• Aucun résultat trouvé

Introduction Consider a complete Riemannian manifold (Mn, g) endowed with a Ricci-flat metric g

N/A
N/A
Protected

Academic year: 2022

Partager "Introduction Consider a complete Riemannian manifold (Mn, g) endowed with a Ricci-flat metric g"

Copied!
30
0
0

Texte intégral

(1)

ALIX DERUELLE AND KLAUS KR ¨ONCKE

Abstract. We prove that if an ALE Ricci-flat manifold (M, g) is linearly stable and inte- grable, it is dynamically stable under Ricci flow, i.e. any Ricci flow starting close togexists for all time and converges modulo diffeomorphism to an ALE Ricci-flat metric close to g.

By adapting Tian’s approach in the closed case, we show that integrability holds for ALE Calabi-Yau manifolds which implies that they are dynamically stable.

1. Introduction

Consider a complete Riemannian manifold (Mn, g) endowed with a Ricci-flat metric g.

As such, it is a fixed point of the Ricci flow and therefore, it is tempting to study the stability of such a metric with respect to the Ricci flow. Whether the manifold is closed or noncompact makes an essential difference in the analysis. In both cases, if (Mn, g) is Ricci-flat, the linearized operator is the so called Lichnerowicz operator acting on symmetric 2-tensors. Nonetheless, the L2 approach differs drastically in the noncompact case. Indeed, even in the simplest situation, that is the flat case, the spectrum of the Lichnerowicz operator is not discrete anymore and 0 belongs to the essential spectrum. In this paper, we restrict to Ricci-flat metrics on noncompact manifolds that are asymptotically locally Euclidean (ALE for short), i.e. that are asymptotic to a flat cone over a space formSn−1/Γ where Γ is a finite group of SO(n) acting freely on Rn\ {0}.

If (Mn, g0) is an ALE Ricci-flat metric, we assume furthermore that it is linearly stable, i.e.

the Lichnerowicz operator is nonpositive in theL2 sense and that the set of ALE Ricci-flat metrics close tog0 is integrable, i.e. has a manifold structure of finite dimension: see Section 2.1.

The strategy we adopt is given by Koch and Lamm [KL12] that studied the stability of the Euclidean space along the Ricci flow in theLsense. The quasi-linear evolution equation to consider here is

tg=−2 Ricg+LV(g,g0)(g), kg(0)−g0kL(g0) small, (1) where (Mn, g0) is a fixed background ALE Ricci-flat metric and LV(g,g0)(g) is the so called DeTurck’s term. Equation (1) is called the Ricci-DeTurck flow: its advantage over the Ricci flow equation is to be a strictly parabolic equation instead of a degenerate one. Koch and Lamm managed to rewrite (1) in a clever way to get optimal results regarding the regularity of the initial condition: see Section 3.

Our main theorem is then:

Theorem 1.1. Let (Mn, g0) be an ALE Ricci-flat space. Assume it is linearly stable and integrable. Then for every >0, there exists a δ >0 such that the following holds: for any metric g∈ BL2∩L(g0, δ), there is a complete Ricci-DeTurck flow (Mn, g(t))t≥0 starting from g converging to an ALE Ricci-flat metric g∈ BL2∩L(g0, ).

1

(2)

Moreover, the L norm of (g(t)−g0)t≥0 is decaying sharply at infinity:

kg(t)−g0kL(M\Bg0(x0,

t)) ≤C(n, g0, )supt≥0kg(t)−g0kL2(M)

tn4 , t >0.

Schn¨urer, Schulze and Simon [SSS11] have proved the stability of the Euclidean space for anL2∩L perturbation as well. The decay obtained in Theorem 1.1 sharpens their result:

indeed, the proof shows that if (Mn, g0) is isometric to (Rn,eucl) then the L decay holds on the whole manifold.

Remark 1.2. It is an open question whether the decay in time obtained in Theorem 1.1 holds on the whole manifold with an exponentα less than or equal to n/4.

From the physicist point of view, the question of stability of ALE Ricci-flat metrics is of great importance when applied to hyperk¨ahler or Calabi-Yau ALE metrics: the Lichnerowicz operator is always a nonnegative operator because of the special algebraic structure of the curvature tensor shared by these metrics. It turns out that they are also integrable: see The- orem 2.17 based on the fundamental results of Tian [Tia87] in the closed case. In particular, it gives us plenty of examples to which one can apply Theorem 1.1.

Another source of motivation comes from the question of continuing the Ricci flow after it reached a finite time singularity on a 4-dimensional closed Riemannian manifold: the works of Bamler [BZ15] and Simon [Sim15] show that the singularities that can eventually show up are exactly ALE Ricci-flat metrics. However, there is no classification available of such metrics in dimension 4 at the moment, except Kronheimer’s classification for hyperk¨ahler metrics [Kro89].

Finally, we would like to discuss some related results especially regarding the stability of closed Ricci-flat metrics. There have been basically two approaches. On one hand, Sesum [Ses06] has proved the stability of integrable Ricci-flat metrics on closed manifolds: in this case, the convergence rate is exponential since the spectrum of the Lichnerowicz operator is discrete. On the other hand, Haslhofer-M¨uller [HM13] and the second author [Kr¨o15] have proved Lojasiewicz inequalities for Perelman’s entropies which are monotone under the Ricci flow and whose critical points are exactly Ricci-flat metrics and Ricci solitons, respectively.

The paper is organized as follows. Section 2.1 recalls the basic definitions of ALE spaces together with the notions of linear stability and integrability of a Ricci-flat metric. Section 2.2 gives a detailed description of the space of gauged ALE Ricci-flat metrics: see Theorem 2.7 and Theorem 2.10. Section 2.3 investigates the integrability of K¨ahler Ricci-flat metrics: this is the content of Theorem 2.17. Section 3 is devoted to the proof of the first part of Theorem 1.1. Section 3.1 discusses the structure of the Ricci-DeTurck flow. Section 3.2 establishes pointwise and integral short time estimates. The core of the proof of Theorem 1.1 is contained in Section 3.3: a priori uniform in timeL2 estimates are proved with the help of a suitable notion of strict positivity for the Lichnerowicz operator developed for Schr¨odinger operators by Devyver [Dev14]. The infinite time existence and the convergence aspects of Theorem 1.1 are then proved in Section 3.4. Finally, Section 4 proves the last part of Theorem 1.1: the decay in time is verified with the help of a Nash-Moser iteration.

Acknowledgements. The authors want to thank the MSRI for hospitality during the re- search program in differential geometry, where part of the work was carried out.

(3)

2. ALE spaces

2.1. Analysis on ALE spaces. We start by recalling a couple of definitions.

Definition 2.1. A complete Riemannian manifold (Mn, g0) is said to be asymptotically locally Euclidean(ALE) with one end of order τ >0 if there exists a compact set K ⊂M, a radiusR and a diffeomorphism such that : φ:M\K →(Rn\BR)/Γ, where Γ is a finite group of SO(n) acting freely onRn\ {0}, such that

|∇eucl,kg0−geucl)|eucl =O(r−τ−k), ∀k≥0.

holds on (Rn\BR)/Γ.

The linearized operator we will focus on is the so called Lichnerowicz operator whose definition is recalled below:

Definition 2.2. Let(M, g)be a Riemannian manifold. Then the operatorLg :C(S2TM)→ C(S2TM), defined by

Lg(h) := ∆gh+ 2 Rm(g)∗h−Ric(g)◦h−h◦Ric(g), (Rm(g)∗h)ij := Rm(g)ikljhmngkmgln,

is called the Lichnerowicz Laplacian acting on the space of symmetric2-tensorsS2TM. In this paper, we consider the following notion of stability:

Definition 2.3. Let (Mn, g0) be a complete ALE Ricci-flat manifold. (Mn, g0) is said to be linearly stable if the (essential) L2 spectrum of the Lichnerowicz operator Lg0 := ∆g0 + 2 Rm(g0)∗ is in(−∞,0].

Equivalently, this amounts to say that σL2(−Lg0) ⊂ [0,+∞). By a theorem due to G.

Carron [Car99], kerL2Lg0 has finite dimension. Denote by Πc the L2 projection on the kernel kerL2Lg0 and Πs the projection orthogonal to Πc so that h = Πch + Πsh for any h∈L2(S2TM).

Let (M, g0) be an ALE Ricci-flat manifold and Ug0 the set of ALE Ricci-flat metrics with respect to the gauge g0, that is :

Ug0 := {g|g ALE metric on M s.t. Ric(g) = 0 and LV(g,g0)(g) = 0}, (2) g0(V(g(t), g0), .) := divg(t)g0− 1

2∇g(t)trg(t)g0, (3)

endowed with the L2∩L topology coming fromg0.

Definition 2.4. (Mn, g0) is said to be integrable if Ug0 has a smooth structure in a neigh- borhood of g0. In other words, (Mn, g0) is integrable if the map

Ψg0 :g∈ Ug0 →Πc(g−g0)∈kerL2(Lg0), is a local diffeomorphism at g0.

If (M, g0) is ALE and Ricci-flat, it is a consequence of [BKN89, Theorem 1.1] that it is already ALE of order n−1. Moreover, if n= 4 or (M, g0) is K¨ahler, it is ALE of order n.

This is due to the presence of Kato inequalities, [BKN89, Corollary 4.10] for the curvature tensor. We will show in Theorem 2.7 that by elliptic regularity, allg∈ Ug0 are ALE of order n−1 with respect to the same coordinates as g0.

(4)

In order to do analysis of partial differential equations on ALE manifolds, one has to work with weighted function spaces which we will define in the following. Fix a point x∈M and define a functionρ :M → R by ρ(y) =p

1 +d(x, y)2. For p ∈[1,∞) andδ ∈R, we define the spacesLpδ(M) as the closure ofC0(M) with respect to the norm

kukLp

δ = ˆ

M

−δu|pρ−n1/p

,

and the weighted Sobolev spacesWδk,p(M) as the closure ofC0(M) under

kukWδk,p =

k

X

l=0

lu

Lp

δ−l

.

The weighted H¨older spaces are defined as the closure ofC0(M) under

kukCδk,α =

k

X

l=0

sup

x∈M

ρ−δ+l(x)|∇lu(x)|

+ sup

x,y∈M 0<d(x,y)<inj(M)

min n

ρ−δ+k+α(x), ρ−δ+k+α(y)

o|τxyku(x)− ∇ku(y)|

|x−y|α ,

whereα∈(0,1] andτxy denotes the parallel transport fromxtoyalong the shortest geodesic joining x and y. All these spaces are Banach spaces, the spaces Hδk(M) := Wδk,2(M) are Hilbert spaces and their definition does not depend on the choice of the base point defining the weight function ρ. All these definitions extend to Riemannian vector bundles with a metric connection in an obvious manner.

In the literature, there are different notational conventions for weighted spaces. We follow the convention of [Bar86]. The Laplacian ∆g is a bounded map ∆g :Wδp,k(M)→Wδ−2p,k−2(M) and there exists a discrete set D ⊂ R such that this operator is Fredholm for δ ∈ R\D.

This is shown in [Bar86] in the asymptotically flat case and the proof in the ALE case is the same up to minor changes. We call the values δ ∈D exceptional and the values δ ∈ R\D nonexceptional. These properties also hold for elliptic operators of arbitrary order acting on vector bundles supposed that the coefficients behave suitable at infinity [LM85, Theorem 6.1].

We will use these facts frequently in this paper.

2.2. The space of gauged ALE Ricci-flat metrics. Fix an ALE Ricci-flat manifold (M, g0). Let M be the space of smooth metrics on the manifold M. For g ∈ M, let V = V(g, g0) be the vector field defined intrinsically by (3) and locally given bygij(Γ(g)kij−Γ(g0)kij) where Γ(g)kij denotes the Christoffel symbols associated to the Riemannian metric g. We call a metricg gauged, if V(g, g0) = 0. Let

F =

g∈ M | −2 Ric(g) +LV(g,g0)g= 0 , (4)

(5)

be the set of stationary points of the Ricci-DeTurck flow. In local coordinates, the above equation (4) can also be written as

0 = gabgab0,2gij −gklgip(g0)pqRm(g0)jklq −gklgjp(g0)pqRm(g0)iklq

+gabgpq 1

2∇gi0gpagj0gqb+∇ga0gjpgq0gib

−gabgpq

ga0gjpgb0giq− ∇gj0gpagb0giq− ∇gi0gpagb0gjq

,

see [Shi89, Lemma 2.1]. By defining h=g−g0, this equation can be again rewritten as 0 =gabgab0,2hij +habgka(g0)lbgip(g0)pqRm(g0)jklq +habgka(g0)lbgjp(g0)pqRm(g0)iklq

+gabgpq 1

2∇gi0hpagj0hqb+∇ga0hjpgq0hib

−gabgpq

ga0hjpgb0hiq− ∇gj0hpagb0hiq− ∇gi0hpagb0hjq ,

(5)

where we used thatg0 is Ricci-flat. The linearization of this equation atg0 is given by d

dt|t=0(−2Ricg0+th+LV(g0+th,g0)(g0+th)) =Lg0h.

A proof of this fact can be found for instance in [Bam11, Chapter 3].

We recall the well-known fact that the L2-kernel of the Lichnerowicz operator consists of transverse traceless tensors:

Lemma 2.5. Let(Mn, g)be an ALE Ricci-flat manifold andh∈kerL2(Lg0). Then,trg0h= 0 anddivg0h= 0.

Proof. Straightforward calculations show that trg0◦Lg0 = ∆g0 ◦trg0 and divg0◦Lg0 = ∆g0 ◦ divg0. Therefore, trg0h ∈ kerL2(∆g0) and divg0h ∈ kerL2(∆g0) which implies the statement

of the lemma.

The next proposition ensures that ALE steady Ricci solitons are Ricci-flat:

Proposition 2.6. Let (Mn, g, X) be a steady Ricci soliton, i.e. Ric(g) = LX(g) for some vector field X onM. Then lim+∞|X|g = 0 implies X = 0. In particular, any steady soliton in the sense of (4) that is ALE with lim+∞V(g, g0) = 0 is Ricci-flat.

Proof. By the contracted Bianchi identity, one has:

1

2∇gRg= divgRic(g) = divgLX(g) = 1

2∇gtrg(LX(g)) + ∆gX+ Ric(g)(X)

= 1

2∇gRg+∆gX+ Ric(g)(X).

Therefore, ∆gX+ Ric(g)(X) = 0. In particular,

g|X|2g+X· |X|2g= 2|∇gX|2g+ 2<∇gXX, X >g −2 Ric(g)(X, X) = 2|∇gX|2g,

which establishes that |X|2g is a subsolution of ∆X := ∆g +X·. The use of the maximum

principle then implies the result in case lim+∞|X|= 0.

Theorem 2.7. Let (Mn, g0) be an ALE Ricci-flat manifold with order τ >0. Let g ∈ F be in a sufficiently small neighbourhood of g0 with respect to the L2∩L-topology. Then gis an ALE Ricci-flat manifold of order n−1 with respect to the same coordinates as g0.

(6)

Remark 2.8. (i) If n = 4 or g0 is K¨ahler, it seems likely that g ∈ F is ALE of order n with respect to the same coordinates as g0. However, we don’t need this decay for further considerations.

(ii) A priori, Proposition 2.7 does not assume any integral or pointwise decay on the difference tensor g−g0 or on the curvature tensor of g. The assumptions on g can be even weakened as follows: If kg−g0kLp(g0) ≤ K < ∞ for some p ∈ [2,∞) and kg−g0kL(g0)< =(g0, p, K), then the conclusions of Theorem 2.7 hold.

Proof of Theorem 2.7. The first step consists in applying a Moser iteration to the norm of the difference of the two metrics : |h|g0 :=|g−g0|g0. Indeed, recall thath satisfies (5) which can also be written as

g−1∗ ∇g0,2h+ 2 Rm(g0)∗h=∇g0h∗ ∇g0h, g−1∗ ∇g0,2hij :=gklgkl0,2hij.

In particular,

g−1∗ ∇g0,2|h|2g0 = 2g−1(∇g0h,∇g0h)−4hRm(g0)∗h, hig0 +h∇g0h∗ ∇g0h, hig0, g−1(∇g0h,∇g0h) := gijgi0hklgj0hkl.

Therefore, as khkL(g0) ≤where >0 is a sufficiently small constant depending onn and g0, we get

g−1∗ ∇g0,2|h|2g

0 ≥ −c(n)|Rm(g0)|g0|h|2g

0.

As |Rm(g0)| ∈ Ln/2(M) and h ∈ L2(S2TM), Lemma 4.6 and Proposition 4.8 of [BKN89]

tell us that |h|2 = O(r−τ) at infinity for any positive τ < n−2, i.e. h = O(r−τ) for any τ < n/2−1. Here, r denotes the distance function on M centered at some arbitrary point x∈M.

The next step is to show that ∇g0h=O(r−τ−1) forτ < n/2−1. By elliptic regularity for weighted spaces and by interpolation inequalities,

khkW2,p

−τ(g0) ≤C(

g−1∗ ∇g0,2h

Lp−τ−2(g0)+khkLp

−τ(g0))

≤C(

|∇g0h|2

Lp−τ−2(g0)+khkLp

−τ(g0))

≤C(

g0,2h

Lp−τ−2(g0)+k∇g0hkLp

−τ−1(g0))khkL(g0)+khkLp

−τ(g0)), which implies

khkC1,α

−τ(g0)≤CkhkW2,p

−τ(g0)≤CkhkLp

−τ(g0)≤CkhkL

−τ(g0)<∞

for allp∈(n,∞) andτ+ < n2−1 provided that theL-norm is small enough. Consequently,

g0h=O(r−1−τ).

In the following we will further improve the decay order and show that h = O(r−n+1).

As a consequence of elliptic regularity for weighted H¨older spaces, we will furthermore get

g0,kh = O(r−n+1−k) for any k ∈ N. To prove these statements we adapt the strategy in [BKN89, p. 325-327].

As Rm(g0) =O(r−n−1),h=O(r−τ) and∇g0h=O(r−τ−1) for some fixedτ slightly smaller than n2 −1, equation (5) implies that ∆g0h =O(r−2τ−2). Thus, ∆g0h =O(r−µ) for some µ slightly smaller than n.

(7)

Let ϕ :M \BR(x) → (Rn\BR)/Γ be coordinates at infinity with respect to which g0 is ALE of order n−1. Let furthermore Π : Rn\BR → (Rn\BR)/Γ be the projection map.

From now on, we consider the objects onM as objects on Rn\BR by identifying them with the pullbacks under the map Π◦ϕ−1. To avoid any confusion, we denote the pullback of ∆g0

by ∆. ∆0 denotes the Euclidean Laplacian onRn\BR.

Letr0 =|z|be the euclidean norm as a function onRn\BR. Then we have, for anyβ∈R,

∆r−β0 = ∆0r0−β+ (∆−∆0)r−β0 =β(β−n+ 2)r−β−20 +O(r0−β−n−1).

Letu=hij for any i, j. For any constantA >0, we have

∆(A·r−β0 ±u) = (A·β(β−n+ 2) +C1·r0−n+1+C2r−µ+β+20 )r−β−20 . If we chooseβ >0 so thatβ+ 2< µ < n, we can choose A so large that

∆(A·r−β0 ±u)<0 A·r0−β±u >0 if r0 =R.

The strong maximum principle then implies thatu=O(r−β0 ) and we also get ∆u=O(r0−β−2).

By elliptic regularity, u ∈ W−β2,p(Rn\ BR). By (5), ∆0u ∈ Lp−2β−2(Rn \BR) and for all nonexceptional n−1 < γ < 2β, there exists a function vijγ ∈ W−γ2,p(Rn \BR) such that

0vijγ = ∆0hij. By the expansion of harmonic functions onRn, we have hij =vijγ +Aijr−n+20 +O(r0−n+1) =Aijr−n+20 +O(r0−n+1),

where the decay ofvijγ follows from Sobolev embedding. Proposition 2.6 now implies that the equations Ric(g) = 0 and V(g, g0) = 0 hold individually. Therefore,

0 =gij(Γ(g)kij−Γ(g0)kij) = (Az−1

2tr(A)z)k(2−n)|z|−n+O(r−n0 ),

which implies that Aij = 0 and thus, h = O(r−n+1). As hij = vijγ +O(r−n+10 ) with vγij ∈ W−γ2,p(Rn\BR) and an harmonic remainder term, Sobolev embedding and elliptic regularity implies that hij ∈ C1−n1,α (Rn\BR), so that ∇g0h = O(r−n). Elliptic regularity for weighted H¨older spaces implies that∇g0,kh=O(r−n+1−k) for all k∈N. Remark 2.9. By almost the same proof as above, one shows that h = O(r−n−1) if h ∈ kerL2(Lg0). To proveAij = 0 in this case, one uses the condition 0 = divg0h−12trg0h which is guaranteed by Lemma 2.5.

Theorem 2.10. Let (M, g0) be an ALE Ricci-flat metric and F as above. Then there exists anL2∩L-neighbourhoodU ofg0in the space of metrics and a finite-dimensional real-analytic submanifold Z ⊂ U withTg0Z =kerL2(Lg0) such that U ∩ F is an analytic subset of Z. In particular if g0 is integrable, we haveU ∩ F =Z.

Proof. Let Φ : g 7→ −2 Ric(g) +LV(g,g0)(g) and BL2∩L(g0, ) be the -ball with respect to theL2∩L-norm induced byg0 and centered atg0. By Theorem 2.7, we can choose >0 be so small that any g ∈ F ∩ BL2∩L(g0, ) satisfies the conditiong−g0 =O(r−n+1) so that kg−g0kHk

δ <∞ for any k∈Nand δ >−n+ 1.

Suppose now in addition thatk > n/2 + 2 andδ≤ −n/2 and letV be aHδk-neighbourhood of g0 with V ⊂ BL2∩L(g0, 1). Then the map Φ, considered as a map Φ : Hδk(S2TM) ⊃ V →Hδ−2k−2(S2TM) is a real analytic map between Hilbert manifolds. Ifδ is nonexceptional, the differential dΦg0 =Lg0 :Hδk(S2TM)→Hδ−2k−2(S2TM) is Fredholm. By [Koi83, Lemma

(8)

13.6], there exists (possibly after passing to a smaller neighbourhood) a finite-dimensional real- analytic submanifoldW ⊂ V withg0 ∈ W andTg0W = kerHk

δ(Lg0) such thatV ∩Φ−1(0)⊂ W is a real-analytic subset.

By the proof of Theorem 2.7, we can choose anL2∩L-neighbourhoodUg0 ⊂ BL2∩L(g0, ) of g0 so small thatUg0∩ F ⊂ V (provided thatV is small enough). Then the setZ =Ug0∩ W fulfills the desired properties because Tg0Z = Tg0W = kerHk

δ(Lg0) = kerL2(Lg0). Here, the

last equation holds by Remark 2.9.

Proposition 2.11. Let (Mn, g0) be an ALE Ricci-flat manifold and let k > n/2 + 1 and δ ≤ −n/2 nonexceptional. Then there exists a Hδk-neighbourhood Uδk of g0 in the space of metrics such that the set

Gδk:=n

g∈ Uδk|gij(Γ(g)kij −Γ(g0)kij) = 0o ,

is a smooth manifold. Moreover, for any g ∈ Uδk, there exists a unique diffeomorphism ϕ which is Hδ+1k+1-close to the identity such that ϕg∈ Gδk.

Proof. Let U be a Hδk-neighbourhood of g0 in the space of metrics such that the map V : Hδk(S2TM) ⊃ U → Hδ−1k−1(T M), given by V(g)k = V(g, g0)k = gij(Γ(g)kij −Γ(g0)kij) is well-defined. Linearization at g0 yields the map F :Hδk(S2TM)→ Hδ−1k−1(T M), defined by F(h) = (divg0h)]12g0trg0h. To prove the theorem, it suffices to prove that F is surjective and that the decomposition

Hδk(S2TM) = kerF⊕L(g0)(Hδ+1k+1(T M))

holds (here, LX(g0) denotes the Lie-Derivative of g0 along X). In fact, a calculation shows that F ◦ L(g0) = ∆g0 + Ric(g0)(·) = ∆g0 since g0 is Ricci-flat. Since the map ∆g0 : Hδ+1k+11M) → Hδ−1k−11M) is an isomorphism, it follows that F is surjective and kerF ∩ L(g0)(Hδ+1k+1(T M)) ={0}. To show that

Hδk(S2TM)⊂kerL2(F)⊕L(g0)(Hδ+1k+1(T M)),

leth∈Hδk(S2TM) and X∈Hδ+1k+1(T M) the unique solution ofF(h) = ∆g0X =F(LX(g0)).

Then,h = (h−LX(g0)) +LX(g0) is the desired decomposition. By surjectivity of F,Gδk is a manifold. The second assertion follows because the map

Φ :Gδk×Hδ+1k+1(Diff(M))→ Mkδ =:M ∩Hδk(S2TM), (g, ϕ)7→ϕg

is a local diffeomorphism around g0 due to the implicit function theorem and the above

decomposition.

Remark 2.12. The construction in Proposition 2.11 is similar to the slice provided by Ebin’s slice theorem [Ebi70] in the compact case. The set F is similar to the local premoduli space of Einstein metrics defined in [Koi83, Definition 2.8]. In contrast to the compact case, the elements in F close to g0 can all be homothetic. In fact, this holds for the Eguchi-Hanson metric, see [Pag78]. More generally, any four-dimensional ALE hyperk¨ahler manifold (M, g) admits a three-dimensional subspace of homothetic metrics inF: see [Via14, p. 52–53].

(9)

2.3. ALE Ricci-flat K¨ahler spaces.

Lemma 2.13 (∂∂-Lemma for ALE manifolds).¯ Let (M, g, J) be an ALE K¨ahler manifold, δ≤ −n/2 nonexceptional, k≥1 andα∈Hδkp,qM). Suppose that

• α=∂β for some β∈Hδ+1k+1p−1,qM) and ∂α¯ = 0 or

• α= ¯∂β for some β∈Hδ+1k+1p,q−1M) and ∂α= 0.

Then there exists a formγ ∈Hδ+2k+2p−1,q−1M) such thatα=∂∂γ. Moreover, we can choose¯ γ to satisfy the estimate kγkHk+2

δ+2

≤C· kαkHk

δ. for someC >0.

Proof. This follows along the lines of Lemma 5.50 in [Bal06]. Let d = ∂ or d = ¯∂ and

∆ = ∆ = ∆¯ Consider ∆ as an operator ∆ : HδkM) → Hδ−2k−2M). Because of the assumption onδ, it is Fredholm and we have theL2-orthogonal decomposition

HδkM) = kerL2(∆)⊕∆(Hδ+2k+2M)).

We define the Green’s operator G to be zero on kerL2(∆) and to be the inverse of ∆ on kerL2(∆). This defines a continuous linear operator G : HδkM) → Hδ+2k+2M). By Hodge theory and becaused+d :Hδ+1k+1M)→HδkM) is also Fredholm,

d(Hδ+1k+1M))⊕d(Hδ+1k+1M)) = ∆(Hδ+2k+2M)).

and it is straightforward to see that G is self-adjoint and commutes with d and d. As in Ballmann’s book, one shows thatγ =−∂G∂¯Gα does the job in both cases. The estimate

onγ follows from construction.

Let (M, g, J) be a K¨ahler manifold. An infinitesimal complex deformation is an endomor- phismI :T M → T M that anticommutes withJ and satisfies ¯∂I = 0 and ¯∂I = 0. By the relationIJ+J I= 0, I can be viewed as a section of Λ0,1M⊗T1,0M.

Theorem 2.14. Let (Mn, g, J)be an ALE K¨ahler manifold with a holomorphic volume form, k > n/2 + 1, δ ≤ −n/2 nonexceptional and I ∈Hδk0,1M⊗T1,0M) such that ∂I¯ = 0 and

∂¯I = 0. Then there exists a smooth family of complex structures J(t) with J(0) =J such thatJ(t)−J ∈Hδk(TM⊗T M) and J0(0) =I.

Proof. The proof follows along the lines of Tian’s proof by the power series approach [Tia87]:

We write J(t) =J(1−I(t))(1 +I(t))−1, where I(t) ∈ Hδk0,1M ⊗T1,0M) and I(t) has to solve the equation

∂I¯ (t) +1

2[I(t), I(t)] = 0,

where [., .] denotes the Fr¨olicher-Nijenhuis bracket. If we writeI(t) as a formal power series I(t) =P

k≥1Iktk, the coefficients have to solve the equation

∂I¯ N+ 1 2

N−1

X

k=1

[Ik, IN−k] = 0,

inductively for allN ≥2. As Λn,0M is trivial, there is a natural identification of the bundles Λ0,1M ⊗T1,0M = Λn−1,1M by using the holomorphic volume form and we now think of the Ik as being (n−1,1)-forms. Initially, we have chosen I1 ∈ Hδk0,1M ⊗T1,0M), given by I = 2I1J. By the multiplication property of weighted Sobolev spaces [CB09, p. 538], [I1, I1]∈Hδ−1k−1n−1,2M). Using ∂I1 = 0 and ¯∂I1 = 0, one can now show that ¯∂[I1, I1] = 0

(10)

and [I1, I1] is ∂-exact. The ∂∂-lemma now implies the existence of a¯ ψ ∈ Hδ+1k+1n−2,1M) such that

∂∂ψ¯ =−1

2[I1, I1],

and so, I2 =∂ψ∈Hδkn−1,1M) does the job. Inductively, we get a solution of the equation

∂∂ψ¯ = 1 2

N−1

X

k=1

[Ik, IN−k],

by the ∂∂-lemma since the right hand side is ¯¯ ∂-closed and ∂-exact (which in turn is true because∂Ik= 0 for 1≤k≤N−1). Now we can chooseIN =∂ψ ∈Hδkn−1,1M).

Let us prove the convergence of the above series: Let D1 be the constant in the estimate of the∂∂-lemma and¯ D2 be the constant such that

k[φ, ψ]kHk−1 δ−1

≤D2kφkHk δ kψkHk

δ . Then one can easily show by induction that

kINkHk δ

≤C(N)·[1

2D1·D2]N−1(kI1kHk δ)N

forN ≥1, whereC(N) is the sequence defined byC(1) = 1 andC(N) =PN−1

i=1 C(i)·C(N−i) for N >1. By defining D:= 2/(D1·D2) ands= 12D1·D2· kI1kHk

δ ·t, we get kI(t)kHk

δ

X

i=1

kIikHk

δ ti≤D·

X

i=1

C(i)·si =D· 1 2 −

r1 4 −s

!

ifs <1/4 which shows that the series converges. Thus I(t)∈Hδkn−1,1M) and J(t)−J =

−2J I(t)(1 +I(t))−1 ∈Hδkn−1,1M)∼=Hδk0,1M ⊗T1,0M).

The proof of the above theorem provides an analytic immersion Θ :Hδk0,1M⊗T1,0M)∩ kerL2(∆)⊃U → Hδk(TM⊗T M) whose image is a smooth manifold of complex structures which we denote byJδk and whose tangent map at J is just the injection.

Proposition 2.15. Let(M, g0, J0)be an ALE Calabi-Yau manifold,δ <2−nnonexceptional and Jδk be as above. Then there exists a Hδk-neighbourhood U of J and a smooth map Φ : Jδk∩ U → Mkδ which associates to each J ∈ Jδk∩ U sufficiently close to J0 a metric g(J) which is Hδk-close to g0 and K¨ahler with respect to J. Moreover, we can choose the map Φ such that

J0(I)(X, Y) = 1

2(g0(IX, J0Y) +g0(J0X, IY)).

Proof. We adapt the strategy of Kodaira and Spencer [KS60, Section 6]. Let Jt be a family in Jδk and define Jt-hermitian forms ωt by Π1,1t ω0(X, Y) = 120(X, Y) + ω0(JtX, JtY)).

Let∂t,∂¯t the associated Dolbeaut operators and ∂t,∂¯t their formal adjoints with respect to the metric gt(X, Y) := ωt(X, JtY). We now define a forth-order linear differential operator Et:Hδkp,qt M)→Hδ−4k−4p,qt M) by

Et=∂t∂¯t∂¯tt+ ¯∂ttt∂¯t+ ¯∂ttt∂¯t+∂t∂¯t∂¯tt+ ¯∂t∂¯t+∂tt.

It is straightforward to see that Et is formally self-adjoint and strongly elliptic. Moreover, α ∈ kerHk

δ(Et) if and only if ∂tα = 0, ¯∂tα = 0 and ¯∂ttα = 0, i.e. dα = 0 and ¯∂ttα = 0 hold simultaneously. Ifδ is nonexceptional,Et is Fredholm which allows to define for each t

(11)

its Greens operatorGt:Hδ−4k−4p,qt M)→Hδkp,qt M). As in [KS60, Proposition 7], one now shows that

kerL2(d)∩Hδkp,qt M) =∂t∂¯t(Hδ+2k+2p−1,q−1t M))⊕kerL2(Et)∩Hδkp,qt M),

is anL2(gt) orthogonal decomposition. The dimension of kerL2(Et)∩Hδk1,1t M) is constant for small t which implies thatGt depends smoothly on t. The proof of this fact is exactly as in [KS60, Proposition 8].

Now observe thatEtωt∈Hδ−4k−41,1t M) ifωtandJtareHδk-close toω0 andJ0, respectively.

This allows us to define

˜

ωt:=ωt−GtEtωt+∂t∂¯tut= (1−GtEt1,1t ω0+∂t∂¯tut,

where ut∈Hδ+2k+2(M) is a smooth family of functions such thatu0 = 0 which will be defined later. Clearly,

¯

ωt:=ωt−GtEtωt∈kerEt.

As ¯ωt is Hδk-close to ω0, ∇gtω¯t ∈ Hδ−1k−1(gt), since ω0 is g0-parallel. Therefore, ∇gtω¯t = O(r−α−1) and ∇gt,2ω¯t = O(r−α−2) for any α < −δ. Thus, if we choose the nonexceptional value δ so that δ <−n+ 2, integration by parts implies that

k∂tω¯tk2L2(gt)+ ∂¯tω¯t

2

L2(gt)≤(Etω¯t,ω¯t)L2(gt)= 0.

Therefore, ¯ωt and hence also ˜ωt is closed. Differentiating at t= 0 yields

˜

ω00 = (1−G0E000−G0(E00ω0) +∂0∂¯0u0000 −G0(E00ω0) +∂0∂¯0u00

Because d˜ωt = 0, we have d˜ω00 = 0 and since J00 is an infinitesimal complex deformation, E0ω00 = 0 anddω00 = 0 which implies that

G0(E00ω0)∈kerL2(E0)∩kerL2(d)∩Hδk1,10 M) =∂0∂¯0(Hδ+2k+2(M)).

Let nowv∈Hδ+2k+2(M) so that ∂0∂¯0v=G0(E00ω0).Then, defineut∈Hδ+2k+2(M) by ut:=tv.

By this choice, ˜ω00 = ω00 and the assertion for dΦJ0(J00) = ˜g00 follows immediately. Finally,

˜

gt(X, Y) := ˜ωt(X, JtY) is a Riemannian metric for t small enough and it is K¨ahler with respect toJt.

Remark 2.16. Let Jt is a smooth family of complex structures in Jδk∩ U and gt = Φ(Jt).

Then the construction in the proof above shows thatI =J00 and h=g00 are related by h(J X, Y) =−1

2(g(X, IY) +g(IX, Y)).

Before we state the next theorem, recall the notationGδk we used in Proposition 2.11.

Theorem 2.17. Let (M, g0, J0) be an ALE Calabi-Yau manifold and δ ∈ (1−n,2 −n) nonexceptional. Then for anyh ∈kerL2(Lg0), there exists a smooth family g(t) of Ricci-flat metrics in Gδk with g(0) =g0 and g00 = h. Each metric g(t) is ALE and K¨ahler with respect to some complex structure J(t) which isHδk-close toJ0. In particular,g0 is integrable.

(12)

Proof. We proceed similarly as in [Bes08, Chapter 12], except the fact that we use weighted Sobolev spaces. Given a complex structureJ close to J0 and aJ-(1,1)-form ω which is Hδk- close to ω0, we seek a Ricci-flat metric in the cohomology class [ω]∈ H1,1J (M). As the first Chern class vanishes, there exists a functionfω ∈Hδk(M), such that i∂∂f¯ ω is the Ricci form ofω. If ¯ω∈[ω] and ¯ω−ω∈Hδk1,1J M), the∂∂-lemma implies that there is a¯ u∈Hδ−2k−2(M) such that ¯ω=ω+i·∂∂u. Ricci-flatness of ¯¯ ω is now equivalent to the condition

fω = log(ω+i∂∂u)¯ n

ωn =:Cal(ω, u).

LetJδk be as above and ∆J the Dolbeaut Laplacian of J and the metric g(J). Then all the (L2δ)-cohomologies H1,1J,δ(M) = kerL2

δ(∆J)∩L2δ1,1M) are isomorphic for J ∈ Jδk if we Jδk is small enough: We have H2δ(M) = H2,0J,δ(M)⊕ H1,1J,δ(M)⊕ H0,2J,δ(M). The left hand side is independent of J and the metric g(J) is provided by Proposition 2.15. The spaces on the right hand side are kernels ofJ-dependent elliptic operators whose dimension depends upper- semicontinuously on J. However the sum of the dimensions is constant and so the dimensions must be constant as well.

Thus, there is a natural projectionprJ : kerL2(∆J0)→kerL2(∆J) which is an isomorphism.

We now want to apply the implicit function theorem to the map G:Jδk× HJ1,1

0(M)×Hδk(M)→Hδ−2k−2(M)

(J, κ, u)7→Cal(ω(J) +prJ(κ), u)−fω(J)+prJ(κ),

where ω(J)(X, Y) := g(J)(J X, Y) and g(J) is the metric constructed in Proposition 2.15.

We haveG(J0,0,0) = 0 and the differential restricted to the third component is just given by

∆ :Hδk(M) → Hδ−2k−2(M), which is an isomorphism [Bes08, p. 328], therefore we find a map Ψ such thatG(J, κ,Ψ(J, κ)) = 0.

Let now h ∈kerL2(Lg0) and let h =hH+hA its decomposition into a J0-hermitian and a J0-antihermitian part. We want to show that h is tangent to a family of Ricci-flat metrics.

We have seen in Theorem 2.7 that h ∈ Hδk(S2TM) for all δ > 1−n) and we can define I ∈Hδk(TM⊗T M) and κ∈Hδk1,1J0)(M) by

g(X, IY) =−hA(X, J0Y), κ(X, Y) =hH(J0X, Y). (6) It is easily seen that I is a symmetric endomorphism satisfying IJ0+J0I = 0 and thus can be viewed as I ∈ Hδk0,1M ⊗T1,0M). Moreover, because hA is a T T-tensor, ¯∂I = 0 and

∂¯I = 0. In additionκ∈ HJ1,1

0(M). The proof of this facts is as in [Koi83]. LetJ(t) = Θ(t·I) be a family of complex structures tangent to I and ˜ω(t) = ˜Φ(J(t)) be the associated family of K¨ahler forms. We consider the family ω(t) = ˜ω(t) +prJ(t)(t·κ) +i∂∂Ψ(J¯ (t), t·κ) and the associated family of Ricci-flat metrics ˜g(t)(X, Y) =ω(t)(X, J(t)Y). It is straightforward that ˜g0(0) =h. By Proposition 2.11, there exist diffeomorphisms ϕt withϕ0 =id such that g(t) = ϕtg(t)˜ ∈ Gδk. We obtain g0(0) = h+LXg0 for some X ∈ Hδ+1k+1(T M). Since h is a TT-tensor due to Lemma 2.5, h ∈ Tg0Gkδ. On the other hand, g0(0) ∈ Tg0Gδk as well which implies thatg0(0) =h due to the decomposition in Proposition 2.11.

By Theorem 2.10, the set of stationary solutions of the Ricci-DeTurck flow F close to g0 is an analytic set contained in a finite-dimensional manifoldZ withTg0Z = kerL2(Lg0). The above construction provides a smooth map Ξ : kerL2(Lg0) ⊃ U → F ⊂ Z whose tangent

Références

Documents relatifs

The solution of the geodesic equation (Theorem 4.1) gives a precise sense to how geodesics in ( M , g N ) generalize those discovered by Calabi [4,5] for the subspace of Kähler

We characterize the conjugate linearized Ricci flow and the associated backward heat kernel on closed three-manifolds of bounded geometry.. We dis- cuss their properties, and

In the remaining of this section, we determine when this criterion is satisfied in the case where the equation exactly encodes the existence of a (singular) Kähler-Einstein metrics on

In Section 5, we build an asymptotic solution to the Ricci flat equation on a rank two symmetric space, using as essential ingredients Stenzel’s metrics and the positive

In fact, on a manifold with a cylindri- cal end, the Gauss–Bonnet operator is non-parabolic at infinity; harmonic forms in the Sobolev space W are called, by Atiyah–Patodi–Singer, L

Without the compactness assumption, irreducible K¨ ahler manifolds with Ricci tensor having two distinct constant eigenvalues are shown to exist in various situations: there

In a neighborhood of c , com- puting distances using the metric measured at c is a better approximation of the perceptual distance than using the canonical euclidean distance of the

In particular, since there is an infinite-dimensional space of harmonic functions on any nonempty open subset M of R 3 , there are many nonhomothetic conformal metrics on such M