorm on
$\tilde {Q}_{O(\varepsilon _{\ell })}$
, hence on
$\tilde {Q}_{2}$
, and hence also an
$o(1)$
bound for the
$C^{{\alpha },{\alpha }/2}$
seminorm on
$\tilde {Q}_{2}$
; that is,

Next, we claim that

The argument is identical to the deduction of (4.181) except we now use the improved parabolic Hölder seminorm from (4.281) instead of that from blowup argument. By applying (3.1), we have

so that

To estimate it, we need the following simple bounds from [Reference Hein and Tosatti21, (4.302)–(4.303)]

for all
$\iota \geqslant 0$
,
${\alpha }\in (0,1)$
and fixed
$S>1$
.
Transferring (4.178) to the tilde picture,

for all
$0\leqslant \iota \leqslant 2k+2+2j$
. Putting
${\beta }>{\alpha }$
in (4.281) yields

for all
$\iota \geqslant 2j$
and thus implies

for all
$2j\leqslant \iota \leqslant 2j+2+2k$
while for
$0\leqslant \iota \leqslant 2j$
, we have from (4.163) that

since
$\varepsilon _{\ell }\to +\infty $
.
We now estimate (4.348) using the above estimates of each terms:

which proves (4.346). Combining it with (4.345), and using the triangle inequality gives

which kills the contribution of
$\tilde {\psi }_{\ell ,j,k}$
in (4.132). In conclusion, all terms that appear on the right-hand side of (4.132) converge to zero, which gives a contradiction and finally concludes Subcase A.
4.13 Subcase B:
$\varepsilon _{\ell }\to \varepsilon _\infty>0$
Without loss of generality, we will assume that
$\varepsilon _\infty =1$
. By (4.144), (4.145) and [Reference Hein and Tosatti21, Lemma 2.6], we see that
$d_{\ell }^{-2j-{\alpha }}\tilde {\psi }_{\ell ,j,k}$
converges in
$C^{2j+2+{\beta },j+1+{\beta }/2}_{\mathrm {loc}}$
to a function
$\tilde {\psi }_{\infty ,j,k}\in C^{2j+2+{\alpha },j+1+{\alpha }/2}_{\mathrm {loc}}$
defined on
$\mathbb {C}^m\times Y\times (-\infty ,0]$
, for all
$0<{\beta }<{\alpha }$
, while (4.153) gives us that
$d_{\ell }^{-2j-{\alpha }}\partial _{\tilde t}\underline{\tilde {\chi }_{\ell }^{*}}$
and
$d_{\ell }^{-2j-{\alpha }}{i\partial \bar \partial }\underline{\tilde {\chi }_{\ell }^{*}}$
converge in
$C_{\mathrm { loc}}^{2j+{\beta },j+{\beta }/2}$
for all
$0<{\beta }<{\alpha }$
to a function
$u_\infty \in C^{2j+{\alpha },j+{\alpha }/2}_{\mathrm {loc}}$
and to a
$(1,1)$
form
$\eta _\infty \in C^{2j+{\alpha },j+{\alpha }/2}_{\mathrm {loc}}$
, respectively, on
$\mathbb {C}^m\times (-\infty ,0]$
. Using again (4.153), we also have that
$d_{\ell }^{-2j-{\alpha }}e^{-d_{\ell }^2\lambda _{\ell }^{-2}\tilde t}\partial _{\tilde t}\underline{\tilde \chi _{\ell }^{*} }\to u_\infty $
in
$C_{\mathrm {loc}}^{2j+{\beta },j+{\beta }/2}$
and
$d_{\ell }^{-2j-{\alpha }}e^{-d_{\ell }^2\lambda _{\ell }^{-2}\tilde t}{i\partial \bar \partial }\underline{\tilde \chi _{\ell }^{*} }\to \eta _\infty $
in
$C_{\mathrm {loc}}^{2j+{\beta },j+{\beta }/2}$
. Moreover, when
$j\geqslant 1$
, we have

From (4.144), (4.145), (4.153), (4.191), we see that
$\tilde {\psi }_{\infty ,j,k}=O(r^{2j+2+{\alpha }})$
,
$u_\infty =O(r^{2j+{\alpha }})$
,
$\eta _\infty =O(r^{2j+{\alpha }}),$
and
$\tilde {\eta }_\infty ^\circ =O(r^{2j+{\alpha }})$
where
$r=|z|+\sqrt {|t|}$
.
By (4.164), (4.163) and (4.168), the functions
$d_{\ell }^{-2j-{\alpha }}\tilde A_{\ell ,i,p,k}^{*}$
converge to limiting functions
$\tilde A_{\infty ,i,p,k}^{*}$
from the base
$\mathbb {C}^m$
in
$C_{\mathrm {loc}}^{2j+2+2k+{\beta },j+1+2k+{\beta }/2}$
for any
$0<{\beta }<{\alpha }$
while
$\tilde G_{i,p,k}$
converge locally uniformly smoothly to functions
$\tilde G_{\infty ,i,p,k}$
pulled back from Y. Using (4.181), we see that the functions
$d_{\ell }^{-2j-{\alpha }}\tilde {\mathfrak {G}}_{\tilde {t},k}(\tilde {A}_{\ell ,i,p,k}^{*},\tilde {G}_{\ell ,i,p,k})$
converge in
$C^{2j+2+{\beta },j+1+{\beta }/2}_{\mathrm {loc}}$
to a function which, thanks to [Reference Hein and Tosatti21, (4.310)], is given by

By (4.158) and (4.200), the metrics
$\tilde g^\sharp _{\ell }(t)\to g_P=g_{\mathbb {C}^m}+g_{Y,z_\infty =0}$
locally smoothly where
$g_{\mathbb {C}^m}$
equals
$g_{\mathrm {can}}|_{z=z_\infty =0}$
.
Recall from (4.216) that

where
$\tilde {\rho }_{\ell }$
is given by (4.214). From (4.144), (4.145), (4.157) and (4.179), it follows that

in
$C^0_{\mathrm {loc}}$
, so as
$\ell \to +\infty $
the LHS of (4.358) converges to

while at the same time, thanks to (4.219), the LHS of (4.358) is forced to be a polynomial
$\mathcal {F}$
in the
$(z,t)$
variables, of (spacetime) degree at most
$2j$
, with coefficients that are functions on Y, namely

for some smooth functions
$H_{p,q}$
pullback from the fiber Y, where here again we treat z as real variables. Plugging (4.361) into (4.360), and taking the fiberwise average, gives

where Q is a polynomial in the
$(z,t)$
variables of degree at most
$2j$
. But thanks to the jet subtractions,
$u_\infty $
and
$\eta _\infty $
have vanishing
$2j$
parabolic jet at
$(z,t)=(0,0)$
, so

on
$\mathbb {C}^m\times (-\infty ,0]$
.
We now claim that
$u_\infty $
and
$\eta _\infty $
are both identically zero. First, we assume that
$j=0$
. In this case, we can use (4.268) and see that
$u_\infty $
is a constant in space-time, and hence identically zero. At this point, we can argue similarly as in subcase A to show that
$\eta _\infty $
also vanishes identically: first, we show that
$\eta _\infty $
is time-independent, by arguing similarly to (4.270), noting that for any
$\tilde {z}\in \mathbb {C}^m$
and
$\tilde {t},\tilde {t}'\leqslant 0$
,

and since
$\int _{\tilde {t}'}^{\tilde {t}}\left(d_{\ell }^{-{\alpha }}\partial _{\tilde {t}}\underline{\tilde {\chi }^{*}_{\ell }}\right)(\tilde {z},\tilde {s})d\tilde {s}\to 0$
locally uniformly (from
$u_\infty \equiv 0$
), it follows that the RHS of (4.364) goes to
$0$
weakly, and since the LHS converges in
$C^\beta _{\mathrm {loc}}$
to
$\eta _\infty (\tilde {z},\tilde {t})-\eta _\infty (\tilde {z},\tilde {t}')$
, this must be identically zero as claimed.
We can then write
$\eta _\infty ={i\partial \bar \partial } v_\infty $
, where
$v_\infty \in C^{2+{\alpha }}_{\mathrm {loc}}(\mathbb {C}^m)$
is time-independent and
$\Delta _{g_{\mathbb {C}^m}}v_\infty =0$
, so
$v_\infty $
is smooth and
$|{i\partial \bar \partial } v_\infty |=O(|z|^{\alpha })$
. The Liouville Theorem applied to each component of
${i\partial \bar \partial } v_\infty $
then implies that
$\eta _\infty $
has constant coefficients; hence, it vanishes identically since its value at
$(0,0)$
vanishes.
Next, we assume
$j\geqslant 1$
. We can then differentiate (4.363) and use (4.356) to see that

so
$u_\infty $
solves the heat equation on
$(-\infty ,0]\times \mathbb {C}^m$
, and
$|\mathfrak {D}^{\iota }u_\infty |=O(r^{2j+{\alpha }-\iota })$
where
$r=|x|+\sqrt {t}$
for
$0\leqslant \iota \leqslant 2j$
by (4.153). By applying Liouville Theorem for ancient heat equation to
$\mathfrak {D}^{2j}u_\infty $
, we see that
$u_\infty $
must be a space-time polynomial of degree at most
$2j$
, and hence,
$u_\infty \equiv 0$
since its parabolic
$2j$
-jet at
$(0,0)$
vanishes. Going back to (4.356), it then follows that
$\eta _\infty $
is time-independent, and it is clearly of the form
$\eta _\infty ={i\partial \bar \partial } v_\infty $
for some time-independent function
$v_\infty \in C^{2j+2+{\alpha }}_{\mathrm {loc}}(\mathbb {C}^m)$
. From (4.363), we see that
$v_\infty $
is harmonic. Since
$|{\mathbb {D}}^{2j}\eta _\infty |=O(|z|^{\alpha })$
, the Liouville Theorem in [Reference Hein and Tosatti20, Proposition 3.12] then shows that the coefficients of
$\eta _\infty $
are polynomials of degree at most
$2j$
, and since these coefficients have vanishing
$2j$
-jet at the origin, this implies that
$\eta _\infty $
vanishes identically.
Now that we know that
$u_\infty \equiv 0, \eta _\infty \equiv 0$
, we can return to (4.360), (4.361) and get

Recall from (4.358) and the discussion after it that
$d_{\ell }^{-2j-{\alpha }}(\mathcal {E}_1+\mathcal {E}_2+\mathcal {E}_3)\to \mathcal {F}$
locally uniformly. From (4.220), we see that
$d_{\ell }^{-2j-{\alpha }}\mathcal {E}_1\to 0$
locally uniformly, while the term
$-d_{\ell }^{-2j-{\alpha }}(\mathcal {E}_2+\mathcal {E}_3)$
is exactly equal to (3.15) (in this subcase, the check picture equal the tilde picture since
$\varepsilon _{\ell } \to 1$
), so the Selection Theorem 3.1 shows that
$\mathcal {F}$
is also equal to the limit of

where
$\Sigma _{\ell }$
was defined in (3.3), and where
$f_{\ell ,0}, f_{\ell ,i,p}$
are time-dependent functions pulled back from the base, such that
$\hat {f}_{\ell , 0}=\Psi _{\ell }^{*}f_{\ell ,0},\hat {f}_{\ell ,i,p}=\Psi _{\ell }^{*}f_{\ell ,i,p}$
converge locally smoothly to zero. This implies that for any function G on
$\mathbb {C}^m\times Y$
which is fiberwise
$L^2$
orthogonal to the span of the functions
$\{\tilde {G}_{\infty ,i,p,k}\}_{1\leqslant i\leqslant j,1\leqslant p\leqslant N_{i,k}}$
together with the constants, and for any
$z\in \mathbb {C}^m$
and
$t\in (-\infty ,0]$
, we have

which implies that we can write

for some functions
$g_0,g_{i,p,k}$
on
$\mathbb {C}^m\times (-\infty ,0]$
. Since
$\mathcal {F}$
is a polynomial in
$(z,t)$
of degree at most
$2j$
, by fiberwise
$L^2$
projecting
$\mathcal {F}$
onto each
$\tilde {G}_{\infty ,i,p,k}$
and onto the constants, we see that the coefficients
$g_0(z,t),g_{i,p,k}(z,t)$
are also polynomials of degree at most
$2j$
. This shows that
$\mathcal {F}$
is a linear combination of the functions
$\tilde {G}_{\infty ,i,p,k}$
together with the constant
$1$
, with coefficients that are polynomials in
$(z,t)$
of degree at most
$2j$
, that is,

where
$K_0(z,t),K_q(z,t)$
are polynomials of degree at most
$2j$
, and
$H_q(y)$
are functions pulled back from the fiber Y that lie in
$\mathcal {H}$
, the fiberwise span of the functions
$\tilde {G}_{\infty ,i,p,k}, 1\leqslant i\leqslant j, 1\leqslant p\leqslant N_{i,k}$
.
Following the argument in deriving [Reference Hein and Tosatti21, (4.330)] (which holds verbatim here), for each fixed
$t\leqslant 0$
, we obtain

where each quantity is evaluated at
$t\leqslant 0$
. For notational convenience, we denote

which thanks to (4.366) satisfies
$ \left(\partial _{t}-\Delta _{\omega _P}\right)u=\mathcal {F}$
, and so given any
$p,q\geqslant 0$
with
$p+2q=2j+2$
, and given
$v_1,\dots ,v_p$
tangent vectors to
$\mathbb {C}^m$
, from (4.369), we see that

while, thanks to (4.144), (4.145) and (4.185), we also have that
${\mathbb {D}}^p_{v_1\cdots v_p}\partial _t^q u=O(r^{\alpha })$
, where
$r=|z|+\sqrt {|t|}$
. Since
$\omega _P$
is of Ricci-flat, we can apply the Liouville Theorem in [Reference Fong and Lee11, Proposition 2.1] for ancient solutions of the heat equation, and conclude that
${\mathbb {D}}^p_{v_1\cdots v_p}\partial _t^q u$
is a constant in space and time. Since this is true for arbitrary
$p,q$
with
$p+2q=2j+2$
, and for arbitrary
$v_1,\dots ,v_p$
, this means that for every given
$y\in Y$
, the function
$u(z,y,t)$
is a (parabolic) polynomial in
$(z,t)$
of degree at most
$2j+2$
. Thus,
$(\partial _t u)(z,y,t)$
is a polynomial in
$(z,t)$
of degree at most
$2j$
; hence, the fiber integration

is also polynomial of degree at most
$2j$
. Thus, if we insert (4.366) into (4.371), we obtain

where
$Q_{i,p}$
is a (parabolic) polynomial on
$\mathbb {C}^m\times (-\infty ,0]$
of degree at most
$2j$
. We can then use this to show by induction on
$0\leqslant i\leqslant j$
that
$\tilde {A}^{*}_{\infty ,i,p,k}=0$
for all
$i,p$
. Indeed, in the base case of the induction
$i=1$
, the last term in (4.375) is not present, and so
$\tilde {A}^{*}_{\infty ,1,p,k}$
is a polynomial of degree at most
$2j$
, but since it also has vanishing
$2j$
-jet at
$(0,0)$
, it must be identically zero. The induction step is then exactly the same.
Next, since we have shown that
$\tilde A_{\infty ,i,p,k}^{*}\equiv 0$
for all
$i,p$
, from (4.357) and (4.372), we see that
$u=\tilde {\psi }_{\infty ,j,k}$
, which by (4.371) satisfies

for all
$z,i,p$
i.e., we have

the fiberwise
$L^2(\omega _Y)$
-orthogonal space to
$\mathcal {H}$
. Thanks to (4.366) and (4.369), we also have

We then claim that we have
$u\equiv 0$
. To show this, we apply a trick from [Reference Fong and Lee11, Claim 3.2]. We define a function v on
$\mathbb {C}^m\times Y\times (-\infty ,0]$
as the fiberwise
$L^2(\omega _Y)$
-orthogonal projection of u onto
$\mathcal {H}^\perp $
. Recalling that
$|u|=O(r^{2j+2+{\alpha }})$
, where
$r=|z|+\sqrt {|t|}$
, we claim that v satisfies the same growth bound
$|v|=O(r^{2j+2+{\alpha }})$
. To see this, it suffices to prove this bound for the fiberwise
$L^2(\omega _Y)$
-orthogonal projection of u onto
$\mathcal {H}$
, which equals

and whose supremum on
$B_r(0)\times (-r^2,0]$
is clearly bounded by
$C\sup _{B_r(0)\times (-r^2,0]}|u|\leqslant Cr^{2j+2+{\alpha }},$
as claimed.
Projecting (4.378) onto
$\mathcal {H}^\perp $
, and using (4.377), we see that

Given
$R>1$
, which later will be taken sufficiently large, we define a function on
$\mathbb {C}^m\times Y\times (-\infty ,0]$
by

and consider the weighted
$L^2$
energy

which is finite since
$|v|=O(r^{2j+2+{\alpha }})$
and
$\phi _R$
decays fast at spatial and time infinity. Differentiating
$E_R$
in time and using (4.380), we get

Since v is a fiberwise
$L^2$
-projection of u, we have
$\int _Y v^2\omega _Y^n\leqslant \int _Yu^2\omega _Y^n$
. Using this, together with (4.377), we can estimate

where the integration by part is justified by the growth bound of u and the fast decay of
$\phi _R$
. Since u has fiberwise average zero, the Poincaré inequality on
$(Y,\omega _Y)$
implies

where the constant C is independent of R; hence, for all
$t\leqslant 0$
, we have

provided we choose R sufficiently large. Thus, for any
$s<t\leqslant 0$
, we have
$E_R(t)\leqslant E_R(s)$
, but if we let
$s\to -\infty $
, then since u grows at most polynomially while
$\phi _R$
has exponential decay, we see that
$\lim _{s\to -\infty } E_R(s)=0$
, and so
$E_R(t)\equiv 0$
for all
$t\leqslant 0$
, which implies that
$v\equiv 0$
. From (4.378), we then see that
$\Delta _{\omega _P}u\equiv 0$
, so u is harmonic on
$\mathbb {C}^m\times Y$
and
$|u|=O(r^{2j+2+{\alpha }})$
, so [Reference Hein and Tosatti20, Proposition 3.12] implies that for any fixed
$t\leqslant 0$
, the function
$u(\cdot , t)$
is the pullback of a polynomial on
$\mathbb {C}^m$
of degree at most
$2j+2$
, and since it also has fiberwise average zero, it vanishes identically.
To summarize, we have thus proved that
$d_{\ell }^{-2j-{\alpha }}\tilde {\psi }_{\ell ,j,k}$
,
$d_{\ell }^{-2j-{\alpha }}\partial _{\tilde t}\underline{\tilde {\chi }_{\ell }^{*}}$
,
$d_{\ell }^{-2j-{\alpha }}{i\partial \bar \partial }\underline{\tilde {\chi }_{\ell }^{*}},$
$d_{\ell }^{-2j-{\alpha }}\tilde {A}^{*}_{\ell ,i,p,k}$
all go to zero locally uniformly in the appropriate topologies, which when
$j\geqslant 1$
shows that the RHS of (4.132) goes to zero, and gives a contradiction. As for the case
$j=0$
, by definition, we have

and we have just shown that
$d_{\ell }^{-{\alpha }}{i\partial \bar \partial }\tilde {\psi }_{\ell ,0,k}$
and
$d_{\ell }^{-{\alpha }}e^{-d_{\ell }^2\lambda _{\ell }^{-2}\tilde {t}}{i\partial \bar \partial }\underline{\tilde {\chi }_{\ell }^{*}}$
both go to zero locally uniformly, so using also (4.279), we see that the RHS of (4.110) also converges to zero, contradiction. This concludes Subcase B.
4.14 Subcase C:
$\varepsilon _{\ell }\to 0$
In this subcase, using repeatedly [Reference Hein and Tosatti21, Lemma 2.6], thanks to (4.144) and (4.145) we have that
$d_{\ell }^{-2j-{\alpha }}\tilde {\psi }_{\ell ,j,k}$
converges to zero in
$C^{2j+2+{\beta },j+1+{\beta }/2}_{\mathrm { loc}}$
. Also, using (4.194), we see that
$d_{\ell }^{-2j-{\alpha }}\tilde {\eta }^\circ _{\ell ,j,k}$
converges in
$C^{2j+{\beta },j+{\beta }/2}_{\mathrm { loc}}$
to a
$(1,1)$
-form
$\tilde {\eta }^\circ _{\infty ,j,k}$
on
$\mathbb {C}^m\times Y\times (-\infty ,0]$
which is weakly closed. The bounds in (4.153) give us that
$d_{\ell }^{-2j-{\alpha }}\partial _{\tilde t}\underline{\tilde {\chi }_{\ell }^{*}}$
and
$d_{\ell }^{-2j-{\alpha }}{i\partial \bar \partial }\underline{\tilde {\chi }_{\ell }^{*}}$
converge in
$C_{\mathrm { loc}}^{2j+{\beta },j+{\beta }/2}$
for all
$0<{\beta }<{\alpha }$
to a function
$u_\infty \in C^{2j+{\alpha },j+{\alpha }/2}_{\mathrm {loc}}$
and to a
$(1,1)$
form
$\eta _\infty \in C^{2j+{\alpha },j+{\alpha }/2}_{\mathrm {loc}}$
, respectively, on
$\mathbb {C}^m\times (-\infty ,0]$
, and again
$\eta _\infty $
is weakly closed, and these satisfy (4.356) when
$j\geqslant 1$
. Using again (4.153), we also have that
$d_{\ell }^{-2j-{\alpha }}e^{-d_{\ell }^2\lambda _{\ell }^{-2}\tilde t}\partial _{\tilde t}\underline{\tilde \chi _{\ell }^{*} }\to u_\infty $
in
$C_{\mathrm {loc}}^{2j+{\beta },j+{\beta }/2}$
and
$d_{\ell }^{-2j-{\alpha }}e^{-d_{\ell }^2\lambda _{\ell }^{-2}\tilde t}{i\partial \bar \partial }\underline{\tilde \chi _{\ell }^{*} }=d_{\ell }^{-2j-{\alpha }}\tilde {\eta }^\diamondsuit _{\ell }\to \eta _\infty $
in
$C_{\mathrm {loc}}^{2j+{\beta },j+{\beta }/2}$
. From (4.160) and (4.203), we see that
$\tilde {\omega }^\sharp _{\ell }\to \omega _{\mathbb {C}^m}$
locally smoothly, and in particular, (as in [Reference Hein and Tosatti21, (4.345)]) this implies that

locally uniformly, and so

locally uniformly. Thanks to (4.170), we also have that for all
$-1\leqslant \iota \leqslant 2k$
,

so when
$j=0$
, from (4.110) (using also (4.160)) we see that

while when
$j\geqslant 1$
, from (4.132), we see that

We can then invoke (4.219), and use again the estimates (4.144) and (4.145), and see that

where we have set

First, we dispose of the case
$j=0$
. Again in this case from (4.268), we see that
$u_\infty $
is a constant in space-time, hence identically zero, and as in subcase B, we see that
$\eta _\infty $
is time-independent (using the argument in (4.364)). Then from (4.393), using also (4.157), we see that the quantity

is becoming asymptotically independent of the base and time directions, and so from (4.389), we conclude that

and since
$\eta _\infty $
vanishes at
$(0,0)$
, we actually have
$\operatorname {tr}_{\omega _{\mathbb {C}^m}}\eta _\infty =0$
. As in Subcase B, we can write
$\eta _\infty ={i\partial \bar \partial } v_\infty $
, where
$v_\infty \in C^{2+{\alpha }}_{\mathrm {loc}}(\mathbb {C}^m)$
is time-independent and
$\Delta _{\omega _{\mathbb {C}^m}}v_\infty =0$
, so
$v_\infty $
is smooth and
$|{i\partial \bar \partial } v_\infty |=O(|z|^{\alpha })$
. The Liouville Theorem applied to each component of
${i\partial \bar \partial } v_\infty $
then implies that
$\eta _\infty $
has constant coefficients; hence, it vanishes identically since its value at
$(0,0)$
vanishes. This means exactly that the RHS of (4.391) converges to zero, which gives us a contradiction when
$j=0$
.
In the remainder of this section, we thus assume that
$j\geqslant 1$
. Given any
$p,q\geqslant 0$
with
$p+2q=2j+2$
, and given
$v_1,\dots ,v_p$
tangent vectors to
$\mathbb {C}^m$
, from (4.393), recalling (4.157) and (4.181), we see that the quantity

is asymptotically independent of the base and time directions. Recall then that from (4.193), we have

where in the second equality, we could exchange
$\operatorname {tr}_{\tilde {\omega }^\sharp _{\ell }}({i\partial \bar \partial })_{\mathbf {ff}}$
with
$\operatorname {tr}_{\varepsilon _{\ell }^2\Theta _{\ell }^{*}\Psi _{\ell }^{*}\omega _{F}}({i\partial \bar \partial })_{\mathbf {ff}}$
with only a
$o(1)$
error since
$d_{\ell }^{-2j-{\alpha }}\|\mathfrak {D}^{2j}\tilde A_{\ell ,i,p,k}^{*}\|_{\infty ,\tilde {Q}_R,\tilde {g}_{\ell }(0)}\leqslant C\varepsilon _{\ell }^2$
by (4.170) and
$\|\tilde {\eta }^\dagger _{\ell }\|_{\infty ,\tilde {Q}_R,\tilde {g}_{\ell }(0)}=o(1)$
by (4.200). Similarly, from (4.186), we see that

while (4.388) implies that

and plugging (4.398), (4.399) and (4.400) into (4.397), we conclude that

is asymptotically independent of the base and time directions. Observe that the first line in (4.401) is a function pulled back from
$\mathbb {C}^m$
, while the second line has fiberwise average zero. Thus, taking the fiberwise average of (4.401), we see that

is approaching a (time-independent) constant locally uniformly on
$\mathbb {C}^m\times (-\infty ,0]$
. Recalling that
$d_{\ell }^{-2j-{\alpha }}\partial _{\tilde {t}}\left(e^{-d_{\ell }^2\lambda _{\ell }^{-2}\tilde t}\underline{\tilde {\chi }_{\ell }^{*}}\right)\to u_\infty $
in
$C^{2j+{\beta },j+{\beta }/2}_{\mathrm {loc}}$
and
$d_{\ell }^{-2j-{\alpha }}\Delta _{\omega _{\mathbb {C}^m}}\left(e^{-d_{\ell }^2\lambda _{\ell }^{-2}\tilde t}\underline{\tilde {\chi }_{\ell }^{*}}\right)\to \operatorname {tr}_{\omega _{\mathbb {C}^m}}{\eta _\infty }$
in
$C^{2j+{\beta },j+{\beta }/2}_{\mathrm {loc}}$
, passing to the limit in (4.402) shows that
$u_\infty -\operatorname {tr}_{\omega _{\mathbb {C}^m}}{\eta _\infty }$
is a parabolic polynomial on
$\mathbb {C}^m\times (-\infty ,0]$
of degree at most
$2j$
, and since by jet subtraction the parabolic
$2j$
-jets of
$u_\infty $
and
$\eta _\infty $
vanish at
$(0,0)$
, we conclude that

on
$\mathbb {C}^m\times (-\infty ,0]$
. Since
$j\geqslant 1$
, we can differentiate (4.403) with respect to t, and use (4.356) to see that

so
$u_\infty $
solves the heat equation on
$(-\infty ,0]\times \mathbb {C}^m$
, and
$|\mathfrak {D}^{\iota }u_\infty |=O(r^{2j+{\alpha }-\iota })$
for
$0\leqslant \iota \leqslant 2j$
by (4.153). By applying Liouville Theorem for ancient heat equation to
$\mathfrak {D}^{2j}u_\infty $
, we see that
$u_\infty $
must be a space-time polynomial of degree at most
$2j$
, and hence,
$u_\infty \equiv 0$
since its parabolic
$2j$
-jet at
$(0,0)$
vanishes. Going back to (4.356), it then follows that
$\eta _\infty $
is time-independent, and it is clearly of the form
$\eta _\infty ={i\partial \bar \partial } v_\infty $
for some time-independent function
$v_\infty \in C^{2j+2+{\beta }}_{\mathrm {loc}}(\mathbb {C}^m)$
. From (4.403), we see that
$v_\infty $
is harmonic. Since
$|{\mathbb {D}}^{2j}\eta _\infty |=O(|z|^{\alpha })$
, the Liouville Theorem in [Reference Hein and Tosatti20, Proposition 3.12] then shows that the coefficients of
$\eta _\infty $
are polynomials of degree at most
$2j$
, and since these coefficients have vanishing
$2j$
-jet at the origin, this implies that
$\eta _\infty $
vanishes identically. This kills the first two terms on the RHS of (4.392).
At this point, we return to (4.401), and subtracting its fiber average, we see that

is asymptotically independent of the base and time directions. Let us then define a function
$u_{\ell }$
on
$\tilde {Q}_{Rd_{\ell }^{-1}}$
by

so that (4.405) is equivalent to the statement that

is asymptotically independent of the base and time directions.
Our next goal is to show that
$u_{\ell }$
itself is asymptotically constant in the base and time directions. Fix any two points
$z,z'\in \mathbb {C}^m$
and times
$t,t'\in (-\infty ,0]$
and consider

so that in
$L^\infty _{\mathrm {loc}}(Y)$
, we have

and we can schematically estimate the difference of Laplacians by

and since

we can use (4.163) with
$\iota =2j$
to estimate
$|({i\partial \bar \partial } u_{\ell })_{\mathbf {ff}}|\leqslant C\varepsilon _{\ell }^2$
, and so (4.410) can be estimated by

so that (4.409) shows that

locally uniformly. So the functions
$v_{\ell }$
are approximate solutions of a fiberwise heat equation, with time parameter
$\tilde {t}\in (-\infty ,0]$
. We then employ another energy argument on the given fiber
$\{z\}\times Y$
. First, observe that (4.163) with
$\iota =2j$
implies

For notational convenience, we denote by
$\omega _Y:=\Theta _{\ell }^{*}\Psi _{\ell }^{*}\omega _F|_{\{z\}\times Y}$
and consider then the energy

which satisfies
$E_{\ell }(\tilde {t})\leqslant C$
for all
$\tilde {t}\leqslant 0$
, and denote by

where
$R>0$
is fixed so that
$(z,t),(z',t')\in Q_{2R}$
. Thanks to (4.413), we have
$\sigma _{\ell }\to 0$
. We can then compute, using the Poincaré inequality on
$(Y,\omega _Y)$
(recall that
$v_{\ell }$
has fiberwise average zero),

and using the Young inequality
$\sigma _{\ell } E_{\ell }(\tilde {t})^{\frac {1}{2}}\leqslant C^{-1}E_{\ell }(\tilde {t})+C\sigma _{\ell }^2,$
we can bound

We will compare
$E_{\ell }$
with the real-variable function
$F_{\ell }$
which solves the ODE

which is given explicitly by

and if we choose A large enough so that
$E_{\ell }(\tilde {t})\leqslant A$
for all
$\tilde {t}\leqslant 0$
(which is possible, as shown above), then we conclude that for all
$\tilde {t}\in [-R^2,0]$
, we have

and hence in particular
$E_{\ell }(0)\to 0$
as
$\ell \to +\infty $
. This means that the functions
$\varepsilon _{\ell }^{-2}v_{\ell }(\cdot , 0)$
, which are defined on Y and are uniformly bounded by (4.414), converge to
$0$
in
$L^2(Y,\omega _Y^n)$
. Recalling the definitions (4.406), (4.408), this means that the function of
$\tilde {y}$
given by

converges to zero in
$L^2(Y,\omega _Y^n)$
. However, we have

where in the second equality we used that since
$G_{\ell ,i,r,k}$
is a smooth function on the total space with fiberwise average zero, the function
$(\Delta _{\omega _{F}|_{ \{\cdot \}\times Y}})^{-1}G_{\ell ,i,r,k}$
is also smooth on the total space (by standard Schauder theory fiber-by-fiber, with continuous dependence on the base variables). Using (4.423) together with the bound
$d_{\ell }^{-2j-{\alpha }}\varepsilon _{\ell }^{-2}|{\mathbb {D}}^p_{v_1\cdots v_p}\partial _t^q\tilde A_{\ell ,i,r,k}^{*}|\leqslant C$
, which comes from (4.163) with
$\iota =2j$
, we see that

so from (4.422), we see that

also converges to zero in
$L^2(Y,\omega _Y^n)$
. Now, the functions
$\{(\Delta _{\Theta _{\ell }^{*}\Psi _{\ell }^{*}\omega _{F}|_{ \{z\}\times Y}})^{-1} \tilde G_{\ell ,i,r,k}\}_{i,r}$
are fiberwise linearly independent (since so are
$\{\tilde G_{\ell ,i,r,k}\}_{i,r}$
), and the function in (4.425) along the fiber
$\{z\}\times Y$
is expressed as a linear combination of these functions with coefficients (which are constants on Y) given by

Since the
$L^2$
norm of (4.425) is going to zero, these coefficients must be going to zero too, which means that the functions

are approximately constant (in space and time) as
$\ell \to +\infty $
. This kills the last term on the RHS of (4.392) and gives the final contradiction, thus completing the proof of Subcase C and of Theorem 4.2.
5 Proof of the main theorem
In this final section, we give the proof of our main Theorem 1.3; namely, we prove Conjectures 1.1 and 1.2. The asymptotic expansion in Theorem 4.2 will play a crucial role.
5.1 Higher order estimates
To prove the higher order estimates in Conjecture 1.1 from the expansion in Theorem 4.2, we follow the arguments in [Reference Hein and Tosatti21, Proof of Theorem A], but since our estimate (4.16) is weaker than the corresponding [Reference Hein and Tosatti21, (4.12)], we will have to deal with some new difficulties. As explained in the Introduction, in this section we work locally on the base (away from the image of the singular fibers), and the Kähler-Ricci flow that we analyze thus lives on
$B\times Y\times [0,+\infty )$
(with a non-product complex structure) for some Euclidean ball
$B\subset \mathbb {C}^m$
. For brevity, in this section, all norms and seminorms will be tacitly taken on
$B\times Y\times [t-1,t]$
(or
$B\times [t-1,t]$
for objects that live on the base), without making this explicit in the notation. The ball B and the interval
$[t-1,t]$
will also be shrunk slightly every time we use interpolation.
Given an even integer
$k\geqslant 2$
, we want to show that
$\omega ^{\bullet }(t)$
is uniformly bounded in
$C^k(g_X)$
. Applying Theorem 4.2 with
$j:=\frac {k}{2}$
, up to shrinking B, we can write

and in this decomposition,
$\omega ^\natural (t)$
is clearly bounded in
$C^k(g_X)$
,
$\gamma _0(t)$
has a similar bound by (4.12),
$\eta _{\frac {k}{2},k}(t)$
is bounded in the shrinking
$C^k$
norm by (4.9) (hence in the regular
$C^k(g_X)$
norm by [Reference Hein and Tosatti21, Lemma 2.6]), so we are left with dealing with the terms
$\gamma _{i,k}(t),1\leqslant i\leqslant \frac {k}{2}$
. By definition and using (3.1), we have

and so, as in [Reference Hein and Tosatti21, (5.10)], for
$0\leqslant q\leqslant k$
,

and using the fixed metric
$g_X$
, we can estimate
$|{\mathbb {D}}^{i_1}\Phi _{\iota ,r}(G_{i,p,k})|_{g_X}\leqslant C$
and
$|({\mathbb {D}}^{q+1-s}J)|_{g_X}\leqslant C$
, while from (4.16), we see that
$|\mathfrak {D}^{i_2+\iota } A_{i,p,k}|=o(1)$
when
$i_2+\iota \leqslant k+2$
and from (4.17) that
$|\mathfrak {D}^{i_2+\iota } A_{i,p,k}|=o(e^{(i_2+\iota -k-2)\frac {t}{2}})$
when
$k+2<i_2+\iota \leqslant k+2+2k$
, and so

since
$i_2\leqslant q+2$
and
$q\leqslant k$
. This completes the proof of Conjecture 1.1.
5.2 Ricci curvature bounds
Next, we prove Conjecture 1.2 – namely, the Ricci curvature bound for
$\omega ^{\bullet }(t)$
on compact subsets of
$X\backslash f(S)$
, which in our setting translates to

The argument is similar to [Reference Hein and Tosatti21, Proof of Theorem B], but there are some crucial differences coming from the time evolution in the Monge-Ampère equation, and from the fact that the bounds in (4.16) are worse than those in [Reference Hein and Tosatti21, (4.12)].
We will use the expansion (4.8) with
$j=1$
and
$k\geqslant 4$
(arbitrary), and with
$\alpha $
close to
$1$
, and our first task is to improve the estimates (4.16), (4.17). These give us

and we can interpolate between
$|A_{1,p,k}|\leqslant e^{-(2+\alpha )\frac {t}{2}}$
and
$[\mathfrak {D}^2A_{1,p,k}]_{C^{\alpha }}\leqslant Ce^{-t}$
from (4.18) and get

so in particular, we have
$|\dot {A}_{1,p,k}|=o(e^{-t}).$
The next claim is that

Indeed, using the decomposition in (5.3),

and we bound the RHS of (5.9) by
$o(e^{-t})$
by considering the possible values of
$i_2+\iota \in \{0,\dots ,2k+2\}$
: if
$i_2+\iota =0,1,2$
, then (5.7) in particular gives
$|{\mathbb {D}}^{i_2+\iota } A_{1,p,k}|=o(e^{-t})$
, which is acceptable. If
$i_2+\iota =3,4$
, then necessarily
$\iota \geqslant 1$
and so
$r\geqslant 1$
, while (5.6) in particular gives
$|{\mathbb {D}}^{i_2+\iota } A_{1,p,k}|=o(1)$
so the RHS of (5.9) is again
$o(e^{-t})$
. And if
$i_2+\iota \geqslant 5$
, then we use (5.6) exactly as in (5.4) to bound the RHS of (5.9) by

since
$i_2\leqslant 2$
, which concludes the proof of (5.8). Next, we want to show that

where the
$o(e^{-2t})$
is in
$L^\infty _{\mathrm {loc}}(g_X)$
. Indeed from (5.2), we can write

and we can estimate each term as follows. For
$\iota \geqslant 4$
, we have
$|{\mathbb {D}}^\iota A_{1,p,k}|=o(e^{-(4-\iota )\frac {t}{2}})$
from (5.6), and so

since
$r\geqslant \frac {\iota }{2}$
. For
$\iota =3$
, we have
$r\geqslant 2$
and
$|{\mathbb {D}}^\iota A_{1,p,k}|=o(1)$
from (5.6), so the term is again
$o(e^{-2t})$
. For
$\iota =1,2$
, we have
$r\geqslant 1$
and
$|{\mathbb {D}}^\iota A_{1,p,k}|=o(e^{-t})$
from (5.7), so the term is again
$o(e^{-2t})$
. And for
$\iota =0$
, let us first look at the terms with
$r\geqslant 1$
. For these, we have
$|A_{1,p,k}|=O(e^{-(2+\alpha )\frac {t}{2}})$
, and so when multiplied by
$e^{-r t},r\geqslant 1$
, these terms are indeed
$o(e^{-2t})$
. So we are only left with the terms where
$\iota =r=0$
which equal

since
$\Phi _{0,0}(G)=(\Delta ^{\omega _F|_{\{\cdot \}\times Y}})^{-1}G$
by (3.2), thus proving (5.11). In particular, using the bound (5.6) in (5.11) gives

while tracing (5.11) fiberwise gives

Before we continue with the proof of (5.5), recall that from [Reference Fong and Zhang12, p.110] (see also [Reference Tosatti39, Lemma 5.13]) we know that

for all
$t\geqslant 0$
. The argument in [Reference Tosatti, Weinkove and Yang43, Lemma 3.1 (iv)] then allows one to deduce from this that

As a consequence of our asymptotic expansion, we can now improve both of these:
Proposition 5.1. On
$B\times Y$
, we have

Proof. Recall that by definition, we can write

and we want to bound the
$L^\infty $
norm of the two terms on the RHS of (5.20) and their time derivatives. For
$\psi _{1,k}$
, from (4.9), we have

so restricting this to a fiber
$\{z\}\times Y$
gives

and since
$\psi _{1,k}$
has fiberwise average zero, applying Moser iteration on
$\{z\}\times Y$
(with constant that does not depend on
$z\in B$
) gives

Arguing similarly for
$\dot {\psi }$
, which by (4.9) satisfies (since
$k\geqslant 4$
)

we get

As for the term
$\mathfrak {G}_{t,k}(A_{1,p,k},G_{1,p,k})$
, using again (5.2), we can write it as

and using (5.6) and (5.7), we can argue similarly to the proof of (5.8) and bound

by
$o(e^{-t})$
by considering the possible values of
$\iota \in \{0,\dots ,2k\}$
: if
$\iota =0$
, then (5.6) in particular gives
$|A_{1,p,k}|=o(e^{-t})$
, which is acceptable. If
$1\leqslant \iota \leqslant 4$
, then necessarily
$r\geqslant 1$
, while (5.6) in particular gives
$|{\mathbb {D}}^{\iota } A_{1,p,k}|=o(1)$
so the RHS of (5.27) is again
$o(e^{-t})$
. And if
$\iota \geqslant 5$
, then we use (5.6) to bound the RHS of (5.27) by

and so

Similarly,

and we argue as above to show that this is
$o(e^{-t})$
: if
$\iota =0$
, then (5.7) in particular gives
$|\dot {A}_{1,p,k}|=o(e^{-t})$
, which is acceptable. If
$1\leqslant \iota \leqslant 2$
, then
$r\geqslant 1$
, while (5.6) in particular gives
$|{\mathbb {D}}^{\iota } \dot {A}_{1,p,k}|=o(1)$
so the RHS of (5.30) is again
$o(e^{-t})$
. And if
$\iota \geqslant 3$
, then we use (5.6) to bound the RHS of (5.30) by

and so

Combining (5.20) with (5.23), (5.25), (5.29) and (5.32) we see that (5.19) holds.
Remark 5.2. If we use the stronger bounds (5.55), (5.56) for derivatives of
$A_{1,p,k}$
, which will be established below, we can then repeat the proof of Proposition 5.1, and we get the better bounds

where
$\gamma =\frac {\alpha }{4+\alpha }>0.$
Now that (5.19) is established, we can use it to prove the next claim, which is the analog of [Reference Hein and Tosatti21, (5.26)]:
Proposition 5.3. We have

Proof. First, observe that thanks to (4.9), we have

which implies that

and combining this with (5.8) and (5.15), we get

The Monge-Ampère equation (4.1) that describes the flow can be written as

and multiplying this by
$\frac {e^{nt}}{\binom {m+n}{n}}$
, expanding it out, and using that
${i\partial \bar \partial } \underline{\varphi }(t)$
is small in
$C^2$
by (4.12), as well as (5.36), (5.37), we see that

where the error terms
$o(e^{-t})$
are in
$L^\infty _{\mathrm {loc}}(g_X)$
. We used here that
$\alpha>\frac {2}{3}$
, so that
$|(e^t\gamma _{1,k})_{\mathbf {ff}}^p|=o(e^{-t})$
for
$p\geqslant 3$
by (5.15). As in [Reference Hein and Tosatti21, (5.29)], we define a function
$\mathcal {S}$
by

so that dividing (5.39) by the volume form
$\omega _{\mathrm {can}}^m\wedge \omega _F^n$
and multiplying by
$e^{-t}$
gives

and subtracting from (5.41) its fiber average (the two terms with
$\gamma _{1,k}$
have fiberwise average zero since
$\gamma _{1,k}$
is
$\partial \overline {\partial }$
-exact) and using (5.16) gives

To bound the LHS of (5.42), use the Taylor expansion of the exponential, together with Lemma 4.1 (ii) and (5.19) to bound

while on the RHS of (5.42), we can bound

by (5.15). Thus, going back to (5.42), we learn that

and taking the fiberwise
$L^2$
inner product with each
$G_{1,p,k}$
, we see that

which improves over the bound
$A_{1,p,k}=O(e^{-(1+\frac {\alpha }{2})t})$
from (5.6). We can then go back to (5.11) and see that

which allows us to improve (5.37) to

and so as long as we choose
$\alpha>\frac {1}{2}$
, we see that

and so in (5.42), we have

and returning to (5.42) again, we see that

and again taking the fiberwise inner product with each
$G_{1,p,k}$
concludes the proof of (5.34).
Next, observe that from (4.9) we have in particular

We want to show a similar estimate for
$\gamma _{1,k}$
which is only slightly worse:
Proposition 5.4. We have

and also

where
$\gamma =\frac {\alpha }{4+\alpha }>0$
.
Proof. The first step is to use (5.34) to improve the estimates in (5.6), by interpolating it with
$[\mathfrak {D}^4 A_{1,p,k}]_{C^{\alpha }}\leqslant C$
from (4.18) and get

where
$\gamma =\frac {\alpha }{4+\alpha }>0$
. As for higher order derivatives of
$A_{1,p,k}$
, given
$1\leqslant \iota \leqslant 2k$
, we interpolate between
$|\mathfrak {D}^4A_{1,p,k}|\leqslant Ce^{-\frac {\gamma }{2}t}$
from (5.55) and
$[\mathfrak {D}^{4+\iota }A_{1,p,k}]_{C^{\alpha }}\leqslant Ce^{\frac {\iota }{2}t}$
from (4.18) and letting
$i=4+\iota $
, we get

Finally, using (5.55) and (5.56), we can bound
$|\mathfrak {D}^i\gamma _{1,k}|_{g(t)}, 0\leqslant i\leqslant 2,$
by going back to (5.3) and bounding

and using also
$|{\mathbb {D}}^{i_1}\Phi _{\iota ,r}(G_{1,p,k})|_{g(t)}\leqslant Ce^{i_1\frac {t}{2}}$
,
$|{\mathbb {D}}^\iota J|_{g(t)}\leqslant Ce^{\iota \frac {t}{2}},$
we get

for
$0\leqslant i\leqslant 2$
, which proves (5.53). As for (5.54), using (5.3) again, we can bound

After these preparations, we can finally give the proof of the Ricci curvature bound (5.5). For this, we take
${i\partial \bar \partial }\log $
of the Monge-Ampère equation (4.1) and get

where we used the known relation
$-{i\partial \bar \partial }\log (\omega _{\mathrm { can}}^m\wedge \omega _F^n)=-\omega _{\mathrm {can}}$
(see, for example, [Reference Tosatti39, Proposition 5.9]). Since
$\omega ^{\bullet }(t)$
is uniformly equivalent to
$\omega ^\natural (t)$
(by Lemma 4.1 (i)), it suffices to bound
$|\textrm {Ric}(\omega ^{\bullet }(t))|_{g^\natural (t)}$
on compact subsets of
$X\backslash S$
. We can then work on
$B\times Y$
as before, where we may assume that we have the expansion (4.8) with
$j=1$
and
$k\geqslant 4$
,

Using (4.12), (5.52) and (5.53), we thus see that

Taking a time derivative of (5.61), we obtain

and using (4.12), (5.52), (5.54), we see that

Thus, going back to (5.60), we can write

where thanks to (5.62) and (5.64), we have

and since we clearly have
$|\omega _{\mathrm {can}}|_{g^\natural (t)}\leqslant C,$
the Ricci bound
$|\textrm {Ric}(\omega ^{\bullet }(t))|_{g^\natural (t)}\leqslant C$
follows.
Observe that (5.65) and (5.66) give us very detailed information about the Ricci curvature of
$\omega ^{\bullet }(t)$
(on compact subsets of
$X\backslash S$
), showing that it is asymptotic to
$-\omega _{\mathrm {can}}$
in a strong sense.
Remark 5.5. Since
$\omega _{\mathrm {can}}$
is pulled back from the base, and since
$\omega ^{\bullet }(t)\to \omega _{\mathrm {can}}$
locally uniformly (from Lemma 4.1 (i), (iii)), it follows that

locally uniformly on
$X\backslash S$
. Thus, taking the trace of (5.65) with respect to
$\omega ^{\bullet }(t)$
, we see that

in
$C^0_{\mathrm {loc}}(X\backslash S)$
, which recovers the main theorem of [Reference Jian22].
Remark 5.6. Continuing the above arguments along the lines of [Reference Hein and Tosatti21, (5.37)–(5.86)], one can also identify the first nontrivial term in the expansion (4.8) of
$\omega ^{\bullet }(t)$
, whose shape is identical to the one in the elliptic setting; see [Reference Hein and Tosatti21, Theorem B]. For the sake of brevity, we leave the straightforward proof to the interested reader.
Acknowledgements
Part of this work was carried out during visits of the second-named author to the Courant Institute of Mathematical Sciences, and of the third-named author to the Center of Mathematical Sciences and Applications at Harvard University, which we would like to thank for the hospitality. We thank J. Cheng, C. Hacon, J. Lott, S. Karigiannis and M. Păun for discussions about Remarks 1.4 and 1.5, and the referee for helpful comments on a previous version.
Competing interest
The authors have no competing interests to declare.
Funding statement
The first-named author was partially supported by the German Research Foundation (DFG) under Germany’s Excellence Strategy EXC 2044-390685587 ‘Mathematics Münster: Dynamics-Geometry-Structure’ and by the CRC 1442 ‘Geometry: Deformations and Rigidity’ of the DFG. The second-named author was partially supported by Hong Kong RGC grant (Early Career Scheme) of Hong Kong No. 24304222, No. 14300623, and an NSFC grant No. 12222122. The third-named author was partially supported by NSF grants DMS-2231783 and DMS-2404599.