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

Uniqueness of shrinking gradient Kähler-Ricci solitons on non-compact toric manifolds

Charles Cifarelli charles_cifarelli@berkeley.edu
Abstract

We show that, up to biholomorphism, there is at most one complete TnT^{n}-invariant shrinking gradient Kähler-Ricci soliton on a non-compact toric manifold MM. We also establish uniqueness without assuming TnT^{n}-invariance if the Ricci curvature is bounded and if the soliton vector field lies in the Lie algebra 𝔱\mathfrak{t} of TnT^{n}. As an application, we show that, up to isometry, the unique complete shrinking gradient Kähler-Ricci soliton with bounded scalar curvature on 1×\mathbb{C}\mathbb{P}^{1}\times\mathbb{C} is the standard product metric associated to the Fubini-Study metric on 1\mathbb{C}\mathbb{P}^{1} and the Euclidean metric on \mathbb{C}.

1 Introduction

A Ricci soliton (M,g,X)(M,g,X) is a Riemannian manifold (M,g)(M,g) together with a vector field XX satisfying

Ricg+12Xg=λ2g\text{Ric}_{g}+\frac{1}{2}\mathcal{L}_{X}g=\frac{\lambda}{2}g (1.1)

for λ\lambda\in\mathbb{R}. By a simultaneous rescaling of XX and gg, we can always assume that a Ricci soliton is normalized so that λ{1,0,+1}\lambda\in\{-1,0,+1\}. We will always assume that the metric gg is complete, which in turn forces the vector field XX to be complete [53]. A Ricci soliton is said to be gradient if the vector field XX is the gradient of a smooth function ff, usually called the soliton potential. In this case the equation becomes

Ricg+g2f=λ2g.\text{Ric}_{g}+\nabla_{g}^{2}f=\frac{\lambda}{2}g. (1.2)

If gg is a Kähler metric on MM with Kähler form ω\omega, we say that (M,ω,X)(M,\omega,X) is a Kähler-Ricci soliton if ω\omega satisfies the equation

Ricω+12Xω=λω,\text{Ric}_{\omega}+\frac{1}{2}\mathcal{L}_{X}\omega=\lambda\omega, (1.3)

where Ricω\text{Ric}_{\omega} is the Ricci form and λ{1,0,+1}\lambda\in\{-1,0,+1\}. The coefficients appearing in (1.3) are chosen to be different from those in (1.1); this choice being more natural from the perspective of the Kähler-Ricci flow. Ricci solitons and Kähler-Ricci solitons are called expanding, steady, and shrinking, respectively when λ{1,0,+1}\lambda\in\{-1,0,+1\}. In this paper we will only consider shrinking solitons and so we will always assume that λ=1\lambda=1. As for Ricci solitons, we say that a shrinking Kähler-Ricci soliton is gradient if X=gfX=\nabla_{g}f, in which case (1.3) takes the form

Ricω+i¯f=ω.\text{Ric}_{\omega}+i\partial\bar{\partial}f=\omega. (1.4)

Ricci solitons are interesting both from the perspective of canonical metrics and of Ricci flow. On the one hand, they represent one direction in which one can generalize the concept of an Einstein manifold. On compact manifolds, shrinking solitons are known to exist in several situations where there are obstructions to the existence of Einstein metrics; see for example [50]. By the maximum principle, there are no nontrivial expanding or steady solitons on compact manifolds. There are many examples on noncompact manifolds, however; see for example [12, 13, 27] and the references therein. On the other hand, one can associate to a Ricci soliton a self-similar solution of the Ricci flow, and gradient shrinking Ricci solitons in particular provide models for finite-time Type I singularities along the flow [24, 40]. Even in complex dimension two, however, it is not known which shrinking Ricci solitons arise in this way. From this perspective, it is an important problem to classify shrinking gradient Kähler-Ricci solitons in order to better understand the singularity development along the Kähler-Ricci flow.

In this paper we study Kähler-Ricci solitons on non-compact complex manifolds MM under the additional assumption that MM is toric. For the purposes of this paper, a complex toric manifold is a smooth nn-dimensional complex manifold (M,J)(M,J) together with an effective holomorphic action of the complex torus ()n(\mathbb{C}^{*})^{n}. In such a setting there always exists an orbit UMU\subset M of the ()n(\mathbb{C}^{*})^{n}-action which is open and dense in MM. Moreover, we always assume that there are only finitely many points which are fixed by the ()n(\mathbb{C}^{*})^{n}-action. The ()n(\mathbb{C}^{*})^{n}-action of course determines the action of the real torus Tn()nT^{n}\subset(\mathbb{C}^{*})^{n}, and our main theorem is a uniqueness result for complete shrinking gradient Kähler-Ricci solitons which are invariant under this action.

Theorem A.

Suppose that (M,J)(M,J) is a non-compact complex toric manifold and that the fixed point set of the ()n(\mathbb{C}^{*})^{n}-action is finite. Then, up to biholomorphism, there is at most one complete TnT^{n}-invariant shrinking gradient Kähler-Ricci soliton (g,X)(g,X) on (M,J)(M,J).

As we will see, TnT^{n}-invariance implies that the holomorphic vector field JXJX associated to the soliton vector field XX lies in the Lie algebra 𝔱\mathfrak{t} of the real torus TnT^{n}. There is also a notion of a toric manifold coming purely from symplectic geometry. To distinguish this from the definition above, we say that a symplectic toric manifold is an nn-dimensional symplectic manifold (M,ω)(M,\omega) together with an effective Hamiltonian action of the real torus TnT^{n}. As before, we will always assume in this paper that the fixed point set of the TnT^{n}-action is finite. We remark here that this assumption is non-trivial; see for example [35, Example 6.9].

Of course, the intersection of these ideas naturally lies in the realm of Kähler geometry. In particular, if (M,J)(M,J) is a complex toric manifold as above and ω\omega is the Kähler form of a compatible Kähler metric gg on MM with respect to which the real TnT^{n}-action is Hamiltonian, then the symplectic manifold (M,ω)(M,\omega) is naturally a symplectic toric manifold. When (M,J,ω)(M,J,\omega) is a compact Kähler manifold, then the two definitions are equivalent in the following sense. Suppose that (M,J,ω)(M,J,\omega) admits an effective Hamiltonian and holomorphic action of the real nn-dimensional torus TnT^{n}, so that (M,ω)(M,\omega) in particular carries the structure of a symplectic toric manifold. Then this action can always be complexified to an action of the full complex torus ()n(\mathbb{C}^{*})^{n}, giving (M,J)(M,J) the structure of a complex toric manifold. This can be done essentially because any vector field on MM is complete. Of course in the non-compact setting this is no longer the case, and so it makes sense to ask if Theorem A can be extended to the more general setting of symplectic toric manifolds. We prove this under the additional assumption that the Ricci curvature of gg is bounded, i.e. supxM|Ricg|g(x)<\sup_{x\in M}|\text{Ric}_{g}|_{g}(x)<\infty.

Theorem B.

Suppose that (M,J)(M,J) is a non-compact complex manifold with dimM=n\dim_{\mathbb{C}}M=n, together with an effective holomorphic action of a real torus TnT^{n} with Lie algebra 𝔱\mathfrak{t} and finite fixed point set. Then, up to biholomorphism, there is at most one complete shrinking gradient Kähler-Ricci soliton (g,X)(g,X) on (M,J)(M,J) with JX𝔱JX\in\mathfrak{t} and with bounded Ricci curvature.

Notice that in this case we do not need to assume that gg is TnT^{n}-invariant, only that the Ricci curvature is bounded and JX𝔱JX\in\mathfrak{t}. In fact, we will see in Section 4 that any Kähler-Ricci soliton satifsying these hypotheses is isometric to a TnT^{n}-invariant one. When MM is compact, these results are special cases of the general uniqueness theorem of Tian-Zhu [46, 47]. The non-compact case is generally much more delicate. Typically, one needs to prescribe the asymptotics of the metric, for example by imposing a fixed model metric at infinity, in order to work in well-behaved function spaces. An important feature of this work is that we do not impose any assumptions on the specific behavior of the metric at infinity. Instead, a generalization of the setup of Berman-Berndtsson [8] allows us to work with the Ding functional on broadly defined L1L^{1}-type spaces; see Section 3 for details. For a result of even greater generality in the special case when M=M=\mathbb{C}, see also [49].

As an application of Theorem B, we prove a stronger uniqueness result for the special case of M=1×M=\mathbb{CP}^{1}\times\mathbb{C}.

Corollary C.

Up to isometry, the standard product of the Fubini-Study and Euclidean metrics is the unique complete shrinking gradient Kähler-Ricci soliton with bounded scalar curvature on 1×\mathbb{CP}^{1}\times\mathbb{C}.

The point is that, if one has a complete shrinking gradient Kähler-Ricci soliton with bounded scalar curvature on MM, then it suffices to assume that JXJX lies in the Lie algebra of the standard torus acting on 1×\mathbb{CP}^{1}\times\mathbb{C}, in which case Theorem B applies directly. This is achieved in Section 4 by a Morse theoretic argument similar to the one implemented in [13, Proposition 2.27].

Other well-known examples of complete shrinking gradient Kähler-Ricci solitons include those constructed by Feldman-Ilmanen-Knopf [25] on the total spaces of the line bundles 𝒪(k)n1\mathcal{O}(-k)\to\mathbb{C}\mathbb{P}^{n-1} for 0<k<n0<k<n, which were recently shown to be the unique such metrics on these manifolds with bounded Ricci curvature in [13, Theorem E]. In fact, there are recent examples of Futaki [27] which generalize this construction on the total space of any root of the canonical bundle of a compact toric Fano manifold (see also [29, 52] for the case where the soliton vector field generates an S1S^{1}-action). All of the aforementioned examples are toric in the sense that underlying manifold is always a complex toric manifold and the metric is invariant under the action of the corresponding real torus TnT^{n}. As before, we denote the Lie algebra of this fixed real torus by 𝔱\mathfrak{t}. As a direct consequence of Theorem B, we have that these are the only examples of shrinking gradient Kähler-Ricci solitons on these manifolds with bounded Ricci curvature and with JX𝔱JX\in\mathfrak{t}.

Corollary D.

Let NN be an (n1)(n-1)-dimensional toric Fano manifold, LNL\to N be a holomorphic line bundle such that Lp=KNL^{p}=K_{N} with 0<p<n0<p<n, and let MM denote the total space of LL. There is a natural action of the real torus TnT^{n} on MM, and we denote the Lie algebra by 𝔱\mathfrak{t}. Then, up to biholomorphism, there is a unique complete shrinking gradient Kähler-Ricci soliton with bounded Ricci curvature and JX𝔱JX\in\mathfrak{t} on MM, namely the one constructed by Futaki in [27].

We also study the weighted volume functional FF on a complex toric manifold. This was introduced by Tian-Zhu [47] for compact manifolds, and is by definition a convex function on the space 𝔥\mathfrak{h} of all real holomorphic vector fields on (M,J)(M,J). As in [47], the derivative of FF at a given holomorphic vector field can be viewed as a generalization of the Futaki invariant. The upshot is that if (g,X)(g,X) is a complete shrinking gradient Kähler-Ricci soliton on MM, then JXJX is necessarily the unique critical point of FF. As a result, the vector field XX associated to a complete shrinking gradient Kähler-Ricci soliton on (M,J)(M,J) is unique. It was shown in [13] using the Duistermaat-Heckman theorem [22, 23, 42] that FF can be defined in the non-compact setting in the presence of a holomorphic TkT^{k}-action when the metric gg has bounded Ricci curvature. More precisely, there is an open cone Λ𝔱\Lambda\subset\mathfrak{t}, comprising those holomorphic vector fields which admit Hamiltonian potentials which are proper and bounded from below, on which FF is well-defined. Just as in [42], we will see in Section 3 that, in the toric setting, there is a natural identification of Λ\Lambda with a certain open convex cone C𝔱C^{*}\subset\mathfrak{t} determined by the ()n(\mathbb{C}^{*})^{n}-action on (M,J)(M,J). Furthermore, any soliton vector field XX with JX𝔱JX\in\mathfrak{t} necessarily has the property that JXΛJX\in\Lambda and is the unique critical point of FF, which in turn gives uniqueness among all holomorphic vector fields YY with JY𝔱JY\in\mathfrak{t} [13, Theorem D].

We show that on a complex toric manifold, the weighted volume functional FF is proper on Λ\Lambda, and therefore that there exists a unique candidate holomorphic vector field XX with JX𝔱JX\in\mathfrak{t} that could be associated to a complete shrinking gradient Kähler-Ricci soliton. Here we make no assumptions on the curvature. Thus, we recover an analog of [13, Theorem D] when the torus is full-dimensional, without having to assume a Ricci curvature bound; see Theorem 4.6 below for the precise statement.

The main theorems here also give partial answers to some open questions raised in [13, Section 7.2]. Namely, we obtain a positive answer to question 7 assuming that the torus is the real torus underlying an effective holomorphic and full-dimensional ()n(\mathbb{C}^{*})^{n}-action with finite fixed point set, and a positive answer to question 2 with the same assumption on the torus as well as the assumption that either gg is invariant or that gg has a Ricci curvature bound. We also show that any symplectic toric manifold with finite fixed point set admitting a compatible complete shrinking gradient Kähler-Ricci soliton is quasiprojective, which gives a positive answer to question 1 in the toric setting. Finally, we show that the weighted volume functional FF is proper on a complex toric manifold with finite fixed point set, which gives a positive answer to question 9 when the real torus is full-dimensional and admits a complexification. As we will see, this is always the case in the presence of an invariant solution to (1.4).

Since the foundational work of Delzant [16] and Guillemin [30] (which themselves relied on the earlier foundational work of Atiyah [6] and Guillemin-Sternberg [31]), toric manifolds have played a key role in the study of special Kähler metrics on compact Kähler manifolds; see [1, 18, 50] and many others. As a consequence of this setup, many aspects of the Kähler geometry of TnT^{n}-invariant metrics on MM reduce to questions about convex functions on a given polytope PP in n\mathbb{R}^{n}. We show that under certain mild hypotheses, much of the structure from the compact setting carries over, replacing the bounded polytopes with potentially unbounded polyhedra. In the purely symplectic setting, there has been much work done in this direction, spanning many years; see [7, 32, 35, 36, 42]. There has been somewhat less attention focused on the Kähler case, and our work draws significantly on the notable exceptions of [4, 10, 37, 48]. There has also been recent progress in the Kähler setting on singular toric varieties; see [10] and of particular relevance to this paper [8].

The paper is organized as follows. In Section 2 we recall some of the basics of toric geometry from both the algebraic and symplectic perspectives. We show that the Abreu-Guillemin setup can be extended with the appropriate assumptions to non-compact manifolds. Much of this material seems to be fairly well-known in the symplectic setting, and we simply provide a rephrasing particularly suited for Kähler geometry. In particular, we give conditions under which the familiar Delzant classification holds in the non-compact setting. In Section 3 we study properties of some real Monge-Ampère equations on unbounded convex domains in n\mathbb{R}^{n}, and explain how these relate to the Kähler-Ricci soliton equation on toric manifolds. We introduce a Ding-type functional 𝒟\mathcal{D} on the appropriate space of symplectic potentials and use its convexity to determine uniqueness. Much of what appears here is drawn from [8] and [20]. A result of Wylie [51] implies that any complete shrinking gradient Kähler-Ricci soliton admits a moment map. In Section 4, we use this to apply the results of the previous sections to complete the proofs of Theorem A and Theorem B. We also include in Section 4 a proof of Corollary D, which amounts to demonstrating that the examples constructed in [27] indeed have bouned Ricci curvature. We conclude with an application of our work to the special case of M=1×M=\mathbb{CP}^{1}\times\mathbb{C}, and show that a complete shrinking gradient Kähler-Ricci soliton on MM is isometric to the standard product metric. This is the content of Corollary C.

Acknowledgements

I would like to thank Song Sun for suggesting this problem, as well as for his continued support and many useful discussions. I would like to thank Ronan Conlon for his interest, useful discussions, and detailed commentary on the preliminary versions of this article. I would also like to thank Vestislav Apostolov for his invaluable suggestions and Akito Futaki for his interest and comments. I am also grateful for the many conversations with Chris Eur in which I learned much of what I know about the algebraic geometry of toric varieties, and to the referee for their insightful suggestions. This work was partially supported by the National Science Foundation RTG grant DMS-1344991.

2 Kähler geometry on non-compact toric manifolds

2.1 Algebraic preliminaries

We begin by recalling some basics from algebraic toric geometry that we will use later on. The main reference here is [15]. Fix an algebraic torus ()n(\mathbb{C}^{*})^{n} and let 𝔱\mathfrak{t} be the Lie algebra of the real torus Tn()nT^{n}\subset(\mathbb{C}^{*})^{n}. Fix an integer lattice Γ𝔱\Gamma\subset\mathfrak{t} so that ()n𝔱i𝔱/Γ(\mathbb{C}^{*})^{n}\cong\mathfrak{t}\oplus i\mathfrak{t}/\Gamma acting only in the second factor. Let Γ\Gamma^{*} denote the corresponding dual lattice in 𝔱\mathfrak{t}^{*}.

Definition 1.

A toric variety MM is an algebraic variety together with the effective algebraic action of the complex torus ()n(\mathbb{C}^{*})^{n} with a dense orbit. More precisely, this means that the action ()n×MM(\mathbb{C}^{*})^{n}\times M\to M is a morphism of algebraic varieties, and there exists a point pMp\in M such that the orbit ()npM(\mathbb{C}^{*})^{n}\cdot p\subset M is Zariski open and dense in MM.

We emphasize that, contrary to the definitions presented in the introduction, a toric variety MM is always assumed to be algebraic. As we will see, the fixed point set of the ()n(\mathbb{C}^{*})^{n}-action associated to a toric variety is necessarily finite. In particular, the underlying complex manifold of a smooth toric variety is always a complex toric manifold as defined in the introduction.

The algebraic geometry of toric varieties has a rich interplay with combinatorics, which is integral to many of the constructions that follow. We begin by introducing the relevant combinatorial objects.

Definition 2.

A polyhedron is any finite intersection of affine half spaces Hν,a={x𝔱|ν,xa}H_{\nu,a}=\{x\in\mathfrak{t}^{*}\>|\>\langle\nu,x\rangle\geq a\} with ν𝔱,a\nu\in\mathfrak{t},a\in\mathbb{R}. A polytope is a bounded polyhedron.

We will often not distinguish between a polyhedron PP and its interior, but where confusion may arise we will denote by P¯\overline{P} the closed object and PP the interior. The intersection of PP with the plane ν,x=a\langle\nu,x\rangle=a is a polyhedron FνF_{\nu} of one less dimension and is called a facet of PP. The intersections of any number of the FνF_{\nu}’s form the collection of faces of PP.

Definition 3.

Let PP be a polyhedron given by the intersection of the half spaces Hνi,aiH_{\nu_{i},a_{i}}. We define the recession cone (or asymptotic cone) CC of PP by

C={x𝔱|νi,x0}.C=\left\{x\in\mathfrak{t}^{*}\>|\>\langle\nu_{i},x\rangle\geq 0\right\}.

Given any convex cone C𝔱C\subset\mathfrak{t}, the dual cone C𝔱C^{*}\subset\mathfrak{t} is defined by

C={ξ𝔱|ξ,x0 for all xC}.C^{*}=\{\xi\in\mathfrak{t}\>|\>\langle\xi,x\rangle\geq 0\text{ for all }x\in C\}. (2.1)

Note that (the interior of) CC^{*} is necessarily an open cone in 𝔱\mathfrak{t}, even when CC is not full-dimensional.

Definition 4.

Let PP be a polyhedron. If the vertices of PP lie in the dual lattice Γ𝔱\Gamma^{*}\subset\mathfrak{t}^{*}, then we say that PP is rational.

Rational polyhedra play an important role in the algebraic geometry of toric varieties, in that each such PP determines a unique quasiprojective toric variety P\mathcal{M}_{P}. This procedure is constructive and can be understood via the introduction of a fan. A rational polyhedral cone σ\sigma is by definition a convex subset of 𝔱\mathfrak{t} of the form

σ={λiνi|λi+},\sigma=\left\{\sum\lambda_{i}\nu_{i}\>|\>\lambda_{i}\in\mathbb{R}_{+}\right\},

where ν1,,νkΓ\nu_{1},\dots,\nu_{k}\in\Gamma is a fixed finite collection of lattice points. The recession cone CC of a rational polyhedron PP is always a rational polyhedral cone [15, Chapter 7].

Definition 5.

A fan Σ\Sigma in 𝔱\mathfrak{t} is a finite set consisting of rational polyhedral cones σ\sigma satisfying

  1. 1.

    For every σΣ\sigma\in\Sigma, each face of σ\sigma also lies in Σ\Sigma.

  2. 2.

    For every pair σ1,σ2Σ\sigma_{1},\sigma_{2}\in\Sigma, σ1σ2\sigma_{1}\cap\sigma_{2} is a face of each.

We will also assume that the support of Σ\Sigma is full-dimensional, that is to say, there exists at least one nn-dimensional cone σΣ\sigma\in\Sigma. To every fan Σ\Sigma there is an associated toric variety Σ\mathcal{M}_{\Sigma}. We will give a very brief summary of this construction below; for more details see [15, Chapter 3]. For us the main point is the following corollary of a result of Sumihiro [45]:

Proposition 2.1 ([15, Corollary 3.1.8]).

Let MM be a toric variety. Then there exists a fan Σ\Sigma such that MΣM\cong\mathcal{M}_{\Sigma}.

To construct Σ\mathcal{M}_{\Sigma} from Σ\Sigma, one begins by taking each nn-dimensional cone σΣ\sigma\in\Sigma and constructing an affine toric variety UσU_{\sigma}. We define the dual cone σ\sigma^{*} of σ\sigma by (2.1):

σ={x𝔱|x,ξ0 for all ξσ}.\sigma^{*}=\left\{x\in\mathfrak{t}^{*}\>|\>\langle x,\xi\rangle\geq 0\text{ for all }\xi\in\sigma\right\}.

Let SσS_{\sigma} be the semigroup of those lattice points which lie in σ\sigma^{*} under addition. Then one defines the semigroup ring, as a set, as all finite sums of the form

[Sσ]={λss|sSσ}.\mathbb{C}[S_{\sigma}]=\left\{\sum\lambda_{s}s\>|\>s\in S_{\sigma}\right\}.

The ring structure is then defined on monomials by λs1s1λs2s2=(λs1λs2)(s1+s2)\lambda_{s_{1}}s_{1}\cdot\lambda_{s_{2}}s_{2}=(\lambda_{s_{1}}\lambda_{s_{2}})(s_{1}+s_{2}) and extended in the natural way. The basic example is σ=+n\sigma=\mathbb{R}^{n}_{+}, where [Sσ]\mathbb{C}[S_{\sigma}] is naturally isomorphic to [z1,,zn]\mathbb{C}[z_{1},\dots,z_{n}]. Then the affine variety UσU_{\sigma} is defined to be Spec([Sσ])\text{Spec}(\mathbb{C}[S_{\sigma}]). This is automatically endowed with a ()n(\mathbb{C}^{*})^{n}-action with an open dense orbit. This construction of course can be implemented on the lower-dimensional cones τΣ\tau\in\Sigma. If σ1σ2=τ\sigma_{1}\cap\sigma_{2}=\tau, then there is a natural way to map UτU_{\tau} into Uσ1U_{\sigma_{1}} and Uσ2U_{\sigma_{2}} isomorphically. Thus one constructs Σ\mathcal{M}_{\Sigma} by declaring the collection of all UσU_{\sigma} to be an open affine cover with transition data determined by UτU_{\tau}. An important property of this construction is the Orbit-Cone correspondence.

Proposition 2.2 (Orbit-Cone correspondence, [15, Theorem 3.2.6]).

Let Σ\Sigma be a fan and Σ\mathcal{M}_{\Sigma} be the associated toric variety. The kk-dimensional cones σΣ\sigma\in\Sigma are in natural one-to-one correspondence with the (nk)(n-k)-dimensional orbits OσO_{\sigma} of the ()n(\mathbb{C}^{*})^{n}-action on Σ\mathcal{M}_{\Sigma}. Moreover, given a kk-dimensional cone σΣ\sigma\in\Sigma and a corresponding orbit OσΣO_{\sigma}\subset\mathcal{M}_{\Sigma}, we have that σ\sigma lies as an open subset of the Lie algebra 𝔱σ\mathfrak{t}_{\sigma} of the kk-dimensional real subtorus TσTnT_{\sigma}\subset T^{n} that stabilizes the points on OσO_{\sigma}.

In particular, the fixed point set of the ()n(\mathbb{C}^{*})^{n}-action is in natural bijection with the full-dimensional cones in Σ\Sigma, and is therefore always finite. At the other extreme, each ray σΣ\sigma\in\Sigma determines a unique torus-invariant divisor DσD_{\sigma}. As a consequence, a torus-invariant Weil divisor DD on Σ\mathcal{M}_{\Sigma} naturally determines a polyhedron PD𝔱P_{D}\subset\mathfrak{t}^{*} as follows. We can decompose DD uniquely as D=i=1NaiDσiD=\sum_{i=1}^{N}a_{i}D_{\sigma_{i}}, where σiΣ\sigma_{i}\in\Sigma, i=1,,Ni=1,\dots,N is the collection of rays. By assumption, there exists a unique minimal νiσiΓ\nu_{i}\in\sigma_{i}\cap\Gamma. Then set

PD={x𝔱|νi,xai for all i=1,,N}.P_{D}=\left\{x\in\mathfrak{t}^{*}\>|\>\langle\nu_{i},x\rangle\geq-a_{i}\text{ for all }i=1,\dots,N\right\}. (2.2)

The importance of polyhedra for our purposes lies in the fact that this procedure is partially reversible. That is, given a suitable polyhedron PP, one can determine a unique toric variety P\mathcal{M}_{P} through its normal fan ΣP\Sigma_{P}. To form ΣP\Sigma_{P}, one starts with a vertex vPv\in P and considers those facets FF containing vv. This determines a cone σv\sigma_{v} spanned by the inner normals νF\nu_{F} corresponding to each such FF. Then there is a unique fan ΣP\Sigma_{P} which consists of the collection of σv\sigma_{v} along with all of each of their faces. Finally, P\mathcal{M}_{P} is defined to be the toric variety associated to ΣP\Sigma_{P}. As we will see, the variety P\mathcal{M}_{P} comes naturally equipped with a divisor DD whose corresponding polyhedron is precisely PP. Moreover,

Proposition 2.3 ([15, Theorem 7.1.10]).

Let PP be a full-dimensional rational polyhedron in 𝔱\mathfrak{t}^{*}. Then the variety P\mathcal{M}_{P} constructed above is quasiprojective.

2.2 Complex coordinates

Let MM be a complex manifold together with an effective holomorphic ()n(\mathbb{C}^{*})^{n}-action. Such an action always has an open and dense orbit. Indeed, let Tn()nT^{n}\subset(\mathbb{C}^{*})^{n} be the real torus with Lie algebra 𝔱\mathfrak{t}. Choose a basis (X1,,Xn)(X_{1},\dots,X_{n}) for 𝔱\mathfrak{t}. Then each XiX_{i} is a holomorphic vector field on MM, and thus vanishes along an analytic subvariety. In particular, there is a fixed analytic subvariety VMV\subset M such that on U=MVU=M-V, none of the vector fields XiX_{i} vanish. Clearly XiX_{i} and JXiJX_{i} are complete and commute, and so the vector fields (X1,JX1,,Xn,JXn)(X_{1},JX_{1},\dots,X_{n},JX_{n}) can be integrated to determine an isomorphism U()nU\cong(\mathbb{C}^{*})^{n}. Throughout the remainder of the paper we will make heavy use of this natural coordinate system, which we usually just denote by ()nM(\mathbb{C}^{*})^{n}\subset M. In particular, we fix once and for all such a basis (X1,,Xn)(X_{1},\dots,X_{n}) for 𝔱\mathfrak{t}. This induces a background coordinate system (ξ1,,ξn)(\xi^{1},\dots,\xi^{n}) on 𝔱\mathfrak{t}. We use the natural inner product on 𝔱\mathfrak{t} to identify 𝔱𝔱\mathfrak{t}\cong\mathfrak{t}^{*} and thus can also identify 𝔱n\mathfrak{t}^{*}\cong\mathbb{R}^{n}. For clarity, we will denote the induced coordinates on 𝔱\mathfrak{t}^{*} by (x1,,xn)(x^{1},\dots,x^{n}). Let (z1,,zn)(z_{1},\dots,z_{n}) be the natural coordinates on ()n(\mathbb{C}^{*})^{n} as an open subset of n\mathbb{C}^{n}. There is a natural diffeomorphism Log:()n𝔱×Tn\text{Log}:(\mathbb{C}^{*})^{n}\to\mathfrak{t}\times T^{n}, which provides a one-to-one correspondence between TnT^{n}-invariant smooth functions on ()n(\mathbb{C}^{*})^{n} and smooth functions on 𝔱\mathfrak{t}. Explicitly, Log(z1,,zn)=(log(r1),,log(rn),θ1,,θn)\text{Log}(z_{1},\dots,z_{n})=(\log(r_{1}),\dots,\log(r_{n}),\theta_{1},\dots,\theta_{n}), where zj=rjeiθjz_{j}=r_{j}e^{i\theta_{j}}. Given a function H(ξ)H(\xi) on 𝔱\mathfrak{t}, we can extend HH trivially to 𝔱×Tn\mathfrak{t}\times T^{n} and pull back by Log to obtain a TnT^{n}-invariant function on ()n(\mathbb{C}^{*})^{n}. Clearly, any TnT^{n}-invariant function on ()n(\mathbb{C}^{*})^{n} can be written in this form.

Definition 6.

Let ω\omega be a TnT^{n}-invariant Kähler metric on MM. We say that the TnT^{n}-action is Hamiltonian with respect to the ω\omega if there exists a moment map μ\mu. This by definition is a smooth function μ:M𝔱\mu:M\to\mathfrak{t}^{*} satisfying

dμ,v=ivω,d\langle\mu,v\rangle=-i_{v}\omega,

for each v𝔱v\in\mathfrak{t} where ivi_{v} denotes the interior product and ,\langle\cdot\,,\cdot\rangle is the dual pairing.

The Kähler metrics on the complex torus ()n(\mathbb{C}^{*})^{n} itself with respect to which the standard TnT^{n}-action is Hamiltonian have a natural characterization due to Guillemin.

Proposition 2.4 ([30, Theorem 4.1]).

Let ω\omega be any TnT^{n}-invariant Kähler form on ()n(\mathbb{C}^{*})^{n}. Then the action is Hamiltonian with respect to ω\omega if and only if there exists a TnT^{n}-invariant potential ϕ\phi such that ω=2i¯ϕ\omega=2i\partial\bar{\partial}\phi.

Suppose that (M,J,ω)(M,J,\omega) admits an effective and holomorphic ()n(\mathbb{C}^{*})^{n}-action and that ω\omega is the Kähler form of a TnT^{n}-invariant compatible Kähler metric. In this context, Proposition 2.4 implies that if the TnT^{n}-action on MM is Hamiltonian with respect to ω\omega, then restriction of ω\omega to the dense orbit is ¯\partial\bar{\partial}-exact. As before, let (z1,,zn)(z_{1},\dots,z_{n}) denote the standard coordinates on ()n(\mathbb{C}^{*})^{n}. Choose any branch of log\log and write w=log(z)w=\log(z). Then clearly w=ξ+iθw=\xi+i\theta (or, more precisely, there is a corresponding lift of θ\theta to the universal cover with respect to which the equality holds), and so if ϕ\phi is TnT^{n}-invariant and ω=2i¯ϕ\omega=2i\partial\bar{\partial}\phi, we have that

ω=2i2ϕwiw¯jdwidw¯j=2ϕξiξjdξidθj.\omega=2i\frac{\partial^{2}\phi}{\partial w^{i}\partial\bar{w}^{j}}dw_{i}\wedge d\bar{w}_{j}=\frac{\partial^{2}\phi}{\partial\xi^{i}\partial\xi^{j}}d\xi^{i}\wedge d\theta^{j}.

In this setting, the metric gg corresponding to ω\omega is given on 𝔱×Tn\mathfrak{t}\times T^{n} by

g=ϕij(ξ)dξidξj+ϕij(ξ)dθidθj.g=\phi_{ij}(\xi)d\xi^{i}d\xi^{j}+\phi_{ij}(\xi)d\theta^{i}d\theta^{j}.

The moment map μ\mu as a map μ:𝔱×Tn𝔱\mu:\mathfrak{t}\times T^{n}\to\mathfrak{t}^{*} is defined by the relation

μ(ξ,θ),b=ϕ(ξ),b\langle\mu(\xi,\theta),b\rangle=\langle\nabla\phi(\xi),b\rangle

for all b𝔱b\in\mathfrak{t}, and where ϕ\nabla\phi is the Euclidean gradient of ϕ\phi. Since the Hessian of ϕ\phi is positive-definite, it follows that ϕ\phi is strictly convex on 𝔱\mathfrak{t}. In particular, ϕ\nabla\phi is a diffeomorphism onto its image. Using the identifications mentioned at the beginning of this section, we view ϕ\nabla\phi as a map from 𝔱\mathfrak{t} into an open subset of 𝔱\mathfrak{t}^{*}.

2.3 Setup of the equation

Suppose now that (g,X)(g,X) is a shrinking gradient Kähler-Ricci soliton on a complex toric manifold MM and that gg is TnT^{n}-invariant. Restricting to the dense orbit, we see that gg is determined by a convex function ϕ\phi on 𝔱\mathfrak{t}. We wish therefore to write equation (1.4) as an equation for ϕ\phi. From (1.4), we can assume by averaging that the soliton potential ff, and therefore the vector field XX, must also be TnT^{n}-invariant. Writing f=f(ξ,θ)f=f(\xi,\theta) in the real coordinate system (ξ,θ)(\xi,\theta) above, it follows that ff is independent of θ\theta. Therefore we have that

X=gf=ϕijfξiξj.X=\nabla_{g}f=\phi^{ij}\frac{\partial f}{\partial\xi^{i}}\frac{\partial}{\partial\xi^{j}}. (2.3)

In fact, the coefficients ϕijfξi\phi^{ij}\frac{\partial f}{\partial\xi^{i}} must be constant. Indeed, let w=log(z)w=\log(z) as above, where zz is the standard coordinate on ()n(\mathbb{C}^{*})^{n}, so that w=ξ+iθw=\xi+i\theta. In these coordinates we can write

X1,0=ϕijfξiwj,\displaystyle X^{1,0}=\phi^{ij}\frac{\partial f}{\partial\xi^{i}}\frac{\partial}{\partial w_{j}},

where the coefficients ϕijfξi\phi^{ij}\frac{\partial f}{\partial\xi^{i}} depend only on the real part ξ\xi of ww. Since XX is holomorphic, it follows that

ξk(ϕijfξi)=2w¯k(ϕijfξi)=0.\displaystyle\frac{\partial}{\partial\xi^{k}}\left(\phi^{ij}\frac{\partial f}{\partial\xi^{i}}\right)=2\frac{\partial}{\partial\bar{w}_{k}}\left(\phi^{ij}\frac{\partial f}{\partial\xi^{i}}\right)=0.

In particular, it follows that JX𝔱JX\in\mathfrak{t}. We will denote the coefficients ϕijfξj=bXi\phi^{ij}\frac{\partial f}{\partial\xi^{j}}=b_{X}^{i}, so that JX=bXiθiJX=b_{X}^{i}\frac{\partial}{\partial\theta^{i}} is determined by the constant bX𝔱b_{X}\in\mathfrak{t}.

Lemma 2.5.

Suppose that ω\omega is a TnT^{n}-invariant Kähler metric on MM and that the TnT^{n}-action is Hamiltonian with respect to ω\omega, so that there exists a Kähler potential ϕ\phi for ω\omega on the dense orbit ()nM(\mathbb{C}^{*})^{n}\subset M. If YY is any real holomorphic vector field such that JY𝔱JY\in\mathfrak{t}, let θYC(M)\theta_{Y}\in C^{\infty}(M) be the Hamiltonian potential θY=μ(JY)\theta_{Y}=\mu(JY) corresponding to JYJY. Then θY\theta_{Y} also satisfies Yω=2i¯θY\mathcal{L}_{Y}\omega=2i\partial\bar{\partial}\theta_{Y}. Moreover, up to a constant, the restriction of θY\theta_{Y} to the dense orbit is given by θY(ξ,θ)=Y(ϕ)\theta_{Y}(\xi,\theta)=Y(\phi).

Proof.

By Cartan’s formula it suffices to show that

iYω=JiJYω=Jdμ(JY)=dcμ(JY),i_{Y}\omega=-Ji_{JY}\omega=-Jd\mu(JY)=d^{c}\mu(JY),

which proves the first statement. The second statement follows immediately from the fact that the restriction of ω\omega to the dense orbit is given by 2i¯ϕ2i\partial\bar{\partial}\phi. ∎

On the dense orbit then, the term Xω\mathcal{L}_{X}\omega in (1.4) is given by

Xω=2i¯X(ϕ).\mathcal{L}_{X}\omega=2i\partial\bar{\partial}X(\phi).

Hence, up to a constant, the soliton potential ff is given in real logarithmic coordinates on the dense orbit by

f=X(ϕ)=bXjϕξj.f=X(\phi)=b_{X}^{j}\frac{\partial\phi}{\partial\xi^{j}}. (2.4)

Since the Ricci form of ω\omega is given by

Ricω=i¯logdet(ϕij),\text{Ric}_{\omega}=-i\partial\bar{\partial}\log\det(\phi_{ij}),

we can succinctly rewrite (1.4) in terms of ϕ\phi alone.

Proposition 2.6.

Suppose that MM is a complex toric manifold and (ω,X)(\omega,X) is a shrinking gradient Kähler-Ricci soliton. If the TnT^{n}-action is Hamiltonian with respect to ω\omega, then ω\omega has a Kähler potential ϕ\phi on the dense orbit, which can be viewed via the identification 𝔱×Tn()n\mathfrak{t}\times T^{n}\cong(\mathbb{C}^{*})^{n} as a convex function on n\mathbb{R}^{n}. Then there exists a unique affine function a(ξ)a(\xi) on n\mathbb{R}^{n} such that ϕa=ϕa\phi_{a}=\phi-a satisfies the real Monge-Ampère equation

det(ϕa)ij=e2ϕa+bX,ϕa.\det(\phi_{a})_{ij}=e^{-2\phi_{a}+\langle b_{X},\nabla\phi_{a}\rangle}. (2.5)
Proof.

In light of the above discussion, the soliton equation (1.3)

ωRicω12Xω=0\omega-\text{Ric}_{\omega}-\frac{1}{2}\mathcal{L}_{X}\omega=0

can be rewritten as

0\displaystyle 0 =i¯(2ϕ+logdet(ϕij)X(ϕ))\displaystyle=i\partial\bar{\partial}\left(2\phi+\log\det(\phi_{ij})-X(\phi)\right)
=22ξiξj(2ϕ+logdet(ϕij)bX,ϕ)dξidθj,\displaystyle=2\frac{\partial^{2}}{\partial\xi^{i}\partial\xi^{j}}\left(2\phi+\log\det(\phi_{ij})-\langle b_{X},\nabla\phi\rangle\right)d\xi^{i}\wedge d\theta^{j},

and so the function 2ϕ+logdet(ϕij)bX,ϕ2\phi+\log\det(\phi_{ij})-\langle b_{X},\nabla\phi\rangle on n\mathbb{R}^{n} has vanishing Hessian, and is therefore equal to an affine function a(ξ)a(\xi). Define

ϕa(ξ)=ϕ(ξ)12a(ξ)\phi_{a}(\xi)=\phi(\xi)-\frac{1}{2}a(\xi)

and let cc be the constant c=12bX,ac=\frac{1}{2}\langle b_{X},\nabla a\rangle. Then it is clear that

2ϕa+logdet(ϕa,ij)bX,ϕa=c.2\phi_{a}+\log\det(\phi_{a,ij})-\langle b_{X},\nabla\phi_{a}\rangle=c.

Thus, by modifying aa by the addition of a constant, we have that ϕa\phi_{a} satisfies (2.5).

As we have seen, the metric gg depends only on the Hessian of ϕ\phi. Part of the content of Proposition 2.6 therefore is a normalization for the potential ϕ\phi, and we will make use of this later on.

2.4 Polyhedra and symplectic coordinates

Definition 7.

Let PP be a full-dimensional polyhedron in 𝔱\mathfrak{t}^{*}. Then PP is called Delzant if, for each vertex vPv\in P, there are exactly nn edges eie_{i} stemming from pp which can be written ei=v+λiεie_{i}=v+\lambda_{i}\varepsilon_{i} for λi\lambda_{i}\in\mathbb{R} and (εi)(\varepsilon_{i}) a \mathbb{Z}-basis of Γ\Gamma^{*}.

This says that each vertex of a Delzant polyhedron, when translated to the origin, can be made to look locally like standard +n\mathbb{R}^{n}_{+} via an element of GL(n,)\text{GL}(n,\mathbb{Z}). It follows from the definition that there is a well-defined normal fan Σ\Sigma associated to any Delzant polyhedron PP. One only needs to check that the relevant cones are rational polyhedral cones. This can be shown by induction, for example, since any face of a Delzant polyhedron must itself be Delzant. Therefore, given any Delzant polyhedron PP, there is an associated toric variety P=Σ\mathcal{M}_{P}=\mathcal{M}_{\Sigma}. The condition on the vertices of PP is precisely what is required to ensure that P\mathcal{M}_{P} is smooth; see [15, Theorem 3.1.19] and the preceding statements there.

In Section 2.1, we encountered a purely algebraic construction which produced a toric variety, and therefore a complex toric manifold, P\mathcal{M}_{P} from the data of a Delzant polyhedron. We now introduce a different construction, this time coming from symplectic geometry, which will produce a symplectic toric manifold from the data of PP. The idea is to construct a complex symplectic manifold (MP,ωP,JP)(M_{P},\omega_{P},J_{P}) as a Kähler quotient of N\mathbb{C}^{N} by a subgroup GG_{\mathbb{C}} of the standard torus ()N(\mathbb{C}^{*})^{N}. The next proposition is standard for compact symplectic toric manifolds, and in the more general setting of potentially singular and non-compact varieties it is essentially proved in [10, Lemma 2.1], and earlier in [7, Chapter VI, Proposition 3.1.1]. We could not find the precise statement that we use in the literature, and so we briefly outline the proof below.

Proposition 2.7.

Let PP be a Delzant polyhedron in 𝔱\mathfrak{t} with NN facets. Then there exists a Kähler manifold (MP,ωP,JP)(M_{P},\omega_{P},J_{P}) with an effective JPJ_{P}-holomorphic ()n(\mathbb{C}^{*})^{n}-action on MPM_{P} associated to PP, obtained as a Kähler quotient of N\mathbb{C}^{N} by a complex subgroup G()NG\subset(\mathbb{C}^{*})^{N} acting in the usual way. The TnT^{n}-action is Hamiltonian with respect to ωP\omega_{P}, and the moment map μP:MP𝔱\mu_{P}:M_{P}\to\mathfrak{t}^{*} has image P¯\overline{P}. If PP is rational, then ωP\omega_{P} is the curvature form of a hermitian metric on an equivariant line bundle LPMPL_{P}\to M_{P} determined by PP.

Proof.

This is a direct consequence of [10, Lemma 2.1]. In particular there is a complex subgroup G()NG\subset(\mathbb{C}^{*})^{N}, a corresponding maximal compact subgroup KGK\subset G, and a moment map μK\mu_{K} for the KK-action on N\mathbb{C}^{N}. Then MPM_{P} is defined as the symplectic quotient Z/KZ/K, where ZNZ\subset\mathbb{C}^{N} is the preimage of a particular regular value of μK\mu_{K}. Denote the quotient map by π:ZMP\pi:Z\to M_{P}. The symplectic form ωP\omega_{P} is induced by the symplectic quotient by restricting the standard Euclidean symplectic form ωE\omega_{E} to ZZ. The complex structure JPJ_{P} on MPM_{P} is determined via the usual Kähler quotient construction. In particular, there is a closed analytic subset VV in N\mathbb{C}^{N} where GG acts freely, and we can equivalently define MP=(NV)/GM_{P}=(\mathbb{C}^{N}-V)/G.

Now, if the vertices of PP lie on the integer lattice, then the group GG is algebraic and the construction of MPM_{P} in [10] becomes a GIT quotient (see for example [15, Chapter 14] for details on this point). In particular, PP determines a character χP:G\chi_{P}:G\to\mathbb{C}^{*} which gives rise to an action of GG on the trivial line bundle 𝒪N\mathcal{O}\to\mathbb{C}^{N}, and the quotient of 𝒪\mathcal{O} by GG is a well-defined line bundle LPL_{P} on MPM_{P} [15, Theorem 14.2.13]. The fact that ωP2πc1(LP)\omega_{P}\in 2\pi c_{1}(L_{P}) follows directly from the explicit Guillemin formula [10, Theorem 5.1] for ωP\omega_{P}, [15, Theorem 14.2.13], and the following proposition, which we state separately below for emphasis. ∎

In particular, given the data of a rational polyhedron PP, we have two constructions, each associating to PP a toric geometric object in the appropriate category. These turn out, after making the relevant identifications, to be equivalent. Let PP be a rational polyhedron and P\mathcal{M}_{P} be the toric variety constructed in Section 2.1.

Proposition 2.8.

The complex manifold (MP,JP)(M_{P},J_{P}) is equivariantly biholomorphic to P\mathcal{M}_{P}.

We omit the proof here, but this is essentially proven in [10, Lemma 2.1] (c.f. [7, Chapter VI, Proposition 3.2.1]). From the description of MPM_{P} given there, one simply applies the main theorem in [14] to deduce the proposition.

In sum, given the data of a rational polyhedron PP, we have two constructions, each associating to PP a toric geometric object in the appropriate category, and these constructions are compatible up to an appropriate identification. For the remainder of this section we work with a given Delzant polyhedron PP and denote M=PMPM=\mathcal{M}_{P}\cong M_{P}. In particular, we have a canonical Kähler metric ωP\omega_{P} on MM. The induced TnT^{n}-action is Hamiltonian by construction, so that by Proposition 2.4 there is a Kähler potential ωP=2i¯ϕP\omega_{P}=2i\partial\bar{\partial}\phi_{P} on the dense orbit.

We move on to consider an arbitrary Kähler metric ω\omega on MM with respect to which the TnT^{n}-action is Hamiltonian, not necessarily equal to ωP\omega_{P}. We impose the additional assumption that the corresponding moment map μ\mu also has image equal to P¯\overline{P}. Recall from Proposition 2.4 that there then exists a potential ϕ\phi on the dense orbit ()nM(\mathbb{C}^{*})^{n}\subset M. We introduce logarithmic coordinates (ξj,θj)(\xi^{j},\theta^{j}) as in the previous section so that the moment map μ\mu is determined by the diffeomorphism ϕ:𝔱P\nabla\phi:\mathfrak{t}\to P. We can then use the moment map to introduce a change of coordinates ϕ=x\nabla\phi=x, and thereby view ()nP×Tn(\mathbb{C}^{*})^{n}\cong P\times T^{n}. In these coordinates the Kähler form ω\omega is standard, i.e.

ω=dxjdθj.\omega=dx^{j}\wedge d\theta^{j}.

So the moment map μ=ϕ\mu=\nabla\phi induces a natural choice of Darboux coordinates, and for this reason (xj,θj)(x^{j},\theta^{j}) are typically referred to as symplectic coordinates on MM. This is only a real coordinate system, and hence the coefficients of the Kähler form do not determine those of the corresponding Riemannian metric. One can still determine the metric gg by introducing a smooth function uu on PP which is related to ϕ\phi by the Legendre transform:

ϕ(ξ)+u(x)=ξ,x.\phi(\xi)+u(x)=\langle\xi,x\rangle. (2.6)

Then the metric gg is given by

g=uij(x)dxidxj+uij(x)dθidθj.g=u_{ij}(x)dx^{i}dx^{j}+u^{ij}(x)d\theta_{i}d\theta_{j}. (2.7)

Thus the metric structure is determined by the Hessian of the function uu, and so by analogy with the complex case this function is sometimes called the symplectic potential for gg. Although we will not use this here, it is worth noting that it is more natural to view the function uu as determining the complex structure JJ, from which the formula (2.7) for the metric is a consequence. The Legendre transform will be used heavily in the remainder of the paper, and so for convenience we collect some basic properties here. For references focusing on aspects most closely related to the situation here; see for example [8, 20, 30].

Lemma 2.9.

Let VV be a real vector space and ϕ\phi be a smooth and strictly convex function on a convex domain ΩV\Omega^{\prime}\subset V. Then there is a unique function L(ϕ)=uL(\phi)=u defined on Ω=ϕ(Ω)V\Omega=\nabla\phi(\Omega^{\prime})\subset V^{*} by (2.6):

ϕ(ξ)+u(x)=ξ,x\phi(\xi)+u(x)=\langle\xi,x\rangle

for x=ϕ(ξ)x=\nabla\phi(\xi). The function uu is smooth and strictly convex on Ω\Omega. Moreover, LL has the following properties:

  1. 1.

    L(L(ϕ))=ϕL(L(\phi))=\phi,

  2. 2.

    ϕ:ΩΩ\nabla\phi:\Omega^{\prime}\to\Omega and u:ΩΩ\nabla u:\Omega\to\Omega^{\prime} are inverse to each other,

  3. 3.

    ϕij(u(x))=uij(x)\phi_{ij}(\nabla u(x))=u^{ij}(x),

  4. 4.

    L((1t)ϕ+tϕ)(1t)L(ϕ)+tL(ϕ)L((1-t)\phi+t\phi^{\prime})\leq(1-t)L(\phi)+tL(\phi^{\prime}),

  5. 5.

    L(ϕ)(x)=supξΩ{x,ξϕ(ξ)}L(\phi)(x)=\sup_{\xi\in\Omega^{\prime}}\{\langle x,\xi\rangle-\phi(\xi)\}.

The third item can be understood to mean that the Euclidean Hessians 2ϕ\nabla^{2}\phi and 2u\nabla^{2}u are inverse to each other, under the appropriate change of coordinates. In most situations, we will use the shorthand ϕu=L(u)\phi_{u}=L(u). One application that will be used throughout the paper is the following. The last item is often taken as the definition of the Legendre transform (the so called Legendre-Fenchel transform), as it can be used to define L(ϕ)L(\phi) for ϕ\phi merely continuous. Henceforth we will take (v)(v) as the definition of L(ϕ)L(\phi) in any case where ϕ\phi is not necessarily C1C^{1}.

Lemma 2.10 (c.f. [8, Lemma 2.6]).

Let ϕ\phi be any strictly convex function on an open convex domain Ωn\Omega^{\prime}\subset\mathbb{R}^{n}. Let uu be its Legendre transform defined on Ω\Omega. If 0Ω0\in\Omega, then there exists a C>0C>0 such that

ϕ(ξ)C1|ξ|C.\phi(\xi)\geq C^{-1}|\xi|-C. (2.8)

In particular, ϕ\phi is proper. Moreover, we can estimate CC the following way. Let ε>0\varepsilon>0 be sufficiently small so that Bε(0)ΩB_{\varepsilon}(0)\subset\Omega, then

ϕ(ξ)ε|ξ|supBε(0)L(ϕ).\phi(\xi)\geq\varepsilon|\xi|-\sup_{B_{\varepsilon}(0)}L(\phi). (2.9)

The estimate (2.9) is an immediate corollary of Lemma 2.9. For smooth functions one can see (2.8) directly; since 0 is in the domain of uu, there is some ξ\xi such that ϕu(ξ)=0\nabla\phi_{u}(\xi)=0. Then ϕu\phi_{u} is a strictly convex function with a minimum, and hence must grow at least linearly. However in what follows we will need to make use of (2.9) even for smooth functions. Although the notation is suggestive of the situation where ϕ\phi is the Kähler potential of a toric metric, it is worth noting, and will be used later on, that this is completely symmetric in ϕ\phi and uu. That is to say, if 0 lies in the domain Ω\Omega^{\prime} of ϕ\phi, it follows that uu must also satisfy (2.8) (with respect to the coordinate xx in Ω\Omega). Indeed the entirety of Lemma 2.10 is entirely symmetric in uu and ϕ\phi.

We collect some further elementary properties of the behavior of convex functions under the Legendre transform, all consequences of the properties laid out in Lemma 2.9. As we will see, these in turn give rise to interesting geometric consequences when interpreted in the context of Kähler geometry on complex toric manifolds.

Lemma 2.11.

Let ϕ\phi be a strictly convex function on 𝔱\mathfrak{t} and u=L(ϕ)u=L(\phi) be its Legendre transform. Let Ω\Omega denote the image of the gradient ϕ:𝔱𝔱\nabla\phi:\mathfrak{t}\to\mathfrak{t}^{*}.

  1. 1.

    For BGL(n,)B\in\text{GL}(n,\mathbb{Z}), set ϕB(ξ)=ϕ(Bξ)\phi_{B}(\xi)=\phi\left(B\xi\right). Then L(ϕB)(x)=u((BT)1x)L(\phi_{B})(x)=u((B^{T})^{-1}x), and the image of ϕB:𝔱𝔱\nabla\phi_{B}:\mathfrak{t}\to\mathfrak{t}^{*} is equal to BT(Ω)B^{T}(\Omega).

  2. 2.

    For b1𝔱b_{1}\in\mathfrak{t}, set ϕb1(ξ)=ϕ(ξb1)\phi_{b_{1}}(\xi)=\phi(\xi-b_{1}). Then L(ϕb1)(x)=u(x)+b1,xL(\phi_{b_{1}})(x)=u(x)+\langle b_{1},x\rangle. Clearly, the image of ϕb1\nabla\phi_{b_{1}} is also equal to Ω\Omega.

  3. 3.

    Symmetrically, for b2𝔱b_{2}\in\mathfrak{t}^{*}, set ϕb2(ξ)=ϕ(ξ)+b2,ξ\phi^{b_{2}}(\xi)=\phi(\xi)+\langle b_{2},\xi\rangle. Then L(ϕb2)(x)=u(xb2)L(\phi^{b_{2}})(x)=u(x-b_{2}) and the image of ϕb2\nabla\phi^{b_{2}} is equal to Ωb2\Omega-b_{2}.

Let MM be a complex toric manifold together with a Kähler metric ω\omega with respect to which the real TnT^{n} action is Hamiltonian, and let ϕ\phi be a strictly convex function on the dense orbit ()nM(\mathbb{C}^{*})^{n}\subset M such that ω=2i¯ϕ\omega=2i\partial\bar{\partial}\phi. Let μ:M𝔱\mu:M\to\mathfrak{t}^{*} denote the corresponding moment map, normalized so that μ,b=ϕ,b\langle\mu,b\rangle=\langle\nabla\phi,b\rangle on the dense orbit as in Section 2.2, and suppose that the image of μ\mu is equal to a Delzant polyhedron PP. Recall also from Section 2.2 that we fix a basis X1,,XnX_{1},\dots,X_{n} for 𝔱\mathfrak{t}. Then the action of GL(n,)\text{GL}(n,\mathbb{Z}) on ϕ\phi corresponds simply to changing this basis by an automorphism of ()n(\mathbb{C}^{*})^{n}. This will be useful to simplify calculations later on, since by the Delzant condition we can use this to assume that PP locally coincides with a translate of the positive orthant +n\mathbb{R}^{n}_{+} near any vertex. The action of 𝔱\mathfrak{t} on ϕ\phi given in (ii)(ii) corresponds to composing the ()n(\mathbb{C}^{*})^{n}-action on MM with an element of the form eb1()ne^{-b_{1}}\in(\mathbb{C}^{*})^{n}. Notice that this is is always induced from the global automorphism eb1:MMe^{-b_{1}}:M\to M of MM. The 𝔱\mathfrak{t}^{*}-action of (iii)(iii) is most naturally viewed as a modification of the moment map μ\mu by the action of 𝔱\mathfrak{t}^{*} on itself by translation.

Recall from Proposition 2.6 that we are interested in the case where ϕ\phi is a solution to (2.5) on 𝔱\mathfrak{t}. Since ϕ\phi uniquely determines and is uniquely determined by its Legendre transform u=L(ϕ)u=L(\phi), we can once again make use of the properties laid out in Lemma 2.9 to rewrite (2.5) as a real Monge-Ampère equation for the convex function uu, defined on the interior of the image of the moment map. We assume as above that this image is equal to a Delzant polyhedron PP.

Proposition 2.12.

Suppose that MM is a complex toric manifold and (ω,X)(\omega,X) is a shrinking gradient Kähler-Ricci soliton on MM. Suppose that the TnT^{n}-action is Hamiltonian with respect to ω\omega, so that, by Proposition 2.4, ω\omega admits Kähler potential determined by a strictly convex function ϕ\phi on 𝔱\mathfrak{t} satisfying (2.5). Let u=L(ϕ)u=L(\phi) be the Legendre transform, which we assume is defined on the Delzant polyhedron PP. Then uu satisfies the real Monge-Ampère equation

2(uixiu(x))logdet(uij)=bX,x.2\left(u_{i}x^{i}-u(x)\right)-\log\det(u_{ij})=\langle b_{X},x\rangle. (2.10)

We return now to the canonical metric ωP\omega_{P} defined in Proposition 2.7. We have just seen that there is a corresponding symplectic potential uPu_{P} on PP. The main result of [30] is an explicit formula for uPu_{P}, only in terms of the data of PP, in the case that P¯\overline{P} (and therefore MM) is compact. This has been generalized in [10, Theorem 5.2] to (essentially) arbitrary polyhedra, the Delzant case included. Let FiF_{i}, i=1,,di=1,\dots,d denote the (n1)(n-1)-dimensional facets of PP with inward-pointing normal vector νiΓ\nu_{i}\in\Gamma, normalized so that νi\nu_{i} is the minimal generator of σi=+νi\sigma_{i}=\mathbb{R}_{+}\cdot\nu_{i} in Γ\Gamma. Let i(x)=νi,x\ell_{i}(x)=\langle\nu_{i},x\rangle, so that P¯\overline{P} is defined by the system of inequalities i(x)ai\ell_{i}(x)\geq-a_{i}, i=1,,Ni=1,\dots,N, aia_{i}\in\mathbb{R}. Then from [10] we have the following explicit formula for uPu_{P}:

uP(x)=12i=1d(i(x)+ai)log(i(x)+ai).u_{P}(x)=\frac{1}{2}\sum_{i=1}^{d}(\ell_{i}(x)+a_{i})\log\left(\ell_{i}(x)+a_{i}\right). (2.11)

2.5 Equivalences

Thus far, we have shown that associated to any Delzant polyhedron PP there is a toric Kähler manifold (MP,JP,ωP)(M_{P},J_{P},\omega_{P}). We begin this subsection by giving conditions under which we can extend the Delzant classification to the non-compact setting. In brief, we would like to understand the answers to the following questions. First, given a toric Kähler manifold (M,J,ω)(M,J,\omega), under what conditions is the image of the moment map equal to a Delzant polyhedron PP? Second, given a toric Kähler manifold (M,J,ω)(M,J,\omega) whose moment image is equal to a Delzant polyhedron PP, under what conditions can we say that (M,J)(MP,JP)(M,J)\cong(M_{P},J_{P}) and (M,ω)(MP,ωP)(M,\omega)\cong(M_{P},\omega_{P})?

To a large extent these questions have already been studied, and much of what appears below is simply a collection of existing results, rephrased in order to better suit the current setup. The answer to the first question and part of the second comes from the work of [35, 42].

Lemma 2.13.

Let (M,ω)(M,\omega) be any symplectic toric manifold with finite fixed point set. Suppose that there exists b𝔱b\in\mathfrak{t} such that the function μ,b:M\langle\mu,b\rangle:M\to\mathbb{R} is proper and bounded from below. Then the image of the moment map μ\mu is a Delzant polyhedron PP, and moreover (M,ω)(M,\omega) is equivariantly symplectomorphic to (MP,ωP)(M_{P},\omega_{P}).

Proof.

Since the fixed point set of the TnT^{n}-action is finite, it follows from [32, Theorem 4.1] (c.f. [42, Proposition 1.4] and the preceeding remarks) that the existence of such a b𝔱b\in\mathfrak{t} is sufficient to show that the image of the moment map μ\mu is a polyhedral set in 𝔱\mathfrak{t}^{*}. This means by definition that μ(M)\mu(M) is equal to the intersection of finitely many half spaces. It then follows immediately from [35, Proposition 1.1] that PP is a Delzant (unimodular) polyhedron. Finally, [35, Theorem 1.3, c.f. Theorem 6.7] furnishes the desired equivariant symplectomorphism. ∎

Given a general symplectic toric manifold (M,ω)(M,\omega) satisfying the conditions of Lemma 2.13, let PP be the corresponding polyhedron in 𝔱\mathfrak{t}^{*}. Suppose that there is a compatible complex structure JJ such that TnT^{n} acts holomorphically. When MM is compact, JJ is determined up to biholomorphism by PP. This follows in part since we can always use JJ to complexify the TnT^{n}-action to an action of the full ()n(\mathbb{C}^{*})^{n}. In general, the issue is more subtle. The following example illustrates the problem.111We thank Vestislav Apostolov for providing this example.

Example 1.

Let (𝔻,ω)(\mathbb{D},\omega) denote the Poincaré model of the hyperbolic metric on the unit disc in \mathbb{C}. The standard S1S^{1}-action on \mathbb{C} restricts to an action on 𝔻\mathbb{D}, but clearly this does not admit a complexified action of \mathbb{C}^{*} on 𝔻\mathbb{D}. The symplectic form ω\omega is S1S^{1}-invariant and, with an appropriate normalization, the moment map μ:𝔻\mu:\mathbb{D}\to\mathbb{R} has image equal to the unbounded closed interval P=[0,)P=[0,\infty). Thus, the image is the Delzant polyhedron PP, but 𝔻≇MP\mathbb{D}\not\cong M_{P}\cong\mathbb{C}.

However, if we assume a priori that there exists a complexified action, then it does indeed follow that the complex structure must be biholomorphic to the standard one JPJ_{P} on MPM_{P}. Let (M,J)(M,J) be a complex toric manifold, so that there exists an effective holomorphic ()n(\mathbb{C}^{*})^{n}-action. Suppose that ω\omega is the Kähler form of a compatible Kähler metric such that the TnT^{n}-action is Hamiltonian.

Lemma 2.14 (c.f. [3, Proposition A.1]).

Let (M,J,ω)(M,J,\omega) be as above, and assume that the image of the moment map is equal to a Delzant polyhedron PP. Then MM is equivariantly biholomorphic to MPM_{P}. In particular, (M,J)(M,J) is quasiprojective.

Proof.

As usual, choose a point pp in the interior of the dense orbit ()nM(\mathbb{C}^{*})^{n}\subset M, and further choose points xix_{i} in the interior of each kk-dimensional face FiF_{i} of PP. By [32, Theorem 4.1, part (v)] (c.f. [31, 41]), each point qμ1(Fi)q\in\mu^{-1}(F_{i}) is stabilized by a common torus TFinkTnT^{n-k}_{F_{i}}\subset T^{n} with Lie algebra 𝔱i\mathfrak{t}_{i}, and moreover FiF_{i} lies as an open subset of the dual kk-plane 𝔱Fi𝔱\mathfrak{t}_{F_{i}}^{\perp}\subset\mathfrak{t}^{*}. By the holomorphic slice theorem [44, Theorem 1.24], there exists a ()n(\mathbb{C}^{*})^{n}-invariant open neighborhood UiMU_{i}\subset M of the orbit ()npiM(\mathbb{C}^{*})^{n}\cdot p_{i}\subset M and an equivariant biholomorphism Φi:Ui()k×nk\Phi_{i}:U_{i}\to(\mathbb{C}^{*})^{k}\times\mathbb{C}^{n-k}, with the standard ()n(\mathbb{C}^{*})^{n}-action such that Φi(xi)=(1,0)\Phi_{i}(x_{i})=(1,0) and Φi(μ1(Fi)Ui)=()k×{0}\Phi_{i}(\mu^{-1}(F_{i})\cap U_{i})=(\mathbb{C}^{*})^{k}\times\{0\}. We see that the stabilizer TFinkT^{n-k}_{F_{i}} acts in the coordinates induced by Φi\Phi_{i} by the standard action on nk\mathbb{C}^{n-k}. Note that the equivariance of Φi\Phi_{i} ensures that entire dense orbit lies in UiU_{i}, and hence we can modify the map Φi\Phi_{i} by the ()n(\mathbb{C}^{*})^{n}-action to ensure that Φi(p)=(1,,1)\Phi_{i}(p)=(1,\dots,1). In this way, we produce an equivariant holomorphic coordinate covering of MM by running through each FiF_{i}. Suppose now that F1,F2F_{1},F_{2} are two kk-dimensional faces that which lie on the boundary of a higher-dimensional face EE of PP, and let ΦF1:UF1()k×nk,ΦF2:UF2()k×nk,ΦE:UE()l×nl\Phi_{F_{1}}:U_{F_{1}}\to(\mathbb{C}^{*})^{k}\times\mathbb{C}^{n-k},\Phi_{F_{2}}:U_{F_{2}}\to(\mathbb{C}^{*})^{k}\times\mathbb{C}^{n-k},\Phi_{E}:U_{E}\to(\mathbb{C}^{*})^{l}\times\mathbb{C}^{n-l} denote the corresponding maps as above. By equivariance, the transition map ΦF2ΦF11\Phi_{F_{2}}\circ\Phi_{F_{1}}^{-1} is uniquely determined by the inclusions of ()l×nl()k×nk(\mathbb{C}^{*})^{l}\times\mathbb{C}^{n-l}\subset(\mathbb{C}^{*})^{k}\times\mathbb{C}^{n-k} given by ΦE\Phi_{E}, as EE varies across all faces containing F1F_{1} and F2F_{2}. These in turn are determined uniquely by the inclusions of the stabilizer algebra 𝔱E𝔱F1,𝔱F2\mathfrak{t}_{E}\subset\mathfrak{t}_{F_{1}},\mathfrak{t}_{F_{2}}. As we have seen, the stabilizer algebras 𝔱E,𝔱F1,𝔱F2\mathfrak{t}_{E},\mathfrak{t}_{F_{1}},\mathfrak{t}_{F_{2}} comprise the normal directions to the faces E,F1,F2E,F_{1},F_{2} in 𝔱\mathfrak{t}^{*}, respectively. In particular, the transition data of this covering is determined uniquely by the normal fan ΣP\Sigma_{P} of PP. Now let (Wi,Ψi)(W_{i},\Psi_{i}) be a cover of MPM_{P} constructed in the same way. For each face FiF_{i} of PP, we have maps Ψi1Φi:UiWi\Psi_{i}^{-1}\circ\Phi_{i}:U_{i}\to W_{i}. Since the transition data for each covering is uniquely determined by ΣP\Sigma_{P}, we see that these local maps patch together to form a well-defined biholomorphism MMPM\to M_{P}. ∎

We have thus far met several inequivalent definitions of what it means for a non-compact manifold to be “toric.” To avoid confusion, we introduce the following definition, which lies at the intersection of all of the previously introduced notions.

Definition 8.

We say that (M,J,ω)(M,J,\omega), together with a given ()n(\mathbb{C}^{*})^{n}-action is algebraic-Kähler toric (AK-toric) if the following conditions are met:

  1. 1.

    The ()n(\mathbb{C}^{*})^{n}-action is effective and holomorphic with respect to JJ.

  2. 2.

    The symplectic form ω\omega is the Kähler form of a compatible, TnT^{n}-invariant Kähler metric on MM.

  3. 3.

    The TnT^{n}-action is Hamiltonian with respect to ω\omega, and the moment map μ:M𝔱\mu:M\to\mathfrak{t}^{*} has image equal to a Delzant polyhedron PP.

Such an MM is always equivariantly biholomorphic to the algebraic toric variety P\mathcal{M}_{P} by Lemma 2.14 and Proposition 2.8. When (M,ω)(M,\omega) is a compact toric manifold, the polytope PP is determined up to translation in 𝔱\mathfrak{t}^{*} by the cohomology class [ω][\omega] [1, 3, 30]. We show this is true in the case that there is an action of the full ()n(\mathbb{C}^{*})^{n}.

Proposition 2.15.

If (M,J,ω)(M,J,\omega) be AK-toric, then the moment polyhedron PP is determined up to translation by the cohomology class [ω][\omega].

Proof.

The polyhedron PP determines a torus-invariant divisor DωD_{\omega} on (M,J)(M,J) as follows. Since (M,J)(M,J) is biholomorphic to (MP,JP)(M_{P},J_{P}), we use this biholomorphism and assume without loss of generality that (M,J,ω)=(MP,JP,ω)(M,J,\omega)=(M_{P},J_{P},\omega) with ω\omega not necessarily equal to ωP\omega_{P}. Recall that (MP,JP)(M_{P},J_{P}) naturally carries the structure of the algebraic toric variety P\mathcal{M}_{P}. Thus, we can identify the normal fan Σ\Sigma of PP with the fan corresponding to P\mathcal{M}_{P}. Let νi\nu_{i} be the minimal generator in Γ\Gamma of the ray σiΣ\sigma_{i}\in\Sigma corresponding to the direction normal to each facet FiF_{i} of PP. Then each FiF_{i} of PP has the local defining equation i(x)+ai=0\ell_{i}(x)+a_{i}=0, where i(x)=νi,x\ell_{i}(x)=\langle\nu_{i},x\rangle for some aia_{i}\in\mathbb{R}. Recall that σi\sigma_{i} defines via the Orbit-Cone correspondence an irreducible Weil divisor DiD_{i}. The divisor DωD_{\omega} is then given by

Dω=aiDi.D_{\omega}=\sum a_{i}D_{i}.

We can assume without loss of generality that the irreducible component D1D_{1} of DωD_{\omega} is compact. If there is no such D1D_{1}, then it follows that there is a bnb\in\mathbb{R}^{n} and AGL(n,)A\in\text{GL}(n,\mathbb{Z}) such that the affine transformation Ax+bAx+b takes PP to the positive orthant +n\mathbb{R}^{n}_{+}, and so MnM\cong\mathbb{C}^{n}. Note that the entire construction behaves well with respect to restriction, so that D1=F1D_{1}=\mathcal{M}_{F_{1}}. Since PP is Delzant, so is F1F_{1}, and so it follows that D1D_{1} is a nonsingular projective variety. If we restrict ω\omega to D1D_{1}, we obtain a moment map for the Tn1T^{n-1}-action μ1:D1𝔱1\mu_{1}:D_{1}\to\mathfrak{t}_{1}, where 𝔱1𝔱\mathfrak{t}_{1}\subset\mathfrak{t} is the orthogonal complement of the stabilizer algebra of D1D_{1}. Then the image of μ1\mu_{1} is the face F1F_{1} of PP corresponding to D1D_{1}. After potentially acting by an element of GL(n,)(n,\mathbb{Z}), we can assume that ν1,x=x1\langle\nu_{1},x\rangle=x_{1}, so that 𝔱1\mathfrak{t}_{1} can be identified with the subspace x1=0x_{1}=0. Inside of 𝔱1\mathfrak{t}_{1}, F1F_{1} is then defined by ηi,(x2,,xn)αi\langle\eta_{i},(x_{2},\dots,x_{n})\rangle\geq-\alpha_{i} for some ηi\eta_{i} in the lattice and αi\alpha_{i}\in\mathbb{R}. Thus, the Delzant polytope F1F_{1} determines a divisor Δ=αiΔi\Delta=\sum\alpha_{i}\Delta_{i} on D1D_{1}, where Δi\Delta_{i} are the torus-invariant divisors on D1D_{1} corresponding to ηi\eta_{i} through the Orbit-Cone correspondence.

Since (D1,ω|D1)(D_{1},\omega|_{D_{1}}) is itself a compact symplectic toric manifold, we can now appeal to the well-established theory in the compact setting [6, 31, 16, 30]. Specifically, we have that the cohomology class of the symplectic form ω|D1\omega|_{D_{1}} is given by [16, 30]

[ω|D1]=αi[Δi].[\omega|_{D_{1}}]=\sum\alpha_{i}[\Delta_{i}].

The coefficients αi\alpha_{i}, by definition, fix the defining equations of F1F_{1} inside 𝔱1\mathfrak{t}_{1}. Thus, we see that the facet F1F_{1} is uniquely determined by [ω][\omega] up to translation in 𝔱1\mathfrak{t}_{1}. By the Orbit-Cone correspondence, the subspace 𝔱1\mathfrak{t}_{1} on which F1F_{1} lies is uniquely determined by the fixed fan Σ\Sigma, up to translation in its normal direction. We see then that the set of vertices {v1,,vk}\{v_{1},\dots,v_{k}\} of F1F_{1}, which is the image under μ\mu of the set of fixed points TnT^{n}-action that lie in μ1(F1)\mu^{-1}(F_{1}), is determined uniquely up to a translation in 𝔱\mathfrak{t}^{*} by [ω][\omega]. Now each vertex of PP lies on at least one compact facet, again unless MnM\cong\mathbb{C}^{n} and P=+nP=\mathbb{R}^{n}_{+}. Hence, we can repeat this process for each compact torus-invariant divisor to see that the set of all vertices {v1,,vK}\{v_{1},\dots,v_{K}\} of PP is determined up to translation in 𝔱\mathfrak{t}^{*} by [ω][\omega]. It is clear then that the same is true of PP. ∎

Corollary 2.16.

Let MM be AK-toric with polyhedron P={x𝔱|νi,xai for all i=1,,N}P=\{x\in\mathfrak{t}^{*}\>|\>\langle\nu_{i},x\rangle\geq-a_{i}\text{ for all }i=1,\dots,N\}, and suppose that ω\omega is the curvature of an equivariant hermitian holomorphic line bundle (L,h)(L,h). Then L𝒪(Dω)L\cong\mathcal{O}(D_{\omega}) is the line bundle associated to the divisor Dω=aiDiD_{\omega}=\sum a_{i}D_{i}.

Proof.

Recall that an AK-toric manifold with polyhedron PP is biholomorphic to the toric variety P\mathcal{M}_{P}. Let Σ\Sigma be the normal fan of PP so that P=Σ\mathcal{M}_{P}=\mathcal{M}_{\Sigma}. Since MM is smooth we have by [15, Proposition 4.2.6] that L𝒪(D)L\cong\mathcal{O}(D) for some torus-invariant divisor D=βiDiD=\sum\beta_{i}D_{i} with βi\beta_{i}\in\mathbb{Z}. We let PDP_{D} denote the polyhedron associated to DD given by (2.2), i.e.

PD={x𝔱|x,νiβi, for all i=1,,N},P_{D}=\left\{x\in\mathfrak{t}^{*}\>|\>\langle x,\nu_{i}\rangle\geq-\beta_{i},\text{ for all }i=1,\dots,N\right\},

where ν1,,νN\nu_{1},\dots,\nu_{N} are the minimal generators of the rays σiΣ\sigma_{i}\in\Sigma. If DD and DD^{\prime} are any two torus-invariant divisors on MM with integer coefficients, we define an equivalence relation by declaring that DDD\sim D^{\prime} if and only if there exists some νΓ\nu\in\Gamma^{*} such that PD=PD+νP_{D^{\prime}}=P_{D}+\nu, where PDP_{D} and PDP_{D^{\prime}} are the polyhedra defined in (2.2). By [15, Theorem 4.1.3], DDD\sim D^{\prime} if and only if 𝒪(D)𝒪(D)\mathcal{O}(D)\cong\mathcal{O}(D^{\prime}). Suppose that D1D_{1} is a compact torus-invariant Weil divisor in MM. As before, such a D1D_{1} must exist unless MnM\cong\mathbb{C}^{n} and P=+nP=\mathbb{R}^{n}_{+}. Perhaps by modifying DD by the equivalence relation, we can assume that the coefficient β1\beta_{1} corresponding to D1D_{1} is zero. In other words, there is a section s1s_{1} of LL which does not vanish identically on D1D_{1}. Let F1P¯F_{1}\subset\overline{P} be the facet corresponding to D1D_{1}. As before, the Delzant polyhedron F1F_{1} determines a unique torus-invariant Weil divisor Δ=αiΔi\Delta=\sum\alpha_{i}\Delta_{i} on D1D_{1}. The restriction of s1s_{1} to D1D_{1} is a section of L|D1L|_{D_{1}} which vanishes along Δi=D1Di\Delta_{i}=D_{1}\cap D_{i} to order αi\alpha_{i}. In particular, we see that the coefficients αi\alpha_{i} of Δi\Delta_{i} are equal to those βi\beta_{i} such that DiD1D_{i}\cap D_{1}\neq\emptyset. Recall that Dω=aiDiD_{\omega}=\sum a_{i}D_{i}. We claim that DωDD_{\omega}\sim D. As before, we can act by GL(n,)(n,\mathbb{Z}) so that ν1,x=x1\langle\nu_{1},x\rangle=x_{1}. Write P1=Pω+ν1P_{1}=P_{\omega}+\nu_{1} so that the face F1+ν1F_{1}+\nu_{1} corresponding to D1D_{1} now lies on the hyperplane x1=0x_{1}=0, and in general P1P_{1} is defined by x,νiν1,νiai=a~i\langle x,\nu_{i}\rangle\geq\langle\nu_{1},\nu_{i}\rangle-a_{i}=-\tilde{a}_{i}. Then it is straightforward to compute that the coefficients αi\alpha_{i} are equal to those a~i\tilde{a}_{i} such that DiD1D_{i}\cap D_{1}\neq\emptyset. Running across all compact divisors of MM, we see that the coefficients aia_{i} in the defining equations for PωP_{\omega} are uniquely determined by βi\beta_{i} up to equivalence. In particular, DωDD_{\omega}\sim D. ∎

Proposition 2.17.

Let (M,J,ω)(M,J,\omega) be AK-toric with moment polyhedron PP. Then ω\omega admits a strictly convex symplectic potential uu on PP, unique up to the addition of an affine function on PP. Moreover, the function uu takes a special form. Recall that MMPM\cong M_{P}, so that PP in particular determines a Kähler form ωP\omega_{P} on MM with symplectic potential uPu_{P} defined by (2.11). Then there exists a function vC(P¯)v\in C^{\infty}(\overline{P}) such that

u=uP+v.u=u_{P}+v. (2.12)
Proof.

By Proposition 2.4, the restriction of ω\omega to the dense orbit ()nM(\mathbb{C}^{*})^{n}\subset M is determined by a strictly convex function ϕ\phi on 𝔱\mathfrak{t}. The moment map μ:M𝔱\mu:M\to\mathfrak{t}^{*} is then determined by the Euclidean gradient ϕ\nabla\phi on 𝔱\mathfrak{t}. Thus, there is a symplectic potential u=L(ϕ)u=L(\phi) defined by the Legendre transform (2.6). That uu satisfies the boundary condition (2.12) follows from [5, Proposition 1]. Indeed, the key point is that the Hessian (uij)(u_{ij}) of uu determines a natural complex structure JuJ_{u} on the dense open subset μ1(P)M\mu^{-1}(P)\subset M. In the compact setting, it was proved by Abreu [2] using the global symplectic slice theorem that the boundary conditions (2.12) are equivalent to the fact that the complex structure JuJ_{u} extends to all of MM. Passing via the Legendre transform to complex coordinates on ()nM(\mathbb{C}^{*})^{n}\subset M, we see that the extension of JuJ_{u} to all of MM is then equivalent to the extension of the symplectic form ω=2i¯ϕ\omega=2i\partial\bar{\partial}\phi to MM.

These arguments were then improved by Apostolov-Calderbank-Gauduchon-Tønnesen-Friedman [5] and independently by Donaldson [19] who derived the boundary conditions (2.12) from purely local considerations. Indeed, the proof of [5, Proposition 1] proceeds by showing that if FF is any kk-dimensional face of PP, then for each point yFy\in F, the Hessian (uuP)ij(u-u_{P})_{ij} extends smoothly in a neighborhood of yy. This is achieved by choosing an arbitrary point qμ1(y)q\in\mu^{-1}(y) and a local symplectic slice (for the TnT^{n}-action), and then using the Taylor expansion for the metric around the point qq to prove that the complex structure JuJ_{u} defined by uu extending smoothly to qq is equivalent to the boundary condition (2.12) at yFy\in F, which applies verbatim in our setting. Translating back to the complex picture via the Legendre transform we see that this in turn is equivalent to the condition that the Kähler metric ω=2i¯ϕ\omega=2i\partial\bar{\partial}\phi extends to the subvariety VFV_{F} corresponding to the face FF given by the Orbit-Cone correspondence (Proposition 2.2), using the fact that VFV_{F} is naturally identified with μ1(F)\mu^{-1}(F) by Lemma 2.14. As in the compact case, since the boundary of PP is piecewise linear and since uuPu-u_{P} is smooth on the interior, this implies that uuPu-u_{P} itself extends smoothly to a neighborhood of each point yFy\in F. This completes the proof noting that FF and yy are arbitrary. ∎

Remark 1.

It should be noted, although it is not needed for our purposes, that this also holds under somewhat more general conditions. In particular let (M,ω)(M,\omega) be a 2n2n-dimensional symplectic toric manifold together with a compatible complex structure JJ, making (M,J,ω)(M,J,\omega) into a Kähler manifold (recall that this means that (M,ω)(M,\omega) admits a Hamiltonian TnT^{n}-action, but not necessarily a corresponding ()n(\mathbb{C}^{*})^{n}-action). Suppose that the moment map μ:M𝔱\mu:M\to\mathfrak{t}^{*} is proper, and as usual denote by P¯\overline{P} the image μ(M)𝔱\mu(M)\subset\mathfrak{t}^{*}. Then one can still define a symplectic potential uu by considering the complex structure JuJ_{u} associated to uu on PP, but it is no longer evident a priori in this setting that the metric gg associated to (M,J)(M,J) can be written in the form (2.7) since we do not have a corresponding Kähler potential ϕ\phi furnished by Proposition 2.4. However, using the properness of μ\mu, one can apply Lemma 2.13 to show that there is a globally defined isometry between gg and a metric gg^{\prime} which is defined on the interior of PP by (2.7), and then correspondingly deduce the boundary conditions (2.12) from [5, Proposition 1] as above. This was the approach of the recent work of Sena-Dias in [43, Section 3] to prove a uniqueness result for scalar-flat metrics on non-compact toric 4-manifolds which are not necessarily complex toric.

3 Convexity properties

3.1 The weighted volume functional

Let (N,ω)(N,\omega) be a Fano manifold with a given Kähler metric ω2πc1(N)\omega\in 2\pi c_{1}(N), and let 𝔥\mathfrak{h} be the space of all holomorphic vector fields on NN. Given v𝔥v\in\mathfrak{h}, let θv\theta_{v} be a Hamiltonian potential for JvJv with respect to the TkT^{k}-action generated by the flow of JvJv, which exists because in the compact manifolds with c1>0c_{1}>0 always satisfy H1(N)=0H^{1}(N)=0. Then set F(v)F(v) as

F(v)=Neθvωn.F(v)=\int_{N}e^{-\theta_{v}}\omega^{n}.

In order for this to be well-defined of course one must normalize θv\theta_{v}. With an appropriate choice, it turns out that F(v)F(v) is independent of choice of the metric ω\omega in its cohomology class [47]. The modified Futaki invariant of [47] is then defined as the derivative FX:𝔥F_{X}:\mathfrak{h}\to\mathbb{C} of FF at a given holomorphic vector field XX. Then FXF_{X} is independent of the choice of reference metric, and in [47] it is shown that FXF_{X} must therefore vanish identically if XX is the vector field corresponding to a Kähler-Ricci soliton on NN. A necessary condition therefore for XX to occur as the vector field of a shrinking gradient Kähler-Ricci soliton on NN is that FX0F_{X}\equiv 0.

It is shown in [13] that these ideas can be generalized to the non-compact setting in the presence of a complete shrinking gradient Kähler-Ricci soliton with bounded Ricci curvature. As in [13], we refer to FF as the weighted volume functional. Suppose that a real torus TkT^{k} acts on MM holomorphically and effectively with Lie algebra 𝔱\mathfrak{t}, and that the soliton vector field XX satisfies JX𝔱JX\in\mathfrak{t}. By the Duistermaat-Heckman theorem [22, 23, 42], there is an open cone Λ𝔱𝔥\Lambda\subset\mathfrak{t}\subset\mathfrak{h} where the weighted volume functional FF, and thereby the Futaki invariant, can be defined. Moreover, the domain Λ\Lambda can be naturally identified with the dual asymptotic cone of μ(M)𝔱\mu(M)\subset\mathfrak{t}^{*} (see [42, Definition A.2, Definition A.6]). Just as in [42], we will see that Λ\Lambda is in natural bijection with the space of Hamiltonian potentials which are proper and bounded below on MM. In this setting, the soliton vector field XX has the property that JXΛJX\in\Lambda and is the unique critical point of FF [13, Lemma 5.17]. This is analogous to the volume minimization principle of [38] for the Reeb vector field of a Sasaki-Einstein metric.

We show that on an AK-toric manifold MM with moment polyhedron PP, the weighted volume functional FF is proper, convex, and bounded from below. It is clear from the definitions that the asymptotic cone of PP is equal to its recession cone CC. Thus, there is a natural identification of the domain Λ\Lambda of FF with the dual recession cone C𝔱C^{*}\subset\mathfrak{t}. Fix a Delzant polyhedron PP and let MMPM\cong M_{P}. Throughout this section we make the extra assumption that PP contains the point zero in its interior. This of course can always be achieved by a translation, which corresponds to a modification of the moment map by a constant; see Lemma 2.11. Suppose that there exists an AK-toric metric ω\omega on MM with PP as its moment polyhedron. Then there is a potential ϕ\phi for ω\omega on the dense orbit. For any v𝔱v\in\mathfrak{t}, we know from Lemma 2.5 that there is a fixed bvnb_{v}\in\mathbb{R}^{n} such that the restriction of the Hamiltonian potential θv\theta_{v} to the dense orbit is determined by the function bv,ϕ\langle b_{v},\nabla\phi\rangle on n\mathbb{R}^{n}. Then passing to symplectic coordinates via the Legendre transform (2.6), we then see that θv\theta_{v} is determined by the linear function bv,x\langle b_{v},x\rangle on PP. The next proposition can be interpreted as the existence and uniqueness of a vector field in 𝔱\mathfrak{t} with vanishing Futaki invariant.

Proposition 3.1.

Let P𝔱P\subset\mathfrak{t}^{*} be a Delzant polyhedron containing zero in its interior. Then there exists a unique linear function P(x)\ell_{P}(x) determined by PP such that

P(x)eP(x)𝑑x=0\int_{P}\ell(x)e^{-\ell_{P}(x)}dx=0 (3.1)

for any linear function \ell on PP.

Proof.

Of course here 𝔱\mathfrak{t}^{*} can be any real vector space, although our only application is when 𝔱\mathfrak{t}^{*} is the dual Lie algebra of a real torus TnT^{n}. Let C𝔱C\subset\mathfrak{t}^{*} be the recession cone of PP. It follows immediately from the definition that the interior of CC^{*} is characterized by those b𝔱b\in\mathfrak{t} such that the linear function b,x\langle b,x\rangle on PP is positive outside of a compact set. Indeed, for each b𝔱b\in\mathfrak{t}, set

Hb={xn|b,x0},H_{b}=\{x\in\mathbb{R}^{n}\>|\>\langle b,x\rangle\leq 0\},

and

Qb=HbP¯.Q_{b}=H_{b}\cap\overline{P}.

We see from the definition (Definition 3) that an element y𝔱y\in\mathfrak{t}^{*} lies in CC if and only if x+λyPx+\lambda y\in P for all xPx\in P, λ0\lambda\geq 0. Thus QbQ_{b} is compact if and only if for each xQbx\in Q_{b}, and for each yCy\in C, there exists a λ>0\lambda>0 such that x+λy,b=0\langle x+\lambda y,b\rangle=0. Since x,b0\langle x,b\rangle\leq 0, it follows that QbQ_{b} is compact if and only if bCb\in C^{*}. Thus eb,xe^{-\langle b,x\rangle} is integrable on PP, and so there is a well-defined function F:CF:C^{*}\to\mathbb{R} given by

F(b)=Peb,x𝑑x.F(b)=\int_{P}e^{-\langle b,x\rangle}dx.

Then

bjF=(Pxjeb,x𝑑x).\frac{\partial}{\partial b^{j}}F=-\left(\int_{P}x^{j}e^{-\langle b,x\rangle}dx\right).

Moreover, the critical points of FF are precisely solutions P\ell_{P} to (3.1). The function FF is convex which immediately gives uniqueness. To show existence, it suffices to show that FF is proper. That is, given a sequence bjb_{j} in the interior of CC^{*} such that either |bj||b_{j}|\to\infty or the sequence {bj}\{b_{j}\} approaches a point on the boundary, we need to show that F(bj)F(b_{j})\to\infty. Consider the former case first. Using the natural inner product on 𝔱\mathfrak{t}, we can view the dual recession cone CC^{*} as sitting inside of 𝔱\mathfrak{t}^{*}. Since 0P0\in P, the intersection Q=CPQ=-C^{*}\cap P has positive measure in n\mathbb{R}^{n}. Now suppose that {bj}\{b_{j}\} is any sequence in CC^{*} such that |bj||b_{j}|\to\infty. Let yQy\in Q be a fixed point in the interior and choose ε\varepsilon sufficiently small so that Bε(y)QB_{\varepsilon}(y)\subset Q has strictly positive Euclidean distance to the boundary Q\partial Q. In particular, we then have that infvSn1C¯v,y>0\inf_{v\in S^{n-1}\cap\overline{C}^{*}}\langle v,-y\rangle>0. We choose ε\varepsilon sufficiently small so that δ=infvSn1C¯v,yε>0\delta=\inf_{v\in S^{n-1}\cap\overline{C}^{*}}\langle v,-y\rangle-\varepsilon>0. For any xBε(y)x\in B_{\varepsilon}(y), write x=y+rwx=y+rw for r[0,ε)r\in[0,\varepsilon) and wSn1w\in S^{n-1}. Then we have, for any (b,x)C×Bε(y)(b,x)\in C^{*}\times B_{\varepsilon}(y),

b,xb,yr|b||v|(b|b|,yε)|b|δ|b|.-\langle b,x\rangle\geq\langle b,-y\rangle-r|b||v|\geq\left(\left\langle\frac{b}{|b|},-y\right\rangle-\varepsilon\right)|b|\geq\delta|b|.

Therefore, we see immediately that

F(bj)=Pebj,x𝑑xBε(y)ebj,x𝑑xBε(y)eδ|bj|𝑑x.F(b_{j})=\int_{P}e^{-\langle b_{j},x\rangle}dx\geq\int_{B_{\varepsilon}(y)}e^{-\langle b_{j},x\rangle}dx\geq\int_{B_{\varepsilon}(y)}e^{\delta|b_{j}|}dx.

Since |bj||b_{j}|\to\infty, we have then that F(bj)F(b_{j})\to\infty.

Consider now the latter case. The key point is that C\partial C^{*} is defined by those b¯n\bar{b}\in\mathbb{R}^{n} such that there exists at least one c¯C¯\bar{c}\in\overline{C} with b¯,c¯=0\langle\bar{b},\bar{c}\rangle=0. Choose b¯C\bar{b}\in\partial C^{*}. The result essentially follows from the fact that the polyhedron Qb¯Q_{\bar{b}} defined above is unbounded. More explicitly, if c¯\bar{c} is a point with b¯,c¯=0\langle\bar{b},\bar{c}\rangle=0, then for any x0Qb¯x_{0}\in Q_{\bar{b}} we have that x0+λcQb¯x_{0}+\lambda c\in Q_{\bar{b}} for any λ0\lambda\geq 0. If we then fix a small (n1)(n-1)-disc Dε(x0)Qb¯D_{\varepsilon}(x_{0})\subset Q_{\bar{b}} perpendicular to cc, consider the tubes Tλ={x+rc|xDε(x0),r(0,λ)}Qb¯T_{\lambda}=\{x+rc\>|\>x\in D_{\varepsilon}(x_{0}),r\in(0,\lambda)\}\subset Q_{\bar{b}}. Take a sequence of points bib¯b_{i}\to\bar{b} with bjb_{j} in the interior of CC^{*}, and define QbjQ_{b_{j}} and HbjH_{b_{j}} as above. Recall that each QbjQ_{b_{j}} is bounded. Choosing ε\varepsilon small enough, and perhaps after removing finitely many terms from {bj}\{b_{j}\}, we can assume that Dε(x0)D_{\varepsilon}(x_{0}) is contained in Qb1Q_{b_{1}}. Let λj\lambda_{j} be the largest positive number such that TλjQbjT_{\lambda_{j}}\subset Q_{b_{j}}. Since QbjQb¯Q_{b_{j}}\to Q_{\bar{b}}, we see that λj\lambda_{j}\to\infty. Then we have

F(bj)=Pebj,x𝑑xTλjebj,x𝑑x=λjDε(x0)ebj,y𝑑y,F(b_{j})=\int_{P}e^{-\langle b_{j},x\rangle}dx\geq\int_{T_{\lambda_{j}}}e^{-\langle b_{j},x\rangle}dx=\lambda_{j}\int_{D_{\varepsilon}(x_{0})}e^{-\langle b_{j},y\rangle}dy,

where yy are the coordinates on Dε(x0)D_{\varepsilon}(x_{0}). Clearly F(bj)F(b_{j})\to\infty. ∎

Corollary 3.2.

Let PnP\subset\mathbb{R}^{n} be a Delzant polyhedron, M=MPM=M_{P}, and suppose that ω\omega is a TnT^{n}-invariant Kähler metric with PP as its moment polyhedron. Let vv be the holomorphic vector field on MM determined by bv𝔱b_{v}\in\mathfrak{t} and θv\theta_{v} be a Hamiltonian potential for JvJv. Then

Meθvωn<\int_{M}e^{-\theta_{v}}\omega^{n}<\infty

if and only if bvb_{v} lies in the dual recession cone CC^{*}.

Proof.

We work on the dense orbit in symplectic coordinates ()nP×Tn(\mathbb{C}^{*})^{n}\cong P\times T^{n}. We have seen in Section 2.4 that in these coordinates ω\omega is given simply by ω=dxidθi\omega=\sum dx^{i}\wedge d\theta^{i} so that the integral above becomes

()neθvωn=P×Tnebv,x𝑑x𝑑θ=(2π)nPebv,x𝑑x.\int_{(\mathbb{C}^{*})^{n}}e^{-\theta_{v}}\omega^{n}=\int_{P\times T^{n}}e^{-\langle b_{v},x\rangle}dxd\theta=(2\pi)^{n}\int_{P}e^{-\langle b_{v},x\rangle}dx.

As we have seen, this is finite precisely when bvCb_{v}\in C^{*}. ∎

As a consequence, we recover the result of [42] that domain the Λ\Lambda of the weighted volume functional FF can be identified with the dual asymptotic cone CC^{*}.

3.2 The soliton equation

Let PP be a Delzant polyhedron containing zero in its interior and MMPM\cong M_{P}. Suppose that there is a complete TnT^{n}-invariant shrinking gradient Kähler-Ricci soliton ω\omega with PP as its moment polyhedron, and whose soliton vector field XX satisfies JX𝔱JX\in\mathfrak{t}. From Proposition 2.12, we know that there is a corresponding symplectic potential uC(P)u\in C^{\infty}(P) which satisfies

2(uixiu(x))logdet(uij)=bX,x,2\left(u_{i}x^{i}-u(x)\right)-\log\det(u_{ij})=\langle b_{X},x\rangle,

where the linear function bX,x\langle b_{X},x\rangle on PP corresponds via the Legendre transform to the Hamiltonian potential θX=μ(JX)\theta_{X}=\mu(JX) for JXJX. We adopt the following simplification of notation from [20]. For a given uC(P)u\in C^{\infty}(P), set

ρu=2(uixiu(x))logdet(uij)\rho_{u}=2\left(u_{i}x^{i}-u(x)\right)-\log\det(u_{ij}) (3.2)

so that the soliton equation can once again be rewritten as

ρu=bX,x.\rho_{u}=\langle b_{X},x\rangle. (3.3)

The function eρe^{-\rho} is natural to study in the context of integration over PP. In particular,

Corollary 3.3.

Let PP be a Delzant polyhedron containing zero in its interior. For any smooth and convex function uu on PP, we have that

Peρu𝑑x<.\int_{P}e^{-\rho_{u}}dx<\infty.
Proof.

To prove the corollary, we let ϕu(ξ)=L(u)\phi_{u}(\xi)=L(u) be the Legendre transform and apply the change of coordinates x=ϕu(ξ)x=\nabla\phi_{u}(\xi), where ξ\xi denotes coordinates on the domain Ωn\Omega\subset\mathbb{R}^{n} of ϕu\phi_{u}. Then from Lemma 2.9 we have

det(uij)dx=dξ,\det(u_{ij})dx=d\xi,

and

uu,x=ϕu(ξ).u-\langle\nabla u,x\rangle=-\phi_{u}(\xi).

Therefore

Peρu𝑑x=Ωe2ϕu𝑑ξ.\int_{P}e^{-\rho_{u}}dx=\int_{\Omega}e^{-2\phi_{u}}d\xi.

Then from Lemma 2.10 we know that e2ϕue^{-2\phi_{u}} is integrable on Ω\Omega. ∎

Remark 2.

We emphasize at this stage the statement of Lemma 2.10; simply by asserting that zero lies in the domain of uu, it follows automatically that the Legendre transform ϕu\phi_{u} of uu is proper.

Corollary 3.4.

Let PP be a Delzant polyhedron containing zero in its interior, and suppose that there exists a solution uC(P)u\in C^{\infty}(P) to (3.3). Then the element bX𝔱b_{X}\in\mathfrak{t} determining JXJX lies in CC^{*}.

Proof.

Since PP contains zero in its interior, we have by Corollary 3.3 that

Peρu𝑑x<.\int_{P}e^{-\rho_{u}}dx<\infty.

Since uu satisfies (3.3), we have

PebX,x𝑑x<.\int_{P}e^{-\langle b_{X},x\rangle}dx<\infty.

Since the restriction of the Hamiltonian potential θX\theta_{X} for JXJX to P×TnP\times T^{n} is given by θX|P×Tn=bX,x\theta_{X}|_{P\times T^{n}}=\langle b_{X},x\rangle, it follows from Corollary 3.2 that bXCb_{X}\in C^{*}. ∎

Lemma 3.5.

Let PP be a Delzant polyhedron containing zero in its interior, and suppose that there exists a solution uC(P)u\in C^{\infty}(P) to (3.3). Then the linear function bX,x\langle b_{X},x\rangle on PP satisfies

P(x)ebX,x𝑑x=0\int_{P}\ell(x)e^{-\langle b_{X},x\rangle}dx=0

for any linear function (x)\ell(x) on PP.

Proof.

First, we claim that any function uC(P)u\in C^{\infty}(P) which is the Legendre transform of a smooth convex function ϕ\phi on n\mathbb{R}^{n} satisfies

P(x)eρu𝑑x=0\int_{P}\ell(x)e^{-\rho_{u}}dx=0

for any linear function (x)\ell(x) on PP. Pick any coordinate xjx^{j} and compute

Pxjeρu𝑑x=nϕje2ϕ𝑑ξ=12n(e2ϕ)j𝑑ξ.\int_{P}x^{j}e^{-\rho_{u}}dx=\int_{\mathbb{R}^{n}}\phi_{j}e^{-2\phi}d\xi=-\frac{1}{2}\int_{\mathbb{R}^{n}}\left(e^{-2\phi}\right)_{j}d\xi.

By Lemma 2.10, we know that eϕe^{-\phi} decays at least exponentially in |x||x|. Thus, integration by parts yields that the term on the right-hand side is zero. Then if uu satisfies ρu=bX,x\rho_{u}=\langle b_{X},x\rangle, it follows that

PxjebX,x𝑑x=12n(e2ϕ)j𝑑ξ=0\displaystyle\int_{P}x^{j}e^{-\langle b_{X},x\rangle}dx=-\frac{1}{2}\int_{\mathbb{R}^{n}}\left(e^{-2\phi}\right)_{j}d\xi=0

for each jj. ∎

Therefore, the linear function bX,x\langle b_{X},x\rangle on PP must be equal to the unique linear function P\ell_{P} determined by Proposition 3.1. We will henceforth denote

bX,x=P(x)\langle b_{X},x\rangle=\ell_{P}(x)

since whenever both sides exist, they must coincide.

3.3 Real Monge-Ampère equations on unbounded convex domains

In this section we study the analytic properties of some real Monge-Ampère equations of the same form as (3.3). More precisely, we will consider equations of the form

ρu=A,\rho_{u}=A, (3.4)

where now the right-hand side A(x)C(P¯)A(x)\in C^{\infty}(\overline{P}) can be any smooth function satisfying some fixed hypotheses which we will discuss below. When PP is bounded, this is also the approach taken in [8] and [20]. Let PP be a Delzant polyhedron defined by the system of inequalities i(x)+ai0\ell_{i}(x)+a_{i}\geq 0, and suppose that PP contains zero in its interior. Define uPu_{P} as in (2.11) by

uP(x)=12i=1d(i(x)+ai)log(i(x)+ai),u_{P}(x)=\frac{1}{2}\sum_{i=1}^{d}(\ell_{i}(x)+a_{i})\log\left(\ell_{i}(x)+a_{i}\right),

recalling that uPu_{P} is the symplectic potential of the canonical Kähler metric ωP\omega_{P} on MPM_{P}. Let A(x)C(P¯)A(x)\in C^{\infty}(\overline{P}). We will say that AA is admissible if each of the following conditions hold:

  1. 1.

    VA=PeA𝑑x<V_{A}=\int_{P}e^{-A}dx<\infty,

  2. 2.

    P(x)eA𝑑x=0 for any linear function \int_{P}\ell(x)e^{-A}dx=0\text{ for any linear function }\ell,

  3. 3.

    PuPeA𝑑x<\int_{P}u_{P}e^{-A}dx<\infty.

For an admissible function AA. In analogy with Proposition 2.17, we set A1,\mathcal{E}_{A}^{1,\infty} to be the set

A1,={u=uP+v|PueA(x)𝑑x<,(u)ij>0,vC(P¯)},\mathcal{E}_{A}^{1,\infty}=\left\{u=u_{P}+v\,\bigg{|}\,\int_{P}ue^{-A(x)}dx<\infty\,,\,(u)_{ij}>0\,,\,v\in C^{\infty}(\overline{P})\right\},

and similarly

A1,0={u=uP+v|P|u|eA(x)𝑑x<,u is convex,vC0(P¯)}.\mathcal{E}_{A}^{1,0}=\left\{u=u_{P}+v\,\bigg{|}\,\int_{P}|u|e^{-A(x)}dx<\infty\,,\,u\textnormal{ is convex}\,,\,v\in C^{0}(\overline{P})\right\}.

The space 𝒫\mathcal{P} of symplectic potentials is then

𝒫={uA1,|u:Pn is surjective}.\mathcal{P}=\left\{u\in\mathcal{E}_{A}^{1,\infty}\,\bigg{|}\,\nabla u:P\to\mathbb{R}^{n}\textnormal{ is surjective}\right\}.

In fact we have 𝒫A1,A1,0\mathcal{P}\subset\mathcal{E}_{A}^{1,\infty}\subset\mathcal{E}_{A}^{1,0}. The first inclusion is clear, and to see the second we proceed as follows. If uA1,u\in\mathcal{E}_{A}^{1,\infty}, we can modify by a linear function to ensure that u(0)=0\nabla u(0)=0, and since AA is admissible this does not affect the value of PueA𝑑x\int_{P}ue^{-A}dx. By Lemma 2.10 we can add a constant to uu to ensure that u0u\geq 0, and again the admissibility of AA ensures that this only affects the value of PueA𝑑x\int_{P}ue^{-A}dx by a the addition of a constant. Hence we see that P|u|eA𝑑x<\int_{P}|u|e^{-A}dx<\infty for any uA1,u\in\mathcal{E}_{A}^{1,\infty}. The space 𝒫\mathcal{P} can be naturally viewed as a convex subset C(P¯)C^{\infty}(\overline{P}).

Lemma 3.6.

Suppose that u0,u1𝒫u_{0},u_{1}\in\mathcal{P} and set ut=tu1+(1t)u0u_{t}=tu_{1}+(1-t)u_{0}. Then ut:Pn\nabla u_{t}:P\to\mathbb{R}^{n} is surjective for all t[0,1]t\in[0,1].

Proof.

We first observe that this is true when n=1n=1. Indeed, in this case PP can be taken to be an interval (a,b)(a,b) for a<0a<0 and b(0,]b\in(0,\infty]. Then by convexity utx:(a,b)\frac{\partial u_{t}}{\partial x}:(a,b)\to\mathbb{R} will be surjective if and only if limxbutx=\lim_{x\to b}\frac{\partial u_{t}}{\partial x}=\infty and limxautx=\lim_{x\to a}\frac{\partial u_{t}}{\partial x}=-\infty. This is a property that utu_{t} clearly inherits from u0u_{0} and u1u_{1}.

In general, we suppose for the sake of contradiction that there is some time tt such that ut(P)=Ωn\nabla u_{t}(P)=\Omega\subsetneq\mathbb{R}^{n}. Choose ξΩ\xi^{*}\in\partial\Omega and a sequence xiPx_{i}\in P such that ut(xi)ξ\nabla u_{t}(x_{i})\to\xi^{*}. By potentially adding a linear function, we assume without loss of generality that u0(0)=u1(0)=0\nabla u_{0}(0)=\nabla u_{1}(0)=0. By passing to a subsequence then we can assume that either xix_{i} accumulate in P\partial P, |xi||x_{i}|\to\infty, or indeed both. In either case, it follows from the choice of normalization together with the one-dimensional case that the radial derivative utr\frac{\partial u_{t}}{\partial r} satisfies |utr(xi)|\left|\frac{\partial u_{t}}{\partial r}(x_{i})\right|\to\infty, and hence |ut|(xi)|\nabla u_{t}|(x_{i})\to\infty. This is a contradiction with the assumption that ut(xi)ξ\nabla u_{t}(x_{i})\to\xi^{*}. ∎

Lastly we define 𝒫0𝒫\mathcal{P}_{0}\subset\mathcal{P} to be the space of normalized symplectic potentials; these will be those u𝒫u\in\mathcal{P} such that

PueA(x)𝑑x=0.\int_{P}ue^{-A(x)}dx=0. (3.5)

Clearly for any u𝒫u\in\mathcal{P}, we can find a constant cc such that u+c𝒫0u+c\in\mathcal{P}_{0}.

Definition 9.

Given any u0,u1𝒫u_{0},u_{1}\in\mathcal{P}, we say that the linear path ut=(1t)u0+tu1u_{t}=(1-t)u_{0}+tu_{1} joining u0u_{0} and u1u_{1} is a geodesic.

We will see that, as a consequence of an elementary local argument, geodesics in this sense have the property that their Legendre transforms define geodesics in the space of Kähler metrics on MMPM\cong M_{P} in the usual sense. The interpretation is that if u0,u1C(P)u_{0},u_{1}\in C^{\infty}(P) are the Legendre transforms of two Kähler potentials ϕ0,ϕ1\phi_{0},\phi_{1} on MM, then the path ϕt=L(ut)\phi_{t}=L(u_{t}) solves the pointwise equation

ϕ¨t12|ωtϕ˙t|ωt2=0,\ddot{\phi}_{t}-\frac{1}{2}\left|\nabla_{\omega_{t}}\dot{\phi}_{t}\right|^{2}_{\omega_{t}}=0, (3.6)

and can thus be considered a geodesic in the space of Kähler metrics in the sense of [17]. This is a simple exercise in the basic properties of the Legendre transform. We will only make use of a small piece of the computation, but for completeness we include the proof below.

Lemma 3.7.

Let utu_{t} be any path in 𝒫\mathcal{P} and ϕt=L(ut)\phi_{t}=L(u_{t}). Then the time derivatives satisfy

u˙t=ϕ˙t.\dot{u}_{t}=-\dot{\phi}_{t}. (3.7)

Consequently, if u¨t=0\ddot{u}_{t}=0 then ϕt\phi_{t} satisfies (3.6).

Proof.

We have

u˙t=tut(x)\displaystyle\dot{u}_{t}=\frac{\partial}{\partial t}u_{t}(x) =t(ut,xϕt(ut))\displaystyle=\frac{\partial}{\partial t}\left(\big{\langle}\nabla u_{t},x\big{\rangle}-\phi_{t}(\nabla u_{t})\right)
=tut,xϕ˙(ut)ϕt,tut\displaystyle=\bigg{\langle}\frac{\partial}{\partial t}\nabla u_{t},x\bigg{\rangle}-\dot{\phi}(\nabla u_{t})-\bigg{\langle}\nabla\phi_{t},\frac{\partial}{\partial t}\nabla u_{t}\bigg{\rangle}
=ϕ˙,\displaystyle=-\dot{\phi},

which is the first statement. For the second, note that it follows from Lemma 2.9 that

ϕ˙tξj=utiju˙txi.\displaystyle\frac{\partial\dot{\phi}_{t}}{\partial\xi^{j}}=-u_{t}^{ij}\frac{\partial\dot{u}_{t}}{\partial x^{i}}.

Now compute

u¨t\displaystyle\ddot{u}_{t} =tϕ˙t(ut)=ϕ¨t(ut)mϕ˙tξmu˙txm\displaystyle=-\frac{\partial}{\partial t}\dot{\phi}_{t}(\nabla u_{t})=-\ddot{\phi}_{t}(\nabla u_{t})-\sum_{m}\frac{\partial\dot{\phi}_{t}}{\partial\xi^{m}}\frac{\partial\dot{u}_{t}}{\partial x^{m}}
=ϕ¨t(ut)+utlmu˙txlu˙txm,\displaystyle=-\ddot{\phi}_{t}(\nabla u_{t})+u_{t}^{lm}\frac{\partial\dot{u}_{t}}{\partial x^{l}}\frac{\partial\dot{u}_{t}}{\partial x^{m}},

so that

ϕ¨t12|ωtϕ˙t|t2\displaystyle\ddot{\phi}_{t}-\frac{1}{2}|\nabla_{\omega_{t}}\dot{\phi}_{t}|_{t}^{2} =ϕ¨tϕtijϕ˙tξiϕ˙tξj\displaystyle=\ddot{\phi}_{t}-\phi_{t}^{ij}\frac{\partial\dot{\phi}_{t}}{\partial\xi^{i}}\frac{\partial\dot{\phi}_{t}}{\partial\xi^{j}}
=u¨t+utlmu˙txlu˙txm(ut)ijutilutmju˙txlu˙txm\displaystyle=-\ddot{u}_{t}+u_{t}^{lm}\frac{\partial\dot{u}_{t}}{\partial x^{l}}\frac{\partial\dot{u}_{t}}{\partial x^{m}}-(u_{t})_{ij}u_{t}^{il}u_{t}^{mj}\frac{\partial\dot{u}_{t}}{\partial x^{l}}\frac{\partial\dot{u}_{t}}{\partial x^{m}}
=u¨t=0.\displaystyle=-\ddot{u}_{t}=0.

Remark 3.

While the proof of (3.6) requires two spacial derivatives of uu and ϕ\phi, the proof of the simpler equality (3.7) works at any point where uu and ϕ\phi are C1C^{1}. Since any convex function is C1C^{1} outside of a set of measure zero, it follows that (3.7) actually holds almost everywhere (in the sense that u˙t(x)=ϕ˙t(ut(x))\dot{u}_{t}(x)=-\dot{\phi}_{t}(\nabla u_{t}(x))) for any uA1,0u\in\mathcal{E}^{1,0}_{A}, a fact that we will make use of later on.

We introduce a Ding-type functional 𝒟\mathcal{D} defined on A1,0\mathcal{E}^{1,0}_{A} whose critical points, at least formally, are solutions to (3.4). Define 𝒟1\mathcal{D}_{1} on A1,0\mathcal{E}^{1,0}_{A} by setting

𝒟1(u)=ne2ϕu𝑑ξ,\mathcal{D}_{1}(u)=\int_{\mathbb{R}^{n}}e^{-2\phi_{u}}d\xi, (3.8)

where ϕu=L(u)\phi_{u}=L(u). This is well-defined on A1,0\mathcal{E}^{1,0}_{A} by Lemma 2.10, since the domain PP of uu contains zero by assumption. In particular, we can extend e2ϕue^{-2\phi_{u}} continuously by zero outside of the domain of ϕu\phi_{u} to make sense of the integral (3.8) over all of n\mathbb{R}^{n}.

Remark 4.

Whenever uA1,0u\in\mathcal{E}^{1,0}_{A} is C2C^{2} in the interior of PP, in particular when uA1,u\in\mathcal{E}^{1,\infty}_{A}, it follows that

𝒟1(u)=Peρu𝑑x.\mathcal{D}_{1}(u)=\int_{P}e^{-\rho_{u}}dx. (3.9)

The Ding functional 𝒟\mathcal{D} on A1,0\mathcal{E}^{1,0}_{A} is then defined to be

𝒟(u)=1VAPueA𝑑x12log𝒟1(u).\mathcal{D}(u)=\frac{1}{V_{A}}\int_{P}ue^{-A}dx-\frac{1}{2}\log{\mathcal{D}_{1}(u)}. (3.10)
Lemma 3.8.

Suppose that u𝒫u\in\mathcal{P} satisfies (3.4) and that wC00(n)w\in C^{0}_{0}(\mathbb{R}^{n}) is a continuous and compactly supported (as a function on n\mathbb{R}^{n}), such that ut=u+twA1,0u_{t}=u+tw\in\mathcal{E}^{1,0}_{A} for sufficiently small tt. Then the first variation of 𝒟1\mathcal{D}_{1} at uu in the direction ww is given by

δu𝒟1(w)=2PweA(x)𝑑x,\delta_{u}\mathcal{D}_{1}(w)=2\int_{P}we^{-A(x)}dx,

and consequently

t|t=0𝒟(u+tw)=0.\left.\frac{\partial}{\partial t}\right|_{t=0}\mathcal{D}(u+tw)=0.
Proof.

Let ϕt=L(ut)\phi_{t}=L(u_{t}). Since ww is compactly supported, u:Pn\nabla u:P\to\mathbb{R}^{n} is surjective, and the domain of the Legendre transform is convex, it follows that the domain of ϕt\phi_{t} is the whole of n\mathbb{R}^{n}, and moreover that ϕt=ϕ0=L(u)\phi_{t}=\phi_{0}=L(u) outside of a fixed compact set independent of tt. Moreover, by Lemma 2.10 we have that

ϕt(ξ)ε|ξ|supBε(0)|ut|ε|ξ|C,\phi_{t}(\xi)\geq\varepsilon|\xi|-\sup_{B_{\varepsilon}(0)}|u_{t}|\geq\varepsilon|\xi|-C,

for ε,C>0\varepsilon,C>0 independent of tt. By (3.7), we know that there is a set EPE\subset P of measure zero and a compact subset KtnK_{t}\subset\mathbb{R}^{n} (which does not necessarily have zero measure) such that ut(P\E)=n\Kt\nabla u_{t}(P\backslash E)=\mathbb{R}^{n}\backslash K_{t} and supn\Kt|ϕ˙t|=supP|w|<\sup_{\mathbb{R}^{n}\backslash K_{t}}|\dot{\phi}_{t}|=\sup_{P}|w|<\infty. This tells us two things. First, since u0=uu_{0}=u is smooth, we see that K0=u(E)nK_{0}=\nabla u(E)\subset\mathbb{R}^{n} has measure zero. Moreover, as we have seen the family KtK_{t} is contained in a fixed ball BnB\subset\mathbb{R}^{n} independent of tt, so that in fact supn|ϕ˙t|C(supP|w|+1)\sup_{\mathbb{R}^{n}}|\dot{\phi}_{t}|\leq C(\sup_{P}|w|+1). Thus

|ϕ˙t|eϕtC(supP|w|+1)eϕtC(supP|w|+1)eε|ξ|+CL1(n).\begin{split}|\dot{\phi}_{t}|e^{-\phi_{t}}\leq C(\sup_{P}|w|+1)e^{-\phi_{t}}\leq C(\sup_{P}|w|+1)e^{-\varepsilon|\xi|+C}\in L^{1}(\mathbb{R}^{n}).\end{split}

Hence, by the mean value theorem and the dominated convergence theorem, it follows that

t𝒟1(ut)|t=0=2nϕ˙0e2ϕ0𝑑ξ=2PweA𝑑x,\left.\frac{\partial}{\partial t}\mathcal{D}_{1}(u_{t})\right|_{t=0}=-2\int_{\mathbb{R}^{n}}\dot{\phi}_{0}e^{-2\phi_{0}}d\xi=2\int_{P}we^{-A}dx,

using (3.7), (3.4) and that K0K_{0} has measure zero. So

t𝒟1(ut)|t=0\displaystyle\left.\frac{\partial}{\partial t}\mathcal{D}_{1}(u_{t})\right|_{t=0} =1VAPweA𝑑xδu𝒟1(w)2𝒟1(u)\displaystyle=\frac{1}{V_{A}}\int_{P}we^{-A}dx-\frac{\delta_{u}\mathcal{D}_{1}(w)}{2\mathcal{D}_{1}(u)}
=1VAPweA𝑑x1ne2ϕ0𝑑ξPweA𝑑x=0,\displaystyle=\frac{1}{V_{A}}\int_{P}we^{-A}dx-\frac{1}{\int_{\mathbb{R}^{n}}e^{-2\phi_{0}}d\xi}\int_{P}we^{-A}dx=0,

since ne2ϕ0𝑑ξ=Peρu𝑑x=PeA𝑑x\int_{\mathbb{R}^{n}}e^{-2\phi_{0}}d\xi=\int_{P}e^{-\rho_{u}}dx=\int_{P}e^{-A}dx by (3.4). ∎

Proposition 3.9 (c.f. [8, Proposition 2.15]).

The Ding functional 𝒟\mathcal{D} is convex on A1,0\mathcal{E}^{1,0}_{A}. It is invariant under the action of n×\mathbb{R}^{n}\times\mathbb{R} given by addition of affine-linear functions, and it is strictly convex modulo this action. In particular, suppose that u0,u1𝒫0u_{0},u_{1}\in\mathcal{P}_{0}. Then if 𝒟(tu1+(1t)u0)=t𝒟(u1)+(1t)𝒟(u0)\mathcal{D}(tu_{1}+(1-t)u_{0})=t\mathcal{D}(u_{1})+(1-t)\mathcal{D}(u_{0}), there exists a linear function (x)\ell(x) on PP such that u1=u0+u_{1}=u_{0}+\ell.

Proof.

If u0,u1A1,0u_{0},u_{1}\in\mathcal{E}^{1,0}_{A} satisfy u1=u0+(x)+au_{1}=u_{0}+\ell(x)+a with aa\in\mathbb{R} and \ell any linear function, then by Lemma 2.9 we see that neϕ1𝑑ξ=e2aneϕ0𝑑ξ\int_{\mathbb{R}^{n}}e^{-\phi_{1}}d\xi=e^{2a}\int_{\mathbb{R}^{n}}e^{-\phi_{0}}d\xi. Therefore we see directly from the definition (3.10) that the fact that 𝒟\mathcal{D} is invariant is equivalent to the statement that P(x)eA𝑑x=0\int_{P}\ell(x)e^{-A}dx=0 for any linear function \ell on PP, which AA satisfies by definition. We prove convexity directly, and show that

𝒟(ut)t𝒟(u1)+(1t)𝒟(u0),\mathcal{D}(u_{t})\leq t\mathcal{D}(u_{1})+(1-t)\mathcal{D}(u_{0}),

where ut=tu1+(1t)u0u_{t}=tu_{1}+(1-t)u_{0} for any u0,u1A1,0u_{0},u_{1}\in\mathcal{E}^{1,0}_{A}. Set ϕt=L(ut)\phi_{t}=L(u_{t}). First notice that the functional uPueA𝑑xu\mapsto\int_{P}ue^{-A}dx is clearly affine on A1,0\mathcal{E}^{1,0}_{A}. Therefore it suffices to show that the function

tlogne2ϕt𝑑ξt\mapsto-\log\int_{\mathbb{R}^{n}}e^{-2\phi_{t}}d\xi

is convex in tt. This follows from the fact that the Legendre transform is itself a convex mapping, i.e.

ϕt(ξ)tϕ1(ξ)+(1t)ϕ0(ξ),\phi_{t}(\xi)\leq t\phi_{1}(\xi)+(1-t)\phi_{0}(\xi), (3.11)

which is the fourth item in Lemma 2.9. It then follows immediately from the Prékopa-Leindler inequality [21] that this is convex in tt. This says precisely that any family ϕt\phi_{t} of convex functions satisfying (3.11) has the property that the function of one variable ne2ϕt𝑑ξ\int_{\mathbb{R}^{n}}e^{-2\phi_{t}}d\xi is log-concave (i.e. tlogne2ϕt𝑑ξt\mapsto-\log\int_{\mathbb{R}^{n}}e^{-2\phi_{t}}d\xi is convex). The strict convexity follows from the equality case of the Prékopa-Leindler inequality, which was also studied in [21]. If the function ne2ϕt𝑑ξ\int_{\mathbb{R}^{n}}e^{-2\phi_{t}}d\xi is affine in tt, then by [21, Theorem 12] there exists mm\in\mathbb{R} and ana\in\mathbb{R}^{n} such that

ϕ1(ξ)=ϕ0(mξ+a)nlog(m)log(ne2ϕ1𝑑ξne2ϕ0𝑑ξ).\phi_{1}(\xi)=\phi_{0}(m\xi+a)-n\log(m)-\log\left(\frac{\int_{\mathbb{R}^{n}}e^{-2\phi_{1}}d\xi}{\int_{\mathbb{R}^{n}}e^{-2\phi_{0}}d\xi}\right).

Firstly, we see that mm must be equal to 1 since u0,u1𝒫u_{0},u_{1}\in\mathcal{P}. Indeed L(ϕ0(mξ))=u0(m1x)L(\phi_{0}(m\xi))=u_{0}(m^{-1}x). If u0𝒫u_{0}\in\mathcal{P}, then u0(m1x)uP(x)C(P¯)u_{0}(m^{-1}x)-u_{P}(x)\in C^{\infty}(\overline{P}) if and only if m=1m=1. Then we have that ϕ1(ξ)=ϕ0(ξ+a)C\phi_{1}(\xi)=\phi_{0}(\xi+a)-C for some CC. Again passing to the Legendre transform, we have that

u1(x)=L(ϕ1(ξ))=L(ϕ0(ξ+a)C)=u0(x)+a(x)+C.u_{1}(x)=L(\phi_{1}(\xi))=L(\phi_{0}(\xi+a)-C)=u_{0}(x)+\ell_{a}(x)+C.

Finally, the normalization condition (3.5) implies that in fact C=0C=0. ∎

To prove that solutions to (3.4) in 𝒫0\mathcal{P}_{0} are unique, we would like to make use of this strict convexity. To do this, we need to ensure that, if u0,u1𝒫0u_{0},u_{1}\in\mathcal{P}_{0} are two solutions, the Ding functional 𝒟\mathcal{D} is minimized along the geodesic ut=tu1+(1t)u0u_{t}=tu_{1}+(1-t)u_{0} at the endpoints t=0,1t=0,1. This would be clear from Lemma 3.8 if the variation v=u1u0v=u_{1}-u_{0} were compactly supported, but there is no reason a priori why this should be the case. To this end, we have

Lemma 3.10.

Suppose that u𝒫0u\in\mathcal{P}_{0} and vC(P¯)v\in C^{\infty}(\overline{P}) is such that uv:=u+v𝒫0u_{v}:=u+v\in\mathcal{P}_{0}. Then there exists a sequence of compactly supported functions wiC00(n)w_{i}\in C^{0}_{0}(\mathbb{R}^{n}) such that Ui:=u+wiA1,0U_{i}:=u+w_{i}\in\mathcal{E}^{1,0}_{A} and that

𝒟(Ui)𝒟(uv)\mathcal{D}(U_{i})\to\mathcal{D}(u_{v})

as ii\to\infty.

Proof.

Let 𝔱Cn\mathfrak{t}^{*}_{C}\subset\mathbb{R}^{n} be the linear subspace spanned by recession cone CC of PP. We can see from the definition of CC (Definition 3) that there exists some point qnq\in\mathbb{R}^{n}, not necessarily unique, such that the translate CqC-q coincides with the intersection PCP_{C} of PP with 𝔱C\mathfrak{t}^{*}_{C}. for each k0k\geq 0 set BkCB_{k}^{C} to be the cylinder

BkC={xn|xq𝔱C<k},B_{k}^{C}=\{x\in\mathbb{R}^{n}\>|\>||x-q||_{\mathfrak{t}_{C}}<k\},

where ||||𝔱C||\cdot||_{\mathfrak{t}_{C}} denotes the norm of the induced inner product on 𝔱C\mathfrak{t}_{C}. As a shorthand we will denote r(x)=xq𝔱Cr(x)=||x-q||_{\mathfrak{t}_{C}}. Note that, if we set Ωk=P¯BkC\Omega_{k}=\overline{P}\cap B_{k}^{C}, then any point in Ωk\Omega_{k} can be joined to Ωk~\Omega_{\tilde{k}} by a line emanating from qq for any k,k~k,\tilde{k} sufficiently large. Now, uvu_{v} is proper, so we can choose a k10k_{1}\geq 0 sufficiently large such that the set Ωk1\Omega_{k_{1}} contains the unique critical point of uvu_{v}. Let α1=supΩk1uvr+1\alpha_{1}=\sup_{\partial\Omega_{k_{1}}}\frac{\partial u_{v}}{\partial r}+1, noting that this is finite by the choice of Ωk1\Omega_{k_{1}}. Indeed, by construction we have that r\frac{\partial}{\partial r} is tangent to any face of PP, and hence the corresponding quantity supΩk1uPr\sup_{\partial\Omega_{k_{1}}}\frac{\partial u_{P}}{\partial r} for uPu_{P} is finite. Set u~v,1\tilde{u}_{v,1} to be continuous convex function on P¯\overline{P} defined by setting u~v,1=uv\tilde{u}_{v,1}=u_{v} on Ωk1\Omega_{k_{1}} and extending continuously linearly with slope α1\alpha_{1}, i.e.

u~v,1(y)={uv(y)yΩk1uv(πk1(y))+α1r(yπk1(y))yP¯\Ωk1,\tilde{u}_{v,1}(y)=\left\{\begin{array}[]{cl}u_{v}(y)&y\in\Omega_{k_{1}}\\ u_{v}\left(\pi_{k_{1}}(y)\right)+\alpha_{1}r(y-\pi_{k_{1}}(y))&y\in\overline{P}\backslash\Omega_{k_{1}}\end{array}\right.,

where πk1(y)=y(1k1r(y))(yq)\pi_{k_{1}}(y)=y-(1-\frac{k_{1}}{r(y)})(y-q) is the linear projection onto Ωk1\partial\Omega_{k_{1}} relative to the base point qq. Since uu grows faster than linearly in |x||x| by Lemma 3.6, we can choose k2k_{2} sufficiently large such that uu~v,1+1u\geq\tilde{u}_{v,1}+1 on P¯\Ωk2\overline{P}\backslash\Omega_{k_{2}}, infΩk2urα1+1\inf_{\partial\Omega_{k_{2}}}\frac{\partial u}{\partial r}\geq\alpha_{1}+1. We then choose β1=infΩk21\beta_{1}=\inf_{\partial\Omega_{k_{2}}}-1 and set u~1\tilde{u}_{1} to be

u~1(y)={u(πk2(y))β1r(yπk2(y))yΩk2u(y)yP¯\Ωk2.\tilde{u}_{1}(y)=\left\{\begin{array}[]{cl}u\left(\pi_{k_{2}}(y)\right)-\beta_{1}r(y-\pi_{k_{2}}(y))&y\in\Omega_{k_{2}}\\ u(y)&y\in\overline{P}\backslash\Omega_{k_{2}}\end{array}\right..

By construction, u~1(y)u~v,1(y)\tilde{u}_{1}(y)\geq\tilde{u}_{v,1}(y) on Ωk2\partial\Omega_{k_{2}}. As a consequence of the tangent plane property of convexity, the properness of uu, together with the monotonicity of ur\frac{\partial u}{\partial r}, we see that the norm |y||y| (equivalently yC||y||_{C}) of any point satisfying u(πk2(y))β1r(yπk2(y))=uv(πk1(y))+α1r(yπk1(y))u\left(\pi_{k_{2}}(y)\right)-\beta_{1}r(y-\pi_{k_{2}}(y))=u_{v}\left(\pi_{k_{1}}(y)\right)+\alpha_{1}r(y-\pi_{k_{1}}(y)) can be made to strictly increase by sufficiently increasing the value of k2k_{2}. Hence after perhaps making an even larger choice for k2k_{2} we can ensure that the set of points yy such that u~1(y)=u~v,1(y)\tilde{u}_{1}(y)=\tilde{u}_{v,1}(y) lies inside (the closure of) of Ωk2\Ωk1\Omega_{k_{2}}\backslash\Omega_{k_{1}}. Thus, if we set U1=max{u~1,u~v,1}U_{1}=\textnormal{max}\{\tilde{u}_{1},\tilde{u}_{v,1}\}, then U1U_{1} is convex and

U1(x)={uv(x)xΩk1u(x)xP¯\Ωk2.U_{1}(x)=\left\{\begin{array}[]{cl}u_{v}(x)&x\in\Omega_{k_{1}}\\ u(x)&x\in\overline{P}\backslash\Omega_{k_{2}}\end{array}\right..

In particular, if we set w1=U1uw_{1}=U_{1}-u, we see that w1C00(n)w_{1}\in C^{0}_{0}(\mathbb{R}^{n}) has support in Ωk2\Omega_{k_{2}}. Continuing in this way, we produce a sequence of functions wiC00(n)w_{i}\in C^{0}_{0}(\mathbb{R}^{n}) together with a sequence of compact convex sets Ωki\Omega_{k_{i}} such that Ui=u+wiU_{i}=u+w_{i} is convex, wi=vw_{i}=v on Ωi\Omega_{i} and wi=0w_{i}=0 on P¯\Ωi+1\overline{P}\backslash\Omega_{i+1}. Moreover, it follows from the construction that in fact Uimax{u,uv}U_{i}\leq\textnormal{max}\{u,u_{v}\} everywhere.

Now since Ui=uU_{i}=u outside of a compact set, we see that P|Ui|eA𝑑x<\int_{P}|U_{i}|e^{-A}dx<\infty, and consequently UiA1,0U_{i}\in\mathcal{E}^{1,0}_{A}. In order to deduce that limi𝒟(Ui)=𝒟(uv)\lim_{i\to\infty}\mathcal{D}(U_{i})=\mathcal{D}(u_{v}), we first argue that limiPUieA𝑑x=0\lim_{i\to\infty}\int_{P}U_{i}e^{-A}dx=0. For any ε>0\varepsilon>0, let i0i_{0} be sufficiently large such that

|ΩiuveA𝑑x|+|P\ΩiuveA𝑑x|+|P\ΩiueA𝑑x|<ε,\left|\int_{\Omega_{i}}u_{v}e^{-A}dx\right|+\left|\int_{P\backslash\Omega_{i}}u_{v}e^{-A}dx\right|+\left|\int_{P\backslash\Omega_{i}}ue^{-A}dx\right|<\varepsilon,

for all ii0i\geq i_{0}. Clearly we can increase i0i_{0} if necessary to ensure that Ui,u,uv0U_{i},u,u_{v}\geq 0 on P\ΩiP\backslash\Omega_{i} for all ii0i\geq i_{0}. Hence for ii0i\geq i_{0} we have

|PUieA𝑑x||ΩiuveA𝑑x|+|Ωi+1\ΩiUieA𝑑x|+|P\Ωi+1ueA𝑑x|ε+|Ωi+1\ΩiUieA𝑑x|=ε+Ωi+1\ΩiUieA𝑑xε+P\Ωimax{u,uv}eA𝑑x=ε+(P\Ωi){uuv}uveA𝑑x+(P\Ωi){uuv}ueA𝑑xε+P\ΩiueA𝑑x+P\ΩiuveA𝑑x2ε.\begin{split}\left|\int_{P}U_{i}e^{-A}dx\right|&\leq\left|\int_{\Omega_{i}}u_{v}e^{-A}dx\right|+\left|\int_{\Omega_{i+1}\backslash\Omega_{i}}U_{i}e^{-A}dx\right|+\left|\int_{P\backslash\Omega_{i+1}}ue^{-A}dx\right|\\ &\leq\varepsilon+\left|\int_{\Omega_{i+1}\backslash\Omega_{i}}U_{i}e^{-A}dx\right|=\varepsilon+\int_{\Omega_{i+1}\backslash\Omega_{i}}U_{i}e^{-A}dx\\ &\leq\varepsilon+\int_{P\backslash\Omega_{i}}\max\{u,u_{v}\}e^{-A}dx=\varepsilon+\int_{(P\backslash\Omega_{i})\cap\{u\leq u_{v}\}}\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!u_{v}e^{-A}dx+\int_{(P\backslash\Omega_{i})\cap\{u\geq u_{v}\}}\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!ue^{-A}dx\\ &\leq\varepsilon+\int_{P\backslash\Omega_{i}}ue^{-A}dx+\int_{P\backslash\Omega_{i}}u_{v}e^{-A}dx\leq 2\varepsilon.\end{split}

Next, we claim that limine2ϕi𝑑ξ=ne2ϕv𝑑ξ\lim_{i\to\infty}\int_{\mathbb{R}^{n}}e^{-2\phi_{i}}d\xi=\int_{\mathbb{R}^{n}}e^{-2\phi_{v}}d\xi. Once again fix some ε>0\varepsilon>0, and set ϕi=L(Ui)\phi_{i}=L(U_{i}), ϕ=L(u),ϕv=L(uv)\phi=L(u),\phi_{v}=L(u_{v}). By Lemma 2.10, we have that

ϕi(ξ)δ|ξ|supBδ(0)|Ui|=δ|ξ|supBδ(0)uvδ|ξ|C,\phi_{i}(\xi)\geq\delta|\xi|-\sup_{B_{\delta}(0)}|U_{i}|=\delta|\xi|-\sup_{B_{\delta}(0)}u_{v}\geq\delta|\xi|-C,

for some fixed δ>0\delta>0 sufficiently small, and uniformly for all ii sufficiently large. Since ϕv\phi_{v} is proper, perhaps after modifying CC we can ensure that ϕvδ|ξ|C\phi_{v}\geq\delta|\xi|-C for the same choice of δ\delta and CC. Next choose R>0R>0 sufficiently large such that e2Cn\BR(0)e2δ|ξ|𝑑ξ<εe^{2C}\int_{\mathbb{R}^{n}\backslash B_{R}(0)}e^{-2\delta|\xi|}d\xi<\varepsilon, and then i0i_{0} sufficiently large that BR(0)uv(Ωi0)B_{R}(0)\subset\nabla u_{v}(\Omega_{i_{0}}), which we can achieve by Lemma 3.6. Then since ui=uvu_{i}=u_{v} on Ωi\Omega_{i}, it follows that ui=uv\nabla u_{i}=\nabla u_{v} on the interior of Ωi\Omega_{i} and hence ϕi=ϕv\phi_{i}=\phi_{v} on uv(Ωi)\nabla u_{v}(\Omega_{i}). Thus

|ne2ϕi𝑑ξne2ϕv𝑑ξ||uv(Ωi)(e2ϕie2ϕv)𝑑ξ|+n\uv(Ωi)(e2ϕi+e2ϕv)𝑑ξn\BR(0)(e2ϕi+e2ϕv)𝑑ξ2e2Cn\BR(0)e2δ|ξ|𝑑ξ2ε,\begin{split}\left|\int_{\mathbb{R}^{n}}e^{-2\phi_{i}}d\xi-\int_{\mathbb{R}^{n}}e^{-2\phi_{v}}d\xi\right|&\leq\left|\int_{\nabla u_{v}(\Omega_{i})}\left(e^{-2\phi_{i}}-e^{-2\phi_{v}}\right)d\xi\right|+\int_{\mathbb{R}^{n}\backslash\nabla u_{v}(\Omega_{i})}\left(e^{-2\phi_{i}}+e^{-2\phi_{v}}\right)d\xi\\ &\leq\int_{\mathbb{R}^{n}\backslash B_{R}(0)}\left(e^{-2\phi_{i}}+e^{-2\phi_{v}}\right)d\xi\leq 2e^{2C}\int_{\mathbb{R}^{n}\backslash B_{R}(0)}e^{-2\delta|\xi|}d\xi\leq 2\varepsilon,\end{split}

for all ii0i\geq i_{0}. Thus ne2ϕi𝑑ξne2ϕv𝑑ξ\int_{\mathbb{R}^{n}}e^{-2\phi_{i}}d\xi\to\int_{\mathbb{R}^{n}}e^{-2\phi_{v}}d\xi as desired, and finally we conclude that 𝒟(ui)𝒟(uv)\mathcal{D}(u_{i})\to\mathcal{D}(u_{v}). ∎

Theorem 3.11.

Let PP be a polyhedron containing zero in its interior, and suppose that AC(P¯)A\in C^{\infty}(\overline{P}) is admissible. Then up to the action of the linear functions, there is at most one solution uu to (3.4) in 𝒫0\mathcal{P}_{0}.

Proof.

Suppose that we have two solutions u0,u1𝒫0u_{0},u_{1}\in\mathcal{P}_{0}, and let ut=tu1+(1t)u0u_{t}=tu_{1}+(1-t)u_{0}, v=u1u0v=u_{1}-u_{0}. Fix any t(0,1)t\in(0,1). By Lemma 3.10, there exists a sequence of compactly supported functions wiw_{i} such that Ui=u0+wiA1,0U_{i}=u_{0}+w_{i}\in\mathcal{E}^{1,0}_{A} and 𝒟(Ui)𝒟(ut)\mathcal{D}(U_{i})\to\mathcal{D}(u_{t}). By Lemma 3.8 and Proposition 3.9, moreover, we know that 𝒟(Ui)𝒟(u0)\mathcal{D}(U_{i})\geq\mathcal{D}(u_{0}), and therefore by passing to the limit we see that 𝒟(ut)𝒟(u0)\mathcal{D}(u_{t})\geq\mathcal{D}(u_{0}). Of course this is completely symmetric in u0u_{0} and u1u_{1} and independent of the choice of tt, and hence it follows that 𝒟(ut)\mathcal{D}(u_{t}) is minimized at t=0t=0 and t=1t=1. Now let \mathcal{H} denote the space of equivalence classes [u][u] in 𝒫0\mathcal{P}_{0} under the action of n\mathbb{R}^{n} by the addition of linear functions. By Proposition 3.9, 𝒟\mathcal{D} descends to a strictly convex functional on \mathcal{H}, and we have just seen that the convex function of one variable

t𝒟([ut])t\mapsto\mathcal{D}\left([u_{t}]\right)

is minimized at both t=0t=0 and t=1t=1, and hence is constant. Since 𝒟\mathcal{D} is strictly convex, it follows that [u0]=[u1][u_{0}]=[u_{1}]. ∎

4 Proofs of the main theorems

4.1 Preliminaries

Let (M,J)(M,J) be a complex manifold with a fixed effective and holomorphic action of the real torus TnT^{n} with finite fixed point set. Suppose that ω\omega is the Kähler form of a complete shrinking gradient Kähler-Ricci soliton (g,X)(g,X) on MM with JX𝔱JX\in\mathfrak{t}. By [51, Theorem 1.1] it follows that any manifold which admits a complete shrinking Ricci soliton must satisfy H1(M)=0H^{1}(M)=0. It is an immediate consequence that the TnT^{n} action is Hamiltonian with respect to the Kähler form ω\omega of gg. Indeed, let X1,,XnX_{1},\dots,X_{n} be any basis for 𝔱\mathfrak{t}, and θjC(M)\theta_{j}\in C^{\infty}(M) satisfy iXjω=dθj-i_{X_{j}}\omega=d\theta_{j}. Then one defines a moment map explicitly by the formula μ(x)=(θ1,,θn)\mu(x)=\left(\theta_{1},\dots,\theta_{n}\right). There is of course an ambiguity in the choice of each θj\theta_{j} of the addition of a constant. Put together, this corresponds to a translation of the image μ(M)𝔱\mu(M)\subset\mathfrak{t}^{*}. We begin by showing that if we assume that the Ricci curvature of gg is bounded, we can fit this situation into the general framework of the previous sections.

Lemma 4.1.

Let (M,J,ω)(M,J,\omega) be as above, and suppose that gg has bounded Ricci curvature and that JX𝔱JX\in\mathfrak{t}. Then there exists a complexification of the TnT^{n}-action, i.e. an action of ()n(\mathbb{C}^{*})^{n} whose underlying real torus corresponds to the original TnT^{n}-action. Furthermore, there exists an automorphism α\alpha of (M,J)(M,J) such that αg\alpha^{*}g is TnT^{n}-invariant.

To prove this, we make use of the general structure theory for holomorphic vector fields on manifolds admitting Kähler-Ricci solitons from [13]. Let 𝔞𝔲𝔱X\mathfrak{aut}^{X} be the space of holomorphic vector fields commuting with the soliton vector field XX and 𝔤X\mathfrak{g}^{X} be those real holomorphic killing fields commuting with XX.

Theorem 4.2 ([13, Theorem 5.1]).

Let (M,J,g,X)(M,J,g,X) be a complete shrinking gradient Kähler-Ricci soliton with bounded Ricci curvature such that JX𝔱JX\in\mathfrak{t}. Then

𝔞𝔲𝔱X=𝔤XJ𝔤X\mathfrak{aut}^{X}=\mathfrak{g}^{X}\oplus J\mathfrak{g}^{X} (4.1)

Furthermore, 𝔞𝔲𝔱X\mathfrak{aut}^{X} and 𝔤X\mathfrak{g}^{X} are the Lie algebras of finite-dimensional Lie groups AutXAut^{X} and GXG^{X} corresponding to holomorphic automorphisms and holomorphic isometries commuting with the flow of XX.

of Lemma 4.1.

Let (X1,,Xn)(X_{1},\dots,X_{n}) be a basis for 𝔱\mathfrak{t}. Since JX𝔱JX\in\mathfrak{t}, it is clear that [X,Xi]=[X,JXi]=0[X,X_{i}]=[X,JX_{i}]=0 for any ii. In particular, 𝔱𝔞𝔲𝔱X\mathfrak{t}\subset\mathfrak{aut}^{X}. Since the scalar curvature of gg is bounded by assumption, we have by [13, Lemma 2.26] that the zero set of XX is compact. Therefore by [13, Lemma 2.34], it follows that for each ii, XiX_{i} and JXiJX_{i} are complete. In particular, the flow of (Xi,JXi)(X_{i},JX_{i}) determines a unique effective and holomorphic action of \mathbb{C}^{*}. Thus we can complexify the TnT^{n} action, and moreover the corresponding ()n(\mathbb{C}^{*})^{n}-action satisfies 𝔱=𝔱J𝔱𝔞𝔲𝔱X\mathfrak{t}_{\mathbb{C}}=\mathfrak{t}\oplus J\mathfrak{t}\subset\mathfrak{aut}^{X}. Since then XX and JXJX lie in 𝔞𝔲𝔱X\mathfrak{aut}^{X}, we have that the ()n(\mathbb{C}^{*})^{n}-action on MM embeds ()nAutX(\mathbb{C}^{*})^{n}\subset Aut^{X}, and so the real torus Tn()nT^{n}\subset(\mathbb{C}^{*})^{n} lies in some maximal compact subgroup GG of AutXAut^{X}. Since any two maximal compact subgroups of a reductive group are conjugate by Iwasawa’s theorem [34], it follows such that there exists an automorphism α\alpha such that the group GG, and therefore TnT^{n}, preserves the metric αg\alpha^{*}g. ∎

Thus, for the remainder of this section, we assume that (M,J)(M,J) admits an effective holomorphic ()n(\mathbb{C}^{*})^{n}-action with finite fixed point set, and ω\omega is the Kähler form of a complete TnT^{n}-invariant shrinking gradient Kähler-Ricci soliton (g,X)(g,X). In particular, if there is an element b𝔱b\in\mathfrak{t} such that μ,b\langle\mu,b\rangle is proper and bounded from below, then MM is AK-toric by Lemma 2.13. We have by Proposition 2.4 that there exists a potential ϕ\phi for ω\omega on the dense orbit which can be viewed as a smooth strictly convex function on n\mathbb{R}^{n}. We note also that ω\omega is the curvature form of the TnT^{n}-invariant hermitian metric hX=ef(ωn)1h_{X}=e^{-f}(\omega^{n})^{-1} on KM-K_{M}. From (2.4) we know that the soliton potential ff is given by

f=ϕ,bX=μ,bX.f=\langle\nabla\phi,b_{X}\rangle=\langle\mu,b_{X}\rangle.

We have the following from [11].

Proposition 4.3 ([11, Theorem 1.1]).

Let (M,g,f)(M,g,f) be any non-compact complete shrinking gradient Ricci soliton. The soliton potential ff grows quadratically with respect to the distance function dgd_{g} defined by gg, so there is a constant cfc_{f} such that

14(dpcf)2f14(dp+cf)2.\frac{1}{4}(d_{p}-c_{f})^{2}\leq f\leq\frac{1}{4}(d_{p}+c_{f})^{2}.

Therefore bX𝔱b_{X}\in\mathfrak{t} is an element for which the map μ,bX:M\langle\mu,b_{X}\rangle:M\to\mathbb{R} is proper and bounded from below. Thus μ\mu has image equal to a Delzant polyhedron PP by Lemma 2.13, and therefore MM is AK-toric. Let {Di}i=1,,m\{D_{i}\}_{i=1,\dots,m} be the collection of prime, ()n(\mathbb{C}^{*})^{n}-invariant divisors in MM. Since the anticanonical divisor KM-K_{M} of a toric variety is always given by the simple formula [15, Theorem 8.2.3]

KMi=1mDi,-K_{M}\sim\sum_{i=1}^{m}D_{i},

we can apply Corollary 2.16 to obtain:

Lemma 4.4.

Let (M,J)(M,J) be a complex manifold with an effective holomorphic ()n(\mathbb{C}^{*})^{n}-action with finite fixed point set. Suppose that ω\omega is the Kähler form of a complete TnT^{n}-invariant shrinking gradient Kähler-Ricci soliton (g,X)(g,X) on MM. Then the moment map μ\mu has image equal to a Delzant polyhedron PP. In particular, (M,J,ω)(M,J,\omega) is AK-toric and quasiprojective. Let {Di}\{D_{i}\} be the prime, ()n(\mathbb{C}^{*})^{n}-invariant divisors in MM, and let νin𝔱\nu_{i}\in\mathbb{Z}^{n}\subset\mathfrak{t} be minimal generators of the corresponding rays given by the Orbit-Cone correspondence. Then the image PP of μ\mu is equal up to translation to the polyhedron

PKM={x𝔱|νi,x1}P_{-K_{M}}=\left\{x\in\mathfrak{t}^{*}\>|\>\langle\nu_{i},x\rangle\geq-1\right\} (4.2)

determined by the anticanonical bundle.

In particular, the line bundle LPL_{P} of Proposition 2.7 is equal to KM-K_{M}. Clearly, zero lies in the interior the polyhedron PKMP_{-K_{M}} above whenever it is full-dimensional. For simplicity of notation, we will denote P=PKMP=P_{-K_{M}}.

We emphasize that as yet the image of the moment map is fixed only up to translation in 𝔱\mathfrak{t}^{*}. Recall (Lemma 2.11) that the addition of a linear function to the Kähler potential ϕ=ϕ(ξ)\phi=\phi(\xi) on the dense orbit corresponds to a translation of the image of the moment map. We claim that the normalization determined in Proposition 2.6 fixes the moment image uniquely. Thus, it is the real Monge-Ampère equation (2.5) that fixes which translate of P𝔱P\subset\mathfrak{t}^{*} appears. The argument is local, and is based on the observation of Donaldson [20] that the choice of normalization for ϕ\phi determines the behavior of Kähler-Ricci soliton equation (1.4) in symplectic coordinates as xPx\to\partial P.

Lemma 4.5.

Let (M,J,ω)(M,J,\omega) be AK-toric, and suppose that ω\omega is the Kähler form of a complete shrinking gradient Kähler-Ricci soliton on MM. Then, by Proposition 2.6, there exists a unique smooth convex function ϕ\phi on n\mathbb{R}^{n} such that ϕ\phi determines a Kähler potential for ω\omega on the dense orbit via the identification ()nn×Tn(\mathbb{C}^{*})^{n}\cong\mathbb{R}^{n}\times T^{n} and satisfies the real Monge-Ampère equation

detϕij=e2ϕ+bX,ϕ.\det\phi_{ij}=e^{-2\phi+\langle b_{X},\nabla\phi\rangle}.

Then the image of the moment map μ=ϕ\mu=\nabla\phi is precisely the translate of PP given in (4.2). In particular, zero lies in the interior of PP.

Proof.

We know from Lemma 4.4 that the image ϕ(n)\nabla\phi(\mathbb{R}^{n}) is a Delzant polyhedron PP^{\prime}. Suppose that PP^{\prime} is defined by the linear inequalities i(x)ai\ell_{i}(x)\geq-a_{i}, where i(x)=νi,x\ell_{i}(x)=\langle\nu_{i},x\rangle. As we saw in Proposition 2.17, any such ω\omega determines and is determined by a symplectic potential uC(P)u\in C^{\infty}(P), which is unique up to the addition of an affine function. Passing to the Legendre transform, recall that uu satisfies the real Monge-Ampère equation ρu=bX,x\rho_{u}=\langle b_{X},x\rangle, where

ρu(x)=2(uixiu)logdet(uij).\rho_{u}(x)=2\left(u_{i}x^{i}-u\right)-\log\det(u_{ij}).

In particular, ρu\rho_{u} extends smoothly past P\partial P. By Proposition 2.17, there exists a function vv on PP, extending smoothly across P\partial P, such that u=uP+vu=u_{P}+v, where uPu_{P} is defined as in (2.11) by

uP(x)=12(i(x)+ai)log(i(x)+ai).u_{P}(x)=\frac{1}{2}\sum(\ell_{i}(x)+a_{i})\log(\ell_{i}(x)+a_{i}).

Fix any facet FF of PP^{\prime}. We may assume that FF is given by 1(x)=a1\ell_{1}(x)=-a_{1}. Up to a change of basis in 𝔱\mathfrak{t}^{*}, we may also assume by the Delzant condition that 1(x)=x1\ell_{1}(x)=x_{1}. Choose a point pp in the interior of FF. Near pp, uPu_{P} can therefore be written

uP(x)=12(x1+a1)log(x1+a1)+v1,u_{P}(x)=\frac{1}{2}(x_{1}+a_{1})\log(x_{1}+a_{1})+v_{1},

where v1v_{1} extends smoothly across FF. It then follows that in a small half ball BB in the interior of PP^{\prime} containing pp, ρu\rho_{u} can be expressed as

ρu(x)=x1log(x1+a1)(x1+a1)log(x1+a1)+log(x1+a1)+v2,\rho_{u}(x)=x_{1}\log(x_{1}+a_{1})-(x_{1}+a_{1})\log(x_{1}+a_{1})+\log(x_{1}+a_{1})+v_{2},

where v2v_{2} again extends smoothly across FF in BB. It follows that a1=1a_{1}=1. ∎

In the compact case, the condition that MPM\cong\mathcal{M}_{P} for PP given by (4.2) is equivalent to the condition that MM is Fano. We therefore make the following definition.

Definition 10 (c.f. [13, Definition 7.1]).

Let MM be a complex toric manifold. We say that the pair (M,KM)(M,-K_{M}) is anticanonically polarized if MPKMM\cong\mathcal{M}_{P_{-K_{M}}}.

In particular, an anticanonically polarized toric manifold is quasiprojective.

Theorem 4.6.

There exists a unique holomorphic vector field XX with JX𝔱JX\in\mathfrak{t} on an anticanonically polarized AK-toric manifold (M,KM)(M,-K_{M}) which could be the vector field of a complete TnT^{n}-invariant shrinking gradient Kähler-Ricci soliton.

Proof.

Let ω1\omega_{1} and ω2\omega_{2} be two TnT^{n}-invariant Kähler metrics on MM satisfying (1.4) on MM with vector fields X1X_{1} and X2X_{2}. By Lemma 4.5, we know that each moment map μs\mu_{s}, s=1,2s=1,2, has image equal to P=PKMP=P_{-K_{M}}. Moreover, by Lemma 4.5, we know that ωs\omega_{s} is uniquely determined by a symplectic potential usu_{s} on the fixed polyhedron P=PKMP=P_{-K_{M}} which satisfies the real Monge-Ampère equation ρus=bs,x\rho_{u_{s}}=\langle b_{s},x\rangle. By Lemma 3.5, the function bs,x\langle b_{s},x\rangle satisfies

P(x)ebs,x𝑑x=0\int_{P}\ell(x)e^{-\langle b_{s},x\rangle}dx=0

for each linear function (x)\ell(x) on PP. In particular, bs,x\langle b_{s},x\rangle is equal to the fixed linear function P\ell_{P} determined in Proposition 3.1. Clearly, there is a unique bP𝔱b_{P}\in\mathfrak{t} such that P(x)=bP,x\ell_{P}(x)=\langle b_{P},x\rangle. Let XPX_{P} be the holomorphic vector field on MM which is given by

XP1,0=j=1nbPjzjzjX_{P}^{1,0}=\sum_{j=1}^{n}b_{P}^{j}z_{j}\frac{\partial}{\partial z_{j}}

on the dense orbit. We have in particular that XPωs=Xsωs\mathcal{L}_{X_{P}}\omega_{s}=\mathcal{L}_{X_{s}}\omega_{s}. Since ωs\omega_{s} is TnT^{n}-invariant and JXP,JX1,JX2𝔱JX_{P},JX_{1},JX_{2}\in\mathfrak{t}, this immediately implies that X1=X2=XPX_{1}=X_{2}=X_{P}. ∎

4.2 Proofs of Theorem A and Theorem B

We begin with the proof of Theorem A. Suppose that ω1\omega_{1} and ω2\omega_{2} are two complete TnT^{n}-invariant Kähler metrics on MM satisfying (1.4). By Theorem 4.6, the soliton vector fields are given by X1=X2=XPX_{1}=X_{2}=X_{P}. Recall from the proof of Theorem 4.6 we know that each ωs\omega_{s} is determined uniquely by a symplectic potential usu_{s} on the fixed polyhedron PP. Each usu_{s} itself is unique up to the addition of an affine function, and satisfies the real Monge-Ampère equation

ρus=bP,x,\rho_{u_{s}}=\langle b_{P},x\rangle, (4.3)

where bP𝔱b_{P}\in\mathfrak{t} is the element determining XPX_{P} as in the proof of Theorem 4.6. If we set

A(x)=bP,x,A(x)=\langle b_{P},x\rangle,

then equation (4.3) takes the form ρ=A\rho=A with respect to the fixed function AA on PP. Thus, we are in the setting of Section 3.3. We would then like to apply the uniqueness theorem Theorem 3.11 to conclude that usu_{s} are related via the addition of an affine function. We need to show therefore that AA is admissible and that PueA𝑑x<\int_{P}ue^{-A}dx<\infty, so that usu_{s} lies in the space of symplectic potentials 𝒫\mathcal{P} defined by AA. To see that AA is admissible, first note that by Lemma 3.5, we have

PeA𝑑x=0,\int_{P}\ell e^{-A}dx=0,

which is condition (2)(2) from Section 3.3. Since, by Proposition 4.3,

Mefωn<,\int_{M}e^{-f}\omega^{n}<\infty,

we have that

Mefωn=n×TnebP,ϕdet(ϕij)dξdθ=(2π)nPeA𝑑x.\int_{M}e^{-f}\omega^{n}=\int_{\mathbb{R}^{n}\times T^{n}}e^{-\langle b_{P},\nabla\phi\rangle}\det(\phi_{ij})d\xi d\theta=(2\pi)^{n}\int_{P}e^{-A}dx.

This implies that

PeA𝑑x<,\int_{P}e^{-A}dx<\infty, (4.4)

which is condition (1)(1). Furthermore, from (4.4) it follows from Corollary 3.2 that bPCb_{P}\in C^{*}, and in particular A(x)=O(|x|)A(x)=O(|x|). Since uP=O(|x|log|x|)u_{P}=O(|x|\log|x|) we then have

PuPeA𝑑x<,\int_{P}u_{P}e^{-A}dx<\infty,

which is condition (3)(3). Thus AA is admissible, and it remains only to show that each PuseA<\int_{P}u_{s}e^{-A}<\infty. This follows from an elementary calculation.

Lemma 4.7 (c.f. [20, Lemma 1]).

Let PP be a polyhedron containing zero in the interior and uC(P)u\in C^{\infty}(P) be any strictly convex function such that the gradient u\nabla u maps PP diffeomorphically onto n\mathbb{R}^{n}. Then

Pueρu𝑑x<.\int_{P}ue^{-\rho_{u}}dx<\infty.
Proof.

Let ϕ(ξ)=L(u)\phi(\xi)=L(u). Recall that by Lemma 2.10, ϕ\phi grows at least linearly in |ξ||\xi|, and in particular is necessarily bounded from below. Then

Pueρu𝑑x=n(ϕ,ξϕ)e2ϕ𝑑ξn(ϕ,ξ+C)e2ϕ𝑑ξ.\int_{P}ue^{-\rho_{u}}dx=\int_{\mathbb{R}^{n}}\left(\langle\nabla\phi,\xi\rangle-\phi\right)e^{-2\phi}d\xi\leq\int_{\mathbb{R}^{n}}\left(\langle\nabla\phi,\xi\rangle+C\right)e^{-2\phi}d\xi.

The second term Ce2ϕ𝑑ξ\int Ce^{-2\phi}d\xi is bounded again by Lemma 2.10, so that

Pueρu𝑑xnϕ,ξe2ϕ𝑑ξ+C.\int_{P}ue^{-\rho_{u}}dx\leq\int_{\mathbb{R}^{n}}\langle\nabla\phi,\xi\rangle e^{-2\phi}d\xi+C.

In polar coordinates we have

nϕ,ξe2ϕ𝑑ξ=Sn10rnϕre2ϕ𝑑r𝑑Θ.\int_{\mathbb{R}^{n}}\langle\nabla\phi,\xi\rangle e^{-2\phi}d\xi=\int_{S^{n-1}}\int_{0}^{\infty}r^{n}\frac{\partial\phi}{\partial r}e^{-2\phi}drd\Theta.

Integrating by parts, we obtain

Sn10rnϕre2ϕ𝑑r𝑑Θ=n2Sn10rn1e2ϕ𝑑r𝑑Θ=nne2ϕ𝑑ξ.\int_{S^{n-1}}\int_{0}^{\infty}r^{n}\frac{\partial\phi}{\partial r}e^{-2\phi}drd\Theta=\frac{n}{2}\int_{S^{n-1}}\int_{0}^{\infty}r^{n-1}e^{-2\phi}drd\Theta=n\int_{\mathbb{R}^{n}}e^{-2\phi}d\xi.

Note that the boundary term converges since ϕ=O(r)\phi=O(r) as rr\to\infty. Thus

Pueρu𝑑xnϕ,ξe2ϕ𝑑ξ+C=nne2ϕ𝑑ξ+C<.\int_{P}ue^{-\rho_{u}}dx\leq\int_{\mathbb{R}^{n}}\langle\nabla\phi,\xi\rangle e^{-2\phi}d\xi+C=n\int_{\mathbb{R}^{n}}e^{-2\phi}d\xi+C<\infty.

Since each usu_{s} satisfies ρus=A\rho_{u_{s}}=A, Lemma 4.7 states that

PueA𝑑x<.\int_{P}ue^{-A}dx<\infty. (4.5)

Each usu_{s} is strictly convex on PP, and by Proposition 2.17 there exists for each usu_{s} a smooth function vsC(P¯)v_{s}\in C^{\infty}(\overline{P}) such that us=uP+vsu_{s}=u_{P}+v_{s}. Strict convexity of usu_{s} along with (4.5) then imply that us𝒫u_{s}\in\mathcal{P}, and so by Theorem 3.11 it follows that there is an affine function a(x)=ba,x+ca(x)=\langle b_{a},x\rangle+c such that u2=u1+au_{2}=u_{1}+a. Let ϕs=L(us)\phi_{s}=L(u_{s}) be the Legendre transform, so that ωs=2i¯ϕs(ξ)\omega_{s}=2i\partial\bar{\partial}\phi_{s}(\xi) on the dense orbit. As we have seen in Lemma 2.11, it follows that ϕ2(ξ)=ϕ1(ξba)c\phi_{2}(\xi)=\phi_{1}(\xi-b_{a})-c, so that 2i¯ϕ2(ξ)=2i¯ϕ1(ξba)2i\partial\bar{\partial}\phi_{2}(\xi)=2i\partial\bar{\partial}\phi_{1}(\xi-b_{a}). Let α:MM\alpha:M\to M denote the automorphism determined by the action of eba()ne^{-b_{a}}\in(\mathbb{C}^{*})^{n}. Then it is clear that ϕ1(ξba)=ϕ1α(ξ)\phi_{1}(\xi-b_{a})=\phi_{1}\circ\alpha(\xi), and therefore that ω2=αω1\omega_{2}=\alpha^{*}\omega_{1}. This concludes the proof of Theorem A.

Theorem B follows immediately from Lemma 4.1, Lemma 4.4, and Theorem A.

4.3 Proof of Corollary D

Recalling the setting, let NN be an (n1)(n-1)-dimensional compact toric Fano manifold, and LNL\to N satisfy Lp=KNL^{p}=K_{N} for 0<p<n0<p<n. By Theorem B, it suffices to show that the metrics have bounded Ricci curvature and that the corrsponding soliton vector fields satisfy JX𝔱JX\in\mathfrak{t}. We first observe that the total space of LL admits an effective and holomorphic ()n(\mathbb{C}^{*})^{n}-action by augmenting the (n1)(n-1)-dimensional action on NN with the natural \mathbb{C}^{*}-action acting on the fibers of LL. It was shown in [28] that the cone formed by contracting the zero section on LL admits a Ricci-flat Kähler cone metric ωRF=i2¯r~2\omega_{RF}=\frac{i}{2}\partial\bar{\partial}\tilde{r}^{2} with Reeb vector field Jr~r~=K𝔱J\tilde{r}\frac{\partial}{\partial\tilde{r}}=K\in\mathfrak{t}. Futaki’s construction begins by deforming ωRF\omega_{RF} to what’s called a Sasaki η\eta-Einstein metric by a choice of reparameterization of the radial function r~r=r~a\tilde{r}\mapsto r=\tilde{r}^{a} for some a>0a>0 (here η=dclogr\eta=d^{c}\log r refers to the contact 1-form associated to the Sasakian structure). Set ω=i2¯r2\omega=\frac{i}{2}\partial\bar{\partial}r^{2} to be this choice and set t=logrt=\log r and ωT=i¯t\omega_{T}=i\partial\bar{\partial}t. Then the metric ωKRS\omega_{KRS} is chosen via the momentum construction (or Calabi Ansatz), and thus splits orthogonally as

ωKRS=ωT+i¯H(t)=(1+τ)ωT+φ(τ)dtdct,\begin{split}\omega_{KRS}=\omega_{T}+i\partial\bar{\partial}H(t)=(1+\tau)\omega_{T}+\varphi(\tau)dt\wedge d^{c}t,\end{split}

where HH is a smooth convex function of one variable, τ=H(t)\tau=H^{\prime}(t), φ(τ)=H′′(t)\varphi(\tau)=H^{\prime\prime}(t). Here τ(0,)\tau\in(0,\infty) and τ0\tau\to 0 corresponds to approaching the zero section of LL whereas τ\tau\to\infty goes off to infinity along the complete end. We refer to [28, 29, 27] (see also [25, 33]) for more details on this construction. In particular, the soliton vector field satisfies JX=rr𝔱JX=r\frac{\partial}{\partial r}\in\mathfrak{t}.

To see that the Ricci curvature of ωKRS\omega_{KRS} is bounded, we use the explicit form [27, Claim 4.4] of φ\varphi

φ(τ)=(κ2)μ(1+τ)+κ2κnμn+1j=0n1n!j!μj(1+τ)j(n1),\varphi(\tau)=\frac{(\kappa-2)}{\mu}(1+\tau)+\frac{\kappa-2-\frac{\kappa}{n}}{\mu^{n+1}}\sum_{j=0}^{n-1}\frac{n!}{j!}\mu^{j}(1+\tau)^{j-(n-1)},

where κ>2,μ>0\kappa>2,\mu>0 are constants determined by the soliton equation. So φ\varphi is a rational function and one sees immediately that φ=O(1+τ),φ=O(1),φ′′=O((1+τ)3)\varphi=O(1+\tau),\varphi^{\prime}=O(1),\varphi^{\prime\prime}=O((1+\tau)^{-3}) as τ\tau\to\infty. Moreover, the Ricci form is also explicit ([27, Equation 3.8])

RicωKRS=(κ((n1)φ1+τ+φ))ωT((n1)φ1+τ+φ)dtdct.\text{Ric}_{\omega_{KRS}}=\left(\kappa-\left(\frac{(n-1)\varphi}{1+\tau}+\varphi^{\prime}\right)\right)\omega_{T}-\left(\frac{(n-1)\varphi}{1+\tau}+\varphi^{\prime}\right)^{\prime}dt\wedge d^{c}t.

Thus we read off that RicωKRS=O(1)ωT+O((1+τ)2)dtdct\text{Ric}_{\omega_{KRS}}=O(1)\omega_{T}+O((1+\tau)^{-2})dt\wedge d^{c}t, whereas the metric ωKRS=O(1+τ)ωT+O(1+τ)dtdct\omega_{KRS}=O(1+\tau)\omega_{T}+O(1+\tau)dt\wedge d^{c}t, from which we see that RicωKRSωKRS||\text{Ric}_{\omega_{KRS}}||_{\omega_{KRS}} actually decays as τ\tau\to\infty.

\square

4.4 Example: 1×\mathbb{C}\mathbb{P}^{1}\times\mathbb{C}

Choose homogeneous coordinates [w1:w2][w_{1}:w_{2}] on 1\mathbb{C}\mathbb{P}^{1}, and let w=w1w2w=\frac{w_{1}}{w_{2}}. We let \mathbb{C}^{*} act on 1\mathbb{C}\mathbb{P}^{1} by λ[w1:w2]=[λw1:w2]\lambda\cdot[w_{1}:w_{2}]=[\lambda w_{1}:w_{2}], which gives 1\mathbb{C}\mathbb{P}^{1} the structure of a toric variety. Let ωFS\omega_{FS} be the Fubini-Study metric associated to [w1:w2][w_{1}:w_{2}]. Let zz be a holomorphic coordinate on \mathbb{C} and ωE\omega_{E} denote the Euclidean metric. If \mathbb{C}^{*} acts on \mathbb{C} in the standard way, then we obtain an effective algebraic action of ()2(\mathbb{C}^{*})^{2} on 1×\mathbb{C}\mathbb{P}^{1}\times\mathbb{C}. The product metric ωstd=ωFS+ωE\omega_{\text{std}}=\omega_{FS}+\omega_{E} on 1×\mathbb{C}\mathbb{P}^{1}\times\mathbb{C} is then a complete T2T^{2}-invariant shrinking gradient Kähler-Ricci soliton with respect to the holomorphic vector field zzz\frac{\partial}{\partial z} (here we suppress the obvious pullbacks). As an application of the results of the previous sections, we show that, up to isometry, this is the unique shrinking gradient Kähler-Ricci soliton on 1×\mathbb{CP}^{1}\times\mathbb{C} with bounded scalar curvature.

Corollary C.

Any complete shrinking gradient Kähler-Ricci soliton (g,X)(g,X) on M=1×M=\mathbb{C}\mathbb{P}^{1}\times\mathbb{C} with bounded scalar curvature is isometric to to the standard product metric ωstd\omega_{\text{std}}.

By the work of [39], in real dimension four we know that the scalar curvature controls the full curvature tensor for shrinking solitons. In particular, it follows from [39, Theorem 1.3] that any such (g,X)(g,X) as above has bounded Ricci curvature. Fix a background product coordinate system ([w1:w2],z)([w_{1}:w_{2}],z) on M1×M\cong\mathbb{CP}^{1}\times\mathbb{C} as above. In what follows, we will ignore the standard ()2(\mathbb{C}^{*})^{2}-action determined by this choice, but we will routinely make use of the corresponding projection onto the \mathbb{C}-factor, which we denote by π:M\pi:M\to\mathbb{C}. Corollary C then follows from Theorem B as soon as we have the following lemma.

Proposition 4.8.

Let (g,X)(g,X) be any complete shrinking gradient Kähler-Ricci soliton on M=1×M=\mathbb{C}\mathbb{P}^{1}\times\mathbb{C} with bounded scalar curvature, and let T()2T\subset(\mathbb{C}^{*})^{2} be the real torus corresponding to the standard ()2(\mathbb{C}^{*})^{2}-action on MM with Lie algebra 𝔱\mathfrak{t}. Then there exists a holomorphic automorphism α\alpha of MM such that J(αX)𝔱J(\alpha^{*}X)\in\mathfrak{t}.

The proof of this proposition will take up the remainder of this section. Let ff denote the soliton potential so that the soliton vector field X=gfX=\nabla_{g}f. As before (c.f. Lemma 4.1), we define GXG^{X} to be the of the group holomorphic isometries of (M,J,g)(M,J,g) that commute with the flow of XX, and we let G0XG^{X}_{0} be the connected component of the identity in GXG^{X}. Then G0XG^{X}_{0} is a compact Lie group by [13, Lemma 5.12]. Clearly the flow of JXJX defines a one-parameter subgroup in G0XG^{X}_{0}, and so the closure in G0XG^{X}_{0} is a real torus TXT^{X} of holomorphic isometries of gg. Let M0M_{0} denote the zero set of XX. Since the scalar curvature is bounded, it follows from [13, Lemma 2.26] that M0M_{0} is a compact analytic subvariety of MM, and hence is equal to a finite collection of points in MM and curves Lz=1×{z}ML_{z}=\mathbb{CP}^{1}\times\{z\}\subset M. Note that the fixed point set of TXT^{X} is equal to M0M_{0}. By Lemma 4.1, there exists a complexification TXAutXT^{X}_{\mathbb{C}}\subset\text{Aut}^{X} of TXT^{X}, which is a complex torus with dimTX=dimTX\dim_{\mathbb{C}}T^{X}_{\mathbb{C}}=\dim_{\mathbb{R}}T^{X}. In what follows we will need to treat the the two possible cases, dimTX=1\dim_{\mathbb{R}}T^{X}=1 and dimTX=2\dim_{\mathbb{R}}T^{X}=2, separately. For the moment, we make no distinction.

We first study M0M_{0}, making use of the fact that ff is a Morse-Bott function on MM [26]. Since MM is Kähler we have moreover that the Morse indices of any critical point must be even. Since M0M_{0} consists of the critical points of ff, we can write

M0=M(0)M(2)M(4),M_{0}=M^{(0)}\cup M^{(2)}\cup M^{(4)},

where M(i)M^{(i)} denotes the connected component with Morse index ii. By [13, Claim 2.30], we know that M(0)M^{(0)} is a nonempty, compact, and connected analytic subvariety of MM, and therefore must either be equal to a single projective line LzL_{z} or an isolated point. We begin with a construction which will be used throughout the rest of the section.

Claim 4.9.

Suppose that xx is a point in M(2)M(4)M^{(2)}\cup M^{(4)}. Then there exists a holomorphic map Rx:1MR_{x}:\mathbb{CP}^{1}\to M with Rx(0)=xR_{x}(0)=x and Rx()M0R_{x}(\infty)\in M_{0} defined by the negative gradient flow of ff. Since MM is a trivial 1\mathbb{CP}^{1}-fibration, the image of RxR_{x} must lie in the unique fiber LzL_{z} of π\pi containing xx.

Proof.

By [9, Proposition 6] there exists a local holomorphic coordinate system (z1,z2)(z_{1},z_{2}) centered at xx such that the holomorphic vector field X1,0=12(XiJX)X^{1,0}=\frac{1}{2}(X-iJX) is given by

X1,0=a1z1z1+a2z2z2X^{1,0}=a_{1}z_{1}\frac{\partial}{\partial z_{1}}+a_{2}z_{2}\frac{\partial}{\partial z_{2}} (4.6)

for a1,a2a_{1},a_{2}\in\mathbb{R}. By assumption, Hess(f)g{}_{g}(f) has at least one negative eigenvalue at xx, and therefore we can assume without loss of generality that a2<0a_{2}<0. Then JXJX is tangential to the z2z_{2}-axis, and the flow of JXJX here is given by regular periodic orbits. We fix any such nontrivial orbit θ:S1M\theta:S^{1}\to M. If we let ψt:MM\psi_{t}:M\to M denote the flow of X=gf-X=-\nabla_{g}f, then we define a holomorphic map r:S1×Mr:\mathbb{C}^{*}\cong S^{1}\times\mathbb{R}\to M by r(s,t)=ψt(θ(s))r(s,t)=\psi_{t}(\theta(s)). It follows immediately from the local form (4.6) that rr extends to a holomorphic map r:Mr:\mathbb{C}\to M with rx(0)=xr_{x}(0)=x. Now ff is bounded from below and decreases along its negative gradient flow, and therefore ff is bounded along the image of rxr_{x}. Since ff is proper, this implies that the image of rxr_{x} lies in the compact set f1((,a])f^{-1}((-\infty,a]), where a=supfrxa=\sup f\circ r_{x}. If π:M\pi:M\to\mathbb{C} denotes the projection onto the second factor of M=1×M=\mathbb{CP}^{1}\times\mathbb{C}, then πrx:\pi\circ r_{x}:\mathbb{C}\to\mathbb{C} is therefore bounded and hence constant. Thus, πrx()=z\pi\circ r_{x}(\mathbb{C})=z for some fixed zz\in\mathbb{C}, so that the image of rxr_{x} lies in Lz=π1(z)L_{z}=\pi^{-1}(z). For each fixed sS1s\in S^{1}, we have by [13, Proposition 2.28] a well-defined limit limtψt(θ(s))\lim_{t\to\infty}\psi_{t}(\theta(s)), also lying in M0M_{0}. In this case, the limits must all coincide with the unique point p=Lz\rx()p=L_{z}\backslash r_{x}(\mathbb{C}). Thus, there is a well-defined holomorphic extension of Rx:1MR_{x}:\mathbb{CP}^{1}\to M of rxr_{x} with Rx()=pR_{x}(\infty)=p. ∎

4.4.1 Case 1: M(0)M^{(0)} is an isolated point

Claim 4.10.

Let yy be any point in M(2)M(4)M^{(2)}\cup M^{(4)}. Let Ry:1MR_{y}:\mathbb{CP}^{1}\to M be a holomorphic map with Ry(0)=yR_{y}(0)=y and Ry()M0R_{y}(\infty)\in M_{0}, which must exist by Claim 4.9. Then Ry()M(0)R_{y}(\infty)\in M^{(0)}.

Proof.

Set p=Ry()p=R_{y}(\infty), and assume without loss of generality that z=0z=0, so that the image of RyR_{y} is the fiber L0=π1(0)L_{0}=\pi^{-1}(0) of π\pi. If pM(4)p\in M^{(4)}, then we choose coordinates centered at pp in which X1,0X^{1,0} takes the form (4.6). This immediately yields a contradiction, since pp is defined as the forward limit point of a flow line of X-X. If both aia_{i} are negative, then no forward flow of X-X near pp converges to pp. Thus, either pM(0)p\in M^{(0)} or pM(2)p\in M^{(2)}. Since L0L_{0} is the image of the map RyR_{y} defined by the flow of (X,JX)(X,JX), it follows that XX is tangential to L0L_{0}. In particular, the restriction X|L0X|_{L_{0}} is a well-defined holomorphic vector field on L0L_{0} and does not vanish identically since the map RyR_{y} is non-constant. It follows that M0L0M_{0}\cap L_{0} consists only of the isolated points xx and pp, and that pp is the point in L0L_{0} at which ff attains its minimum value among all points in L0L_{0}. Suppose that pM(2)p\in M^{(2)}. Then by Claim 4.9, there is a holomorphic embedding Rp:1MR_{p}:\mathbb{CP}^{1}\to M with rp(0)=pr_{p}(0)=p, defined by the negative gradient flow of ff. Thus, once again, the image of RpR_{p} must be equal to L0L_{0}. This is a contradiction, since ff decreases along its negative gradient flow and f(p)=minL0ff(p)=\min_{L_{0}}f. ∎

Claim 4.11.

If we assume that M(0)={p}M^{(0)}=\{p\}, then M0M_{0} lies in a fixed fiber L0L_{0} of π\pi, and consists precisely of the two isolated points M0={x}{p}M_{0}=\{x\}\cup\{p\} with xM(2)x\in M^{(2)}.

Proof.

In this case we have from [9] that M(2)M(4)M^{(2)}\cup M^{(4)} must indeed be nonempty or else M2M\cong\mathbb{C}^{2}, which is clearly a contradiction. Let xM(2)M(4)x\in M^{(2)}\cup M^{(4)} be one such point. By Claim 4.10, there is a map Rx:1MR_{x}:\mathbb{CP}^{1}\to M with Rx(0)=xR_{x}(0)=x and Rq()=pM(0)R_{q}(\infty)=p\in M^{(0)}. In particular, π(x)=π(p)\pi(x)=\pi(p). Suppose that there is another point qM0q\in M_{0} not equal to pp or xx. Then again by Claim 4.10 there is a map Rq:1MR_{q}:\mathbb{CP}^{1}\to M with Rq(0)=qR_{q}(0)=q and Rq()=pM(0)R_{q}(\infty)=p\in M^{(0)}. Thus Rq(1)=L0R_{q}(\mathbb{CP}^{1})=L_{0}, which means in particular that qL0q\in L_{0}. This is a contradiction, since qpq\neq p and qxq\neq x, and a holomorphic vector field on 1\mathbb{CP}^{1} which vanishes at three distinct points must vanish identically. Finally, we claim that the point xM(2)x\in M^{(2)}. If not, then xM(4)x\in M^{(4)}, and both coefficients aia_{i} in the representation (4.6) for XX centered at xx are negative. Thus, there is a distinct holomorphic curve Rx:1MR^{\prime}_{x}:\mathbb{CP}^{1}\to M with Rx(0)=xR^{\prime}_{x}(0)=x, intersecting Rx(1)R_{x}(\mathbb{CP}^{1}) transversely at xx. This is impossible, so we obtain our contradiction. ∎

In particular, we have shown that if M(0)={p}M^{(0)}=\{p\}, then the fixed point set of TXT^{X} is finite. If TXT^{X} is two-dimensional, then TXT^{X} together with the Kähler form ω\omega of gg give MM the structure of a symplectic toric manifold. We are therefore in the setting of the previous sections, and we can deduce Proposition 4.8 from the results there.

Claim 4.12.

Suppose that TXT^{X} is contained in a two-dimensional real torus 𝕋\mathbb{T} acting on MM by holomorphic isometries of ω\omega. Then there exists an equivariant biholomorphism α:M1×\alpha:M\to\mathbb{CP}^{1}\times\mathbb{C}, where 1×\mathbb{CP}^{1}\times\mathbb{C} is endowed with the standard ()2(\mathbb{C}^{*})^{2}-action.

Proof.

As we have seen in Section 4, the fact that ω\omega is the Kähler form of a complete shrinking gradient Kähler-Ricci soliton on MM implies automatically that the 𝕋\mathbb{T}-action is Hamiltonian. Since dim𝕋=2=dimM\dim_{\mathbb{C}}\mathbb{T}_{\mathbb{C}}=2=\dim_{\mathbb{C}}M and the fixed point set is finite, we can apply Lemma 4.4 to deduce that the image of the moment map μ\mu is a Delzant polyhedron PP in Lie(𝕋)\text{Lie}(\mathbb{T})^{*}. Then Lemma 2.14 implies that there exists an equivariant biholomorphism α:(M,J)(MP,JP)\alpha:(M,J)\to(M_{P},J_{P}), where (MP,JP,ωP)(M_{P},J_{P},\omega_{P}) is the AK-toric manifold of Proposition 2.7. By Proposition 2.8, MPM_{P} is equivariantly biholomorphic to the unique algebraic toric variety P\mathcal{M}_{P} associated to PP. It follows that the underlying complex structure of MPM_{P} is biholomorphic to 1×\mathbb{CP}^{1}\times\mathbb{C}. Since the topology of an algebraic toric variety is uniquely characterized by its fan (c.f. [15, Chapter 12]), the only algebraic toric variety with this property is 1×\mathbb{CP}^{1}\times\mathbb{C} with the standard ()2(\mathbb{C}^{*})^{2}-action up to equivariant isomorphism. Thus, α\alpha is the required biholomorphsim α:M1×\alpha:M\to\mathbb{CP}^{1}\times\mathbb{C}. ∎

In particular, if TXT^{X} itself is two-dimensional and M(0)={p}M^{(0)}=\{p\}, then TXT^{X} itself satisfies the hypotheses of Claim 4.12, and we can simply take 𝕋=TX\mathbb{T}=T^{X}. In fact, even when dimTX=1\dim_{\mathbb{R}}T^{X}=1, we can always find a two dimensional torus 𝕋\mathbb{T} satisfying the hypotheses of Claim 4.12.

Claim 4.13.

If M(0)={p}M^{(0)}=\{p\}, then there exists a two-dimensional torus 𝕋\mathbb{T} of biholomorphisms acting on MM such that TX𝕋T^{X}\subset\mathbb{T}.

Proof.

If TXT^{X} is two-dimensional, then there is nothing to prove. Therefore, we can assume that TXT^{X}_{\mathbb{C}} defines an action of \mathbb{C}^{*} on MM. Recall that π\pi denotes the projection π:M\pi:M\to\mathbb{C} under a fixed identification M1×M\cong\mathbb{CP}^{1}\times\mathbb{C}. Let ϖ:M1\varpi:M\to\mathbb{CP}^{1} denote the other projection. Then the (1,0)(1,0) tangent bundle TM1,0T^{1,0}_{M} of MM splits holomorphically as TM1,0ϖT11,0πT1,0T^{1,0}_{M}\cong\varpi^{*}T^{1,0}_{\mathbb{CP}^{1}}\oplus\pi^{*}T^{1,0}_{\mathbb{C}}. In particular, there exist holomorphic projection maps onto the subbundles ϖT11,0\varpi^{*}T^{1,0}_{\mathbb{CP}^{1}} and πT1,0\pi^{*}T^{1,0}_{\mathbb{C}} of TM1,0T^{1,0}_{M}. We can therefore write X1,0=V1,0+W1,0X^{1,0}=V^{1,0}+W^{1,0}, where V1,0,W1,0V^{1,0},W^{1,0} are holomorphic vector fields lying in ϖT11,0\varpi^{*}T^{1,0}_{\mathbb{CP}^{1}} and πT1,0\pi^{*}T^{1,0}_{\mathbb{C}} respectively.

Notice that the coordinate zz on \mathbb{C} defines a global holomorphic coordinate on MM. Since T1,0T^{1,0}_{\mathbb{C}} is trivial, we can write the vector field W1,0=fzW^{1,0}=f\frac{\partial}{\partial z}, where ff is a holomorphic function on M. Now since X1,0X^{1,0} generates \mathbb{C}^{*}-action on MM, W1,0W^{1,0} also generates a \mathbb{C}^{*}-action on \mathbb{C}. In particular, f=f(z)f=f(z) depends only on zz. Now X1,0X^{1,0} is tangential to L0L_{0}, this action fixes 00\in\mathbb{C}. Since the automorphism group of \mathbb{C} consists of linear transformations, it follows that f(z)f(z) is of the form f(z)=kzf(z)=kz.

For each zz\in\mathbb{C}, the restriction of V1,0V^{1,0} to LzL_{z} is a holomrophic vector field on Lz1L_{z}\cong\mathbb{CP}^{1} which we denote by Vz1,0V_{z}^{1,0}. A nonzero holomorphic vector field on 1\mathbb{CP}^{1} vanishes at two points with multiplicity, so that Vz1,0V_{z}^{1,0} either vanishes identically or has zero set equal to a degree 2 divisor in 1\mathbb{CP}^{1}. Recall that X1,0X^{1,0} is tangential to L0L_{0}, and so V01,0V_{0}^{1,0} vanishes only at M0M_{0}, which consists of the two isolated points {x}\{x\} and {p}\{p\}. Thus, by the continuity of the map H0(1,𝒪(2))\mathbb{C}\to H^{0}(\mathbb{CP}^{1},\mathcal{O}(2)) given by zVz1,0z\mapsto V^{1,0}_{z}, the same is true for Vz1,0V_{z}^{1,0} with |z||z| sufficiently small. In particular, there exists a small neighborhood Δ\Delta\subset\mathbb{C} of 0 such that the zero set of Vz1,0V_{z}^{1,0} for zΔz\in\Delta consists of disjoint embedded discs Δp,ΔxM\Delta_{p},\Delta_{x}\subset M, centered at pp and xx respectively, each meeting a given fiber LzL_{z} at a unique point. Let {pz}=ΔpLz\{p_{z}\}=\Delta_{p}\cap L_{z} and {xz}=ΔxLz\{x_{z}\}=\Delta_{x}\cap L_{z}. Let y0L0y_{0}\in L_{0} be a point which does not lie in M0M_{0}, and let Φt\Phi_{t} denote the flow of W1,0W^{1,0}. Since W1,0=kzzW^{1,0}=kz\frac{\partial}{\partial z}, clearly there exists a point yML0y\in M-L_{0} such that the orbit WW-orbit Φt(y)\Phi_{t}(y) of yy under the flow of W1,0W^{1,0} converges to y0y_{0} as t0t\to 0. Let CyMC^{y}\subset M be the closure of the orbit Φt(y)\Phi_{t}(y) and let Δy\Delta_{y} denote the intersection of CyC^{y} with 1×Δ\mathbb{CP}^{1}\times\Delta. Again since WW takes this special form, and perhaps after shrinking Δ\Delta, we can choose y0y_{0} such that Δy\Delta_{y} does not intersect ΔpΔx\Delta_{p}\cup\Delta_{x}. We denote the unique point of ΔyLz\Delta_{y}\cap L_{z} by yzy_{z}. Then there is a unique automorphism AzPGL(2,)A_{z}\in\text{PGL}(2,\mathbb{C}) of 1\mathbb{CP}^{1} such that Az(xz)=A_{z}(x_{z})=\infty, Az(yz)=1A_{z}(y_{z})=1, and Az(pz)=0A_{z}(p_{z})=0. Then we define an automorphism α1:1×Δ1×Δ\alpha_{1}:\mathbb{CP}^{1}\times\Delta\to\mathbb{CP}^{1}\times\Delta by setting α1(,z)=(Az1(),z)\alpha_{1}(\ell,z)=\left(A_{z}^{-1}(\ell),z\right). After changing coordinates on 1×Δ\mathbb{CP}^{1}\times\Delta by α1\alpha_{1}, we can assume that we have a homogeneous coordinate system [w1:w2][w_{1}:w_{2}] on 1\mathbb{CP}^{1} in which the vector field Vz1,0V^{1,0}_{z} vanishes at the points {0}\{0\} and {}\{\infty\}. Up to scale, there is a unique holomorphic vector field on 1\mathbb{CP}^{1} vanishing at two given points. If we set w=w1w2w=\frac{w_{1}}{w_{2}}, it follows then that V1,0=h(z)wwV^{1,0}=h(z)w\frac{\partial}{\partial w}, where h(z)h(z) is a holomorphic function only on Δ\Delta (notice that, although it is defined with respect to a coordinate system, www\frac{\partial}{\partial w} is in fact a global holomorphic vector field on 1\mathbb{CP}^{1}).

Now, X1,0X^{1,0} generates a \mathbb{C}^{*}-action on MM, and moreover each orbit of this action intersects the neighborhood 1×Δ\mathbb{CP}^{1}\times\Delta of L0L_{0}. Therefore, we can use the flow of X1,0X^{1,0} itself to extend this local description. In particular, there is a global holomorphic extension α:MM\alpha:M\to M of α1\alpha_{1} inducing a change of coordinates on MM in which X1,0X^{1,0} takes the form X1,0=h(z)ww+kzzX^{1,0}=h(z)w\frac{\partial}{\partial w}+kz\frac{\partial}{\partial z}, where w=w1w2w=\frac{w_{1}}{w_{2}} with respect to the homogeneous coordinates [w1:w2][w_{1}:w_{2}] on 1\mathbb{CP}^{1} and now h(z)h(z) is an entire holomorphic function on \mathbb{C}. Set Y1,0=wwY^{1,0}=w\frac{\partial}{\partial w}. Then clearly Y=Re(Y1,0)Y=\text{Re}(Y^{1,0}) is complete and [X,Y]=0[X,Y]=0. Furthermore, the flow of (Y,JY)(Y,JY) generates a \mathbb{C}^{*}-action on MM, which in these coordinates is just the standard action on 1\mathbb{CP}^{1} on each fiber of π\pi. Then ()2(\mathbb{C}^{*})^{2} acts on MM via X,JX,Y,JYX,JX,Y,JY, and therefore we can take 𝕋\mathbb{T} to be the underlying real torus of this action. ∎

4.4.2 Case 2: M(0)M^{(0)} is a fiber of π\pi

Claim 4.14.

Suppose that M(0)M^{(0)} is a fiber of π\pi, and so without loss of generality we may assume that M(0)=L0M^{(0)}=L_{0}. Then both M(2)M^{(2)} and M(4)M^{(4)} must be empty.

Proof.

Indeed, suppose that there exists a point qM(2)q\in M^{(2)}. Let z=π(q)z=\pi(q) so that qLzq\in L_{z}. By assumption, z0z\neq 0. By Claim 4.9, there is a holomorphic embedding Rq:1MR_{q}:\mathbb{CP}^{1}\to M defined by flowing along (X,JX)(-X,-JX) with the property that Rq(0)=qR_{q}(0)=q and Rq(1)=LzR_{q}(\mathbb{CP}^{1})=L_{z}. Set q=Rq()q^{\prime}=R_{q}(\infty), then it follows that the tangential component Vz1,0V_{z}^{1,0} of X1,0X^{1,0} to LzL_{z} vanishes precisely at the two points q,qq,q^{\prime} and that qq^{\prime} is the point at which ff achieves minLzf\min_{L_{z}}f. In particular, qq^{\prime} cannot lie in M(0)=L0M^{(0)}=L_{0}, which means that qM(2)q^{\prime}\in M^{(2)}. But then we run the same argument at qq^{\prime} to obtain a contradiction. Therefore M(2)M^{(2)} must be empty. The case qM(4)q\in M^{(4)} is similar. Alternatively, one can see that M(4)M^{(4)} is empty directly by an argument similar to the one in the proof of Claim 4.11. ∎

Claim 4.15.

If M(0)=L0M^{(0)}=L_{0}, then TXT^{X} is necessarily one-dimensional.

Proof.

Let yM(0)y\in M^{(0)}. Choose coordinates (z1,z2)(z_{1},z_{2}) in a neighborhood UyU_{y} centered at yy such that X1,0X^{1,0} takes the form (4.6). Since yM(0)y\in M^{(0)}, we have that a1,a2a_{1},a_{2} are both nonnegative. If a1a_{1} and a2a_{2} are both strictly positive, then it follows that every point yUyy^{\prime}\in U_{y} lies on an orbit which converges as t0t\to 0 to yy. Since M(0)=L0M^{(0)}=L_{0}, we can choose a point yM(0)Uyy^{\prime}\in M^{(0)}\cap U_{y}. Then Φt(y)y\Phi_{t}(y^{\prime})\to y as t0t\to 0, which contradicts the fact that XX vanishes identically on M0M_{0}. Therefore, we may assume without loss of generality that a1=0a_{1}=0 and a2>0a_{2}>0. In particular, L0UyL_{0}\cap U_{y} is given by the z1z_{1}-axis and indeed all of the orbits of (X,JX)(X,JX) in these coordinates are given by the affine lines z2=constz_{2}=const.

If TXT^{X}_{\mathbb{C}} is two-dimensional, then as we have seen at the beginning of Section 2.2 there exists an orbit of TXT^{X}_{\mathbb{C}} which is open and dense in MM. The flow of JXJX determines by assumption a dense subgroup in TXT^{X}, and therefore there must be some point qMq\in M such that the flow of (X,JX)(X,JX) from qq is dense in MM, and in particular is dense in UyU_{y}. But as we have seen, for sufficiently small tt, the Φ\Phi-orbit of any point in UyU_{y} lies on a unique complex submanifold of UyU_{y}, the line z2=constz_{2}=const. If the orbit Φt(q)\Phi_{t}(q) is dense in UyU_{y}, pick two points q1,q2q_{1},q_{2} such that z2(q1)z2(q2)z_{2}(q_{1})\neq z_{2}(q_{2}) and such that q2=Φt(q1)q_{2}=\Phi_{t^{*}}(q_{1}). By ensuring q2q_{2} is close enough to the z1z_{1}-axis, we can futher assume that |t|<1|t^{*}|<1. By the local form (4.6) we can see that the orbit of any point in UyU_{y} of the punctured unit disc 𝔻\mathbb{D}^{*}\subset\mathbb{C}^{*} is contained in UyU_{y}. In particular it follows that z1(q2)=z1(q1)z_{1}(q_{2})=z_{1}(q_{1}), a contradiction. ∎

Claim 4.16.

Let p,qMM(0)p,q\in M-M^{(0)}, and let Φ:×MM\Phi:\mathbb{C}^{*}\times M\to M denote the complex flow of (X,JX)(X,JX). If limt0Φt(p)=limt0Φt(q)M(0)\lim_{t\to 0}\Phi_{t}(p)=\lim_{t\to 0}\Phi_{t}(q)\in M^{(0)}, then q=Φt(p)q=\Phi_{t}(p) for some tt\in\mathbb{C}^{*}, i.e. pp and qq lie on the same orbit.

Proof.

This follows again from the local form (4.6). Since M(i)M^{(i)} are empty for i0i\neq 0 by Claim 4.14, it must be that limt0Φt(p)M(0)\lim_{t\to 0}\Phi_{t}(p)\in M^{(0)} for all pMp\in M. Now suppose that p,qMp,q\in M with limt0Φt(p)=limt0Φt(q)=yM(0)\lim_{t\to 0}\Phi_{t}(p)=\lim_{t\to 0}\Phi_{t}(q)=y\in M^{(0)}. As we have seen, we can choose coordinates (z1,z2)(z_{1},z_{2}) near yy in which X1,0X^{1,0} takes the form (4.6) where a1=0a_{1}=0 and a2>0a_{2}>0. It follows then that for sufficiently small ε\varepsilon, that both Φt(p)\Phi_{t}(p) and Φt(q)\Phi_{t}(q) lie on the line z2=0z_{2}=0 if |t|<ε|t|<\varepsilon. Thus the orbits from pp and from qq intersect, and are thereby equal. ∎

We can now treat the final case that may arise. Together with Claims 4.12 and 4.13, this completes the proof of Proposition 4.8.

Claim 4.17.

If M(0)=L0M^{(0)}=L_{0}, then there exists an equivariant biholomorphism α:M1×\alpha:M\to\mathbb{CP}^{1}\times\mathbb{C}, where 1×\mathbb{CP}^{1}\times\mathbb{C} is endowed with the product \mathbb{C}^{*}-action determined by the trivial action on 1\mathbb{CP}^{1} and the standard one on \mathbb{C}. In particular, under the identification determined by α\alpha, we have that JXJX lies in the Lie algebra 𝔱\mathfrak{t} of the standard T2T^{2}-action on 1×\mathbb{CP}^{1}\times\mathbb{C}.

Proof.

From the proof of Claim 4.15, we know that X1,0X^{1,0} satisfies a1=0,a2>0a_{1}=0,a_{2}>0 with respect to the local form (4.6). From this it is clear that the composition of any orbit Oq:MO_{q}:\mathbb{C}^{*}\hookrightarrow M of X1,0X^{1,0} with the projection π:M\pi:M\to\mathbb{C} defines a surjective map \mathbb{C}^{*}\to\mathbb{C}^{*}. In particular, if we let β=a21\beta=a_{2}^{-1}, then the orbits of (βX,J(βX))(\beta X,J(\beta X)) intersect each fiber of π\pi precisely once. Now choose any fiber Lz1L_{z}\cong\mathbb{CP}^{1} of π\pi in MM which is not equal to L0L_{0}, and let Φβ\Phi^{\beta} denote the flow of (βX,J(βX))(\beta X,J(\beta X)). We define a map α:M1×\alpha:M\to\mathbb{CP}^{1}\times\mathbb{C}^{*} by the formula

α(p)=(Φt1β(p),t),\alpha(p)=(\Phi^{\beta}_{t^{-1}}(p),t),

where tt\in\mathbb{C}^{*} is the unique point such that Φt1β(p)Lz\Phi^{\beta}_{t^{-1}}(p)\in L_{z}. By the previous claim, this extends to a biholomorphism α:M1×\alpha:M\to\mathbb{CP}^{1}\times\mathbb{C} such that αX1,0=a2zz\alpha_{*}X^{1,0}=a_{2}z\frac{\partial}{\partial z}. ∎

5 Discussion

We pose some open questions related to the work here. For the most part, these problems have appeared in [13]. We reproduce them here, partially because they take on a slightly different light in the toric setting, and partially because they may simply be easier to prove in this context.

  1. 1.

    Is the assumption in Theorem B that the Ricci curvature is bounded necessary? More specifically, suppose that (M,J)(M,J) admits an effective and holomorphic action of the real torus TnT^{n}. Given a complete shrinking gradient Kähler-Ricci soliton (g,X)(g,X) on MM, does there exists a complexification of the TnT^{n}-action? We use the bound on the Ricci curvature to apply the work of [13] to show that there exists a complexification if the soliton vector field satisfies JX𝔱JX\in\mathfrak{t}. Alternatively one could attempt to do away with the dependence on the full ()n(\mathbb{C}^{*})^{n}-action and corresponding dense complex coordinate chart. One can still interpret equation (1.4) as an equation for the complex structure JJ and produce a symplectic potential uu as in [5]. Our approach falls short at this stage, since we lack a method to determine good properties of the relevant functionals that appear in Section 3.

  2. 2.

    Suppose that MM is a toric manifold and (g,X)(g,X) is a complete shrinking gradient Kähler-Ricci soliton on MM. Does there always exist an automorphism α\alpha of MM such that αg\alpha^{*}g is invariant under the action of the real torus TnT^{n}? If we assume in addition that gg has bounded Ricci curvature, this is equivalent to the existence of an automorphism α\alpha such that JαX𝔱J\alpha^{*}X\in\mathfrak{t}. If so, then Theorem A (resp. Theorem B) implies that (g,X)(g,X) is the unique complete shrinking gradient Kähler-Ricci soliton on MM (resp. with bounded Ricci curvature). As it stands, we know little about the existence and uniqueness of shrinking solitons on MM without these hypotheses. We establish this in the special case that M=1×M=\mathbb{CP}^{1}\times\mathbb{C} in Proposition 4.8, and Conlon-Deruelle-Sun show this for MM equal to n\mathbb{C}^{n} or the total space of the line bundle 𝒪(k)n1\mathcal{O}(-k)\to\mathbb{CP}^{n-1} for 0<k<n0<k<n [13, Theorem 5.20].

  3. 3.

    Related to the previous question, suppose that MM is an arbitrary non-compact Kähler manifold and XX is a fixed holomorphic vector field. Is there at most one complete shrinking gradient Kähler-Ricci soliton gg on MM with XX as its soliton vector field? What if gg has bounded Ricci curvature? Moreover, is there at most one vector field XX on MM admitting a shrinking gradient Kähler-Ricci soliton? This is established by Tian-Zhu [47] for compact manifolds and by Conlon-Deruelle-Sun [13] for non-compact manifolds among all YY such that JYJY lie in the Lie algebra of a fixed real torus acting on MM, with the estra assumption that the Ricci curvatureis bounded. We recover this result in Theorem 4.6 in the toric setting.

  4. 4.

    In this paper we work exclusively on smooth spaces MM to avoid technical complications. In the compact setting there has also been much interest surrounding weak Kähler-Einstein metrics and Kähler-Ricci solitons on singular spaces. Many of the techniques in this paper are adapted from the paper of Berman-Berndtsson [8], in which such objects are of primary interest. Can the results here be generalized along the lines of [8] to include similar results for weak Kähler-Ricci solitons on non-compact singular toric varieties?

References

  • [1] M. Abreu. Kähler geometry of toric varieties and extremal metrics. Internat. J. Math., 9(6):641–651, 1998.
  • [2] M. Abreu. Kähler metrics on toric orbifolds. J. Differential Geom., 58:151–187, 2001.
  • [3] M. Abreu. Kähler geometry of toric manifolds in symplectic coordinates. In Symplectic and contact topology: interactions and perspectives (Toronto, ON/Montreal, QC, 2001), volume 35 of Fields Inst. Commun., pages 1–24. Amer. Math. Soc., Providence, RI, 2003.
  • [4] M. Abreu and R. Sena-Dias. Scalar-flat Kähler metrics on non-compact symplectic toric 4-manifolds. Ann. Global Anal. Geom., 41(2):209–239, 2012.
  • [5] V. Apostolov, D. M. J. Calderbank, P. Gauduchon, and C. W. Tø nnesen Friedman. Hamiltonian 2-forms in Kähler geometry. II. Global classification. J. Differential Geom., 68(2):277–345, 2004.
  • [6] M. F. Atiyah. Convexity and commuting Hamiltonians. Bull. London Math. Soc., 14(1):1–15, 1982.
  • [7] M. Audin. The topology of torus actions on symplectic manifolds, volume 93 of Progress in Mathematics. Birkhäuser Verlag, Basel, 1991. Translated from the French by the author.
  • [8] R. J. Berman and B. Berndtsson. Real Monge-Ampère equations and Kähler-Ricci solitons on toric log Fano varieties. Ann. Fac. Sci. Toulouse Math. (6), 22(4):649–711, 2013.
  • [9] R. L. Bryant. Gradient Kähler Ricci solitons. Number 321, pages 51–97. 2008. Géométrie différentielle, physique mathématique, mathématiques et société. I.
  • [10] D. Burns, V. Guillemin, and E. Lerman. Kähler metrics on singular toric varieties. Pacific J. Math., 238(1):27–40, 2008.
  • [11] H.-D. Cao and D. Zhou. On complete gradient shrinking Ricci solitons. J. Differential Geom., 85(2):175–185, 2010.
  • [12] R. J. Conlon and A. Deruelle. Steady gradient Kähler-Ricci solitons on crepant resolutions of Calabi-Yau cones. arXiv:2006.03100, 2020.
  • [13] R. J. Conlon, A. Deruelle, and S. Sun. Classification results for expanding and shrinking gradient Kähler-Ricci solitons. arXiv:1904.00147, 2019.
  • [14] D. A. Cox. The homogeneous coordinate ring of a toric variety. J. Algebraic Geom., 4(1):17–50, 1995.
  • [15] D. A. Cox, J. B. Little, and H. K. Schenck. Toric varieties, volume 124 of Graduate Studies in Mathematics. American Mathematical Society, Providence, RI, 2011.
  • [16] T. Delzant. Hamiltoniens périodiques et images convexes de l’application moment. Bull. Soc. Math. France, 116(3):315–339, 1988.
  • [17] S. K. Donaldson. Symmetric spaces, Kähler geometry and Hamiltonian dynamics. In Northern California Symplectic Geometry Seminar, volume 196 of Amer. Math. Soc. Transl. Ser. 2, pages 13–33. Amer. Math. Soc., Providence, RI, 1999.
  • [18] S. K. Donaldson. Scalar curvature and stability of toric varieties. J. Differential Geom., 62(2):289–349, 2002.
  • [19] S. K. Donaldson. Interior estimates for solutions of abreu’s equation. Collect. Math, 56, 2004.
  • [20] S. K. Donaldson. Kähler geometry on toric manifolds, and some other manifolds with large symmetry. In Handbook of geometric analysis. No. 1, volume 7 of Adv. Lect. Math. (ALM), pages 29–75. Int. Press, Somerville, MA, 2008.
  • [21] S. Dubuc. Critères de convexité et inégalités intégrales. Ann. Inst. Fourier (Grenoble), 27(1):x, 135–165, 1977.
  • [22] J. J. Duistermaat and G. J. Heckman. On the variation in the cohomology of the symplectic form of the reduced phase space. Invent. Math., 69(2):259–268, 1982.
  • [23] J. J. Duistermaat and G. J. Heckman. Addendum to: “On the variation in the cohomology of the symplectic form of the reduced phase space”. Invent. Math., 72(1):153–158, 1983.
  • [24] J. Enders, R. Müller, and P. M. Topping. On type-I singularities in Ricci flow. Comm. Anal. Geom., 19(5):905–922, 2011.
  • [25] M. Feldman, T. Ilmanen, and D. Knopf. Rotationally symmetric shrinking and expanding gradient Kähler-Ricci solitons. J. Differential Geom., 65(2):169–209, 2003.
  • [26] T. Frankel. Fixed points and torsion on Kähler manifolds. Ann. of Math. (2), 70:1–8, 1959.
  • [27] A. Futaki. Irregular Eguchi-Hanson type metrics and their soliton analogues. Pure Appl. Math. Q., 17:27–53, 2021.
  • [28] A. Futaki, H. Ono, and M.-T. Wang. Transverse kähler geometry of sasaki manifolds and toric sasaki-einstein manifolds. J. Differential Geom., 83:585–636, 2006.
  • [29] A. Futaki and M.-T. Wang. Constructing kähler-ricci solitons from sasaki-einstein manifolds. Asian J. Math., 15:33–52, 2011.
  • [30] V. Guillemin. Kaehler structures on toric varieties. J. Differential Geom., 40(2):285–309, 1994.
  • [31] V. Guillemin and S. Sternberg. Convexity properties of the moment mapping. Invent. Math., 67(3):491–513, 1982.
  • [32] J. Hilgert, K.-H. Neeb, and W. Plank. Symplectic convexity theorems and coadjoint orbits. Compositio Math., 94(2):129–180, 1994.
  • [33] D. Hwang and M. Singer. A momentum construction for circle-invariant kähler metrics. Trans. Am. Math. Soc., 354:2285–2325, 2002.
  • [34] K. Iwasawa. On some types of topological groups. Ann. of Math. (2), 50:507–558, 1949.
  • [35] Y. Karshon and E. Lerman. Non-compact symplectic toric manifolds. SIGMA Symmetry Integrability Geom. Methods Appl., 11:Paper 055, 37, 2015.
  • [36] E. Lerman. Contact toric manifolds. J. Symplectic Geom., 1(4):785–828, 2003.
  • [37] D. Martelli, J. Sparks, and S.-T. Yau. The geometric dual of aa-maximisation for toric Sasaki-Einstein manifolds. Comm. Math. Phys., 268(1):39–65, 2006.
  • [38] D. Martelli, J. Sparks, and S.-T. Yau. Sasaki-Einstein manifolds and volume minimisation. Comm. Math. Phys., 280(3):611–673, 2008.
  • [39] O. Munteanu and J. Wang. Structure at infinity for shrinking Ricci solitons. Ann. Sci. Éc. Norm. Supér. (4), 52(4):891–925, 2019.
  • [40] A. Naber. Noncompact shrinking four solitons with nonnegative curvature. J. Reine Angew. Math., 645:125–153, 2010.
  • [41] E. Prato. Convexity properties of the moment map for certain non-compact manifolds. Comm. Anal. Geom., 2(2):267–278, 1994.
  • [42] E. Prato and S. Wu. Duistermaat-Heckman measures in a non-compact setting. Compositio Math., 94(2):113–128, 1994.
  • [43] R. Sena-Dias. Uniqueness among scalar‐flat kähler metrics on non‐compact toric 4‐manifolds. J. London Math. Soc., 103:372–397, 2021.
  • [44] R. Sjamaar. Holomorphic slices, symplectic reduction and multiplicities of representations. Ann. of Math. (2), 141(1):87–129, 1995.
  • [45] H. Sumihiro. Equivariant completion. J. Math. Kyoto Univ., 14:1–28, 1974.
  • [46] G. Tian and X. Zhu. Uniqueness of Kähler-Ricci solitons. Acta Math., 184(2):271–305, 2000.
  • [47] G. Tian and X. Zhu. A new holomorphic invariant and uniqueness of Kähler-Ricci solitons. Comment. Math. Helv., 77(2):297–325, 2002.
  • [48] C. van Coevering. Ricci-flat Kähler metrics on crepant resolutions of Kähler cones. Math. Ann., 347(3):581–611, 2010.
  • [49] W. Wang. Rigidity of entire self-shrinking solutions to kähler-ricci flow on the complex plane. Proc. Amer. Math. Soc., 145:3105–3108, 2017.
  • [50] X.-J. Wang and X. Zhu. Kähler-Ricci solitons on toric manifolds with positive first Chern class. Adv. Math., 188(1):87–103, 2004.
  • [51] W. Wylie. Complete shrinking Ricci solitons have finite fundamental group. Proc. Amer. Math. Soc., 136(5):1803–1806, 2008.
  • [52] B. Yang. A characterization of noncompact koiso-type solitons. Int. J. Math., 23:1250054, 2012.
  • [53] Z.-H. Zhang. On the completeness of gradient Ricci solitons. Proc. Amer. Math. Soc., 137(8):2755–2759, 2009.