STABILITY OF THE ALMOST HERMITIAN CURVATURE FLOW

Similar documents
Riemannian Curvature Functionals: Lecture III

BERGMAN KERNEL ON COMPACT KÄHLER MANIFOLDS

Hyperkähler geometry lecture 3

As always, the story begins with Riemann surfaces or just (real) surfaces. (As we have already noted, these are nearly the same thing).

On the Convergence of a Modified Kähler-Ricci Flow. 1 Introduction. Yuan Yuan

MATH 263: PROBLEM SET 1: BUNDLES, SHEAVES AND HODGE THEORY

Metrics and Holonomy

Introduction to Extremal metrics

Topic: First Chern classes of Kähler manifolds Mitchell Faulk Last updated: April 23, 2016

RICCI SOLITONS ON COMPACT KAHLER SURFACES. Thomas Ivey

HYPERKÄHLER MANIFOLDS

Formal Groups. Niki Myrto Mavraki

arxiv:alg-geom/ v1 29 Jul 1993

Geometry of Ricci Solitons

SUMMARY OF THE KÄHLER MASS PAPER

Cohomology of the Mumford Quotient

Gradient Estimates and Sobolev Inequality

A Joint Adventure in Sasakian and Kähler Geometry

The Calabi Conjecture

1. Geometry of the unit tangent bundle

arxiv:math/ v1 [math.dg] 19 Nov 2004

Symplectic critical surfaces in Kähler surfaces

Analysis in weighted spaces : preliminary version

Topics in Geometry: Mirror Symmetry

Critical metrics on connected sums of Einstein four-manifolds

LECTURE 5: COMPLEX AND KÄHLER MANIFOLDS

Published as: J. Geom. Phys. 10 (1993)

Infinitesimal Einstein Deformations. Kähler Manifolds

The oblique derivative problem for general elliptic systems in Lipschitz domains

MILNOR SEMINAR: DIFFERENTIAL FORMS AND CHERN CLASSES

Geometry and the Kato square root problem

ON THE REGULARITY OF SAMPLE PATHS OF SUB-ELLIPTIC DIFFUSIONS ON MANIFOLDS

THE HODGE DECOMPOSITION

Mirror Symmetry: Introduction to the B Model

Homogeneous para-kähler Einstein manifolds. Dmitri V. Alekseevsky

From holonomy reductions of Cartan geometries to geometric compactifications

Flat complex connections with torsion of type (1, 1)

EXISTENCE THEORY FOR HARMONIC METRICS

APPROXIMATE YANG MILLS HIGGS METRICS ON FLAT HIGGS BUNDLES OVER AN AFFINE MANIFOLD. 1. Introduction

LECTURE 6: J-HOLOMORPHIC CURVES AND APPLICATIONS

Geometria Simplettica e metriche Hermitiane speciali

NOTE ON ASYMPTOTICALLY CONICAL EXPANDING RICCI SOLITONS

INSTANTON MODULI AND COMPACTIFICATION MATTHEW MAHOWALD

Geometry and the Kato square root problem

K-stability and Kähler metrics, I

Complex manifolds, Kahler metrics, differential and harmonic forms

AFFINE HERMITIAN-EINSTEIN METRICS

Some brief notes on the Kodaira Vanishing Theorem

GENERALIZED VECTOR CROSS PRODUCTS AND KILLING FORMS ON NEGATIVELY CURVED MANIFOLDS. 1. Introduction

Hard Lefschetz Theorem for Vaisman manifolds

CALIBRATED FIBRATIONS ON NONCOMPACT MANIFOLDS VIA GROUP ACTIONS

DIFFERENTIAL FORMS, SPINORS AND BOUNDED CURVATURE COLLAPSE. John Lott. University of Michigan. November, 2000

Compact 8-manifolds with holonomy Spin(7)

Differential Geometry MTG 6257 Spring 2018 Problem Set 4 Due-date: Wednesday, 4/25/18

1. Plurisubharmonic functions and currents The first part of the homework studies some properties of PSH functions.

THE MODULI SPACE OF TRANSVERSE CALABI YAU STRUCTURES ON FOLIATED MANIFOLDS

Geometry of the Calabi-Yau Moduli

LECTURE 10: THE PARALLEL TRANSPORT

ERRATUM TO AFFINE MANIFOLDS, SYZ GEOMETRY AND THE Y VERTEX

arxiv: v1 [math.dg] 11 Jan 2009

Poisson Equation on Closed Manifolds

Rigidity and Non-rigidity Results on the Sphere

A Bird Eye s view: recent update to Extremal metrics

NOTES ON LINEAR ODES

LECTURE 3 Functional spaces on manifolds

Multi-moment maps. CP 3 Journal Club. Thomas Bruun Madsen. 20th November 2009

THE NEWLANDER-NIRENBERG THEOREM. GL. The frame bundle F GL is given by x M Fx

LECTURE 2: SYMPLECTIC VECTOR BUNDLES

Riemannian Curvature Functionals: Lecture I

On the BCOV Conjecture

UNIVERSAL UNFOLDINGS OF LAURENT POLYNOMIALS AND TT STRUCTURES. Claude Sabbah

arxiv:math/ v1 [math.dg] 29 Sep 1998

Scalar curvature and the Thurston norm

NOTES ON NEWLANDER-NIRENBERG THEOREM XU WANG

Elliptic Regularity. Throughout we assume all vector bundles are smooth bundles with metrics over a Riemannian manifold X n.

The Strominger Yau Zaslow conjecture

LECTURE 8: THE SECTIONAL AND RICCI CURVATURES

4.7 The Levi-Civita connection and parallel transport

Constructing compact 8-manifolds with holonomy Spin(7)

LAGRANGIAN SUBMANIFOLDS IN STRICT NEARLY KÄHLER 6-MANIFOLDS

Hamiltonian Systems of Negative Curvature are Hyperbolic

Hermitian vs. Riemannian Geometry

Invariant Nonholonomic Riemannian Structures on Three-Dimensional Lie Groups

Radial balanced metrics on the unit disk

SYMPLECTIC GEOMETRY: LECTURE 5

MATH 205C: STATIONARY PHASE LEMMA

Geometry of Shrinking Ricci Solitons

Takao Akahori. z i In this paper, if f is a homogeneous polynomial, the correspondence between the Kodaira-Spencer class and C[z 1,...

THE UNIFORMISATION THEOREM OF RIEMANN SURFACES

Almost-Kähler Geometry. John Armstrong

Uniform K-stability of pairs

Compactness of Symplectic critical surfaces

A brief introduction to Semi-Riemannian geometry and general relativity. Hans Ringström

Conformal Killing forms on Riemannian manifolds

LAGRANGIAN HOMOLOGY CLASSES WITHOUT REGULAR MINIMIZERS

The spectral action for Dirac operators with torsion

η = (e 1 (e 2 φ)) # = e 3

LECTURE 9: MOVING FRAMES IN THE NONHOMOGENOUS CASE: FRAME BUNDLES. 1. Introduction

MATH 4030 Differential Geometry Lecture Notes Part 4 last revised on December 4, Elementary tensor calculus

Orientation transport

Transcription:

STABILITY OF THE ALOST HERITIAN CURVATURE FLOW DANIEL J. SITH Abstract. The almost Hermitian curvature flow was introduced in [9] by Streets and Tian in order to study almost Hermitian structures, with a particular interest in symplectic structures. This flow is given by a diffusion-reaction equation. Hence it is natural to ask the following: which almost Hermitian structures are dynamically stable? An almost Hermitian structure (ω, J) is dynamically stable if it is a fixed point of the flow and there exists a neighborhood N of (ω, J) such that for any almost Hermitian structure (ω 0, J 0) N the solution of the almost Hermitian curvature flow starting at (ω 0, J 0) exists for all time and converges to a fixed point of the flow. We prove that on a closed Kähler-Einstein manifold (, ω, J) such that either c 1(J) < 0 or (, ω, J) is a Calabi- Yau manifold, then the Kähler-Einstein structure (ω, J) is dynamically stable. 1. Introduction Let (, J, g) be a closed almost complex manifold such that J is compatible with the Riemannian metric g, that is for any vector fields X and Y we have g(x, Y ) = g(j(x), J(Y )). To the metric g we associate the -form ω defined by ω(x, Y ) = g(j(x), Y ). We call such a pair (ω, J) an almost Hermitian structure. The Ricci flow has proven to be a successful tool in studying the Riemannian geometry of manifolds. Therefore it is natural to attempt to use a parabolic flow to understand the almost Hermitian geometry of almost complex manifolds. However, the Ricci flow does not, in general, preserve the set of almost Hermitian structures. In [9], Streets and Tian introduce the almost Hermitian curvature flow (AHCF), which is a weakly-parabolic flow on the space of almost Hermitian structures. The almost Hermitian curvature flow is a coupled flow of metrics and almost complex structures. It is written t ω = S + H + Q (1.1) J = K + H, t where (ω(0), J(0)) = (ω 0, J 0 ). The highest order term in the evolution of ω(t), denoted S is a Ricci-type curvature. In particular, S ij = ω kl Ω klij and Ω is the curvature of the almost-chern connection. That is, is the unique connection satisfying ω = 0, J = 0 and T 1,1 = 0. T 1,1 is the (1, 1) component of the torsion of. Q is any (1, 1) form that is quadratic in the torsion of. Kj i = ωkl k Nlj i where N is the Nijenhuis tensor with respect to J. H is any endomorphism of T that is quadratic in N and skew commutes with J. The term H(X, Y ) = 1 [ ] ω(( K + H)(X), J(Y )) + ω(j(x), ( K + H)(Y )) is required in order to maintain the compatibility of ω t with J t. In Theorem 1.1 of [9], Streets and Tian prove short-time existence and uniqueness of the flow starting at an Supported in part by Research Training in Geometry and Topology grant DS 073908. 1

DANIEL J. SITH almost Hermitian structure (ω 0, J 0 ). Notice that the generality with which the tensors Q and H are defined implies that (1.1) is in fact a family of geometric flows. The almost Hermitian curvature flow generalizes Kähler Ricci flow in the sense that if the initial structure (ω 0, J 0 ) is Kähler, then the evolution of (ω(t), J(t)) by AHCF coincides with Kähler Ricci flow. oreover AHCF generalizes a couple of geometric flows introduced by Streets and Tian in [11] and [9]. In [11], Streets and Tian construct a parabolic flow on the space of Hermitian metrics, called Hermitian curvature flow (HCF). If J 0 is a complex structure then it follows immediately that under AHCF (1.1), J(t) J 0 and hence H(t) 0. In this case, the appropriate choice of the quadratic torsion term Q yields HCF. It is noteworthy that in addition to HCF, a number of other geometric flows of Hermitian metrics have recently been introduced (e.g. [10], [4], [14], [15], [7]). In [9] Streets and Tian introduce a geometric flow of symplectic structures, called symplectic curvature flow (SCF). In particular, given an almost Hermitian structure (ω 0, J 0 ) such that dω 0 = 0, SCF is a coupled flow (ω(t), J(t)) such that ω(t) is a closed -form as long as the flow exists. Also in [9] Streets and Tian show that the appropriate choices of Q and H in AHCF (1.1) give SCF. Hence we see that AHCF is a very general family of geometric flows. Associated to AHCF is the volume-normalized version of the flow (VNAHCF), the volume-normalized version is the one with which we will work. One natural question to ask is: does admit a Kähler-Einstein structure? If so, is it detected by VNAHCF? The main result of the paper is the following: Theorem 1.1. Let ( n, ω, J) be a closed complex manifold with ( ω, J) a Kähler-Einstein structure such that either c 1 ( J) < 0 or (, ω, J) is a Calabi-Yau manifold. Then there exists ɛ > 0 such that if (ω(0), J(0)) is an almost Hermitian structure with (ω(0) ω, J(0) J) C < ɛ, then the solution to the volume normalized AHCF starting at (ω(0), J(0)) exists for all time and converges exponentially to a Kähler-Einstein structure (ω KE, J KE ). Remark 1.. Theorem 1.1 gives evidence that the almost Hermitian curvature flow reflects the underlying almost Hermitian geometry of. Remark 1.3. In this paper we define a Calabi-Yau manifold (, ω, J) to be a compact Kähler-Einstein manifold with trivial canonical bundle. Remark 1.4. In the case when c 1 ( J) < 0, the Kähler-Einstein structure that the flow starts close to is the same one that the flow converges to, in other words ( ω, J) = (ω KE, J KE ). This is proved in Theorem 3.1. In the Calabi-Yau case we cannot guarantee that ( ω, J) and (ω KE, J KE ) are the same Calabi-Yau structure. The notion of stability in Theorem 1.1 is often referred to as dynamic stability. Dynamic stability has also been studied in the case of the Hermitian curvature flow by Streets and Tian ([11]) and for the Ricci flow by Sesum ([8]) and by Guenther, Isenberg, and Knopf ([5]). The first step in proving Theorem 1.1 is to show that Kähler-Einstein structures behave like sinks of the linear flow associated to VNAHCF, this is done in Section. Also in Section, we derive parabolic estimates for the VNAHCF (see Theorem.8). Next, in Section 3 we prove Theorem 1.1 in the case when c 1 ( J) < 0. Finally, in the last section we complete the proof of Theorem 1.1 by showing how to find a Kähler-Einstein structure (ω KE, J KE ) to which the flow exponentially converges in the Calabi-Yau case.

STABILITY OF THE ALOST HERITIAN CURVATURE FLOW 3 Acknowledgements. The author would like to thank his advisor Jon Wolfson for all of his help and suggestions on this project. The author would also like to thank Tom Parker for suggesting corrections to earlier versions of this paper.. Linear Stability and Parabolic Estimates To prove theorem 1.1 we first show that, on a linear level, any Kähler-Einstein structure ( ω, J) is a degenerate sink with respect to the almost Hermitian curvature flow, meaning that the linear operator associated to the non-linear flow is negative semi-definite at Kähler-Einstein structures. For notation sake write the volume-normalized almost Hermitian curvature flow (VNAHCF): (.1) t ω = F t J = G, with (ω(0), J(0)) = (ω 0, J 0 ). To the operator (F, G), we have the associated linear operator ( F, G). In particular, we consider a one-parameter family of compatible, unit volume, almost Hermitian structures (ω(a), J(a)) and ( F, G) =. a a=0 (F, G)(ω(a), J(a)). Similarly we write ( ω, J) =. a=0 (ω(a), J(a)). a Definition.1. An almost-hermitian structure (ω, J) is called static provided (F(ω, J), G(ω, J)) = 0. oreover, a static structure (ω, J) is linearly stable if the linearization L =. ( F, G) (ω,j) is negative semi-definite, that is L, L (g) 0. Notice that Kähler-Einstein structures are static under VNAHCF. Next, we prove Theorem.. Let (, J) be a closed complex manifold with c 1 ( J) 0, then any Kähler- Einstein structure ( ω, J) on is linearly stable. Proof. Employing the DeTurck trick, Streets and Tian prove that (F, G) is weakly-ellipticity (see Proposition 5.4 and 5.5 of [9]). Furthermore, computing the linearization of (F, G) at a Kähler-Einstein structure, in holomorphic coordinates with respect to J, the linearization is written (.) (.3) (.4) F ij = ω ij + ω kl R klij F ij = ω ij G i j = J i j + J pq R i jpq. Here is the L ( g) adjoint of and R denotes the Riemannian curvature of g. To show that ( ω, J) is linearly stable we have to deal with the fact that in (.) and (.4) the lower-order terms do not have a sign. To see that the linearized operator is negative semi-definite at Kähler-Einstein structures we use a couple of Weitzenböck-Bochner formulas (cf. []).

4 DANIEL J. SITH Lemma.3. Let α and β be (0,) and (1,1) forms respectively. If g is a Kähler-Einstein metric, then we have (.5) ( d α) ij = α ij + s n α ij (.6) ( d β) ij = β ij β kl R klij + s n β ij. Where d is the Hodge Laplacian with respect to g and s is the scalar curvature of g. In addition we will use another Weitzenböck-Bochner formula. Lemma.4. Given a T 1,0 (, ω, J) valued (0, 1)-form, φ and Kähler-Einstein metric g we have: (.7) ( φ) i j = φ i j φpq R i jpq + s n φi j. Where represents the complex laplacian g + g. Therefore using equations (.), (.3), (.5) and (.6), we have that (.8) F = d ω + s n ω. Similarly, using equations (.4) and (.7), we have (.9) Combining (.8) and (.9) we see that (.10) L( ω, J) = G = J + s J. n ( d ω + s n ω, J + s J n ). Finally since c 1 ( J) 0 implies that s 0, the theorem follows by integrating (.10) paired with ( ω, J) over. Notice that if c 1 ( J) < 0, then the scalar curvature of g is negative; that is s < 0. Hence from (.10) it follows that if c 1 ( J) < 0, then L is strictly negative definite with respect to L ( g). Let λ = min{ λ i : λ i is an eigenvalue of L}. Further let C denote the space of almost Hermitian structures modulo diffeomorphism. Therefore we have proved the following corollary. Corollary.5. Let (, ω, J) be a closed complex manifold such that ( ω, J) is a Kähler- Einstein structure and moreover c 1 ( J) < 0. Let ψ T ( ω, J) C, then L( ω, J) ψ, ψ L ( g) λ ψ L ( g). Corollary.5 will be crucial to proving Theorem 1.1 in the case when c 1 ( J) < 0 (see Section 3). Fix a Kähler-Einstein structure ( ω, J) and let (ω(t), J(t)) be a solution of the coupled system (.1). We will quantify the amount by which the solution deviates from ( ω, J) using ρ(t) = (ω(t) ω, J(t) J). Notice that ρ(t) Λ () End(T ). Throughout the paper we use the operator norm on End(T ). As noted in the proof of Theorem 1.1 in [9], the space of almost Hermitian structures C

STABILITY OF THE ALOST HERITIAN CURVATURE FLOW 5 is a non-linear manifold. In the following lemma we will show that ρ(t) / T ( ω, J) C, however ρ(t) can be estimated by an element of the tangent space T ( ω, J) C. Lemma.6. Fix t and let (ω(t), J(t)) be an almost Hermitian structure. Write ω(t) = ω + h(t) and J(t) = J + K(t), in other words ρ(t) = (h(t), K(t)). If ρ(t) C 0 < 1, then there exists ψ(t) T ( ω, C so that J) (.11) (.1) and (.13) ψ(t) C k ρ(t) C k, ρ(t) L ψ(t) L + C 1 ψ(t) L ρ(t) C k ψ(t) C k + C ψ(t) C k where C 1 and C depend on the L and C k norms of ρ(t) respectively. Proof. We will begin by studying the tangent space T ( ω, C. Let (ω J) s, J s ) denote a path of almost Hermitian structures such that (ω s, J s ) s=0 = ( ω, J) and let s s=0 (ω(s), J(s)) =. ( ω, J). Given vector fields X and Y, the compatibility condition is written: ω s (X, Y ) = ω s (J s (X), J s (Y )) and the almost complex condition is written: J s (X) = X. Hence the linearized compatibility and almost complex conditions are given by: (.14) (.15) ω(x, Y ) = ω( J(X), J(Y )) + ω( J(X), J(Y )) + ω( J(X), J(Y )) 0 = J J(X) + J J(X). From (.15) we see that the tangent space to the space of almost complex structures is given by endomorphisms that skew-commute with J. Equivalently, J can be viewed as a section of [ Λ 0,1 T 1,0] [ Λ 1,0 T 0,1]. Here we use J to decompose T = T 1,0 T 0,1. First we will show that the endomorphism K(t) can be estimated by an element of the tangent space to the space of almost complex structures at J. For the sake of notation we will often write K(t) = K. Using J we decompose K = K 1,0 0,1 + K0,1 + K1,0 + K0,1 1,0 where K1,0 0,1 : T 0,1 T 0,1, equivalently K 1,0 0,1 Λ0,1 T 1,0. Take ψ(t) T ( ω, J) C and write ψ(t) = (ψ 1 (t), ψ (t)) Λ () End(T ). We define (.16) ψ (t). = K 1,0 0,1 + K0,1 1,0. That is, ψ (t) is defined to be the projection of K onto [ Λ 0,1 T 1,0] [ Λ 1,0 T 0,1]. Next we will show that K 0,1 0,1 and K1,0 1,0 are quadratic in ψ (t). We will only prove this for K 0,1 0,1 since the same argument applies to K 1,0 1,0. Using that J(t) is an almost complex structure we see that K(t) satisfies: (.17) 0 = K J(X) + J K(X) + K (X).

6 DANIEL J. SITH For K acting on T 0,1 we will write K = K 0,1 have 0,1 + K1,0 0,1. Therefore using (.17), on T 0,1 we 0 = 1K 0,1 0,1 + K0,1 K0,1 + K1,0 0,1 K0,1 + K1,0 K1,0 0,1 + K0,1 1,0 K1,0 0,1 and so by type consideration, (.18) K 0,1 0,1 = 1 ( ) K 0,1 0,1 K0,1 + K0,1 1,0 K1,0 0,1. Notice that on T 0,1, K 0,1 1,0 K1,0 0,1 = ψ (t). Hence we are able to write K 0,1 0,1 in terms of ( K0,1) and a term that is quadratic in ψ (t). (. Next, consider the first term on the right-hand side of (.18), K0,1) We will use ( ) ( 4 (.18) to show that K 0,1 0,1 can be expressed as K0,1) plus terms which are quadratic (, in ψ (t). Plugging (.18) into each factor of K0,1) we see that ( K 0,1 0,1 ) 1 = 4 [ ( ) 4 ( ( K 0,1 0,1 + K0,1) ψ + ψ K 0,1 0,1 ) + ψ 4 which can be substituted into the term K 0,1 0,1 K0,1 in (.18). Iterating this process by successively plugging (.18) into the highest power term in K 0,1 0,1, we see that K0,1 can be expressed as a series. Notice that since ρ C 0 < 1 it follows that K 0,1 0,1 C 0 < 1, and so this series converges. Therefore (.19) K 0,1 0,1 = ψ (t) + [higher-power terms in ψ higher-power terms in K]. Next we will show that the two form h(t) can be estimated by an element of the tangent space to the space of compatible metrics. Notice that for vector fields X and Y, and so by (.14) ω(x, Y ) ω( J(X), J(Y )) = ω (,0)+(0,) (X, Y ), ω (,0)+(0,) (X, Y ) = ω( J(X), J(Y )) + ω( J(X), J(Y )). Using the compatibility of ω(t) and J(t), we see that h(t) = ω(t) ω and K(t) = J(t) J satisfy: (.0) h(x, Y ) = h( J(X), J(Y )) + ω(k(x), J(Y )) + ω( J(X), K(Y )) We define ψ 1 (t) as follows (.1) (.) (.3) + ω(k(x), K(Y )) + h(k(x), J(Y )) + h( J(X), K(Y )) + h(k(x), K(Y )). ψ (1,1) 1. = h (1,1) ψ (,0)+(0,) 1 (X, Y ). = ω(k(x), J(Y )) + ω( J(X), K(Y )) = ω(ψ (X), J(Y )) + ω( J(X), ψ (Y )). ],

STABILITY OF THE ALOST HERITIAN CURVATURE FLOW 7 The last equality follows from the definition of ψ and the fact that ω is of type (1, 1). Next we will show that h (,0)+(0,) ψ (,0)+(0,) 1 can be expressed as terms that are quadratic in ψ(t). Combining (.0), (.) and (.3) we have ( ) (.4) h (,0)+(0,) (X, Y ) ψ (,0)+(0,) 1 (X, Y ) = h(k(x), J(Y )) + h( J(X), K(Y )) ω(k(x), K(Y )) + h(k(x), K(Y )). As we proved above in (.16) and (.19), K can be written in terms of ψ and hence the terms in the second line of (.4) are higher-power in ψ. Next we consider the term h(k(x), J(Y )). Since the left-hand side of (.4) is a section of Λ (,0)+(0,), let X, Y T 0,1. So for X, Y T 0,1 we can write the components of h(k(x), J(Y )) as (.5) K 0,1 0,1 h(0,) + K 1,0 0,1 h(1,1). From (.16) and (.1) we see that the second term in (.5) is quadratic in ψ. By (.19) the first term is quadratic in ψ plus higher-power terms in ψ composed with higherpower terms in ρ. Notice that the same argument can be applied to h( J(X), K(Y )). Abusing notation we let ψ ψ denote terms which are quadratic in ψ plus terms that are higher-power in ψ composed with terms that are higher-power in ρ. Therefore we have (.6) h (,0)+(0,) (X, Y ) ψ (,0)+(0,) 1 (X, Y ) = ψ ψ. Notice that by the definition of ψ(t), given in (.16) (.1) and (.), the inequality ψ(t) C k ρ(t) C k follows immediately. Again using the definition of ψ(t) along with (.19) and (.6) we see that ρ(t) C k ψ(t) C k + C ψ(t) C k where C depends on the C k norm of ρ(t). Notice that (.1) follows analogously. In Theorem. we proved that the linearization of (F, G), denoted L, is negative semidefinite on T ( ω, J) C. The goal is to use the sign on L to prove exponential decay of ρ(t). However as we observed in the previous lemma, ρ(t) / T ( ω, J) C. To deal with this we will prove exponential decay of ψ(t) T ( ω, J) C and exponential decay of ρ(t) will follow by (.13). Next we show that ψ(t) evolves by a parabolic flow equation and moreover that we have estimates on the non-linear part of the flow. Lemma.7. Let L be the differential operator defined by (.10). Then for ψ(t) T ( ω, J) C defined by (.16) (.1) and (.3) we have (1) t ψ(t) = L(ψ(t)) + A(( ω, J), ψ(t)) () A(( ω, J), ψ(t)) C k C( ψ(t) C k ψ(t) C k + ψ(t) C k ) where C depends on the C k norm of ρ(t). Proof. To prove t ψ(t) = L(ψ(t)) + A(( ω, J), ψ(t)) we first study the evolution of ρ(t). Notice that since ( ω, J) is independent of t, (.7) t ρ(t) = t (ω(t), J(t)) = ( F(ω(t), J(t)), G(ω(t), J(t)) ).

8 DANIEL J. SITH Furthermore since ( ω, J) is a static structure, when we linearize (F, G) at ( ω, J) in the direction ψ(t), we have (.8) Hence from (.7) and (.8) it follows that (F, G) = L(ψ(t)) + A(( ω, J), ρ(t)). (.9) t ρ(t) = L(ψ(t)) + A(( ω, J), ρ(t)), where A represents the error in approximating (F, G) by the linearization L. As in [11] and [8] we have the following C k bounds on A: A C k C( ρ C k ρ C k + ρ C k ). Therefore by Lemma.6 we have the following bounds on A: (.30) A C k C( ψ C k ψ C k + ψ C k ). Next we will use the definition of ψ(t) and the evolution of ρ(t) to derive an evolution equation for ψ(t). By the definition of ψ (t) and the (1, 1) part of ψ 1 (t) (see (.16) and (.1) respectively) we have (.31) (.3) t ψ (t) = ( ) K 1,0 0,1 t + K0,1 1,0 t ψ(1,1) 1 (t) = t h(1,1) (t). For t ψ(,0)+(0,) 1 (t), it follows from (.4) that (.33) t ψ(,0)+(0,) 1 (t) = t h(,0)+(0,) (t) + ( ) t ρ ρ, since the (, 0) + (0, ) components of ψ 1 (t) and h(t) differ by terms that are quadratic in ρ(t) = (h(t), K(t)). Notice that by (.9) and (.30) we have tρ is second order in ψ and hence the final term in (.33) may be absorbed in the error estimate A. Therefore from (.31), (.3), (.33) and (.9) it follows that t ψ(t) = L(ψ(t)) + A(( ω, J), ψ(t)), where A is a different tensor than in (.9), but we still have A C k C( ψ C k ψ C k + ψ ). C k Next, we will show roughly that given any finite time T > 0, by starting the flow very close to ( ω, J), the solution (ω(t), J(t)) remains close to ( ω, J) on the interval [0, T ). This theorem is analogous to Lemma 8.5 in [11] and we will employ the same argument here. Theorem.8. Given T > 0, ɛ > 0 and an integer k 0, there exists ɛ = ɛ(t, ɛ, k) > 0 such that if ρ(0) C < ɛ then, (ω(t), J(t)) exists on [0, T ) and moreover ρ(t) C k < ɛ on [0, T ). Proof. First, for ɛ sufficiently small, work of Streets and Tian (see Theorem 1.1 in [9]) shows that there exists T > 0 such that the solution (ω(t), J(t)) exists on [0, T ) and moreover ρ(t) C k < ɛ on [0, T ). Suppose by way of contradiction that there exists a maximal T so that for all ɛ > 0 the solution exists and ρ(t) C k < ɛ on [0, T ) with T < T. Fix T < T. To derive a contradiction we will produce bounds on the C k norm

STABILITY OF THE ALOST HERITIAN CURVATURE FLOW 9 of ρ(t) on [0, T ] in terms of ɛ, independent of T. Recall from Lemma.6 that associated to ρ(t) we have ψ(t) T ( ω, J) C. In order to obtain C k estimates on ρ(t) in terms of ɛ we will produce C k bounds on ψ(t) and employ (.13). To this end, we study the evolution of ψ(t). Recall from Lemma.7 part (1) that (.34) t ψ(t) = L(ψ(t)) + A(( ω, J), ψ(t)) where L is negative semi-definite and A represents the error in approximating (F, G) by L. From part () of Lemma.7 we have (.35) A C k C( ψ C k ψ C k + ψ C k ). Notice that C depends on the C k norm of ρ(t) which we are assuming is bounded by ɛ for t [0, T ) [0, T ]. In the estimates that follow Rm will denote the curvature of the fixed metric g and will denote the Levi-Civita connection of g. oreover we will use the fact that is compact and hence there exists a constant C such that Rm C < C..1. L bounds of ψ in terms of ɛ. The linear stability of Kähler-Einstein structures will allow us to produce L bounds on ψ(t) in terms of ɛ which are independent of T. Indeed, for t [0, T ] by (.34) and using that L ( ω, J), L ( g) 0, we have 1 (.36) ψ g t dvol g = t ψ, ψ A ψ. Now, using the bound on A given in (.35), we see that A ψ = ψ ψ + ψ ψ. Using integration by parts on the second term yields (.37) A ψ ψ ψ. For t [0, T ], by assumption and (.11), ψ(t) C k < ɛ therefore ψ ψ Cɛ ψ. Hence combining (.36) and (.37) we have 1 (.38) ψ g t dvol g C 1 ɛ ψ g dvol g. Therefore for any t [0, T ], (.39) ψ(t) L e C 1ɛ T ψ 0 L ɛe C 1ɛ T... L 1, bounds of ψ in terms of ɛ. Given the L bounds above, we bootstrap to obtain higher-order bounds. Notice that linear stability was only used to start the bootstrapping process. Using (.), (.3), (.4) and (.34) we have 1 ψ g t dvol g = t ψ, ψ (.40) = ψ + Rm ψ + A, ψ.

10 DANIEL J. SITH Since is compact there exists a constant C such that Rm( g) C < C. oreover, we can bound the term associated with A as we did in (.37) and (.38) to get 1 (.41) ψ g t dvol g ψ dvol g + C 3 ψ dvol g. Integrating from 0 to T, we see that T ψ + 1 ψ( T ) 1 0 T ψ 0 + C 3 Now the L bounds from (.39) imply T (.4) ψ C 4 T e C 1ɛ T ψ 0 L C 4 T e C 1ɛ T ɛ. 0.3. L, bounds of ψ in terms of ɛ. Next, we use the L 1, bounds above to produce L, bounds. Similar to (.40), 1 ψ g (.43) t dvol g = ( ψ + Rm ψ + A), ψ. First consider the term ( ψ), ψ above. Commuting covariant derivatives and using integration by parts we get ( ψ), ψ (.44) = ψ + Rm ψ ψ. Next we obtain estimates on the term (Rm ψ), ψ = Rm ψ +Rm ψ, ψ from equation (.43). Since Rm C < C, we can use Young s Inequality, to show (Rm ψ), ψ C (.45) ψ + C ψ. Finally, we consider the final term in (.43), A ψ. Using the estimates on A from (.35), we have A ψ = ψ ψ + ψ ψ 3 ψ. Integration by parts on the last term yields A ψ = ψ ψ + ψ ψ and hence A, ψ C (.46) ψ + C 7 ɛ ψ since ψ C k < ɛ for t [0, T ]. Combining (.43), (.44), (.45), and (.46) we see that (.47) 1 t 0 ψ. ψ g dvol g ψ + C 5 ψ + C 6 ψ + C 7 ɛ ψ. Hence we choose ɛ small enough so that (.48) C 7 ɛ < 1.

STABILITY OF THE ALOST HERITIAN CURVATURE FLOW 11 Integrating (.47) from 0 to T we have (.49) T 0 T T ψ + ψ( T ) ψ 0 + C 5 ψ + C 6 ψ. 0 0 Therefore, using the L estimate from (.39) and the L 1, estimate from (.4) we have T (.50) ψ C 7 T e C 1ɛ T ψ 0 L C 1, 7 T e C 1ɛ T ɛ. 0 Notice that (.49) also gives bounds on ψ( T ) L. oreover by integrating (.47) from 0 to t for t [0, T ] these bounds hold not just at T but for any t [0, T ]. Hence we also have sup ψ [0, T L C 7 T e C 1ɛ T ɛ. ] Now since t ψ is second order in ψ, estimate (.50) also gives T 0 t ψ CT e C 1ɛ T ɛ. By a standard inductive argument, it follows that for any p we have both: T (.51) p ψ C(p)T e C 1ɛ T ψ 0 L C(p)T e C 1ɛ T ɛ p 1, and (.5) 0 sup p 1 ψ [0, T L C(p)T e C 1ɛ T ψ 0 L C(p)T e C 1ɛ T ɛ. p 1, ] Furthermore, since t ψ is second order in ψ, (.51) also implies that T 0 q t q r ψ Cɛ for any q, r > 0, where C is independent of T. Now use the Sobolev Embedding Theorem, with respect to g, to obtain C k bounds on ψ in terms of ɛ. And hence by (.13) we have C k bounds on ρ in terms of ɛ. In [9], Theorem 1.9, Streets and Tian prove that if there is a finite time singularity τ of the flow, then lim t τ sup{ Rm C 0, DT C 0, T C 0 } =. Here D denotes the Levi-Citia connection and Rm is the curvature of D. Therefore, the fact that the estimates above are independent of T implies that the solution exists on [0, T ]. Again, using the short-time existence result of Streets and Tian ([9]), the solution can be extended past time T. oreover, for ɛ sufficiently small, we maintain the C k estimates on ρ(t) past time T. This contradicts the maximality of T. 3. Dynamic Stability when c 1 ( J) < 0 In this section we prove that when c 1 ( J) < 0, the volume-normalized almost Hermitian curvature flow converges exponentially to the Kähler-Einstein structure ( ω, J). As above we let ρ(t) = (ω(t) ω, J(t) J). Theorem 3.1. Let ( n, ω, J) be a closed complex manifold where ( ω, J) is a Kähler- Einstein structure such that c 1 ( J) < 0. Given a positive integer k, there exists ɛ = ɛ(k) > 0 such that if (ω 0, J 0 ) is an almost Hermitian structure with ρ(0) C < ɛ, then the solution

1 DANIEL J. SITH to the volume-normalized AHCF (.1) exists for all time and converges exponentially in C k to ( ω, J). Proof. We prove Theorem 3.1 using two lemmas. As in Section, we have to deal with the non-linearity of the space of almost Hermitian structures. To prove Theorem 3.1 we will show that there exists ɛ such that if ρ(0) C < ɛ, then ψ(t) exponentially decays in C k. Finally employing (.13) exponential C k decay of ρ(t) will follow from exponential C k decay of ψ(t). Lemma 3.. Given δ > 0 and an integer k 0, there exists ɛ 1 = ɛ 1 (δ, k) > 0 such that if ρ(0) C < ɛ 1 then ψ(t) C k < δ for all t 0 and moreover ψ(t) L Ce λt for all t 0. Proof. As in Section, let λ = min{ λ i : λ i is an eigenvalue of L}. Further let ψ(t) T ( ω, J) C be the element of the tangent space, from Lemma.6, associated to ρ(t). Recall that from Corollary.5 that c 1 ( J) < 0 implies that (3.1) L( ω, ψ, ψ g J) λ ψ L ( g). And by (.34), (3.) 1 t ψ g dvol g = t ψ, ψ = Lψ + A, ψ. Then for any t for which ψ(t) C < δ, we can employ the bound (3.3) A, ψ C 1 δ ψ L derived in (.37) and (.38). Combining (3.1), (3.) and (3.3) yields Here we choose δ so that 1 t ψ g dvol g λ ψ L + C 1 δ ψ L. (3.4) C 1 δ < 1 λ. Integrating from 0 to t yields L exponential decay of ψ(t). In particular, (3.5) ψ(t) L e λt ψ(0) L for any t for which ψ(t) C < δ. Therefore to complete the lemma, we will show that given k, there exists ɛ 1 (k, δ) > 0 such that if ρ(0) C < ɛ 1, then ψ(t) C k < δ for all t [0, ). Notice again that from (.11) it follows that ρ(0) C < ɛ 1 implies ψ(0) C < ɛ 1. By Theorem.8 we know that given T > 0, k 0 and ɛ > 0, there exists ɛ > 0 such that ρ(0) C < ɛ implies that ψ(t) C k < ɛ on [0, T ). We apply Theorem.8 with ɛ = δ and δ sufficiently small so that (3.4) holds. Let ɛ denote a constant that is small enough so that if ψ(0) C < ɛ then ψ(t) C k < δ on [0, T ) and assume that (3.6) ψ(0) C < ɛ.

(3.7) STABILITY OF THE ALOST HERITIAN CURVATURE FLOW 13 Given T, let t 0 < t < T, then integrating (.41) from t 0 to t we have t t 0 ψ L 1 ψ(t 0) L + C 3 Furthermore since (3.5) holds on [0, T ), (3.8) t t 0 ψ(s) L t Therefore combining (3.7) and (3.8) yields (3.9) t t t 0 ψ(s) L. t 0 e λs ψ(0) L = 1 λ e λt 0 ψ(0) L. t 0 ψ L 1 ψ(t 0) L + C 3 λ e λt 0 ψ(0) L ( 1 + C ) 3 e λt 0 ψ(0) L. λ The last inequality is again by (3.5). The key observation here is that the L 1, estimate in (3.9) is independent of t. Next, to obtain a similar L, estimate we integrate (.47) from t 0 to t. (3.10) t t 0 ψ L + ψ(t) L ψ(t 0 ) L + C 5 t t 0 ψ(t) L + C 6 Bounding the last two terms of (3.10) using (3.8) and (3.9) yields t t 0 ψ L + ψ(t) L ψ(t 0 ) L + Ce λt 0 ψ(0) L. t t 0 ψ(t) L. Again the key observation is that the estimate above is independent of t. Using the same inductive argument as in the proof of Theorem.8 shows that for any p, (3.11) t t 0 p ψ L + p 1 ψ(t) L C 1 (p) ψ(t 0 ) L p 1, + C (p)e λt 0 where C 1 (p) and C (p) are independent of t. Notice that there exists a constant C, such that ψ(t 0 ) L p 1, C ψ(t 0 ) C p 1. Hence by (3.11) we have p 1 ψ(t) L C 1 (p) ψ(t 0) C p 1+ C (p)e λt 0. Therefore applying the Sobolev Embedding Theorem we have (3.1) ψ(t) C k C 1 (k) ψ(t 0 ) C p 1 + C (k)e λt 0 where C 1 (k) and C (k) are independent of t. Since (3.1) is independent of t, to prove that ψ(t) C k < δ for all t [0, ), it suffices to show that there exists a constant ɛ 1 with 0 < ɛ 1 ɛ and such that ρ(0) C < ɛ 1 implies that the right-hand side of (3.1) is bounded above by δ. First we bound the second term on the right-hand side of (3.1). Notice that, given δ > 0 small enough so that we have (3.4), the argument above which led to inequality (3.1) holds under the assumption (3.6). Furthermore, notice that the estimate in (3.1) holds for t 0 < T, independent of T. Therefore we take T to be sufficiently large so that T > t 0 and (3.13) C (k)e λt 0 < 1 δ.

14 DANIEL J. SITH To bound the first term in (3.1) we again use Theorem.8 with ɛ = 1 C 1 (k) δ and T > t 0. Hence, by Theorem.8 there exists ɛ 3 > 0 such that ρ(0) C < ɛ 3 implies that (3.14) ψ C p 1 < ɛ = 1 C 1 (k) δ for t [0, T ) [0, t 0 ]. Finally, choose ɛ 1 = min{ɛ, ɛ 3 }. Hence combining (3.1), (3.13) and (3.14) proves that if ψ(0) C < ɛ 1, then (3.1) holds independent of t. Therefore it follows that ψ(t) C k < δ for all t 0 and moreover the L decay estimate in (3.5) holds for all t 0. To finish the proof of Theorem 3.1, we show that the L decay estimate above and parabolic theory can be used to prove C k decay of ψ(t). Lemma 3.3. Given an integer k, there exists δ = δ(k) > 0 such that if both ψ(t) C k < δ for all t [0, ) and ψ(t) L Ce λt then ψ(t) C k C(k)e λt. Proof. We begin the proof by deriving an L 1, exponential decay estimate. The same argument that was used to derive (.40) shows that there exists a constant C 1 such that t ψ L ψ L + C 1 ψ L, where C 1 depends on both ( ω, J) and ψ(t) C ; but by assumption ψ C < δ for all t 0. Integrating from t to yields, (3.15) ψ L ψ(t) L + C 1 ψ L Ce λt. t The last inequality follows from the assumed L exponential decay estimate. Next, for a fixed t, let θ(s) be a smooth function which is 0 for s [ t 1, t 1 ], monotonically increasing from 0 to 1 for s [ t 1, t] and 1 for s t. As we shall see below, θ(s) will be used to deal with boundary terms which arise in the parabolic estimates that follow. The same argument that was used to produce (.47) shows that there exist constants such that (3.16) 1 t ψ g dvol g ψ + C ψ + C 3 ψ + C 4 δ ψ, again these constants depend on both ( ω, J) and ψ(t) C. Now we choose δ sufficiently small so that C 4 δ < 1. Hence using (3.16) and that both θ(s) and its derivative are uniformly bounded, ( ) θ(s) ψ(s) (3.17) s L C5 ψ(s) L + C 6 ψ(s) L. t We integrate (3.17) in s from t 1 to t for t 1. Using that θ ( t 1 ) = 0 and θ(t) = 1 we get (3.18) ψ(t) L C 5 t t 1 C 5 t 1 ψ(s) L + C 6 t t 1 ψ(s) L + C 6 t 1 ψ(s) L ψ(s) L.

STABILITY OF THE ALOST HERITIAN CURVATURE FLOW 15 Hence using the L decay assumption and (3.15) it follows from (3.18) that (3.19) ψ(t) L Ce λt. This proves exponential L 1, decay. Next we prove L, decay. By (3.16) with δ small enough so that C 4 δ < 1, it follows that Integrating from t to yields (3.0) t t ψ L ψ L + C 3 ψ L + C ψ L. ψ L ψ(t) L + C 3 ψ L + C ψ L Ce λt. Where (3.19), (3.15), and the L decay assumption were used in the first, second, and third term on the right-hand side respectively. Now, the same parabolic technique that gave rise to (.47) implies that there exists a constant C 5 such that 1 ψ g t dvol g 1 3 ψ + C 5 j ψ L. Hence, there exists a constant C 6 so that (3.1) t ( θ(s) ψ(s) ) L s C6 ψ(s) L + C 6 ψ(s) L + C 6 ψ(s) L. We integrate (3.1) in s from t 1 to t for t 1. Using that θ ( t 1 ) = 0 and θ(t) = 1 we get ψ(t) L C 6 t t 1 j=0 j ψ(s) L C 6 Hence using (3.0), (3.15) and the assumed L decay we get This gives exponential L, decay. Continuing in this way we get ψ(t) L Ce λt. ψ(t) L p, C(p)e λt. j=0 t 1 j=0 t j ψ(s) L. Furthermore by the Sobolev Embedding Theorem we get ψ(t) C k C(k)e λt. By Lemma 3.3 we know that given k, there exists δ > 0 so that if both ψ(t) C k < δ and ψ(t) L Ce λt hold for all t 0, then ψ(t) C k C(k)e λt for all t 0. Furthermore by Lemma 3. we know that there exists ɛ 1 > 0 such that if ψ(0) C < ɛ 1, then both ψ(t) C k < δ and ψ(t) L Ce λt hold for all t 0. Hence let δ be determined by Lemma 3.3. To finish the proof of Theorem 3.1, we apply Lemma 3. with ɛ = ɛ 1 and note that by (.13) exponential decay of ρ(t) follows from exponential decay of ψ(t).

16 DANIEL J. SITH 4. Dynamic Stability in the Calabi-Yau Case In Section 3 we proved Theorem 1.1 when c 1 ( J) < 0 by using that, in this case, the linearization L is negative definite. However in the Calabi-Yau case, the kernel of L is non-trivial and so the non-linear part of the flow is no longer controlled by the linear part. In this section we will show that in the Calabi-Yau case we can find a Calabi-Yau structure to which the flow exponentially converges. In order to find a Calabi-Yau structure to which the flow exponentially converges we will construct a sequence {(ω j, J j )} of successively better Calabi-Yau structures; in the sense that the solution (ω(t), J(t)) to the VNAHCF (.1) converges exponentially on larger and larger intervals. oreover we will prove that each of these Calabi-Yau structures is contained in a fixed neighborhood of the original Calabi-Yau structure (see Theorem 4.5 part ()). This will allow us to extract a limit (ω KE, J KE ) to which {(ω j, J j )} subconverges. One could imagine that if this sequence failed to converge that we would only be able to conclude that the solution becomes asymptotic to the space of Calabi-Yau structures. In order to choose a new Calabi-Yau structure we will use Koiso s Theorem. Before stating Koiso s Theorem we need a definition. Definition 4.1. Let AC denote the space of almost complex structures on modulo diffeomorphism. A complex structure J is unobstructed if for any J T J AC such that Ṅ( J) = 0, there exists a path of complex structures J(a) such that J(0) = J and a=0 J(a) = J. Again, N denotes the Nijenhuis tensor. a Theorem 4.. (Koiso [6]) Let (ω, J) be a Kähler-Einstein structure on. Assume that: (1) the first Chern class of J is zero; () J is unobstructed. Then the space of Kähler-Einstein structures around (ω, J) is a manifold. In addition to [6], Theorem 4. can be found in [] (see Theorem 1.88). In order to make use of Koiso s Theorem we employ a theorem of Tian and Todorov. Theorem 4.3. (Tian [1] and Todorov [13]) Let (, J) be a closed Calabi-Yau manifold. Then J is unobstructed. Next we will describe the tangent space of Calabi-Yau structures at ( ω, J). Using the above two theorems we prove that the kernel of L is isomorphic to the tangent space of Calabi-Yau structures at ( ω, J). Let U denote the space of Calabi-Yau structures near ( ω, J) modulo diffeomorphism. Lemma 4.4. Let (, ω, J) be a closed Calabi-Yau manifold, then T ( ω, J) U = Ker L. Proof. Let (ω(a), J(a)) be a one-parameter family of unit volume almost Hermitian structures and write a a=0 (ω(a), J(a)) = ( ω, J). First since Calabi-Yau structures are static under the system (.1), we have T ( ω, U Ker L. To prove Ker L T J) ( ω, U, let ( ω, J) J) Ker L. We make the following claim. If J ker G then J ker Ṅ. To see this, first notice that from (.9) and using that the scalar curvature s g = 0, if J ker G, then J = 0. By integrating we see that J = 0. On the other hand, in coordinates, the Nijenhuis tensor is written

STABILITY OF THE ALOST HERITIAN CURVATURE FLOW 17 Hence, N i jk = J p j pj i k J p k pj i j J i p j J p k + J i p k J p j. Ṅ i jk = J p j pj i k + J p j p J i k J p k pj i j J p k p J i j J i p j J p k J i p j J p k + J i p k J p j + J i p k J p j. Now each of the terms above of the form J J can be written as J J = J( J + Γ J). So using normal, complex coordinates (with respect to the Calabi-Yau structure ( g, J)), at a point p, we have that each of these terms vanish. Here we also made use of the fact that when ( g, J) is Kähler the Chern connection coincides with the Levi-Civita connection and so J is parallel with respect to the connection. Next, since J Λ 0,1 T 1,0 in these normal, complex coordinates at p, we have, Ṅ i jk = J p j p J i k J p k p J i j J i p j J p k + J i p k J p j = 0. This proves the claim. By Theorem 4.3 there exists a path of complex structures J(a) where J(0) = J and d da J a=0 = J. Next, using (.8) we have that ω ker F implies that d ω = 0, that is ω is harmonic. Therefore, by the Calabi-Yau Theorem ([1], [16], [17], also see Theorem.9 in [3]) ω(a) is a variation through Kähler metrics such that [ω(a)] = [ω KE (a)], where ω KE (a) is Ricci-flat. oreover by the Hodge Decomposition Theorem there is a unique harmonic representative in each cohomology class. Hence ω(a) = ω KE (a) and we have that ω arises as a variation through Calabi-Yau metrics. Notice that Λ () End(T ) is an affine space which can be viewed as a vector space by taking ( ω, J) to be the origin. Throughout this section we will view Λ () End(T ) as a vector space. Let π 0 : Λ () End(T ) Ker L be the projection onto the kernel of L. Let ( ω, J) denote the Calabi-Yau structure from Theorem 1.1. Roughly speaking, we will next prove that there exists a better Calabi-Yau structure (ω I, J I ); in the sense that the solution (ω(t), J(t)) to VNAHCF exponentially converges to (ω I, J I ) on an interval I (see Theorem 4.5 and Lemma 4.6). Throughout this section ρ I (t) =. (ω(t) ω I, J(t) J I ) will quantify the distance the solution is from this new Calabi-Yau structure. As above let ρ(t) = (ω(t) ω, J(t) J). Notice that we may view both ρ(t) and ρ I (t) as elements of Λ () End(T ). As in Section, we have to deal with the non-linearity of the space of almost Hermitian structures modulo diffeomorphism denoted C. Notice that we may write ρ I (t) = ρ(t) ρ I. where ρ I = (ωi ω, J I J). From Lemma.6, associated to ρ(t) we have ψ(t) T ( ω, C J) and analogously associated to ρ I we have ψ I T ( ω, J) C. Hence associated to ρ I (t) we have ψ I (t) T ( ω, J) C defined by ψ I (t). = ψ(t) ψ I. oreover, employing the same argument as in the proof of Lemma.6 we have (4.1) ψ I (t) C k ρ I (t) C k, (4.) ρ I (t) L ψ I (t) L + C ψ I (t) L

18 DANIEL J. SITH and (4.3) ρ I (t) C k ψ I (t) C k + C ψ I (t) C k. Similarly by the proof of Lemma.7 we have (4.4) where (4.5) t ψ I(t) = L(ψ I (t)) + A(( ω, J), ψ I (t)) A(( ω, J), ψ I (t)) C k C( ψ I (t) C k ψ I (t) C k + ψ I (t) C k ). Next we will use the identification of the kernel of L and the tangent space of Calabi- Yau structures at ( ω, J), from Lemma 4.4, to find a new Calabi-Yau structure denoted (ω I, J I ) such that π 0 (ψ I (t)) L is small relative to ψ I (t) L. Furthermore we will show that the new Calabi-Yau structure is contained in a fixed neighborhood of the original Calabi-Yau structure ( ω, J). Theorem 4.5. Given t 0 and T > 0, let I = [t 0, t 0 + T ]. There exists δ(t, g) such that if sup I ψ(t) C k < δ with k, then there exists a Calabi-Yau structure (ω I, J I ) with the following properties: (1) π 0 (ψ I ) L ( g) 1 4 ψ I L ( g) on I () (ω I ω, J I J) C k C sup I ψ C k. Proof. First by Theorem 4. we know that the space of Calabi-Yau structures U has a manifold structure near ( ω, J) and moreover by Lemma 4.4 we have Ker L = T ( ω, J) U. By identifying Ker L and T ( ω, J) U we will view ( ω, J) as the origin of Ker L. Let Φ = Φ ( ω, : Ker L U denote the exponential map at ( ω, J) J). Now since D ( ω, Φ is the J) identity map, the inverse function theorem may be applied to Φ. By the inverse function theorem there exists a neighborhood V Ker L of ( ω, J) on which the exponential map is invertible. Let δ 1 be small enough so that (4.6) π 0 (ψ(t 0 )) C k < δ 1 implies that π 0 (ψ(t 0 )) V. Notice that if sup I ψ C k < δ 1 then since t 0 I, it is clear that π 0 (ψ(t 0 )) C k < δ 1. Hence by the inverse function theorem there is a Calabi-Yau structure (ω I, J I ) U such that (4.7) (Φ V ) 1( (ω I, J I ) ) = π 0 (ψ(t 0 )). Applying Φ to each side of (4.7), it follows from the inverse function theorem that there exists a constant C so that (ω I ω, J I J) C k C π 0 (ψ(t 0 )) C k C sup ψ C k. I This proves (). Next, using that (Φ V ) 1( (ω I, J I ) ) = π 0 ( (ωi ω, J I J) ), from (4.7) we have (4.8) π 0 (ψ I (t 0 )) = 0. In other words, there exists a Calabi-Yau structure (ω I, J I ) such that at time t 0, ψ I (t) is orthogonal, with respect to L ( g), to Ker L.

STABILITY OF THE ALOST HERITIAN CURVATURE FLOW 19 To prove (1) we will carefully study the evolution of ψ I (t) starting at t = t 0. First let ψ I I = I ψ I L ( g) denote the L norm on I. Let {B i } be an orthonormal basis, with respect to L ( g), of T ( ω, J) C determined by the eigenspace decomposition of L. Then there exist constants c i so that {c i B i e λit } is an orthonormal basis, with respect to I, of Ker ( t L) I where λ i is the eigenvalue associated to B i. We let π I( ψ I (t) ) denote the projection of ψ I (t) onto Ker ( t L) I. In other words, (4.9) t πi (ψ I (t)) = L(π I (ψ I (t))). From (4.8), we have π I (ψ I (t 0 )) = λ i 0 k ib i. It then follows that (4.10) π I (ψ I (t)) = λ i 0 k i B i e λ i(t t 0 ). We write (4.11) ψ I (t) = π I (ψ I (t)) + ξ I (t). Since π I( ψ I (t) ) is orthogonal to Ker L on I, it follows that for t I, (4.1) π 0 (ψ I (t)) ξ I (t). Therefore to obtain estimates on π 0 ( ψi (t) ) we compute the evolution of ξ I (t). oreover, from (4.10) and (4.11), since λ i < 0 is bounded away from 0 for all i, we have that ξ I (t) converges exponentially to ψ I (t). Therefore there is a uniform constant C so that on I, (4.13) ξ I (t) C ψ I (t). To compute the evolution of ξ I (t) we compare two evolution equations for ψ I (t). From (4.4), ψ I (t) satisfies t ψ I(t) = L(ψ I (t)) + A(( ω, J), ψ I ) and hence, (4.14) t ψ I(t) = L(π I (ψ I (t))) + L(ξ I (t)) + A(( ω, J), ψ I (t)). On the other hand, π I (ψ I (t)) satisfies (4.9) and so t ψ I(t) = ( (4.15) π I (ψ I (t)) + ξ I (t) ) = L(π I (ψ I (t))) + t t ξ I(t). Combining equations (4.14) and (4.15) we have that ξ I (t) evolves by (4.16) t ξ I(t) = L(ξ I (t)) + A(( ω, J), ψ I (t)). Recall that L is negative semi-definite; and so by (4.16), on I we have ( t ξ I(t) L ( g) = t ξ I(t), ξ I (t) dvol g A ( ω, J), ) ψ I (t) ξ I (t). g Now using the bounds on A from (4.5), the same computation as (.36) shows that t ξ I(t) L ( g) C 1 ψ I ψ I ξ I. Hence by (4.13), (4.17) t ξ I(t) L ( g) C ψ I ψ I.

0 DANIEL J. SITH Next we assume sup I ψ(t) C k < δ with k and δ δ 1 where δ 1 is from (4.6). Using part () of Theorem 4.5 and the triangle inequality, from (4.17) it follows that on I Now since ξ I (t 0 ) = 0, (4.18) t ξ I(t) L C 3 δ ψ I (t) L. t ξ I (t) L ( g) = ξ I (s) s L ( g) ds C 3δ t 0 t t 0 ψ I (s) L ( g) ds. Notice that since t ψ I(t) = t ψ(t) is second order in ψ(t) and sup I ψ(t) C k < δ with k, t ψ I(t) is uniformly bounded in terms of δ and hence each ψ I (s) for s I is uniformly equivalent. Therefore π 0 (ψ I (t)) L ( g) ξ I(t) L ( g) C 4δ t t 0 ψ I (t) L ( g) ds = C 5δ(t t 0 ) ψ I (t) L ( g), where the first inequality follows from (4.1) and the second is from (4.18). To finish the proof we choose δ small enough so that both C 5 T δ < 1 4 and δ δ 1 hold. We will now use part (1) of Theorem 4.5 to prove L exponential decay of ψ I (t) on I. Notice that exponential decay of ρ I (t) = (ω(t) ω I, J(t) J I ) on I will follow immediately from exponential decay of ψ I (t) and (4.3). Lemma 4.6. Let I and (ω I, J I ) be as in Theorem 4.5. There exists β > 0 such that if ψ C < β on I, then ψ I dvol g e T λ sup ψ I dvol g [t 0,t 0 + 1 T] sup [t 0 + 1 T,t 0+T] where λ = min{ λ i : λ i is a non-zero eigenvalue of L } > 0. Proof. We compute the evolution of ψ I L and as in (4.4) we have ( ψ I dvol g = L(ψ I ), ψ I dvol g + A ( ω, t J), ) (4.19) ψ I (t) ψ I dvol g. Recall that by the definition of π 0, ψ I π 0 (ψ I ) is the component of ψ I orthogonal to the kernel of L. Hence by the definition of λ, (4.0) L(ψ I ), ψ I dvol g λ ψ I π 0 (ψ I ) L (4.1) λ ( ψ I L π 0 (ψ I ) L ). Let δ be the constant from Theorem 4.5. By Theorem 4.5 part (1), if sup I ψ(t) C < β 1 with β 1 δ, then from (4.0) and (4.1) it follows that (4.) L(ψ I ), ψ I dvol g 3 λ ψ I L. Next consider the term A ψ I from (4.19). We use the estimate on A from (4.5) to bound A ψ. Notice that if ψ I C < β 3, then as in (.37), ( A ( ω, J), ) (4.3) ψ I (t) ψ I dvol g Cβ 3 ψ I L.

STABILITY OF THE ALOST HERITIAN CURVATURE FLOW 1 Now we choose β 3 small enough so that (4.4) Cβ 3 < λ. Let β be sufficiently small so that ψ C < β on I implies that ψ I C < β 3 on I. Notice that this can be done using the triangle inequality and part () of Theorem 4.5. Finally we choose β = min{β 1, β }. Combining (4.19), (4.), (4.3) and (4.4) it follows that if ψ C < β on I, then (4.5) Integrating from t 0 to t gives ψ I dvol g λ ψ I dvol g. t (4.6) ψ I (t) L e λ(t t 0) ψ I (t 0 ) L. Finally since (4.5) implies that ψ I (t) L is decreasing, plugging t 0 + 1 T into (4.6) proves the lemma. This gives exponential L decay of ψ I (t) on I. Next we prove a general result about parabolic flows (cf. Lemma 8.8 in [11]). The following lemma says roughly that exponential decay at a later time implies exponential decay at an earlier time. Lemma 4.7. There exists ν > 0 so that if κ solves the parabolic flow equation t κ = L(κ) + A(κ) and κ(t) C k < ν for all t [0, t 0 + T ], then (4.7) sup κ e T λ sup κ [t 0 + 1 T,t 0+T] [t 0,t 0 + 1 T] implies that (4.8) sup κ e T λ sup κ. [t 0,t 0 + 1 T] [t 0 1 T,t 0] Proof. Suppose, by way of contradiction, that the lemma fails to hold. Then there is a sequence ν i 0 with κ i (t) solving t κ i = L(κ i ) + A(κ i ) and κ i C k < ν i on [0, t 0 + T ] and moreover (4.7) holds but (4.8) does not. Parabolically rescale the solution κ i ; that is let κ(t) i = νi 1 κ i (ν i t). Now for all i, κ i C k < 1 on [0, νi 1 (t 0 + T )] [0, t 0 + T ] and so by compactness we get a convergent subsequence κ(t) i κ(t) on [0, t 0 + T ] as i. Now since κ i solves t κ i = L(ν i κ i ) + A(ν i κ i ) and A(κ) is quadratic in κ this implies that κ (t) solves the linear equation (4.9) t κ = L( κ ). Furthermore since (4.7) and (4.8) are scale invariant it follows that for κ (4.7) holds but (4.8) does not. This is a contradiction. To see the contradiction, first notice that (4.9) implies that (4.30) t κ (t) L 0

DANIEL J. SITH as L is negative semi-definite. It then follows that (4.31) (4.3) (4.33) κ (t 0 + 1 T ) L = sup κ L [t 0 + 1 T,t 0+T] e T λ sup [t 0,t 0 + 1 T] κ L = e T λ κ (t 0 ) L, where the inequality follows from (4.7). As above, let {B i } be an orthonormal basis, with respect to L ( g), of T ( ω, C determined by the eigenspace decomposition of L. We J) can now write κ (t) with respect to this basis, ( ) κ (t) = κ (t 0 ) B i e λ i(t t 0 ) (4.34). Notice that at time t = t 0 + 1 T we have (4.35) κ (t 0 + 1 T ) L = κ (t 0 ) L ( i i e T λ i By combining (4.31), (4.3), (4.33) and (4.35), it follows that (4.36) e T λ i e T λ. And by (4.34), (4.37) i κ (t 0 1 T ) L = κ (t 0 ) L ( i Finally using (4.37) and the concavity of f(x) = 1 x (4.38) e T λ i we see that κ (t 0 ) L = κ (t 0 1 T 1 ) L ( i e T λ i ) κ (t 0 1 T ) L ( i e T λ κ (t 0 1 T ) L. The last inequality in (4.38) follows from (4.36). Notice that the above inequality along with (4.30) imply that (4.8) holds. This is a contradiction. Corollary 4.8. Given T > 0 and j 1 there exists α = α(t, j) > 0 such that if ψ(t) C < α on [0, (j + 1)T ] then there exists a Calabi-Yau structure (ω j, J j ) so that ρ j (t) = (ω(t) ω j, J(t) J j ) satisfies the following exponential decay estimate: ρ j (t) L Ce λt for t [0, (j + 1)T ] and a constant C independent of j. e T λ i ) )..

STABILITY OF THE ALOST HERITIAN CURVATURE FLOW 3 Proof. First notice that by (4.) it suffices to prove exponential decay of ψ j (t), the tangent vector associated to ρ j (t). We will prove exponential decay of ψ j (t) using the previous two lemmas and Theorem 4.5. Let δ, β and ν be the small constants from Theorem 4.5, Lemma 4.6 and Lemma 4.7 respectively. In order to apply the above lemmas and Theorem 4.5 we let α = min{δ, β, ν} and assume that ψ C < α on [0, (J + 1)T ]. Employing Theorem 4.5 (with t 0 = jt ) and Lemma 4.6, there exists a Calabi-Yau structure, denoted (ω j, J j ), such that (4.39) sup ψ j L e T λ sup ψ j L. [(j+ )T,(j+1)T] 1 [jt,(j+ )T] 1 Now, Lemma 4.7 says that exponential decay at a later time implies exponential decay at an earlier time. In particular, from Lemma 4.7 and (4.39) it follows that for any k { n : n Z and 1 n j + 1}, (4.40) sup ψ j L e T λ sup ψ j L. [kt,(k+ 1 )T] [(k 1 )T,kT] Applying (4.40) with k = 1 implies that for any t [ 1 T, T ], (4.41) Next, using (4.40) with k = 1 ψ j (t) L e T λ sup ψ j [0, 1 T] L e λt sup ψ j [0, 1 T] L. and k = 1 yields sup ψ j [T, 3 T] L e T λ sup ψ j [ 1 T,T] L e T λ sup ψ j [0, 1 T] L. Therefore it follows that for t [ T, 3 T ], (4.4) ψ j (t) L e T λ sup ψ j [0, 1 T] L e λt sup ψ j [0, 1 T] L. Combining (4.41) and (4.4) yields L exponential decay of ψ j (t) on [ 1 T, 3 T ]. Iterating this argument, we see that for t [ 1 T, (j + 1)T ], ψ j (t) L e λt sup ψ j [0, 1 T] L Ce λt. Notice that C is independent of j. Indeed by assumption ψ(t) C < α on [0, (j + 1)T ]. Hence part () of Theorem 4.5 and the triangle inequality imply that ψ j (t) C < C on [0, (j + 1)T ], where C is independent of j. Notice that the decay estimate from Corollary 4.8 may fail to hold for intervals beyond I j. In order to maintain exponential decay we want to choose another Calabi-Yau structure (ω j+1, J j+1 ) to which the solution exponentially converges. To ensure that we can continue this process we need to prove that ψ(t) C is small for all time so that Corollary 4.8 may be applied on any interval. This is the purpose of the following theorem. As a corollary we will prove the existence of a Calabi-Yau structure, denoted (ω KE, J KE ), to which the flow exponentially converges. Theorem 4.9. Let (, ω, J) be a closed complex manifold with ( ω, J) a Calabi-Yau structure. Given ɛ > 0 and k 0 there exists ɛ > 0 so that ρ(0) C < ɛ implies that ψ(t) C k < ɛ on [0, ).

4 DANIEL J. SITH Proof. We will employ Theorem.8. To do this we make explicit T, ɛ, and ɛ. (1) Let T be large enough so that T C 3 (k + ) e T λ 1 + 1 e T λ < 1. Here C 3 (k + ) is a constant that depends only on k and ( ω, J). () Choose ɛ = α, where α is the constant from Corollary 4.8. (3) Choose ɛ sufficiently small so that (ω(t), J(t)) exists on [0, 3T ] and sup ψ(t) C k < ɛ [0,T ] e T λ. We want to prove that ψ(t) C k < ɛ on [0, ). Suppose by way of contradiction there is a finite maximal time T such that ψ(t) C k < ɛ on [0, T ) with k. Write [0, T ) = [0, T ] [T, T ] [NT, T ) and let [jt, (j + 1)T ] = I j. Also let (ω j, J j ) denote the Calabi-Yau structure, from Corollary 4.8, to which (ω(t), J(t)) exponentially converges on I j. Using (4.39) and applying Lemma 4.7 iteratively we have (4.43) sup ψ j L e λt (j 1) sup ψ j I j 1 I j [0, T ] L. We again use a parabolic regularity argument to prove that from (4.43) we can obtain a C k+ decay estimate. Lemma 4.10. There exists a constant α > 0 so that if both ψ C < α on I j and sup Ij 1 I j ψ j L e λt (j 1) sup [0, T ] ψ j L, then there exists a constant C(k + ) such that sup ψ j C k+ < C(k + )T e T λ(j 1) sup ψ j I j [0, 1 T] L. Proof. The proof of Lemma 4.10 uses essentially the same argument as the proof of Lemma 3.3 hence we omit some of the details. From (.41) we have Fix t I j and integrate from (j 1)T to t; (4.44) t (j 1)T t ψ j L ψ j L + C 1 ψ j L. ψ j L ψ j ((j 1)T ) L + C 1 t (j 1)T T λ(j 1) CT e sup ψ j [0, 1 T] L, ψ j L where the second inequality follows from the L exponential decay assumption. Next let θ(s) be a smooth function which is 0 for s (j 1)T, monotonically increasing from 0 to 1 for s [(j 1)T, jt ] and equal to 1 for s jt. As in Lemma 4.10, θ(s) will