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

On conical asymptotically flat manifolds

Mingyang Li Department of Mathematics, University of California, Berkeley, CA 94720 mingyang_li@berkeley.edu  and  Song Sun Institute for Advanced Study in Mathematics, Zhejiang University, Hangzhou 310058, China Department of Mathematics, University of California, Berkeley, CA 94720 songsun@zju.edu.cn Dedicated to Professor Xiaochun Rong for his 70th birthday
Abstract.

We prove a conjecture of Petrunin and Tuschmann on the non-existence of asymptotically flat 4-manifolds asymptotic to the half plane. We also survey recent progress and questions concerning gravitational instantons, which serve as our motivation for studying this question.

1. Introduction

In this paper we study conical asymptotically flat (𝒜\mathcal{AF}) manifolds. By definition, a complete non-compact mm dimensional Riemannian manifold (Mm,g,p)(M^{m},g,p) is called

  • 𝒜\mathcal{AF} if the Riemannian curvature decays at the rate |Rmg|=o(r2)|Rm_{g}|=o(r^{-2}) as rr\rightarrow\infty, where rr denotes the distance to pp; in other words, the asymptotic curvature

    A(M,g)lim suprr2|Rmg|A(M,g)\equiv\limsup_{r\rightarrow\infty}r^{2}|Rm_{g}|

    is zero.

  • conical if it is asymptotic to a unique metric cone 𝒞\mathcal{C}_{\infty} at infinity, i.e., (M,λ2g,p)(M,\lambda^{-2}g,p) converges in the pointed Gromov-Hausdorff topology to 𝒞\mathcal{C}_{\infty} as λ\lambda\rightarrow\infty. Here the asymptotic cone 𝒞\mathcal{C}_{\infty} may have a lower dimension.

We need to make a few remarks about the terminologies here. First the words “asymptotically flat” may refer to properties of varying generality in the literature. The notion of 𝒜\mathcal{AF} we use here is the same as the notion of asymptotically flat used in [35]. A related notion is what we shall call strongly 𝒜\mathcal{AF}, which means that |Rmg||Rm_{g}| is bounded by a positive decreasing function f(r)f(r) with 0rf(r)𝑑r<\int_{0}^{\infty}rf(r)dr<\infty. An even stronger condition is that of faster than quadratic curvature decay, which means that |Rmg|r2ϵ|Rm_{g}|\leq r^{-2-\epsilon} for some ϵ>0\epsilon>0 as rr\rightarrow\infty. We also try to distinguish the notation 𝒜\mathcal{AF} from AF, which usually refers to being asymptotic to a specific flat model end (see Section 4).

There has been extensive work in Riemannian geometry studying topological and geometrical properties of 𝒜\mathcal{AF} manifolds. Recall that flat ends of Riemannian manifolds were completely classified by Eischenburg-Schroeder [18]. Ends of 𝒜\mathcal{AF} manifolds are much more flexible; for example, any 2 dimensional non-compact surface admits an 𝒜\mathcal{AF} metric [1]. It is known [23, 31] that strongly 𝒜\mathcal{AF} manifolds are automatically conical. However there are interesting examples of conical 𝒜\mathcal{AF} manifolds which are not strongly 𝒜\mathcal{AF}. Examples include the family of ALG hyperkähler metrics [21].

Our interest in 𝒜\mathcal{AF} manifolds is motivated by the fact that they provide potential model ends for complete Ricci-flat metrics. In 4 dimensions a complete non-compact Ricci-flat manifold with |Rmg|2<\int|Rm_{g}|^{2}<\infty is called a gravitational instanton. The readers should be warned that there are varying definitions of gravitational instantons in the literature. Some require the extra assumption of being hyperkähler, but we do not make this hypothesis in the current paper. All known examples of gravitational instantons are conical and they are all 𝒜\mathcal{AF} except the family of ALH hyperkähler metrics (see [21, 38]). The latter are known to have nilptoent asymptotic geometries.

For general conical 𝒜\mathcal{AF} manifolds, in 2001 Petrunin and Tuschmann [35] proved a structural result regarding the end structure. In particular, they proved that there are only finitely many ends and a simply connected end must have asymptotic cone the flat m\mathbb{R}^{m} if m4m\neq 4 and the flat 4,3\mathbb{R}^{4},\mathbb{R}^{3} or the half plane ×[0,)\mathbb{H}\equiv\mathbb{R}\times[0,\infty) when m=4m=4. It follows that when m4m\neq 4 the metric has Euclidean volume growth and there is no collapsing phenomena along the convergence to the asymptotic cone. But when m=4m=4 collapsing may indeed occur so the situation is more interesting. Clearly 4\mathbb{R}^{4} can be realized. Also 3\mathbb{R}^{3} can be realized as the asymptotic cone of the Taub-NUT gravitational instanton (see also [39]), whose end is diffeomorphic to S3×[0,)S^{3}\times[0,\infty). Along the convergence to the asymptotic cone 𝒞\mathcal{C}_{\infty}, the S3S^{3} collapses along the Hopf fibration to S2S^{2}. For the half plane \mathbb{H}, we know that it can be realized as the asymptotic cone of the flat space 4/\mathbb{R}^{4}/\mathbb{Z} (and of some members of the Kerr family of gravitational instantons), where the generator of the \mathbb{Z}-action is given by a translation in the first 2\mathbb{R}^{2} factor and an irrational rotation in the second 2\mathbb{R}^{2} factor. The end is diffeomorphic to (S1×S2)×[0,)(S^{1}\times S^{2})\times[0,\infty); in particular it is not simply-connected. Petrunin-Tuschmann further made the following conjecture

Conjecture 1.1 (Petrunin-Tuschmann [35]).

There is no conical 𝒜\mathcal{AF} 4-manifold with a simply-connected end whose asymptotic cone is the half plane.

Notice that there is no curvature equation imposed. It is easy to see that if a conical 𝒜\mathcal{AF} 4-manifold has an end with asymptotic cone \mathbb{H}, then the end must be diffeomorphic to N×[0,)N\times[0,\infty) where NN is diffeomorphic to either S1×S2S^{1}\times S^{2} or S3/ΓS^{3}/\Gamma. So the confirmation of Conjecture 1.1 would give a topological classification of such ends. On the other end, if Conjecture 1.1 is false then potentially there could be new interesting gravitational instantons with this asymptotics.

The results in [35] were proved by investigating the geometry of the rescaled annuli (M,λ2g,p)(M,\lambda^{-2}g,p) as λ\lambda\rightarrow\infty when they collapse to lower dimensions. Foundational results on collapsing were obtained by Cheeger, Fukaya, Gromov, and others (see for example [7, 36]). Specifically, [35] makes use of the continuously collapsing techniques developed by Petrunin-Rong-Tuschmann [34]. In particular, the argument is of local nature, i.e., it focuses on the local collapsing geometry of the family of almost flat annuli of fixed size (with respect to the rescaled metrics), and it does not use the fact that the annuli come from the end of a fixed 𝒜\mathcal{AF} manifold. In fact as mentioned in [35] (see also Section 3), one may construct a sequence of Riemannian metrics gig_{i} on 𝔄=S3×[0,1]\mathfrak{A}=S^{3}\times[0,1] with supA|Rmgi|0\sup_{A}|Rm_{g_{i}}|\rightarrow 0 which collapse to the annulus 𝔸{21r2}\mathbb{A}\equiv\{2^{-1}\leq r\leq 2\} in \mathbb{H}, where rr is the distance function to the origin in \mathbb{H}. Therefore Conjecture 1.1 is of global nature. We remark that even under the stronger assumption of faster than quadratic curvature decay Conjecture 1.1 was unknown.

In this paper we confirm the conjecture of Petrunin-Tuschmann.

Theorem 1.2.

Conjecture 1.1 holds.

We will see that the standard 4\mathbb{R}^{4} admits a complete conical Riemannian metric gjg_{j} with asymptotic curvature A(4,gj)j1A(\mathbb{R}^{4},g_{j})\leq j^{-1} and asymptotic cone \mathbb{H}, but by Theorem 1.2 there is no complete conical 𝒜\mathcal{AF} metric on 4\mathbb{R}^{4} with asymptotic cone \mathbb{H}. This also gives a negative answer to a version of the gap question in [35] (Question 2).

Now we briefly sketch the idea behind the proof of Theorem 1.1. Suppose (M,g)(M,g) is a conical 𝒜\mathcal{AF} 4-manifold with a simply-connected end whose asymptotic cone is \mathbb{H}, we will derive a contradiction. General theory allows us to reduce to the case when the end admits a T2T^{2} action and the metric gg is T2T^{2} invariant. Such a metric gg then yields a family of flat metrics GG on the 2-torus parametrized by (ρ,z)=0×(\rho,z)\in\mathbb{H}=\mathbb{R}_{\geq 0}\times\mathbb{R} for ρ2+z2R2\rho^{2}+z^{2}\geq R^{2}. For fixed zz as ρ0\rho\rightarrow 0 the metric GG degenerates as in a specific model situation that we can understand. Fix an interval I(0,π)I\subset(0,\pi) and consider the sector region z=ρcotθz=\rho\cot\theta for θI\theta\in I. As ρ\rho\rightarrow\infty, the 2-torus endowed with the rescaled metric ρ1G\rho^{-1}G converges to a point in the Gromov-Hausdorff sense. We introduce two quantities σ\sigma and τ\tau to each metric GG. Roughly speaking, σ\sigma measures the area of the 2-torus under the metric ρ1G\rho^{-1}G, and τ\tau measures the deviation of the metric ρ1G\rho^{-1}G from a model flat metric. The fact that the asymptotic cone is \mathbb{H} gives control on the asymptotics of quantities involving τ\tau and σ\sigma as ρ\rho\rightarrow\infty. Then an elementary calculus argument yields a contradiction.

Acknowledgements: Both authors were partially supported by the Simons Collaboration Grant in Special Holonomy. The first author is grateful to the IASM at Zhejiang University for hospitality during his visit in Spring 2024, and would like to thank Prof. Guofang Wei for sharing Abresch’s paper [2]. We thank John Lott for helpful comments that improved the exposition of the paper.

2. Proof of the main result

Suppose (M,g,p)(M,g,p) is a conical 𝒜\mathcal{AF} manifold with a simply-connected end and with asymptotic cone 𝒞=\mathcal{C}_{\infty}=\mathbb{H}. We will derive a contradiction.

First we introduce some notations. Let (ρ,z)(\rho,z) be the standard coordinates on 0×\mathbb{H}\equiv\mathbb{R}_{\geq 0}\times\mathbb{R}. Denote by g0=dρ2+dz2g_{0}=d\rho^{2}+dz^{2} the standard metric on \mathbb{H} and by r0r_{0} the distance function to the origin. For j>0j>0 write AjB¯g(p,2j+1)Bg(p,2j1)A_{j}\equiv\overline{B}_{g}(p,2^{j+1})\setminus B_{g}(p,2^{j-1}) and denote by A~j\widetilde{A}_{j} the manifold AjA_{j} endowed with the rescaled metric gj22jgg_{j}\equiv 2^{-2j}g. By Theorem A of [35] we know AjA_{j} is simply connected for jj large.

As initial steps we make a few reductions using general theory. We first prove a smoothing result.

Lemma 2.1.

For any given k0>0k_{0}>0 there is a conical 𝒜\mathcal{AF} Riemannian metric gg^{\prime} on MM with asymptotic cone \mathbb{H} such that for each kk0k\leq k_{0}, |gkRmg|g=o(rk2)|\nabla_{g^{\prime}}^{k}Rm_{g^{\prime}}|_{g^{\prime}}=o(r^{-k-2}) as r0r\rightarrow 0.

Remark 2.2.

Notice that this higher regularity is automatic if we assume gg satisfies elliptic curvature equations, for example, if we assume gg is Ricci-flat, or if we assume gg is scalar-flat and anti-self-dual (which includes the case of being scalar-flat Kähler).

Proof.

Fix ϵ>0\epsilon>0 small to be determined below. Denote by ηj\eta_{j} the maximum of |Rmgj||Rm_{g_{j}}| on A~jA~j1A~j+1\widetilde{A}_{j}\cup\widetilde{A}_{j-1}\cup\widetilde{A}_{j+1}. By assumption we have ηj0\eta_{j}\rightarrow 0 as jj\rightarrow\infty. By the local smoothing result of Abresch [2] (more precisely, by iterating the proof of Theorem 1.1 in [33] for finitely many steps) to each rescaled annuli A~j\widetilde{A}_{j}, one can find a new Riemannian metric gjg^{\prime}_{j} on A~j\widetilde{A}_{j} with (1ϵ)gjgj(1+ϵ)gj(1-\epsilon)g_{j}\leq g^{\prime}_{j}\leq(1+\epsilon)g_{j}, |Rm(gj)|2ηj|Rm(g^{\prime}_{j})|\leq 2\eta_{j} and |gjkRm(gj)|Ck|\nabla^{k}_{g^{\prime}_{j}}Rm(g^{\prime}_{j})|\leq C_{k} for each kk0k\leq k_{0}. Scale gjg^{\prime}_{j} back to AjA_{j} and glue the adjacent gjg^{\prime}_{j} by suitable cut-off functions (as in [8]) we obtain a metric gg^{\prime} on MM such that (1ϵ)gg(1+ϵ)g(1-\epsilon)g\leq g^{\prime}\leq(1+\epsilon)g and |gkRmg|g=o(r2k)|\nabla^{k}_{g^{\prime}}Rm_{g^{\prime}}|_{g^{\prime}}=o(r^{-2-k}) as rr\to\infty for all kk0k\leq k_{0}.

It remains to show that gg^{\prime} is asymptotic to \mathbb{H} if ϵ\epsilon is chosen to be small. Denote by A~j\widetilde{A}_{j}^{\prime} the manifold AjA_{j} endowed with the rescaled metric gj22jgg^{\prime}_{j}\equiv 2^{-2j}g^{\prime}. Due to the uniform equivalence between gg and gg^{\prime}, by passing to a subsequence we may assume (A~j,gj,p)(\widetilde{A}_{j}^{\prime},g^{\prime}_{j},p) converges to a pointed Gromov-Hausdorff limit space (𝒞,O)(\mathcal{C},O), which is isometric to \mathbb{H} endowed with a possibly different metric dd_{\infty}. In particular 𝒞\mathcal{C} is homeomorphic to \mathbb{H} but a priori we do not know if 𝒞\mathcal{C} is a metric cone.

By the assumption on (M,g)(M,g) locally for any qjA~jq_{j}\in\widetilde{A}_{j} converging to qq_{\infty}\in\mathbb{H} we may find rr small such that the universal cover (B^j,g^j)(\widehat{B}_{j},\widehat{g}_{j}) of BjBgj(qj,r)A~jB_{j}\equiv B_{g_{j}}(q_{j},r)\subset\widetilde{A}_{j} is uniformly non-collapsed and there is an equivariant C1,αC^{1,\alpha} Gromov-Hausdorff convergence of (B^j,g^j,Λj,q^j)(\widehat{B}_{j},\widehat{g}_{j},\Lambda_{j},\widehat{q}_{j}) to a non-collapsing flat limit space (B^,g^,Λ,q^)(\widehat{B}_{\infty},\widehat{g}_{\infty},\Lambda_{\infty},\widehat{q}_{\infty}), where Λj=π1(Bj)\Lambda_{j}=\pi_{1}(B_{j}) such that (B^,g^)/Λ(\widehat{B}_{\infty},\widehat{g}_{\infty})/\Lambda_{\infty} is identified with the ball BBg0(q,r)B_{\infty}\equiv B_{g_{0}}(q_{\infty},r) in \mathbb{H}. We may isometrically immerse B^\widehat{B}_{\infty} into 4\mathbb{R}^{4}. Then the group Λ\Lambda_{\infty} descends to a nilpotent (hence abelian) subgroup Λ\Lambda of the group of Euclidean motions of 4\mathbb{R}^{4}. Since AjA_{j} is simply connected we know by [36] that Λ\Lambda_{\infty} is connected. By the Sublemma in [35] there is an orthogonal decomposition 4=VW\mathbb{R}^{4}=V\oplus W such that ΛRot(W)Trans(V)\Lambda_{\infty}\subset Rot(W)\oplus Trans(V). Using the Key Lemma in [35] (in the statement of the Key Lemma in [35] we know l<ml<m, which is ultimately due to the injectivity radius estimate in [34]) we see that dimW=dimV=2\dim W=\dim V=2, and Λ\Lambda is generated by a rotation in WW and a translation in VV, so it is abstractly isomorphic to S1×S^{1}\times\mathbb{R}. It is also easy to identify the Λ\Lambda for different choices of qq_{\infty}.

Now choose qq_{\infty}\in\partial\mathbb{H}. Then we can identify the corresponding Λ\Lambda_{\infty} with Λ\Lambda. Denote by g^j\widehat{g}_{j}^{\prime} the induced metric on B^j\widehat{B}_{j} from gjg^{\prime}_{j}. Then we may assume the metric g^j\widehat{g}_{j}^{\prime} also converges to a flat metric g^\widehat{g}_{\infty}^{\prime} on B^\widehat{B}_{\infty} which is also Λ\Lambda invariant and the quotient (B^,g^)/Λ(\widehat{B}_{\infty},\widehat{g}_{\infty}^{\prime})/\Lambda_{\infty} is identified with the corresponding ball BB_{\infty}^{\prime} in (,d)(\mathbb{H},d_{\infty}). Similarly by the Sublemma in [35] we may assume that with respect to the metric g^\widehat{g}_{\infty}^{\prime} we also have ΛRot(W)Trans(V)\Lambda\subset Rot(W^{\prime})\oplus Trans(V^{\prime}). Since Λ\Lambda contains a factor isomorphic to S1S^{1} it follows that Λ\Lambda_{\infty} is also generated by a rotation in WW^{\prime} and another element vv in Rot(W)Trans(V)Rot(W^{\prime})\oplus Trans(V^{\prime}). Without loss of generality we may assume vv is a translation in VV^{\prime}. It then follows that BB_{\infty}^{\prime} must be flat hence 𝒞\mathcal{C} is flat and 𝒞\partial\mathcal{C} is totally geodesic (hence a straight line) away from OO. So 𝒞\mathcal{C} is isometric to a wedge. To see 𝒞\mathcal{C} is indeed isometric to \mathbb{H}, we choose a path γ(t)\gamma(t) in the interior of \mathbb{H} connecting qq_{\infty} and q-q_{\infty}. One can cover γ(t)\gamma(t) by finitely many small balls {Dα},α=1,,s\{D_{\alpha}\},\alpha=1,\cdots,s, where qD1q_{\infty}\in D_{1}, qDs-q_{\infty}\in D_{s} and DαDα+1D_{\alpha}\cap D_{\alpha+1}\neq\emptyset. Now each DαD_{\alpha} arises as the quotient of the form D^α,/Λ,α\widehat{D}_{\alpha,\infty}/\Lambda_{\infty,\alpha}, where D^α,\widehat{D}_{\alpha,\infty} is the flat limit of local universal covers of corresponding balls in A~j\widetilde{A}_{j}^{\prime} and Λ,α\Lambda_{\infty,\alpha} can be identified as a subgroup of Rot(Wα)Trans(Vα)Rot(W_{\alpha})\oplus Trans(V_{\alpha}) for an orthogonal decomposition 4=WαVα\mathbb{R}^{4}=W_{\alpha}\oplus V_{\alpha}, under a local isometric immersion of Dα,D_{\alpha,\infty} into 4\mathbb{R}^{4}. Now by considering the intersection DαDα+1D_{\alpha}\cap D_{\alpha+1} and starting from α=1\alpha=1 one can naturally all WαW_{\alpha} and VαV_{\alpha} for different α\alpha, and it follows that Λ,α\Lambda_{\infty,\alpha} is generated by a common rotation and translation. From this it is easy to see that 𝒞\mathcal{C} is isometric to \mathbb{H}.

We will assume from now on that g=gg^{\prime}=g. Next we apply the Cheeger-Fukaya-Gromov theory to reduce the case when the end of gg is T2T^{2} invariant.

Lemma 2.3.

There is an R>0R>0, an effective G=T2G=T^{2} action on MMB(p,R)M^{\circ}\equiv M\setminus B(p,R), a GG-invariant Riemannian metric gˇ\widecheck{g} on MM^{\circ} and a surjective continuous map F:MKF:M^{\circ}\simeq\mathbb{H}^{\circ}\equiv\mathbb{H}\setminus K for some compact KK, such that the following hold

  • gˇ\widecheck{g} is close to gg in the sense that for each kk0k\leq k_{0}, |gk(gˇg)|g=o(rk)|\nabla_{g}^{k}(\widecheck{g}-g)|_{g}=o(r^{-k}) as rr\rightarrow\infty; in particular, gˇ\widecheck{g} is also conical 𝒜\mathcal{AF} with asymptotic cone \mathbb{H}.

  • For each xx\in\mathbb{H}^{\circ}, F1(x)F^{-1}(x) consists of exactly one GG orbit and M/GM^{\circ}/G is a smooth manifold. Moreover, FF induces a diffeomorphism between M/GM^{\circ}/G and \mathbb{H}^{\circ}, and gˇ\widecheck{g} descends to a smooth Riemannian metric gbg_{b} on the quotient M/GM^{\circ}/G (in the sense of manifold with boundary).

  • Denote by PP1P2P\equiv P_{1}\sqcup P_{2} the two boundary components of \mathbb{H}^{\circ}, then the GG action is free over F1(P)F^{-1}(\mathbb{H}^{\circ}\setminus P) and points in F1(Pα)F^{-1}(P_{\alpha}) have stabilizers isomorphic to a 1-dimensional subgroup GαT2G_{\alpha}\subset T^{2} with G1G2={e}G_{1}\cap G_{2}=\{e\}.

  • FF is an asymptotic isometry in the sense that under the above identification between M/GM^{\circ}/G and \mathbb{H}^{\circ}, we have for each kk0k\leq k_{0}, |g0k(gbg0)|g0=o(r0k)|\nabla_{g_{0}}^{k}(g_{b}-g_{0})|_{g_{0}}=o(r_{0}^{-k}) as r0r_{0}\rightarrow\infty.

Proof.

This is essentially a standard application of the Cheeger-Fuakay-Gromov theory, as in [38]. We only give a sketch of the argument here. One first works with the collapsing annuli A~j\widetilde{A}_{j} and apply [7] to obtain an 𝒩\mathcal{N}-structure in A~j\widetilde{A}_{j} (slightly shrinking if necessary) for jj large. The 𝒩\mathcal{N}-structure is pure since we have a diameter bound. By hypothesis each A~j\widetilde{A}_{j} is simply-connected so by Rong [36] the pure 𝒩\mathcal{N}-structure is given by an action of T2T^{2}. Points with codimension 1-stabilizers give rise to boundary in the quotient. Again the simply-connectedness implies that the two boundary components correspond to two different stabilizer groups G1G_{1} and G2G_{2} with G1G2={e}G_{1}\cap G_{2}=\{e\}. Then we can glue together the structures on different A~j\widetilde{A}_{j} to obtain a T2T^{2} action on MM^{\circ} for RR large. The construction and properties of the invariant metric gˇ\widecheck{g} follow from the averaging construction in [7]. ∎

Now we fix k0=10k_{0}=10. Without loss of generality we will assume gˇ=g\widecheck{g}=g in the following. The above discussion only makes use of the topology of the end and of the asymptotic cone. Now we investigate the asymptotical flat geometry in more detail. Introduce the following notations.

  • We write the decomposition G=G1×G2G=G_{1}\times G_{2} and denote by α\partial_{\alpha} the infinitesimal generator of GαG_{\alpha}, normalized to have period 2π2\pi.

  • Fix ϵ>0\epsilon>0 small, we may cover \mathbb{H} by 2 open subsets 𝕌1={0θ<π2+ϵ}\mathbb{U}_{1}=\{0\leq\theta<\frac{\pi}{2}+\epsilon\} and 𝕌2={π2ϵ<θπ}\mathbb{U}_{2}=\{\frac{\pi}{2}-\epsilon<\theta\leq\pi\}. Then we get an open cover of MM^{\circ} by UαF1(𝕌α),α=1,2U_{\alpha}\equiv F^{-1}(\mathbb{U}_{\alpha}),\alpha=1,2.

  • Denote Uα,jUαA~jU_{\alpha,j}\equiv U_{\alpha}\cap\widetilde{A}_{j}, which is endowed with the rescaled metric gjg_{j}, so A~j=U1,jU2,j\widetilde{A}_{j}=U_{1,j}\cup U_{2,j}. Note that Uα,jU_{\alpha,j} is converging to 𝕌α𝔸\mathbb{U}_{\alpha}\cap\mathbb{A}.

  • Denote Hα=π1(Uα,j)H_{\alpha}=\pi_{1}(U_{\alpha,j}). Then HαH_{\alpha} is isomorphic to \mathbb{Z} and is generated by the orbit of the G3αG_{3-\alpha} action. Notice that since the GαG_{\alpha} action on Uα,jU_{\alpha,j} has fixed points the GαG_{\alpha} orbits are homotopically trivial in Uα,jU_{\alpha,j}.

  • Denote 𝕍𝕌1𝕌2\mathbb{V}\equiv\mathbb{U}_{1}\cap\mathbb{U}_{2} and VF1(𝕍)V\equiv F^{-1}(\mathbb{V}). Denote the overlap by VjU1,jU2,jV_{j}\equiv U_{1,j}\cap U_{2,j}, then π1(Vj)=H1×H2\pi_{1}(V_{j})=H_{1}\times H_{2}. Note that VjV_{j} is converging to 𝕍𝔸\mathbb{V}\cap\mathbb{A}.

See Figure 1 and Figure 2 for pictures of the singular fibration FF.

𝕌2\mathbb{U}_{2}𝕌1\mathbb{U}_{1}P1P_{1}P2P_{2}𝕍\mathbb{V}𝔸\mathbb{A}
Figure 1. Cover \mathbb{H}^{\circ} by 𝕌1\mathbb{U}_{1} and 𝕌2\mathbb{U}_{2}
𝕌i\mathbb{U}_{i}PiP_{i}𝕍\mathbb{V}\bullet\bullet\bulletG3iG_{3-i}GiG_{i}\bulletsmooth fibers T2T^{2}singular fibers S1S^{1}isotropy group GiG_{i}
Figure 2. Singular fibration over 𝕌i\mathbb{U}_{i}

Now consider the convergence of A~j\widetilde{A}_{j} to 𝔸={1/4ρ2+z24}\mathbb{A}=\{1/4\leq\rho^{2}+z^{2}\leq 4\}\subset\mathbb{H}. Denote the universal cover of Uα,jU_{\alpha,j} by U^α,j\widehat{U}_{\alpha,j}. Then we see that for each α\alpha, U^α,j\widehat{U}_{\alpha,j} (with the induced metric from gjg_{j}) is volume non-collapsing. Furthermore, 1\partial_{1} and 2\partial_{2} both lift to Killing fields on U^α,j\widehat{U}_{\alpha,j}, which we denote by ^1;α,j\widehat{\partial}_{1;\alpha,j} and ^2;α,j\widehat{\partial}_{2;\alpha,j} respectively. Notice that on U^α,j\widehat{U}_{\alpha,j} the vector field ^α;α,j\widehat{\partial}_{\alpha;\alpha,j} still generates an S1S^{1} action which descends to the action of GαG_{\alpha} on Uα,jU_{\alpha,j}. Passing to subsequences we may assume that (U^α,j,Hα)(\widehat{U}_{\alpha,j},H_{\alpha}) converges equivariantly in the Ck0C^{k_{0}} Cheeger-Gromov topology to a 4-dimensional flat manifold (U^α,,Hα,)(\widehat{U}_{\alpha,\infty},H_{\alpha,\infty}) so that 𝕌α𝔸\mathbb{U}_{\alpha}\cap\mathbb{A} is isometric to U^α,/Hα,\widehat{U}_{\alpha,\infty}/H_{\alpha,\infty}.

Since we know the collapsing to 𝔸\mathbb{A} is along the GG orbits, we see that Hα,H_{\alpha,\infty} is generated by two commuting Killing fields both of which are limits of linear combinations of ^1;α,j\widehat{\partial}_{1;\alpha,j} and ^2;α,j\widehat{\partial}_{2;\alpha,j}. Now since the GαG_{\alpha} action lifts to U^α,j\widehat{U}_{\alpha,j} it also acts on the limit U^α,\widehat{U}_{\alpha,\infty}, hence ^α;α,j\widehat{\partial}_{\alpha;\alpha,j} naturally converges to an element ^α,\widehat{\partial}_{\alpha,\infty} in the Lie algebra 𝔥\mathfrak{h} of Hα,{H}_{\alpha,\infty} with period 2π2\pi. Locally it is given by a rotation field in a 2-dimensional plane W4W\subset\mathbb{R}^{4} after immersion into 4\mathbb{R}^{4}. Given any other Killing field ξ𝔥\xi\in\mathfrak{h}, since it commutes with ^α,\widehat{\partial}_{\alpha,\infty}, locally it must be given by λ^α,+ξ\lambda\widehat{\partial}_{\alpha,\infty}+\xi^{\prime}, where ξ\xi^{\prime} is a Euclidean motion in the plane orthogonal to WW. We claim that ξ\xi^{\prime} can not be a rotation. One way to see this is to apply the Key Lemma in [35] (again, in the statement of the Key Lemma in [35] we know l<ml<m). Therefore we may identify a unique element (up to ±1\pm 1) ξ^α𝔥\widehat{\xi}_{\alpha}\in\mathfrak{h} with ξ^α=1\|\widehat{\xi}_{\alpha}\|=1, which is locally a translation vector field, such that ^α,\widehat{\partial}_{\alpha,\infty} and ξ^α\widehat{\xi}_{\alpha} generate 𝔥\mathfrak{h}.

For xU^α,x\in\widehat{U}_{\alpha,\infty} where ^α,\widehat{\partial}_{\alpha,\infty} does not vanish, we denote by κ(x)\kappa(x) the geodesic curvature of the orbit of the flow of ^α,\widehat{\partial}_{\alpha,\infty} at xx.

Lemma 2.4.

For all such xx, we have

(2.1) ^α,(x)|κ(x)|=1.\|\widehat{\partial}_{\alpha,\infty}(x)\|\cdot|\kappa(x)|=1.
Proof.

Let x0x_{0} be a zero point of ^α,\widehat{\partial}_{\alpha,\infty}. Then we can isometrically embed a neighborhood of x0x_{0} into 4\mathbb{R}^{4} such that ^α,\widehat{\partial}_{\alpha,\infty} is a standard rotation vector field with period 2π2\pi. It is clear that the equality holds for xx close to x0x_{0}. Using the Gauss-Bonnet theorem it is easy to see the same also holds for all xx. ∎

Let Vˇj\widecheck{V}_{j} be the universal cover of VjV_{j}. Passing to subsequences again we may assume (Vˇj,H1×H2)(\widecheck{V}_{j},H_{1}\times H_{2}) converges equivariantly in the Ck0C^{k_{0}} Cheeger-Gromov topology to a 4-dimensional flat manifold (Vˇ,Hˇ)(\widecheck{V}_{\infty},\widecheck{H}), where Hˇ\widecheck{H} is abstractly isomorphic to 2\mathbb{R}^{2}. Denote by V^α,j\widehat{V}_{\alpha,j} the intermediate covering of VjV_{j} associated to the subgroup HαH1×H2H_{\alpha}\subset H_{1}\times H_{2}. Then V^α,j\widehat{V}_{\alpha,j} can be naturally identified as an open subset of U^α,j\widehat{U}_{\alpha,j}. We may assume (V^α,j,Hα)(\widehat{V}_{\alpha,j},H_{\alpha}) converges to (V^α,,Hα,)(U^α,,Hα,)(\widehat{V}_{\alpha,\infty},H_{\alpha,\infty})\subset(\widehat{U}_{\alpha,\infty},H_{\alpha,\infty}) and there is a \mathbb{Z}-covering map from Vˇ\widecheck{V}_{\infty} to V^α,\widehat{V}_{\alpha,\infty}. It follows that Hˇ\widecheck{H} is also generated by a rotation vector field and a translation vector field. See Figure 3 for a topological picture of the intermediate covers and the universal cover for VjV_{j}.

G1G_{1}G2G_{2}G2G_{2}G1G_{1}𝕍𝔸\mathbb{V}\cap\mathbb{A}𝕍𝔸\mathbb{V}\cap\mathbb{A}𝕍𝔸\mathbb{V}\cap\mathbb{A}V^1,j\widehat{V}_{1,j}V^2,j\widehat{V}_{2,j}Vˇj\widecheck{V}_{j}
Figure 3. Intermediate covers V^1,j,V^2,j\widehat{V}_{1,j},\widehat{V}_{2,j} and the universal cover Vˇj\widecheck{V}_{j}

Denote by ˇα,j\widecheck{\partial}_{\alpha,j} the lift to Vˇj\widecheck{V}_{j} of ^α;α,j\widehat{\partial}_{\alpha;\alpha,j}. Since ^α;α,j\widehat{\partial}_{\alpha;\alpha,j} converges to ^α,\widehat{\partial}_{\alpha,\infty} on V^α,\widehat{V}_{\alpha,\infty}, we see that ˇα,j\widecheck{\partial}_{\alpha,j} on Vˇj\widecheck{V}_{j} also converges to the lift ˇα,\widecheck{\partial}_{\alpha,\infty} of ^α,\widehat{\partial}_{\alpha,\infty} on Vˇ\widecheck{V}_{\infty}. Obviously both ˇ1,\widecheck{\partial}_{1,\infty} and ˇ2,\widecheck{\partial}_{2,\infty} belong to the Lie algebra of Hˇ\widecheck{H}. It follows that ˇ2,=λˇ1,\widecheck{\partial}_{2,\infty}=\lambda\widecheck{\partial}_{1,\infty} for some λ0\lambda\neq 0. A crucial observation is that we must have λ=±1\lambda=\pm 1. Indeed, this follows from (2.1), noticing that |κ(x)||\kappa(x)| depends on the local orbit through xx.

We remark that in the above we have passed to subsequences at various stages, but the derived properties hold on all the possible limits. By the continuity of angle between 1,2\partial_{1},\partial_{2}, we may without loss of generality assume that λ=1\lambda=1 on all possible limits. On any limit we also denote ˇˇ1,=ˇ2,\widecheck{\partial}\equiv\widecheck{\partial}_{1,\infty}=\widecheck{\partial}_{2,\infty}. Then Hˇ\widecheck{H} is generated by ˇ\widecheck{\partial} and a translation vector field ξˇ\widecheck{\xi} with ξˇ=1\|\widecheck{\xi}\|=1. Furthermore, the quotient Vˇ/Hˇ\widecheck{V}_{\infty}/\widecheck{H} is the subset 𝕍𝔸\mathbb{V}\cap\mathbb{A} of \mathbb{H}, and ˇ=ρ\|\widecheck{\partial}\|=\rho.

In the rest of the proof we focus on the set V=F1(𝕍)V=F^{-1}(\mathbb{V}). This is a (trivial) T2T^{2} bundle over 𝕍\mathbb{V}. The function ρ\rho on \mathbb{H}^{\circ} can be naturally viewed as a function on VV. The Riemannian metric gg gives rise to a family of flat metrics on G=T2G=T^{2} parametrized by 𝕍\mathbb{V}. We introduce the Gram matrix 𝔾\mathbb{G} with 𝔾αβg(α,β)\mathbb{G}_{\alpha\beta}\equiv g(\partial_{\alpha},\partial_{\beta}) for 1α,β21\leq\alpha,\beta\leq 2. Let τ\tau be the function on VV such that ξ1(1τ)2{\xi}\equiv\partial_{1}-(1-\tau)\partial_{2} is pointwise orthogonal to 2\partial_{2}, and let σξg\sigma\equiv\|\xi\|_{g}. Notice that τ\tau and σ\sigma are both T2T^{2} invariant so can be viewed as functions on 𝕍\mathbb{V}.

As jj\rightarrow\infty, passing to subsequences we know that under the convergence of Vˇj\widecheck{V}_{j} to Vˇ\widecheck{V}_{\infty}, the Killing fields ˇ1,j\widecheck{\partial}_{1,j} and ˇ2,j\widecheck{\partial}_{2,j} both converge naturally to the same vector field ˇ\widecheck{\partial} on Vˇ\widecheck{V}_{\infty}. On Vˇj\widecheck{V}_{j}, the function ρ\rho is uniformly comparable to 2j2^{j}. Denote ξj1,j(1τ)2,j\xi_{j}\equiv\partial_{1,j}-(1-\tau)\partial_{2,j} in the rescaled annulus A~j\widetilde{A}_{j}.

Lemma 2.5.

Passing to subsequences and pulling back to Vˇj\widecheck{V}_{j}, 2jσ1ξj2^{j}\sigma^{-1}\xi_{j} converges in C2C^{2} to ξˇ\widecheck{\xi} as jj\rightarrow\infty.

Proof.

To see this we first choose a vector field ξj\xi_{j}^{\prime} (for j1j\gg 1) in the linear span of 1,j\partial_{1,j} and 2,j\partial_{2,j} on each VjV_{j} such that after passing to subsequences and pulling back to Vˇj\widecheck{V}_{j} it converges in Ck0C^{k_{0}} to the limit ξˇ\widecheck{\xi}. Write ξj\xi_{j}^{\prime} uniquely as

ξj=σj1(1,j(1τj)2,j).\xi_{j}^{\prime}=\sigma_{j}^{-1}(\partial_{1,j}-(1-\tau_{j})\partial_{2,j}).

Then we have

ξj=σj1(ξj+(τjτ)2,j).\xi_{j}^{\prime}=\sigma_{j}^{-1}(\xi_{j}+(\tau_{j}-\tau)\partial_{2,j}).

As ξj\xi_{j}^{\prime} converges in Ck0C^{k_{0}} to ξˇ\widecheck{\xi}, we have

(2.2) gj(ξj,2,j)0,\displaystyle g_{j}(\xi_{j}^{\prime},\partial_{2,j})\to 0, gj(ξj,ξj)1.\displaystyle g_{j}(\xi_{j}^{\prime},\xi_{j}^{\prime})\to 1.

It follows that

σj1|τjτ|0,\displaystyle\sigma_{j}^{-1}{|\tau_{j}-\tau|}\to 0, σj1ξjgj1,\displaystyle\sigma_{j}^{-1}\|\xi_{j}\|_{g_{j}}\to 1,

and σj1ξj\sigma_{j}^{-1}\xi_{j} converges to ξˇ\widecheck{\xi} in Ck0C^{k_{0}}. This implies that ξjgj1ξj=2jσ1ξj\|{\xi}_{j}\|_{g_{j}}^{-1}{\xi}_{j}=2^{j}\sigma^{-1}\xi_{j} also converges in C2C^{2} to ξˇ\widecheck{\xi}. ∎

By construction since ξˇ\widecheck{\xi} is a translation vector field we must have

(2.3) limρτ=0.\lim_{\rho\rightarrow\infty}\tau=0.

Denote f2gf\equiv\|\partial_{2}\|_{g}. Then from the convergence of ˇ2,j\widecheck{\partial}_{2,j} to ˇ\widecheck{\partial} one can see that

(2.4) limρρ1f=1.\lim_{\rho\rightarrow\infty}\rho^{-1}f=1.

Observe that 22jfσ2^{-2j}f\sigma is the area of the T2T^{2} orbit in VjV_{j}. By assumption we have

(2.5) limρρ2fσ=limρρ1σ=0.\lim_{\rho\rightarrow\infty}\rho^{-2}f\sigma=\lim_{\rho\rightarrow\infty}\rho^{-1}\sigma=0.

Next we derive a few key limiting properties.

Proposition 2.6.

We have

(2.6) limρσρτ=0.\lim_{\rho\rightarrow\infty}\frac{\sigma}{\rho\tau}=0.
Proof.

This follows from the fact that the collapsing is of codimension 2; in other words, the diameters of the T2T^{2} orbits are of order o(ρ)o(\rho). Scaling down the metric it suffices to show that a sequence of flat tori given by the quotient of standard 2\mathbb{R}^{2} by the lattice Λ\Lambda generated by two vectors (1,0)(1,0) and (1a,b)(1-a,b) with a,b0a,b\rightarrow 0 has diameter goes to zero if and only if b/a0b/a\rightarrow 0. To see this fact, we simply notice that for x0[0,1]x_{0}\in[0,1], the distance between [(0,0)][(0,0)] and [(x0,0)][(x_{0},0)] in 2/Λ\mathbb{R}^{2}/\Lambda is given by the distance between (0,0)(0,0) and the line bx+ay=bx0bx+ay=bx_{0} in 2\mathbb{R}^{2}. The latter goes to zero if and only if b/a0b/a\rightarrow 0. Write the T2T^{2} orbits in VjV_{j} as quotient of 2\mathbb{R}^{2} by Λ\Lambda in terms of the basis 2,j{\partial}_{2,j} and 2jσ1ξj2^{j}\sigma^{-1}{\xi}_{j} and apply the above. ∎

Proposition 2.7.

We have

(2.7) limρρσρσ=0,\lim_{\rho\rightarrow\infty}\frac{\rho\sigma_{\rho}}{\sigma}=0,
(2.8) limρρ2τρσ=0.\lim_{\rho\rightarrow\infty}\frac{\rho^{2}\tau_{\rho}}{\sigma}=0.
Proof.

Let ρ\partial_{\rho}^{\sharp} be the vector field on VV given by the horizontal lift of the standard vector field ρ\partial_{\rho} on \mathbb{H}^{\circ}. Now consider the convergence of Vˇj\widecheck{V}_{j} to Vˇ\widecheck{V}_{\infty}. We see that the rescaled vector field 2jρ2^{j}\partial_{\rho}^{\sharp} converges in C2C^{2} to the radial vector field ˇρ\widecheck{\partial}_{\rho} in the rotation plane of ˇ\widecheck{\partial} in Vˇ\widecheck{V}_{\infty}. So the norm of the commutator [2jρ,2jσ1ξj]gj\|[2^{j}\partial_{\rho}^{\sharp},2^{j}\sigma^{-1}\xi_{j}]\|_{g_{j}} converges to [ˇρ,ξˇ]=0\|[\widecheck{\partial}_{\rho},\widecheck{\xi}]\|=0 as jj\rightarrow\infty. Notice that since 1\partial_{1} and 2\partial_{2} are Killing fields and [1,2]=ρ,1=ρ,2=0[\partial_{1},\partial_{2}]=\langle\partial_{\rho}^{\sharp},\partial_{1}\rangle=\langle\partial_{\rho}^{\sharp},\partial_{2}\rangle=0, we have for α=1,2\alpha=1,2,

[α,ρ],β=αρ,β=αρ,β(αg)(ρ,β)ρ,αβ=0.\langle[\partial_{\alpha},\partial_{\rho}^{\sharp}],\partial_{\beta}\rangle=\langle\mathcal{L}_{\partial_{\alpha}}\partial_{\rho}^{\sharp},\partial_{\beta}\rangle=\mathcal{L}_{\partial_{\alpha}}\langle\partial_{\rho}^{\sharp},\partial_{\beta}\rangle-(\mathcal{L}_{\partial_{\alpha}}g)(\partial_{\rho}^{\sharp},\partial_{\beta})-\langle\partial_{\rho}^{\sharp},\mathcal{L}_{\partial_{\alpha}}\partial_{\beta}\rangle=0.

Since ξ\xi is orthogonal to 2\partial_{2},

[2jρ,2jσ1ξ]gj2=24j(σρσ2ξ+σ1τρ2)gj2=24j((σρσ)222j+(τρσ)222jf2).\|[2^{j}\partial_{\rho}^{\sharp},2^{j}\sigma^{-1}\xi]\|_{g_{j}}^{2}=2^{4j}\|(-\sigma_{\rho}\sigma^{-2}\xi+\sigma^{-1}\tau_{\rho}\partial_{2})\|_{g_{j}}^{2}=2^{4j}((\frac{\sigma_{\rho}}{\sigma})^{2}2^{-2j}+(\frac{\tau_{\rho}}{\sigma})^{2}2^{-2j}f^{2}).

Thus τρσ1=o(22j)\tau_{\rho}\sigma^{-1}=o(2^{-2j}) and σρσ1=o(2j)\sigma_{\rho}\sigma^{-1}=o(2^{-j}) on VjV_{j}. The conclusion then follows. ∎

Now using (2.3), (2.5), (2.6), (2.7), (2.8) and the L’Hospital’s rule we get

=limρτσ/ρ=limρτρσρ/ρσ/ρ2=limρτρσ/ρ2=0.\displaystyle\infty=\lim_{\rho\to\infty}\frac{\tau}{\sigma/\rho}=\lim_{\rho\to\infty}\frac{\tau_{\rho}}{\sigma_{\rho}/\rho-\sigma/\rho^{2}}=\lim_{\rho\to\infty}\frac{\tau_{\rho}}{-\sigma/\rho^{2}}=0.

This is a contradiction hence completes the proof of Theorem 1.2.

Remark 2.8.

(2.7) has a geometric explanation if we consider the convergence to the asymptotic cone \mathbb{H} as metric measure spaces. On the one hand, we know from the local description of the universal covering geometry that the renormalized limit measure on \mathbb{H} has a density of the form dν=ρdρdzd\nu=\rho d\rho dz. On the other hand, by Fukaya’s theorem [19] we know that up to a multiplicative constant fσf\sigma converges to ρ\rho. Then (2.7) can also be obtained by comparing these two facts.

3. Further discussion

In this section we construct an explicit sequence of Riemannian metrics gig_{i} on 𝔄=S3×[21,2]\mathfrak{A}=S^{3}\times[2^{-1},2] with sup𝔄|Rmgi|0\sup_{\mathfrak{A}}|Rm_{g_{i}}|\rightarrow 0 which collapse to the annulus 𝔸{21r2}\mathbb{A}\equiv\{2^{-1}\leq r\leq 2\} in \mathbb{H}. This shows that Theorem 1.2 is of a global nature. We will also explain the relevance to the asymptotic curvature gap conjecture of Petrunin-Tuschmann [35].

3.1. A model flat metric

Denote the flat product metric 2=×\mathbb{C}^{2}=\mathbb{C}\times\mathbb{C}. Given α,σ[0,1)\alpha,\sigma\in[0,1), we consider the Euclidean motion γ:(z1,z2)(z1+σ,e2π1αz2)\gamma:(z_{1},z_{2})\mapsto(z_{1}+\sigma,e^{2\pi\sqrt{-1}\alpha}z_{2}). Let 𝒳α,σ\mathcal{X}_{\alpha,\sigma} be the quotient of 2\mathbb{C}^{2} by the \mathbb{Z}-action generated by γ\gamma. Then 𝒳α,σ\mathcal{X}_{\alpha,\sigma} is a complete flat manifold with cubic volume growth, but the asymptotic geometry depends on the rationality of α\alpha. When α\alpha is rational we write α=p/q\alpha=p/q for p,qp,q coprime, then 𝒳α,σ\mathcal{X}_{\alpha,\sigma} has asymptotic cone given by the product α×\mathbb{C}_{\alpha}\times\mathbb{R}, where α\mathbb{C}_{\alpha} is a 2 dimensional flat cone with angle 2π/q2\pi/q. When α\alpha is irrational the end geometry is quite different. This is similar to the example of 3 dimensional flat manifolds studied by Gromov (see [32, 13]). One can check that 𝒳α,σ\mathcal{X}_{\alpha,\sigma} is still conical but the asymptotic cone is instead the half plane \mathbb{H}.

Notice that 𝒳α,σ\mathcal{X}_{\alpha,\sigma} admits a T2T^{2} symmetry, with the first S1S^{1} action inherited from the rotation on the second \mathbb{C} factor, and the second S1S^{1} action given by e2π1t(z1,z2)=(z1+tσ,e2π1αtz2)e^{2\pi\sqrt{-1}t}\cdot(z_{1},z_{2})=(z_{1}+t\sigma,e^{2\pi\sqrt{-1}\alpha t}z_{2}). So we may rewrite the flat metric on 𝒳α\mathcal{X}_{\alpha} in terms of the Gram matrix 𝔾\mathbb{G} as

g=dρ2+dz2+α,β𝔾αβdϕαdϕβ,g=d\rho^{2}+dz^{2}+\sum_{\alpha,\beta}\mathbb{G}_{\alpha\beta}d\phi_{\alpha}\otimes d\phi_{\beta},

where (ρ,z)(\rho,z) denote the standard coordinates on \mathbb{H}. One can compute that

(3.1) 𝔾=(ρ2αρ2αρ2α2ρ2+σ2).\mathbb{G}=\begin{pmatrix}\rho^{2}&\alpha\rho^{2}\\ \alpha\rho^{2}&\alpha^{2}\rho^{2}+\sigma^{2}\end{pmatrix}.

Notice det(ρ1σ1𝔾)=1\det(\rho^{-1}\sigma^{-1}\mathbb{G})=1. It is a general fact that 4 dimensional toric Ricci-flat metrics give rise to integrable systems (see for example [26]). One can check that in our setting the normalized Gram matrix σ1ρ1𝔾\sigma^{-1}\rho^{-1}\mathbb{G} defines an axi-symmetric harmonic map from 3\mathbb{R}^{3} (obtained by rotating \mathbb{H} around the zz-axis) into the symmetric space SL(2;)/SO(2)SL(2;\mathbb{R})/SO(2); the latter can be identified with the hyperbolic plane ={(X,Y)|X>0}\mathcal{H}=\{(X,Y)|X>0\} via

(X+X1Y2X1YX1YX1).\left(\begin{matrix}X+X^{-1}Y^{2}&X^{-1}Y\\ X^{-1}Y&X^{-1}\end{matrix}\right).

Following usual terminologies in the literature we call 𝒳α,σ\mathcal{X}_{\alpha,\sigma} an AF model end. An AF manifold is by definition a complete Riemannian manifold that is asymptotic to an AF model end at a polynomial rate. For example, the Schwarzschild gravitational instanton and the Kerr gravitational instantons are AF. More delicate examples of AF gravitational instantons were constructed explicitly by Chen-Teo [14] using the technique of inverse scattering transform.

By definition an AF manifold is conical 𝒜\mathcal{AF}. Notice that an AF end is not simply-connected, so we do not arrive at a contradiction with Theorem 1.2. The manifolds 𝒳α,σ\mathcal{X}_{\alpha,\sigma} for α\alpha irrational serve as local models for the construction in the next subsection.

3.2. Simply-connected domains collapsing to 𝔸\mathbb{A}

We will adopt the notation in Section 2 and consider only T2T^{2} invariant metrics. The question reduces to constructing suitable family of flat metrics on T2T^{2} with prescribed behavior. We introduce polar coordinates (r,θ)(r,\theta) on \mathbb{H} so that ρ=rcosθ\rho=r\cos\theta, z=rsinθz=r\sin\theta with θ[0,π]\theta\in[0,\pi].

Around the boundary when θ\theta is close to 0 and π\pi we can use the annuli in the model flat end 𝒳α,σ\mathcal{X}_{\alpha,\sigma}. In order to create collapsing metrics on 𝔄\mathfrak{A}, we need to glue together annuli in two different such ends.

We introduce the model normalized Gram matrices

(3.2) 𝔾σ,τ1(1σrsinθ1τσrsinθ1τσrsinθ(1τ)2r2sin2θ+σ2σrsinθ),\mathbb{G}^{1}_{\sigma,\tau}\equiv\left(\begin{matrix}\frac{1}{\sigma}r\sin\theta&\frac{1-\tau}{\sigma}r\sin\theta\\ \frac{1-\tau}{\sigma}r\sin\theta&\frac{(1-\tau)^{2}r^{2}\sin^{2}\theta+\sigma^{2}}{\sigma r\sin\theta}\end{matrix}\right),
(3.3) 𝔾σ,τ2((1τ)2r2sin2θ+σ2σrsinθ1τσrsinθ1τσrsinθ1σrsinθ).\mathbb{G}^{2}_{\sigma,\tau}\equiv\left(\begin{matrix}\frac{(1-\tau)^{2}r^{2}\sin^{2}\theta+\sigma^{2}}{\sigma r\sin\theta}&\frac{1-\tau}{\sigma}r\sin\theta\\ \frac{1-\tau}{\sigma}r\sin\theta&\frac{1}{\sigma}r\sin\theta\end{matrix}\right).

Notice that the metric dr2+r2dθ2+σrsinθ𝔾σ,ταdr^{2}+r^{2}d\theta^{2}+\sigma r\sin\theta\mathbb{G}^{\alpha}_{\sigma,\tau} are both isometric to 𝒳1τ,σ\mathcal{X}_{1-\tau,\sigma}. The difference between the two is that for each α\alpha, the rotational Killing field for 𝔾σ,τα\mathbb{G}^{\alpha}_{\sigma,\tau} is given by ϕα\partial_{\phi_{\alpha}}. The idea is to patch these two models together to construct a global normalized Gram matrix 𝔾\mathbb{G}, so that 𝔾=𝔾σ,τ1\mathbb{G}=\mathbb{G}^{1}_{\sigma,\tau} over θ[0,π3]\theta\in[0,\frac{\pi}{3}] and 𝔾=𝔾σ,τ2\mathbb{G}=\mathbb{G}^{2}_{\sigma,\tau^{\prime}} over θ[2π3,π]\theta\in[\frac{2\pi}{3},\pi]. Then by construction the topology will be simply-connected.

The key point is on the choice of the appropriate parameters σ,τ,τ\sigma,\tau,\tau^{\prime}. Viewing 𝔾σ,τα\mathbb{G}^{\alpha}_{\sigma,\tau} as maps into \mathcal{H}, for r[21,2]r\in[2^{-1},2] we have

(3.4) 𝔾σ,τ1(r)\displaystyle\mathbb{G}^{1}_{\sigma,\tau}(r) :[0,π](X,Y)=(σrsinθ(1τ)2r2sin2θ+σ2,(1τ)r2sin2θ(1τ)2r2sin2θ+σ2),\displaystyle:[0,\pi]\mapsto(X,Y)=\left(\frac{\sigma r\sin\theta}{(1-\tau)^{2}r^{2}\sin^{2}\theta+\sigma^{2}},\frac{(1-\tau)r^{2}\sin^{2}\theta}{(1-\tau)^{2}r^{2}\sin^{2}\theta+\sigma^{2}}\right),
(3.5) 𝔾σ,τ2(r)\displaystyle\mathbb{G}^{2}_{\sigma,\tau}(r) :[0,π](X,Y)=(σrsinθ,1τ).\displaystyle:[0,\pi]\mapsto(X,Y)=\left(\frac{\sigma}{r\sin\theta},1-\tau\right).

The images are contained in geodesics in \mathcal{H}. Indeed, the image of 𝔾σ,τ1(r)\mathbb{G}^{1}_{\sigma,\tau}(r) is part of the half circle from (0,0)(0,0) to (0,11τ)(0,\frac{1}{1-\tau}) and the image of 𝔾σ,τ2(r)\mathbb{G}^{2}_{\sigma,\tau}(r) is part of the straight line Y=1τY=1-\tau (see Figure 4).

YYXX
Figure 4. Images of 𝔾σ,τα\mathbb{G}_{\sigma,\tau}^{\alpha}

Now for τ1\tau\ll 1 we set 1τ=11τ1-\tau^{\prime}=\frac{1}{1-\tau} in order to make the half circle and the line tangential. Then we choose στ0\sigma\ll\tau\rightarrow 0 so that each T2T^{2} collapses to a point. We patch 𝔾σ,τ1(r)\mathbb{G}^{1}_{\sigma,\tau}(r) and 𝔾σ,τ2(r)\mathbb{G}^{2}_{\sigma,\tau^{\prime}}(r) by interpolating them over θ[π3,2π3]\theta\in[\frac{\pi}{3},\frac{2\pi}{3}]. Consider the isometry of \mathcal{H} given by

Φ:(X,Y)(1σX,1σ(Y11τ)).\Phi:(X,Y)\mapsto\left(\frac{1}{\sigma}X,\frac{1}{\sigma}(Y-\frac{1}{1-\tau})\right).

Given any normalized Gram matrix 𝔾\mathbb{G} associated to (X,Y)(X,Y)\in\mathcal{H}, we denote by Φ𝔾\Phi\mathbb{G} the normalized Gram matrix corresponding to Φ(X,Y)\Phi(X,Y)\in\mathcal{H}. This is nothing but rewriting 𝔾\mathbb{G} under the basis 1σ(111τ2),σ2\frac{1}{\sqrt{\sigma}}(\partial_{1}-\frac{1}{1-\tau}\partial_{2}),\sqrt{\sigma}\partial_{2}. We have

Φ𝔾σ,τ1(r)\displaystyle\Phi\mathbb{G}_{\sigma,\tau}^{1}(r) =(rsinθ(1τ)2r2sin2θ+σ2,σ(1τ)2r2sin2θ+σ2),\displaystyle=\left(\frac{r\sin\theta}{(1-\tau)^{2}r^{2}\sin^{2}\theta+\sigma^{2}},-\frac{\sigma}{(1-\tau)^{2}r^{2}\sin^{2}\theta+\sigma^{2}}\right),
Φ𝔾σ,τ2(r)\displaystyle\Phi\mathbb{G}_{\sigma,\tau^{\prime}}^{2}(r) =(1rsinθ,0).\displaystyle=\left(\frac{1}{r\sin\theta},0\right).

Now by a straightforward interpolation one can find a smooth family of curves 𝔾σ,τ:{12r2]}\mathbb{G}^{\prime}_{\sigma,\tau}:\{\frac{1}{2}\leq r\leq 2]\}\subset\mathbb{H}\rightarrow\mathcal{H} such that 𝔾σ,τ=Φ𝔾σ,τ1\mathbb{G}_{\sigma,\tau}^{\prime}=\Phi\mathbb{G}^{1}_{\sigma,\tau} for θ[0,π3]\theta\in[0,\frac{\pi}{3}], 𝔾σ,τ=Φ𝔾σ,τ2\mathbb{G}^{\prime}_{\sigma,\tau}=\Phi\mathbb{G}^{2}_{\sigma,\tau^{\prime}} for θ[2π3,π]\theta\in[\frac{2\pi}{3},\pi], and for θ[π3,2π3]\theta\in[\frac{\pi}{3},\frac{2\pi}{3}] we have 𝔾σ,τ\mathbb{G}^{\prime}_{\sigma,\tau} converge smoothly to (1rsinθ,0)(\frac{1}{r\sin\theta},0) as στ0\sigma\ll\tau\rightarrow 0. Then we define 𝔾σ,τΦ1𝔾σ,τ\mathbb{G}_{\sigma,\tau}\equiv\Phi^{-1}\mathbb{G}^{\prime}_{\sigma,\tau}. It induces the metric on S3×[21,2]S^{3}\times[2^{-1},2] via

(3.6) gσ,τ=dr2+r2dθ2+σrsinθ𝔾σ,τ.g_{\sigma,\tau}=dr^{2}+r^{2}d\theta^{2}+\sigma r\sin\theta\mathbb{G}_{\sigma,\tau}.

We claim that as στ0\sigma\ll\tau\to 0, gσ,τg_{\sigma,\tau} is collapsing to 𝔸\mathbb{A} with sup𝔄|Rmgσ,τ|0\sup_{\mathfrak{A}}|Rm_{g_{\sigma,\tau}}|\rightarrow 0. For this we study the limit geometry of local universal covers. It is clear that for θ[0,π3]\theta\in[0,\frac{\pi}{3}] and [2π3,π][\frac{2\pi}{3},\pi] the limit geometry is given by the model flat ends. We only need to consider the region when θ[π3,2π3]\theta\in[\frac{\pi}{3},\frac{2\pi}{3}].

Under the basis {1σ(111τ2),σ2}\{\frac{1}{\sqrt{\sigma}}(\partial_{1}-\frac{1}{1-\tau}\partial_{2}),\sqrt{\sigma}\partial_{2}\} we have

(3.7) σrsinθΦ𝔾σ,τ=σrsinθ(X+X1Y2X1YX1YX1),\sigma r\sin\theta\Phi\mathbb{G}_{\sigma,\tau}=\sigma r\sin\theta\left(\begin{matrix}X^{\prime}+X^{\prime-1}Y^{\prime 2}&X^{\prime-1}Y^{\prime}\\ X^{\prime-1}Y^{\prime}&X^{\prime-1}\end{matrix}\right),

where (X,Y)(X^{\prime},Y^{\prime}) converge smoothly to (1rsinθ,0)(\frac{1}{r\sin\theta},0). This precisely means that under {1σ(111τ2),2}\{\frac{1}{\sigma}(\partial_{1}-\frac{1}{1-\tau}\partial_{2}),\partial_{2}\}, as στ0\sigma\ll\tau\to 0, the metrics gσ,τg_{\sigma,\tau} when pulled back to the universal covers converge smoothly to the standard flat metric in 3×\mathbb{R}^{3}\times\mathbb{R}:

dr2+r2dθ2+(100r2sin2θ).dr^{2}+r^{2}d\theta^{2}+\left(\begin{matrix}1&0\\ 0&r^{2}\sin^{2}\theta\end{matrix}\right).

The condition that στ\sigma\ll\tau ensures that each T2T^{2} collapses to a point as in Proposition 2.6, so the limit is 𝔸\mathbb{A}.

3.3. No gap for asymptotic curvature

It is tempting to apply the construction in the previous subsection to form a counterexample to Conjecture 1.1. However, as Theorem 1.2 shows we know this is impossible – the issue arises when we take the limit as ρ\rho\rightarrow\infty since the asymptotic properties deduced in Section 2 can not be simultaneously achieved (due to the L’Hospital rule!). Nevertheless, if we allow some flexibility then one can indeed construct examples satisfying most of the other asymptotic properties. We may relax the asymptotical flatness property as follows:

Proposition 3.1.

For any ϵ>0\epsilon>0 there is a complete Riemannian metrics gϵg_{\epsilon} on the standard 4\mathbb{R}^{4} with quadratic curvature decay and A(M,gϵ)ϵA(M,g_{\epsilon})\leq\epsilon which is asymptotic to the cone \mathbb{H}.

Remark 3.2.

Since we can not take ϵ=0\epsilon=0, this shows there is no gap property for asymptotic curvature if we fix the asymptotic cone. i.e., gϵg_{\epsilon} can not be perturbed to an 𝒜\mathcal{AF} metric with the same asymptotic cone. In particular, it gives a negative answer to a version of Question 2 in [35].

Proof.

Note the for the construction in Section 3.2 we have chosen σ\sigma and τ\tau to be constants. Now we may choose them to depend on rr. We will still require σ/τ0\sigma/\tau\to 0 as rr\to\infty so that the asymptotic cone is given by collapsing the T2T^{2} orbits. We then obtain the normalized Gram matrix 𝔾σ(r),τ(r)\mathbb{G}_{\sigma(r),\tau(r)}, and define a metric

g=dr2+r2dθ2+σrsinθ𝔾σ,τg=dr^{2}+r^{2}d\theta^{2}+\sigma r\sin\theta\mathbb{G}_{\sigma,\tau}

when r>1r>1. Set σ=ϵrlogr\sigma=\epsilon\frac{r}{\log r} for ϵ>0\epsilon>0 small and τ=1(logr)1/2\tau=\frac{1}{(\log r)^{1/2}}. From the construction in Section 3.2, we see that over r(2j1,2j+1)r\in(2^{j-1},2^{j+1}), the frame {2jr,θ,2j1σ(111τ2),2}\{2^{j}\partial_{r},\partial_{\theta},2^{j}\frac{1}{\sigma}(\partial_{1}-\frac{1}{1-\tau}\partial_{2}),\partial_{2}\} along with the rescaled metric 2jg2^{-j}g converges smoothly to a frame {r,θ,ξ,}\{\partial_{r},\partial_{\theta},\xi_{\infty},\partial_{\infty}\}, under which the metric converges to

(1r21r2sin2θ).\left(\begin{matrix}1&&&\\ &r^{2}&&\\ &&1&\\ &&&r^{2}\sin^{2}\theta\end{matrix}\right).

One can check that the Lie brackets between the limit frames are all zero except

[r,ξ]=1rϵξ.\displaystyle[\partial_{r},\xi_{\infty}]=\frac{1}{r}\epsilon\xi_{\infty}.

In particular, the asymptotic curvature can be made arbitrarily small by taking ϵ\epsilon small. Filling in suitably we get smooth Riemannian metrics on 4\mathbb{R}^{4}. ∎

In the above construction we may also choose functions σ(r)\sigma(r) and τ(r)\tau(r) to converge to zero in a comparable rate as rr\rightarrow\infty. In this way we can achieve asymptotical flatness, but the asymptotic cone will be different. This is indeed the case for some explicit asymptotically flat metrics on 4\mathbb{R}^{4}, for example the Taub-NUT gravitational instantons. The asymptotic cone is given by 3\mathbb{R}^{3} and the collapsing to the asymptotic cone is only of codimension 1.

These examples demonstrate the subtlety in studying Conjecture 1.1.

4. Gravitational instantons

Recall that in this paper by a gravitational instanton we mean a complete noncompact Riemannian 4-manifold (M,g)(M,g) with vanishing Ricci curvature and with finite energy (i.e. M|Rmg|2<\int_{M}|Rm_{g}|^{2}<\infty). For simplicity of discussion we will always assume MM is oriented. Our original motivation to study Conjecture 1.1 arises from the aim of understanding the asymptotic geometry of gravitational instantons. The topic of gravitational instantons has been extensively studied for a long time in both the mathematical and physics literature. In this section we include a brief summary of what is known to date about gravitational instantons and list some open questions.

Recall that we may view the Riemann curvature tensor of a general Riemannian manifold as a self-adjoint endomorphism \mathcal{R} on the bundle Λ2\Lambda^{2} of 2-forms. Standard representation theory of SO(n)SO(n) decomposes \mathcal{R} into the sum of 3 components: one involving the scalar curvature SS, one involving the trace-free Ricci curvature Ric\overset{\circ}{Ric} and one involving the Weyl curvature WW. In dimension 4 since SO(4)SO(4) is a double cover of the product SO(3)×SO(3)SO(3)\times SO(3), we have a refined decomposition. The bundle Λ2\Lambda^{2} splits as the direct sum of the bundle of self-dual and anti-self-dual 2-forms Λ2=Λ+Λ\Lambda^{2}=\Lambda^{+}\oplus\Lambda^{-}. Accordingly we may write \mathcal{R} as

=(W++S12RicRicW+S12).\mathcal{R}=\begin{pmatrix}W^{+}+\frac{S}{12}&\overset{\circ}{Ric}\\ \overset{\circ}{Ric}&W^{-}+\frac{S}{12}\end{pmatrix}.

At each point we may view W+W^{+} as a trace-free symmetric 3×33\times 3 matrix. Following [29] we distinguish gravitational instantons into 3 types.

  • Type I: W+W^{+} vanishes identically. This condition is equivalent to that gg being locally hyperähler. Namely, locally there are 3 compatible complex structures J1,J2,J3J_{1},J_{2},J_{3} satisfying the quaternion relations J1J2=J2J1=J3J_{1}J_{2}=-J_{2}J_{1}=J_{3} which are all parallel with respect to the Levi-Civita connection and which define the given orientation.

  • Type II: W+W^{+} has exactly 2 distinct eigenvalues everywhere. It was first discovered by Derdziński [17] that this condition is equivalent to saying locally there is a complex structure JJ such that gg is hermitian under it. Passing to a double cover if necessary, there exists a global complex structure JJ. Furthermore, on the double cover, the conformal metric g~(24|W+|2)13g\tilde{g}\equiv(24|W^{+}|^{2})^{\frac{1}{3}}g is Kähler extremal in the sense of Calabi and has vanishing Bach tensor. The extremal vector field for g~\tilde{g} is a non-zero Killing field for gg.

  • Type III: W+W^{+} has generically 3 distinct eigenvalues.

Notice that the Type of a gravitational instanton depends on the choice of orientation. For example, the Eguchi-Hanson metric and the Taub-NUT metric are Type I with respect to the standard orientation but are Type II with respect to the reversed orientation.

By definition (global) hyperkähler gravitational instantons are of Type I. These are among the most well-studied classes. In fact, often in the literature the hyperkähler condition is taken for granted in the definition of gravitational instantons. There are many different constructions of hyperkähler gravitational instantons and they exhibit a variety of interesting behavior at infinity.

By Bando-Kasue-Nakajima [5] a gravitational instanton with Euclidean volume growth is ALE, i.e., it is asymptotic to a flat cone 4/Γ\mathbb{R}^{4}/\Gamma for some ΓSO(4)\Gamma\in SO(4) at a polynomial rate. ALE gravitational instantons have finite fundamental groups. All the known examples of ALE gravitational instantons are finite quotients of hyperkähler gravitational instantons; indeed they have been classified [25, 16, 40]. The following is a longstanding question

Conjecture 4.1 (Bando-Kasue-Nakajima [5]).

All ALE gravitational instantons are of Type I up to reversing the orientation.

If there is a hypothetical Type II ALE gravitational instanton, then its structure group Γ\Gamma at infinity is conjugate to a subgroup of U(2)U(2). In [29] it is proved that

Theorem 4.2 ([29]).

The only Type II ALE gravitational instanton with structure group ΓSU(2)\Gamma\subset SU(2) is given by the Eguchi-Hanson metric with the reversed orientation.

This gives supporting evidence to Conjecture 4.1. It is likely that the result holds without the restriction ΓSU(2)\Gamma\subset SU(2). If so then Conjecture 4.1 amounts to ruling out gravitational instantons which are Type III for both orientations.

When a gravitational instanton does not have Euclidean volume growth, the geometry at infinity is much richer. In the hyperkähler case we have

Theorem 4.3 ([38]).

A non-flat hyperkähler gravitational instanton must be conical and asymptotic to one of the following 6 families of model ends: ALE,ALF,ALG,ALH,ALG,ALHALE,ALF,ALG,ALH,ALG^{*},ALH^{*}.

Roughly speaking, the AL()AL\sharp(^{*}) model ends have volume growth of order (4(E))(4-(\sharp-E)). The 4 families without are locally given by flat torus bundles over a domain in the Euclidean space; the 2 families with * have inhomogeneous geometries at infinity. A key common feature is that all the model ends admit local nilpotent symmetries.

The proof of Theorem 4.3 exploits a combination of Riemannian collapsing theory of Cheeger-Fukaya-Gromov and the technique of adiabatic PDE perturbation. It came as a by-product of a more general study of collapsing geometry of 4 dimensional hyperkähler metrics. One simple but central fact is that a hyperkähler metric with a continuous symmetry is locally given by a positive harmonic function on 3\mathbb{R}^{3}. There is an earlier work Chen-Chen [10] proving a special case of theorem under the extra assumption of faster than quadratic curvature decay. Notice that ALE,ALF,ALG,ALHALE,ALF,ALG,ALH have faster than quadratic curvature decay, ALGALG^{*} is 𝒜\mathcal{AF} but not strongly 𝒜\mathcal{AF}, and ALHALH^{*} is not even 𝒜\mathcal{AF}.

Given Theorem 4.3 the next question is to classify all hyperkähler gravitational instantons asymptotic to a given model end. This is essentially completed due to the work of many people, see for example [25, 9, 11, 37, 38, 22, 15, 27, 12]. In particular, Torelli type theorems are proved, which gives a classification of AL\sharp hyperkähler gravitation instantons in terms of the topological data of periods. The strategy is to use the prescribed asymptotics to compactify the underlying complex manifold (for particular choices of the complex structure) into a Kähler surface by adding certain divisor at infinity, and then describe the metric in terms of the (suitably generalized) Tian-Yau construction. One idea to prove the Torelli theorem is to use gluing method to prove that an AL\sharp gravitational instanton (satisfying certain topological bound when =E,F\sharp=E,F) arises as models of singularity formation for a sequence of hyperkähler metrics on the K3 manifold, and then refer to the classical Torelli theorem for K3 manifolds. The original proof of the surjectivity part for K3 Torelli makes use of Yau’s proof of the Calabi conjecture; more recently Liu [30] gave a new proof using instead the compactness result of [38].

The upshot is that we have a relatively complete understanding of hyperkähler gravitational instantons. The next natural question is

Problem 4.4.

Classify Type I gravitational instantons.

Notice that the universal cover of a Type I gravitational instanton is always a complete hyperkähler manifold, but the energy is finite only when the fundamental group is finite. In the case of finite fundamental group the problem reduces to the more manageable

Question 4.5.

Classify free action of finite groups on hyperkähler gravitational instantons.

The case of infinite fundamental group is more subtle. One example is given by the flat manifold Yα,σY_{\alpha,\sigma} introduced in Section 3.1. The asymptotic geometry at infinity depends on the rationality of α\alpha.

Question 4.6.

Is there a non-flat Type I gravitational instanton (X,g)(X,g) with π1(X)\pi_{1}(X) infinite?

If such an example exists, then the universal cover X~\tilde{X} must have infinite energy. The only known examples of infinite energy hyperkähler 4-manifold are constructed using the Gibbons-Hawking ansatz [4, 20].

In general the study of infinite energy complete Ricci-flat metrics is an interesting problem. We have the following conjecture in [38]

Conjecture 4.7 (Finite energy vs finite topology).

A complete Ricci-flat 4-manifold has finite energy if and only if it has finite topological type.

The study of Type II gravitational instantons has only emerged in the recent few years. The known examples are all ALF or AF, except the Eguchi-Hanson metric with the reversed orientation, which is ALE. Recall that AF model ends are introduced in Section 3.1. Known ALF and AF Type II gravitational instantons include

  • the Taub-NUT metric on 4\mathbb{R}^{4} (with the reversed orientation);

  • the Taub-Bolt metric on 2{point}\mathbb{C}\mathbb{P}^{2}\setminus\{point\} (with both orientations);

  • the Kerr metrics on 2×S2\mathbb{R}^{2}\times S^{2} (including the Schwarzschild metrics);

  • the Chen-Teo metrics 2S1\mathbb{C}\mathbb{P}^{2}\setminus S^{1} ([14, 3]).

The first two classes are ALF and the second two classes are AF. These metrics can all be written down explicitly in terms of algebraic formulae and they are all toric (i.e. their isometry group contains a 2 dimensional torus T2T^{2}). It was discovered by Biquard-Gauduchon [6] that all of them can be re-constructed using the LeBrun-Tod ansatz in complex geometry in terms of axi-symmetric harmonic functions in 3\mathbb{R}^{3}, and moreover any toric Type II ALF/AF gravitational instanton must belong to one of the 4 above families.

If (M,g)(M,g) is Type II, then suppose there is a global complex structure JJ, which can always be achieved by passing to a double cover. We denote λ24|W+(g)|>0\lambda\equiv\sqrt{24}|W^{+}(g)|>0, and g~=λ2/3g\tilde{g}=\lambda^{2/3}g the conformal metric. Then g~\tilde{g} is Kähler with respect to an integrable complex structure JJ. In fact, g~\tilde{g} is extremal in the sense of Calabi, namely, the extremal vector field 𝒦Jg~S(g~)\mathcal{K}\equiv J\nabla_{\tilde{g}}S(\tilde{g}) is a non-zero holomorphic Killing field. Moreover, we have S(g~)2=λ2/3S(\tilde{g})^{2}=\lambda^{2/3}. The following is an easy fact.

Lemma 4.8.

S(g~)>0S(\tilde{g})>0.

Proof.

Suppose otherwise, then S(g~)<0S(\tilde{g})<0. By Cheeger-Tian’s ϵ\epsilon-regularity theorem [8] we know |Rm(g)||Rm(g)| decays quadratically fast at infinity. In particular we have S(g~)0S(\tilde{g})\rightarrow 0. So S(g~)S(\tilde{g}) must achieve a global negative minimum at some point, say pp, which is the global maximum of S(g~)1S(\widetilde{g})^{-1}. From the formula of scalar curvature under conformal change we have

S(g~)3(6Δg~+S(g~))S(g~)1=0.S(\widetilde{g})^{3}\left(-6\Delta_{\widetilde{g}}+S(\widetilde{g})\right)S(\widetilde{g})^{-1}=0.

But at pp this gives Δg~(S(g~)1)0\Delta_{\tilde{g}}(S(\tilde{g})^{-1})\geq 0. Contradiction. ∎

Similar to classification of asymptotic geometry of hyperkähler metrics at infinity, Theorem 4.3, we have

Theorem 4.9 ([28]).

A Type II Ricci-flat metric with |Rm|2<\int|Rm|^{2}<\infty on an end that is complete at infinity must be conical and asymptotic to one of the following families of model ends: ALE, ALF, AF, skewed special Kasner, ALH\text{ALH}^{*}, or further 2\mathbb{Z}_{2} quotients of the AF, skewed special Kasner, ALH\text{ALH}^{*} models.

Notice that the ALH\text{ALH}^{*} model ends are Type II under the orientation opposite to their hyperkähler orientation. The skewed special Kasner models are quotients of the following special Kasner metric

(4.1) dρ2+ρ4/3dx12+ρ4/3dx22+ρ2/3dx32d\rho^{2}+\rho^{4/3}dx_{1}^{2}+\rho^{4/3}dx_{2}^{2}+\rho^{-2/3}dx_{3}^{2}

on [1,)×T3[1,\infty)\times T^{3}. These model ends are not 𝒜\mathcal{AF}. The special Kasner metric is Type II under both orientations, since it is hermitian under the complex structures J1:dρρ1/3dx3,dx1dx2J_{1}:d\rho\mapsto\rho^{-1/3}dx_{3},dx_{1}\mapsto dx_{2} and J2:dρρ1/3dx3,dx1dx2J_{2}:d\rho\mapsto-\rho^{-1/3}dx_{3},dx_{1}\mapsto dx_{2}. Note that there is indeed a family of Riemannian Kasner metrics that are Ricci-flat, but only the specific one mentioned above is Type II. That is why we refer to it as special Kasner. Notice that by Lemma 4.8 the skewed special Kasner models and ALH\text{ALH}^{*} models cannot be filled in as Type II gravitational instantons. We moreover have

Theorem 4.10 ([28]).

A Type II gravitational instanton with non-Euclidean volume growth must belong to the above 4 families.

This result particularly confirms a conjecture of Aksteiner-Andersson [3]. The proof depends on studying the compactification of the underlying complex manifold to a log del Pezzo surface, the theory of complex surfaces, and the geometry of complete extremal Kähler metrics. Notice that the conformal Kähler metric associated to the model ALF/AF end is a Bach-flat extremal Kähler metric with a Poincaré type cusp.

In general we expect

Conjecture 4.11 (Type II classification).

All Type II gravitational instantons are toric and must be ALF or AF, except the Eguchi-Hanson metric with the reversed orientation.

By Theorem 4.10, it suffices to study Type II ALE gravitational instantons. Note that the partial conclusion of [29], Theorem 4.2, is that for a Type II ALE gravitational instanton, when the structure group at infinity is in SU(2)SU(2), it must be the reversed Eguchi-Hanson.

The remaining case of Type III gravitational instantons are poorly understood at present. Imposing toric symmetries there are relations to integrable systems and axi-symmetric harmonic maps, but there is no general theory yet. The only example to our knowledge is constructed by Khuri-Reiris-Weinstein-Yamada [24], which is asymptotic to the above special Kasner end (which is Type II). It has quadratic volume growth and admits an effective T2T^{2} action. But these examples are not simply-connected; indeed they have fundamental group \mathbb{Z} with universal cover unwrapping one of the circle direction. They are asymptotic to a ray but they are not 𝒜\mathcal{AF}.

Question 4.12.

Are there simply-connected Type III gravitational instantons?

Question 4.13.

Are there Type III 𝒜\mathcal{AF} gravitational instantons?

As a first step towards the classification of Type III gravitational instantons, it is important to understand their asymptotic behavior. Under the assumption of faster than quadratic curvature decay and a technical assumption regarding asymptotic holonomy, Chen-Li [13] obtained an analogous result to Chen-Chen [10]. These ends are given by flat torus bundles over flat domains. There are still some fundamental questions that remain to be answered.

Question 4.14 (Conicity).

Is every gravitational instanton conical?

Under the assumption of being strongly 𝒜\mathcal{AF} this is known by the general result of Kasue [23]. Even under the assumption of 𝒜\mathcal{AF} this question is open in general.

To connect with topology, recall that even the standard smooth 4\mathbb{R}^{4} admits more than one gravitational instantons, namely, the standard flat metric and the Taub-NUT metric.

Question 4.15.

Classify gravitational instantons on topological 4\mathbb{R}^{4} and other manifolds with small topology.

As a consequence of Theorem 1.2 and the result in [13] we know

Corollary 4.16.

The only gravitational instantons on a topological 4\mathbb{R}^{4} with curvature faster than quadratic decay are given by the flat metric and the Taub-NUT metric up to scaling.

Remark 4.17.

Without the finite energy condition, the question is more complicated. We do not know whether an exotic 4\mathbb{R}^{4} admit an complete metric with vanishing (or non-negative) Ricci curvature.

References

  • [1] U. Abresch. Lower curvature bounds, Toponogov’s theorem, and bounded topology. Ann. Sci. Ec. Norm. Supér., 18(4):651–670, 1985.
  • [2] U. Abresch. Uber das glatten Riemann’scher metriken. Rheinischen Friedrich-Wilhelms-Universitat, Bonn, 1988.
  • [3] S. Aksteiner and L. Andersson. Gravitational instantons and special geometry. arXiv:2112.11863, 2021.
  • [4] M. Anderson, P. Kronheimer, and C. LeBrun. Complete Ricci-flat Kähler manifolds of infinite topological type. Comm. Math. Phys., 125(4):637–642, 1989.
  • [5] Shigetoshi Bando, Atsushi Kasue, and Hiraku Nakajima. On a construction of coordinates at infinity on manifolds with fast curvature decay and maximal volume growth. Invent. Math., 97(2):313–349, 1989.
  • [6] O. Biquard and P. Gauduchon. On Toric Hermitian ALF Gravitational Instantons. Comm. Math. Phys., 399(1):389–422, 2023.
  • [7] J. Cheeger, K. Fukaya, and M. Gromov. Nilpotent structures and invariant metrics on collapsed manifolds. J. Amer. Math. Soc., 5(2):327–372, 1992.
  • [8] J. Cheeger and G. Tian. Curvature and injectivity radius estimates for Einstein 4-manifolds. J. Amer. Math. Soc., 19(2):487–525, 2005.
  • [9] G. Chen and X. Chen. Gravitational instantons with faster than quadratic curvature decay (II). J. Reine Angew. Math., 2019(756):259–284, 2017.
  • [10] G. Chen and X. Chen. Gravitational instantons with faster than quadratic curvature decay (I). Acta Math., 227(2):263–307, 2021.
  • [11] G. Chen and X. Chen. Gravitational instantons with faster than quadratic curvature decay (III). Math. Ann., 380(1-2):687–717, 2021.
  • [12] G. Chen, J. Viaclovsky, and R. Zhang. Torelli-type theorems for gravitational instantons with quadratic volume growth. Duke Math. J., 173(2):227–275, 2024.
  • [13] X. Chen and Y. Li. On the geometry of asymptotically flat manifolds. Geom. Topol., 25(5):2469–2572, 2021.
  • [14] Y. Chen and E. Teo. A new AF gravitational instanton. Physics Letters B, 703(3):359–362, 2011.
  • [15] T. Collins, A. Jacob, and Y. Lin. The Torelli theorem for ALHALH^{*} gravitational instantons. Forum of Mathematics, Sigma, 10(e79), 2022.
  • [16] I. Şuvaina. ALE Ricci-flat Kähler metrics and deformations of quotient surface singularities. Ann. Glob. Anal. Geom., 41(1):109–123, 2021.
  • [17] A. Derdziński. Self-dual Kähler manifolds and Einstein manifolds of dimension four. Compos. Math., 49(3):405–433, 1983.
  • [18] J.-H. Eschenburg and V. Schroeder. Riemannian manifolds with flat ends. Math. Z., 196(4):573–589, 1987.
  • [19] K. Fukaya. Collapsing of Riemannian manifolds and eigenvalues of Laplace operator. Invent. Math., 87(3):517–547, 1987.
  • [20] K. Hattori. The nonuniqueness of the tangent cones at infinity of Ricci-flat manifolds. Geom. Topol., 21(5):2683–2723, 2017.
  • [21] H. Hein. Gravitational instantons from rational elliptic surfaces. J. Amer. Math. Soc., 25(2):355–393, 2012.
  • [22] H. Hein, S. Sun, J. Viaclovsky, and R. Zhang. Asymptotically Calabi metrics and weak Fano manifolds. arXiv:2111.09287, 2022.
  • [23] A. Kasue. A compactification of a manifold with asymptotically nonnegative curvature. Ann. Sci. Ec. Norm. Supér., 21(4):593–622, 1988.
  • [24] M. Khuri, M. Reiris, G. Weinstein, and S. Yamada. Gravitational Solitons and Complete Ricci Flat Riemannian Manifolds of Infinite Topological Type. arXiv:2204.08048, 2022.
  • [25] P. Kronheimer. The construction of ALE spaces as hyper-Kähler quotients. J. Differential Geom., 29(3):665–683, 1989.
  • [26] H. Kunduri and J. Lucietti. Existence and uniqueness of asymptotically flat toric gravitational instantons. Lett. Math. Phys., 111(133), 2021.
  • [27] T. Lee and Y. Lin. Period domains for gravitational instantons. arXiv:2208.13083, 2022.
  • [28] M. Li. Classification results for conformally Kähler gravitational instantons. arXiv:2310.13197, 2023.
  • [29] M. Li. On 4-dimensional Ricci-flat ALE manifolds. arXiv:2304.01609, 2023.
  • [30] H. Liu. A short proof of the surjectivity of the period map on K3 manifolds. arXiv:2304.09501, 2023.
  • [31] Y. Mashiko, K. Nagano, and K. Otsuka. The asymptotic cones of manifolds of roughly non-negative radial curvature. J. Math. Soc. Japan, 57(1):55–68, 2005.
  • [32] V. Minerbe. On the asymptotic geometry of gravitational instantons. Ann. Sci. Ec. Norm. Supér., 43(6):883–924, 2010.
  • [33] P. Petersen, G. Wei, and R. Ye. Controlled Geometry via Smoothing. Comment. Math. Helv., 74(3):345–363, 1999.
  • [34] A. Petrunin, X. Rong, and W. Tuschmann. Collapsing vs. Positive Pinching. Geom. Funct. Anal., 9(4):699–735, 1999.
  • [35] A. Petrunin and W. Tuschmann. Asymptotical flatness and cone structure at infinity. Math. Ann., 321(4):775–788, 2001.
  • [36] X. Rong. On the fundamental groups of manifolds of positive sectional curvature. Ann. of Math., 143(2):397–411, 1996.
  • [37] S. Sun and R. Zhang. Complex structure degenerations and collapsing of Calabi-Yau metrics. arXiv:1906.03368, 2019.
  • [38] S. Sun and R. Zhang. Collapsing geometry of hyperkähler 4-manifolds and applications. arXiv:2108.12991, 2021.
  • [39] S. Unnebrink. Asymptotically flat 4-manifolds. Differ. Geom. Appl., 6(3):271–274, 1996.
  • [40] E. Wright. Quotients of gravitational instantons. Ann. Glob. Anal. Geom., 41(1):91–108, 2021.