
In this paper, we prove an existence result for Kähler-Einstein metrics on Q-Fano compactifications of Lie groups by the variational method, provided their moment polytopes satisfy a fine condition. As an application, we prove that there is no Q-Fano SO4(C)-compactification which admits a Kähler-Einstein metric with the same volume as that of a smooth K-unstable Fano SO4(C)-compactification.
Citation: Yan Li, Gang Tian, Xiaohua Zhu. Singular Kähler-Einstein metrics on Q-Fano compactifications of Lie groups[J]. Mathematics in Engineering, 2023, 5(2): 1-43. doi: 10.3934/mine.2023028
[1] | Tommaso Barbieri . On Kolmogorov Fokker Planck operators with linear drift and time dependent measurable coefficients. Mathematics in Engineering, 2024, 6(2): 238-260. doi: 10.3934/mine.2024011 |
[2] | Luigi Ambrosio, Giuseppe Savaré . Duality properties of metric Sobolev spaces and capacity. Mathematics in Engineering, 2021, 3(1): 1-31. doi: 10.3934/mine.2021001 |
[3] | Rupert L. Frank, Tobias König, Hynek Kovařík . Energy asymptotics in the Brezis–Nirenberg problem: The higher-dimensional case. Mathematics in Engineering, 2020, 2(1): 119-140. doi: 10.3934/mine.2020007 |
[4] | Marco Bramanti, Sergio Polidoro . Fundamental solutions for Kolmogorov-Fokker-Planck operators with time-depending measurable coefficients. Mathematics in Engineering, 2020, 2(4): 734-771. doi: 10.3934/mine.2020035 |
[5] | Francesca G. Alessio, Piero Montecchiari . Gradient Lagrangian systems and semilinear PDE. Mathematics in Engineering, 2021, 3(6): 1-28. doi: 10.3934/mine.2021044 |
[6] | Jérôme Droniou, Jia Jia Qian . Two arbitrary-order constraint-preserving schemes for the Yang–Mills equations on polyhedral meshes. Mathematics in Engineering, 2024, 6(3): 468-493. doi: 10.3934/mine.2024019 |
[7] | Luca Capogna, Giovanna Citti, Nicola Garofalo . Regularity for a class of quasilinear degenerate parabolic equations in the Heisenberg group. Mathematics in Engineering, 2021, 3(1): 1-31. doi: 10.3934/mine.2021008 |
[8] | José M. Mazón . The total variation flow in metric graphs. Mathematics in Engineering, 2023, 5(1): 1-38. doi: 10.3934/mine.2023009 |
[9] | Connor Mooney, Arghya Rakshit . Singular structures in solutions to the Monge-Ampère equation with point masses. Mathematics in Engineering, 2023, 5(5): 1-11. doi: 10.3934/mine.2023083 |
[10] | Nguyen Cong Phuc, Igor E. Verbitsky . Uniqueness of entire solutions to quasilinear equations of p-Laplace type. Mathematics in Engineering, 2023, 5(3): 1-33. doi: 10.3934/mine.2023068 |
In this paper, we prove an existence result for Kähler-Einstein metrics on Q-Fano compactifications of Lie groups by the variational method, provided their moment polytopes satisfy a fine condition. As an application, we prove that there is no Q-Fano SO4(C)-compactification which admits a Kähler-Einstein metric with the same volume as that of a smooth K-unstable Fano SO4(C)-compactification.
Let G be an n-dimensional connected, linear complex reductive Lie group which is the complexification of a compact Lie group K. Let T be a maximal torus of K, which has dimension r and Lie algebra t. Then TC is a maximal complex torus of G. Denote by Φ+ a chosen positive roots system associated to TC. Put
ρ=12∑α∈Φ+α. | (1.1) |
It can be regarded as a character in a∗, where a∗ is the dual space of the non-compact part a=√−1t of tC. Let π be a function on a∗ defined by
π(y)=∏α∈Φ+⟨α,y⟩2,y∈a∗, |
where ⟨⋅,⋅⟩* denotes the Cartan-Killing inner product on a∗.
* Without of confusion, we also write it as α(y) for simplicity.
Let M be a Q-Fano G-compactification (cf. [4]). Denote by Z the closure of TC in M. Then (Z,−KM|Z) is a polarized toric variety. Hence there is an associated moment polytope P of (Z,−KM|Z) induced by (M,−KM) [3,4]. Let P+ be the positive part of P defined by P+=P∩¯a∗+, where
a+={y∈a∗|⟨α,y⟩>0,∀α∈Φ+} |
is the positive Weyl chamber in a∗ defined by Φ+. Denote by 2P+ a dilation of P+ at rate 2. We define the barycenter of 2P+ with respect to the weighted measure π(y)dy by
bar(2P+)=∫2P+yπ(y)dy∫2P+π(y)dy. |
In [20], Delcroix proved the following existence theorem for Kähler-Einstein metrics on smooth Fano G-compactifications.
Theorem 1.1. Let M be a smooth Fano G-compactification. Then M admits a Kähler-Einstein metric if and only if
bar(2P+)∈4ρ+Ξ, | (1.2) |
where Ξ is the relative interior of the cone generated by Φ+.
There is an alternative proof of Theorem 1.1 given by Li, Zhou and Zhu via the variation method [36]. They also showed that (1.2) is actually equivalent to the K-stability condition in terms of [44] and [24] by constructing C∗-action through piecewisely rationally linear function which is invariant under the Weyl group action. In particular, it implies that M is K-unstable if bar(2P+)∉¯4ρ+Ξ. A more general construction of C∗-action was also discussed in [21].
In the present paper, we extend the above theorem to Q-Fano compactifications of G which may be singular. It is well known that any Q-Fano G-compactification has klt-singularities [4,Section 5]. For a Q-Fano variety M with klt-singularities, there is naturally a class of admissible Kähler metrics induced by the Fubini-Study metric (cf. [23]). In [10], Berman, Boucksom, Eyssidieux, Guedj and Zeriahi introduce a class of Kähler potentials associated to admissible Kähler metrics and refer it as the E1(M,−KM) space. Then they define the singular Kähler-Einstein metric on M with the Kähler potential in E1(M,−KM) via the complex Monge-Ampère equation, which is the usual Kähler-Einstein metric on the smooth part of M. It is a natural problem to establish an extension of the Yau-Tian-Donaldson conjecture we have solved for smooth Fano manifolds [44,46], that is, an equivalence relation between the existence of such singular Kähler-Einstein metrics and the K-stability on a Q-Fano variety M with klt-singularities. Actually, there are many recent works on this fundamental problem in terms of uniform K-stability. We refer the reader to [10,11,32,33,34], etc..
We assume that the associated polytope P of (Z,K−1M|Z) is fine in sense of [25], namely, each vertex of P is the intersection of precisely r facets. Then we prove
Theorem 1.2. Let M be a Q-Fano G-compactification such that the associated polytope P is fine.Then M admits a singular Kähler-Einstein metric if and only if (1.2) holds.
By a result of Abreu [1], the assumption that the polytope P being fine is equivalent to that the metric induced by the Guillemin function can be extended to a Kähler orbifold metric on Z.† It follows that the Guillemin function of 2P in Theorem 1.2 induces a K×K-invariant singular metric ω2P in E1(M,−KM) (cf. Lemma 3.4). Moreover, we can prove that the Ricci potential of ω2P on M is uniformly bounded above. We note that P is always fine when rank(G)=2. Thus for a Q-Fano compactification of G with rank(G)=2, M admits a singular Kähler-Einstein metric if and only if (1.2) holds. As an application of Theorem 1.2, we show that there is only one example of non-smooth Gorenstein Fano SO4(C)-compactifications which admits a singular Kähler-Einstein metric (cf. Section 7.1).
† It can not be guaranteed that the G–compactification is smooth even if Z is smooth, see an examples for G=SO2n+1(C) in [47,Section 11].
On the other hand, it has been shown in [41] that there are only three smooth Fano compactifications of SO4(C), i.e., Case-1.1.2, Case-1.2.1 and Case-2 in Section 7.1.
The first two manifolds do not admit any Kähler-Einstein metric [20,36]. By Theorem 1.2, we further prove
Theorem 1.3. There is no Q-Fano compactification of SO4(C) which admits a singular Kähler-Einstein metric with the same volume as Case-1.1.2 or Case-1.2.1 in Section 7.1.
Theorem 1.3 gives a partial answer to a question proposed in [38] about limit of Kähler-Ricci flow on either Case-1.1.2 or Case-1.2.1. It has been proved there that the flow has type II singularities on each of Case-1.1.2 and Case-1.2.1. By the Hamilton-Tian conjecture [7,16,44,48], the limit should be a singular Kähler-Ricci soliton on a Q-Fano variety with the same volume as that of initial metric. However, by Theorem 1.3, the limit can not be a Q-Fano compactification of SO4(C) with a singular Kähler-Einstein metric. This implies that the limiting soliton of flow on either Case-1.1.2 or Case-1.2.1 will have less symmetric than the initial one, which is totally different from the situation of smooth convergence of K×K-invariant metrics on a smooth compactification of Lie group [38].
As in [9,36,49], we use the variational method to prove Theorem 1.2. More precisely, we will prove that a modified version of the Ding functional D(⋅) is proper under the condition (1.2). This functional is defined for a class of convex functions E1K×K(2P) associated to K×K-invariant metrics on the orbit of G (cf. Section 4, 6). The key point is that the Ricci potential h0 of the Guillemin metric ω2P is bounded from above when P is fine (cf. Proposition 5.1). This enables us to control the nonlinear part F(⋅) of D(⋅) by modifying D(⋅) as done in [24,37] (cf. Section 6.1). We shall note that it is in general impossible to get a lower bound of h0 if the compactification is a singular variety (cf. Remark 5.2). However, by a recent deep result of Li in [32], the additional fine condition in Theorem 1.2 can be actually dropped. For the completeness, we will also improve the theorem at the end of paper, Appendix 2.
The minimizer of D(⋅) corresponds to a singular Kähler-Einstein metric. We will prove the semi-continuity of D(⋅) and derive the Kähler-Einstein equation for the minimizer (cf. Proposition 6.6). Our proof is similar with what Berman and Berndtsson studied on toric varieties in [9].
The proof of the necessity part of Theorem 1.2 is the same as one in Theorem 1.1. In fact, a Q-Fano compactification of G is not K-polystable if (1.2) is not satisfied [36,Proposition 3.4] (also see [21]). This will be a contradiction to the K-polystability of Q-Fano variety with a singular Kähler-Einstein metric (cf. [8,33]). We omit this part.
The organization of paper is as follows. In Section 2, we recall some notations in [10] for singular Kähler-Einstein metrics on Q-Fano varieties. In Section 3, we introduce a subspace E1K×K(M,−KM) of E1(M,−KM) and prove that the Guillemin function lies in this space (cf. Lemma 3.4). In Section 4, we prove that E1K×K(M,−KM) is equivalent to a dual space E1K×K(2P) of Legendre functions (cf. Theorem 4.2). In Section 5, we compute the Ricci potential h0 of ω2P and show that it is bounded from above (cf. Proposition 5.1). The sufficient part of Theorem 1.2 will be proved in Section 6. In Section 7, we construct many Q-Fano compactifications of SO4(C) and in particular, we will prove Theorem 1.3.
For a Q-Fano variety M, by Kodaira's embedding, there is an integer ℓ>0 such that we can embed M into a projective space CPN by a basis of H0(M,K−ℓM), for simplicity, we assume M⊂CPN and K−ℓM=OCPN(1). Then we have a metric
ω0=1ℓωFS|M∈2πc1(M), |
where ωFS is the Fubini-Study metric of CPN. Moreover, there is a Ricci potential h0 of ω0 such that
Ric(ω0)−ω0=√−1∂ˉ∂h0,onMreg. |
In the case that M has only klt-singularities, eh0 is Lp-integrable for some p>1 (cf. [10,23]). We call ω an admissible Kähler metric on M if there are an embedding M⊂CPN as above and a smooth function ϕ on CPN such that
ω=ωϕ=ω0+√−1∂ˉ∂ϕ|M. |
In particular, ϕ is a function on M with ϕ∈L∞(M)∩C∞(Mreg), called an admissible Kähler potential associated to ω0.‡
‡ For simplicity, we will denote a Kähler metric by its Kähler form thereafter.
For a general (possibly unbounded) Kähler potential ϕ, we define its complex Monge-Ampère measure ωnϕ by
ωnϕ=limj→∞ωnϕj, |
where ϕj=max{ϕ,−j}. According to [10], we say that ϕ (or ωnϕ) has full Monge-Ampère (MA) mass if
∫Mωnϕ=∫Mωn0. |
The MA-measure ωnϕ with a full MA-mass has no mass on the pluripolar set of ϕ in M. Thus we need to consider the measure on Mreg. Moreover, e−ϕ is Lp-integrable for any p>0 associated to ωn0.
Definition 2.1. We call ωϕ a (singular) Kähler-Einstein metric on M with full MA-mass if ϕ satisfies the following complex Monge-Ampère equation,
ωnϕ=eh0−ϕωn0. | (2.1) |
It has been shown in [10] that ϕ is C∞ on Mreg if it is a solution of (2.1). Thus ωϕ satisfies the Kähler-Einstein equation on Mreg,
Ric(ωϕ)=ωϕ. |
On a smooth Fano manifold, there is a well-known Euler-Lagrange functional for Kähler potentials associated to (2.1), often referred as the Ding functional or F-functional, defined by (cf. [22,43]),
F(ϕ)=−1(n+1)Vn∑k=0∫Mϕωkϕ∧ωn−k0−log(1V∫Meh0−ϕωn0). | (2.2) |
In case of Q-Fano manifold with klt-singularities, Berman-Boucksom-Eyssidieux-Guedj-Zeriahi [10] extended F(⋅) to the space E1(M,−KM) defined by
E1(M,−KM)={ϕ|ϕhasfullMAmassandsupMϕ=0,I(ϕ)=∫M−ϕωnϕ<∞}. |
They showed that E1(M,−KM) is compact in certain weak topology. By a result of Darvas [18], E1(M,−KM) is in fact compact in the topology of L1-distance. It provides a variational approach to study (2.1).
Definition 2.2. [10,44] The functional F(⋅) is called proper if there is a continuous function p(t) on R with the property limt→+∞p(t)=+∞, such that
F(ϕ)≥p(I(ϕ)),∀ϕ∈E1(M,−KM). | (2.3) |
In [10], Berman-Boucksom-Eyssidieux-Guedj-Zeriahi proved the existence of solutions for (2.1) under the properness assumption (2.3) of F(⋅). However, this assumption does not hold in the case of existence of non-zero holomorphic vector fields such as in our case of Q-Fano G-compactifications. So we need to consider a modified version of properness of the Ding functional instead to overcome this new difficulty as done on toric varieties [9,37].
Let M be a Q-Fano compactification of G with Z being the closure of a maximal complex torus TC-orbit. We first characterize the polytope P of Z associated to (M,−KM). Denote by M the lattice of G-weights and W the Weyl group of (G,TC). Then P is an r-dimensional W-invariant, convex, rational polytope in a∗=MR. Let {FA}A=1,...,d0 be the facets of P and {FA}A=1,...,d+ be those whose relative interior intersects a∗+. Suppose that
P=∩d0A=1{loA:=λA−uA(y)≥0} | (3.1) |
for some primitive vector uA∈N and the facet
FA⊆{loA=0},A=1,...,d0. |
By the W-invariance, for each A∈{1,...,d0}, there is some wA∈W such that wA(FA)∈{FB}B=1,...,d+. Denote by ρA=w−1A(ρ), where ρ∈a∗+ is given by (1.1). Then ρA(uA) is independent of the choice of wA∈W and hence it is well-defined.
The following Lemma can be derived from a general result [26,Theorem 1.9] on polytopes of Q-Fano spherical varieties. For readers' convenience, we sketch a direct proof in cases of group compactifications below by using [15,Section 4].
Lemma 3.1. Let M be a Q-Fano compactification of G with P being the associated moment polytope. Then for each A=1,...,d0, it holds
λA=1+2ρA(uA). | (3.2) |
Conversely, a W-invariant convex polytope P given by (3.1) is the associated polytope of some Q-Fano G-compactification if (3.2) holds.
Sketch of proof. Suppose that −mKM is a Cartier divisor for some m∈N+. Up to a dilation of the polytope, it suffices to consider the case when m=1. Denote by B+ the (positive) Borel subgroup of G corresponding to (TC,Φ+) and B− be the opposite one. Then by [15,Section 4], there exists a B+×B−-semiinvariant section of −KM whose divisor is
d=d+∑A=1XA+2∑αi∈Φ+,sYαi, | (3.3) |
where {XA′} is the set of G×G-invariant prime divisors and Yαi is the prime B+×B−-invariant divisor corresponds to the weight αi in Φ+,s, the set of simple roots in Φ+. Note that the corresponding B+×B−-weight of this divisor is 2ρ (cf. [21,Section 3.2.4]). Thus by adding the divisor of a B+×B−-semi-invariant rational function fo with weight −2ρ, we have
d+div(fo)=d+∑A=1(1+2ρ(uA))XA | (3.4) |
is a G×G-invariant divisor. On the other hand, by [5,Theorem 2.4 (3)], the prime G×G-invariant divisors of M are in bijections with W-orbits of prime toric divisors of Z. Restricting the above divisor to Z, we get (3.2).
Conversely, suppose that there is a W-invariant polytope P given by (3.1) and (3.2). Then P is rational and there is an m∈N+ so that mP is integral. The support function of mP is W-invariant and strictly convex. Hence it corresponds to an ample Cartier divisor md′, where (cf. [14,Section 3.3])
d′=d+∑A=1(1+2ρ(uA))XA. |
Obviously d′−div(fo) equals to the divisor d defined by (3.3), which is a divisor of −KM by [15,Section 4]. We conclude the Lemma.
On a Q-Fano compactification of G, we may regard the G×G-action as a subgroup of PGLN+1(C) which acts holomorphically on the hyperplane bundle L=OCPN(1). Then any admissible K×K-invariant Kähler metric ωϕ∈2πℓc1(L) can be regarded as a restriction of K×K-invariant Kähler metric of CPN. Thus the moment polytope P associated to (Z,L|Z) is a W-invariant rational polytope in a∗. By the K×K-invariance, the restriction of ωϕ on TC is an open toric Kähler metric. Hence, it induces a strictly convex, W-invariant function ψϕ on a [6] (also see Lemma 3.3 below) such that
ωϕ=√−1∂ˉ∂ψϕ,onTC. | (3.5) |
By the KAK-decomposition ([31,Theorem 7.39]), for any g∈G, there are k1,k2∈K and x∈a such that g=k1exp(x)k2. Here x is uniquely determined up to a W-action. This means that x is unique in ˉa+. Thus there is a bijection between K×K-invariant functions Ψ on G and W-invariant functions ψ on a which is given by
Ψ(exp(⋅))=ψ(⋅):a→R. |
Without of confusion, we will not distinguish ψ and Ψ, and call Ψ (or ψ) convex on G if ψ is convex on a.
The following KAK-integration formula can be found in [31,Proposition 5.28].
Proposition 3.2. Let dVG be a Haar measure on G and dx the Lebesgue measureon a.Then there exists a constant CH>0 such that for anyK×K-invariant, dVG-integrable function ψ on G,
∫Gψ(g)dVG=CH∫a+ψ(x)J(x)dx, |
where
J(x)=∏α∈Φ+sinh2α(x). |
Without loss of generality, we may normalize CH=1 for simplicity.
Next we recall a local holomorphic coordinate system on G used in [20]. By the standard Cartan decomposition, we can decompose g as
g=(t⊕a)⊕(⊕α∈ΦVα), |
where t is the Lie algebra of T and
Vα={X∈g|adH(X)=α(H)X,∀H∈t⊕a} |
is the root space of complex dimension 1 with respect to α. By [28], one can choose Xα∈Vα such that X−α=−ι(Xα) and [Xα,X−α]=α∨, where ι is the Cartan involution and α∨ is the dual of α by the Killing form. Let Eα=Xα−X−α and E−α=J(Xα+X−α). Denoted by kα,k−α the real line spanned by Eα,E−α, respectively. Then we get the Cartan decomposition of Lie algebra k of K as follows,
k=t⊕(⊕α∈Φ+(kα⊕k−α)). |
Choose a real basis {E01,...,E0r} of t, where r is the dimension of T. Then {E01,...,E0r} together with {Eα,E−α}α∈Φ+ forms a real basis of k, which is indexed by {E1,...,En}. We can also regard {E1,...,En} as a complex basis of g. For any g∈G, we define local coordinates {zi(g)}i=1,...,n on a neighborhood of g by
(zi(g))→exp(zi(g)Ei)g. |
It is easy to see that θi|g=dzi(g)|g, where the dual θi of Ei is a right-invariant holomorphic 1-form. Thus
dVG|g:=∧ni=1(dzi(g)∧dˉzi(g))|g,∀g∈G | (3.6) |
is also a right-invariant (n,n)-form, which defines a Haar measure.
For a K×K-invariant function ψ, Delcroix computed the Hessian of ψ in the above local coordinates as follows [20,Theorem 1.2].
Lemma 3.2. Let ψ be a K×K invariant function on G.Then for any x∈a+, the complex Hessian matrix of ψ in the above coordinates is diagonal by blocks, and equals to
HessC(Ψ)(exp(x))=(14HessR(ψ)(x)000Mα(1)(x)000⋱⋮⋮⋮⋱000Mα(n−r2)(x)), | (3.7) |
where Φ+={α(1),...,α(n−r2)} is the set of positive roots and
Mα(i)(x)=12α(i)(∂ψ(x))(cothα(i)(x)√−1−√−1cothα(i)(x)). |
By (3.7) in Lemma 3.3, we see that ψϕ induced by an admissible K×K-invariant Kähler form ωϕ is convex on a. The complex Monge-Ampère measure is given by
ωnϕ=:(√−1∂ˉ∂ψϕ)n=MAC(ψϕ)dVG. | (3.8) |
By (3.6), for any x∈a+ we have
MAC(ψϕ)(exp(x))=12r+nMAR(ψϕ)(x)1J(x)∏α∈Φ+α2(∂ψϕ(x)) | (3.9) |
in (3.8). In particular, by Proposition 3.2,
Vol(M)=∫Mωnϕ=∫2P+πdy=V. | (3.10) |
Clearly, (3.9) also holds for any Kähler potential in E1(M,−KM), which is smooth and K×K-invariant on G. For the completeness, we introduce a subspace of E1(M,−KM) by
E1K×K(M,−KM)={ϕ∈E1(M,−KM)|ψ0+ϕisK×K-invariantandconvexonG}. | (3.11) |
Here ψ0 is a convex function on a associated to a background admissible K×K-invariant metric ω0 as in (3.5). E1K×K(M,−KM) is locally precompact in terms of convex functions on a. In Sections 4 and 6, we will prove its completeness by using the Legendre dual.
In this subsection, we show that the Legendre dual of Guillemin function u2P on 2P lies in E1K×K(M,−KM) when P is fine.
Recall (3.1). For convenience, we set
lA(y)=2λA−uA(y). |
Then
2P=∩d0A=1{lA(y)≥0}. |
Thus, u2P is given by (cf. [1])
u2P=12d0∑A=1lAloglA(y). |
Clearly, it is W-invariant, so its Legendre function ψ2P is also W-invariant, where
ψ2P(x)=supy∈2P(⟨x,y⟩−u2P(y)),∀x∈a. | (3.12) |
Hence, by [1,Theorem 2] and [6] (also see Lemma 3.3),§ we can extend
§ The corresponding moment map is given by 12∇ψ2P, whose image is P.
ω2P=√−1∂ˉ∂ψ2P,ona, |
to a K×K-invariant metric on G.
Lemma 3.4. Let ψ0 be the background K×K-invariant Kähler potential in (3.11). Assume that P is fine. Then the Kähler potential (ψ2P−ψ0) of ω2P lies in ϕ∈L∞(M)∩C∞(Mreg). In particular, (ψ2P−ψ0)∈E1K×K(M,−KM).
Proof. Fix an m0∈Z+ such that −m0KX is very ample. We consider the projective embedding
ι:M→CPN |
given by |−m0KM|, where N=h0(M,−m0KM)−1. By [39,Section 2.3], the pull back of the Fubini-Study metric on CPN gives a K×K-invariant, Hermitian metric h on L=OCPN(−1)|M. Moreover, we have
h|TC(x)=∑λ∈mP∩Mˉn(λ)e2λ(x), | (3.13) |
where ˉn(λ)∈Z+. Thus we have a Kähler potential on TC by
ψFS=1mlogh|TC. |
Since P is fine, one can show directly that
ψFS∈V(2P)={ψ∈C0(a)|ψ is convex, W-invariantand maxa|v2P−ψ|<∞}, |
where v2P(⋅) is the support function on a defined by
v2P(x)=supy∈2P⟨x,y⟩. | (3.14) |
Recall that the Legendre function uψ of ψ is defined as in (3.12) by
uψ(y)=supx∈a(⟨x,y⟩−ψ(x)),y∈2P. | (3.15) |
It is known that ψ∈V(2P) if and only if uψ is uniformly bounded on 2P since the Legendre function of v2P is zero (cf. [42]). Thus the Legendre function uh of h|TC(x) is uniformly bounded on 2P. It follows that
|uh−v2P|≤C. |
Hence, we get
maxa|ψFS−ψ2P|<+∞. |
Consequently,
maxa|ψ2P−ψ0|<+∞. |
By (3.10), (ψ2P−ψ0) has full MA-mass, so we have completed the proof.
In this section, we describe the space E1K×K(M,−KM) in (3.11) via Legendre functions as in [17] for Q-Fano toric varieties. Recall the background K×K-invariant Kähler potential ψ0 in Lemma 3.4. Then we can normalize ψ0 up to an action of the centre Z(G) of G as follows (cf. [38]),
infaψ0=ψ0(O)=0, | (4.1) |
where O is the origin of a. Similarly, for any ϕ∈E1K×K(M,−KM), ψϕ=ψ0+ϕ can be also normalized as in (4.1).
The following lemma is elementary.
Lemma 4.1. For any K×K-invariant potential ϕ normalized as in (4.1), it holds
∂(ψϕ)⊆2P,andψϕ≤v2P, |
where ∂(ψϕ)(⋅) is the normal mapping of ψϕ.
Proof. We choose a sequence of decreasing and uniformly bounded K×K-invariant potential ϕi normalized as in (4.1) such that
ω0+√−1∂ˉ∂ϕi>0,onMreg |
and
ϕi→ϕ,asi→+∞. |
Then
√−1∂ˉ∂ψϕi>0 in G. |
It follows that
∂ψϕi⊆2P. |
This implies that ∂ψϕ⊆2P. By the convexity, we also get ψϕ≤v2P.
It is easy to see that the Legendre function uϕ of ψϕ with ϕ∈E1K×K(M,−KM) satisfies
inf2Puϕ=uϕ(O)=0. | (4.2) |
We set a class of W-invariant convex functions on 2P by
E1K×K(2P)={u|u is convex,W−invariant on 2P which satisfies (4.2) and ∫2P+uπdy<+∞}. |
Our main goal in this section is to prove
Theorem 4.2. A Kähler potential ϕ∈E1K×K(M,−KM) with normalized ψϕ satisfying (4.1) if and only if the Legendre function uϕ of ψϕ lies in E1K×K(2P). In particular, uϕ is locally bounded in Int(2P) if ϕ∈E1K×K(M,−KM).
As in [17], we need to establish a comparison principle for the complex Monge-Ampère measure in E1K×K(M,−KM). For our purpose, we will introduce a weighted Monge-Ampère measure on a in the following.
Definition 4.3. Let Ω⊆a be a W-invariant domain and ψ any W-invariant convex function on Ω. Define a weighted Monge-Ampère measure on Ω by
∫Ω′MAR;π(ψ)dx=∫∂ψ(Ω′)πdy,∀Ω′⋐Ω, |
where ∂ψ(⋅) is the normal mapping of ψ.
Remark 4.4. Let {ψk} be a sequence of convex functions which converges locally uniformly to ψ on Ω, then MAR;π(ψk) converges to MAR;π(ψ) (cf. [2,Section 15]).
We have the following KAK-integration for the measure ωnϕ with ϕ∈E1K×K(M,−KM).
Lemma 4.5. Let ωϕ=√−1∂ˉ∂ψϕ with ϕ∈E1K×K(M,−KM). Then for any K×K-invariant continuous uniformly bounded function f on G, it holds
∫Mfωnϕ=∫a+fMAR;π(ψϕ)dx. | (4.3) |
Proof. First we assume that f is a K×K-invariant continuous function with compact support on a. We take a sequence of smooth W-invariant convex functions ψk↘ψ and let ωk=√−1∂ˉ∂ψk. Then for any W-invariant Ω′⋐a, it holds
∫Ω′MAR;π(ψk)dx:=∫Ω′det(∇2ψk)π(∇ψk)dy. |
By the standard KAK-integration formula in Proposition 3.2, it follows that
∫Mfωnk=∫a+fdet(∇2ψk)π(∇ψk)dx=∫a+fMAR;π(ψk)dx. |
Since
∫Mfωnk→∫Mfωn, |
we get (4.3) by Remark 4.4.
Next we choose a sequence of exhausting W-invariant convex domains Ωk in a and a sequence of W-invariant convex functions with the support on Ωk+1 such that fk=f|Ωk. Since ωn has full MA-mass, we get
∫Mfωn=limk∫Mfkωn=limk∫a+fkMAR;π(ψϕ)dx=∫a+fMAR;π(ψϕ)dx. |
In this subsection, we establish an ordinary comparison principle for the weighted Monge-Ampère measure MAR;π(ψ). As showed in [17], this will then lead a global comparison principle (see Proposition 4.7 below) which can be used to estimate the MA mass of Kähler potential. We will only prove Proposition 4.6 and omit other proofs, since the others follow directly from Proposition 4.6 by corresponding arguments in [17].
Proposition 4.6. Let Ω⊆a be a W-invariant domain and φ,ψ be two convex functions on Ω such that
φ≥ψ and (φ−ψ)|∂Ω=0. | (4.4) |
Then
∫ΩMAR;π(φ)dx≤∫ΩMAR;π(ψ)dx. | (4.5) |
Proof. It is sufficient to prove (4.5) when φ and ψ are smooth, since we can approximate general φ and ψ by smooth W-invariant convex functions by Lemma 4.5. Let
φt=tφ+(1−t)ψ. |
Then
MAR;π(φt)=det(∇2φt)∏α∈Φ+α2(∇φt) |
and
ddt∫Ωdet(∇2φt)∏α∈Φ+α2(∇φt)dx=∫Ω(∇2φt)−1,ij∇2˙φt,ijdet(∇2φt)∏α∈Φ+α2(∇φt)dx+∫Ω(∑α∈Φ+2α(∇˙φt)α(∇φt))det(∇2φt)∏α∈Φ+α2(∇φt)dx. | (4.6) |
Using the fact that
(det(∇2φt)(∇2φt)−1,ij),j=0 |
and integration by parts, we have
∫Ω(∇2φt)−1,ij∇2˙φt,ijdet(∇2φt)∏α∈Φ+α2(∇φt)dx=∫∂Ω(∇2φt)−1,ij∇˙φt,iνjdet(∇2φt)∏α∈Φ+α2(∇φt)dσ−∫∂Ω[(∇2φt)−1,ijdet(∇2φt)∏α∈Φ+α2(∇φt)],jνi˙φtdσ+∫Ω(∇2φt)−1,ij˙φtdet(∇2φt)(∏α∈Φ+α2(∇φt)),ijdx. | (4.7) |
Also
∫Ω(∑α∈Φ+2α(∇˙φt)α(∇φt))det(∇2φt)∏α∈Φ+α2(∇φt)dx=2∫∂Ω∑α∈Φ+αiνiα(∇φt)det(∇2φt)∏α∈Φ+α2(∇φt)˙φtdσ=−2∫Ω(det(∇2φ)∏α∈Φ+α2(∇φt)∑α∈Φ+αiα(∇φ)),i˙φtdx. | (4.8) |
Note that
(∏α∈Φ+α2(∇φt)),ij∏α∈Φ+α2(∇φt)=−2∑α∈Φ+αkαlφt,ikφt,jlα2(∇φt)+2∑α∈Φ+αkφt,ijkα(∇φt)+4∑α,β∈Φ+αkβlφt,ikφt,jlα(∇φt)β(∇φt)=2(det(∇2φ)∏α∈Φ+α2(∇φt)∑α∈Φ+αiα(∇φ)),idet(∇2φ)∏α∈Φ+α2(∇φt). | (4.9) |
Plugging (4.7)–(4.9) into (4.6) and using the boundary condition (4.4), we have
ddt∫Ωdet(∇2φt)∏α∈Φ+α2(∇φt)dx=∫∂Ω∇˙φt,iνidet(∇2φt)∏α∈Φ+α2(∇φt)dσ≤0. |
Hence we get (4.5).
By the above Proposition, we get the following analogue of [17,Lemma 2.3], which gives a global comparison principle for the weighted Monge-Ampère measures.
Proposition 4.7. Let φ,ψ be two W-invariant convex functions on a so that
φ≥ψ |
and
lim|x|→+∞φ(x)=+∞. |
Then
∫a+MAR;π(φ)dx≥∫a+MAR;π(ψ)dx. |
As an application of Proposition 4.7, we get by the argument of [17,Lemma 2.7],
Lemma 4.6. Let ψ be a W-invariant convex function on a and u its Legendre function. Suppose that for some constant C,
ψ≤v2P+C, | (4.10) |
where v2P is the support function of 2P. Then
∫a+MAR;π(ψ)dx=∫2Pπdy, | (4.11) |
if u<+∞ everywhere in the interior of 2P.
The inverse of Lemma 4.8 is also true as an analogue of [17,Theorem 3.6]. In fact, we have
Proposition 4.9. Let ϕ be a K×K-invariant potential. Then ψϕ satisfies (4.11) if and only ifuϕ is finite everywhere in Int(2P).
Proposition 4.9 will be used in the proof of Theorem 4.2 in next subsection.
It is easy to see that (4.1) is equivalent to (4.2). Thus, to prove Theorem 4.2, we only need to show that
ϕ∈E1K×K(M,−KM)⟺∫2P+|uϕ|πdy<+∞. |
The following lemma can be found in [9,Lemma 2.7] (proved in [9,Appendix]).
Lemma 4.10. Let ψ be a convex function on a and uψ its Legendre dual on P.
(1) uψ is differentiable at p if and only if the sup defining uψ is attained at a unique point xp∈a and xp=∇uψ(p);
(2) Suppose that (ψ−ψ0)∈E1K×K(M,−KM). Let p∈P at which uψ is differentiable. Then for any continuous uniformly bounded function von a, it holds
ddt|t=0uψ+tv(p)=−v(∇uψ(p)), | (4.12) |
where uψ+tv is the Legendre function of ψ+tv as in (3.15) which is well-defined since v is continuous and uniformly bounded on a.
Remark 4.11. By Lemma 4.5 and Part (1) in Lemma 4.7, we can prove the following: Let ϕ∈E1K×K(M,−KM), then for any K×K-invariant continuous uniformly bounded function f on G, it holds
∫Mfωnϕ=∫2Pf(∂uϕ)πdy. | (4.13) |
Proof of Theorem 4.2. We will follow the arguments in [17,Proposition 3.9] to prove the theorem.
Necessary part. First we show that ϕ has full MA-mass by Proposition 4.9. In fact, by a result in [36,Lemma 4.5], we see that for any W-invariant convex polytope 2P′⊆2P, there is a constant C=C(P′) such that for any W-invariant convex uϕ≥0,
∫2P′uϕdy≤C∫2Puϕπdy<+∞. |
This implies that uϕ is finite everywhere in Int(2P) by the convexity of uϕ. Thus we get what we want from Proposition 4.9.
Next we prove that ϕ is L1-integrate associated to the MA-measure ωnϕ. Let ψ1=ψ0+ϕ (ϕ may be different to a constant) be satisfying (4.1). We define a distance between ψ0 and ψ1 for p≥1,
dp(ψ0,ψ1)=infϕt∫10(∫M|˙ϕt|pωnϕt)1pdt, |
where ϕt∈E1(M,−KM) (t∈[0,1]) runs over all curves joining 0 and ϕ with ωϕt≥0. Choose a special path ϕt such that the corresponding Legendre functions of ψt=ψ0+ϕt are given by
ut=tu1+(1−t)u0, | (4.14) |
where u1 and u0 are the Legendre functions of ψ1 and ψ0, respectively. Note that by Lemma 4.10,
˙ψt=−˙ut=u0−u1,almosteverywhere. |
Then by Lemma 4.5 (or Remark 4.11), we get
dp(ψ0,ψ1)≤∫10(∫2P+|˙ut|pπdy)1pdt≤C(p)(∫2P+|u1|pπdy)1p+C′(p.ψ0). | (4.15) |
On the other hand, by a result of Darvas-Rubinstein [19], there are uniform constant C0 and C1 such that for any Kähler potential ϕ with full MA-measure it holds,
−∫Mϕωnϕ≤C0d1(ψ0,ψ1)+C1. |
Thus we obtain
−∫Mϕωnϕ≤C. |
Hence, ϕ∈E1K×K(M,−KM).
Sufficient part. Assume that ϕ∈E1K×K(M,−KM). We first deal with the case of ϕ∈L∞(M)∩C∞(G). Then
v2P−C≤ψϕ≤v2P≤ψ0+C, | (4.16) |
and
∇ψϕ:a→2P |
is a bijection. Thus
−ϕ=(ψ0−ψϕ)(∇uϕ)≥v2P(∇uϕ)−ψϕ(∇uϕ)−C2≥−C2. |
Moreover,
(ψ0−ψϕ)(∇uϕ)≥v2P(∇uϕ)−ψϕ(∇uϕ)−C=supy′∈2P⟨∇uϕ,y′⟩−ψ(∇uϕ)−C≥⟨∇uϕ,y⟩−ψ(∇uϕ)−C=uϕ(y)−C. |
Hence,
∫2P+uϕπdy≤∫2P+(ψ0−ψϕ)(∇uϕ)πdy+C=∫M|ϕ|ωnϕ+C<+∞. | (4.17) |
Next for an arbitrary ϕ∈E1K×K(M,−KM), we choose a sequence of smooth K×K-invariant functions {ϕj} decreasing to ϕ
such that ϕj∈C∞(G) and
√−1∂ˉ∂(ψ0+ϕj)>0,inG. |
Then as in (4.17), we have
∫2P+ujπdy≤∫2P+(ψ0−ψj)(∇uj)πdy=∫M|ϕj|ωnj+C, |
where uj is the Legendre function of ψj=ψ0+ϕj. Note that
∫M|ϕj|ωnj→∫M|ϕ|ωnϕ |
and uj↗uϕ. Thus by taking the above limit as j→+∞, we also get (4.17).
In this section, we assume that the associated polytope P is fine. Then by Lemma 3.4, (ψ2P−ψ0)∈E1K×K(M,−KM) is a smooth K×K-invariant Kähler potential on G. It follows that
−logdet(∂ˉ∂ψ2P)−ψ2P=h0 | (5.1) |
gives a Ricci potential h0 of ω2P, which is smooth and K×K-invariant on G.
The following proposition gives an upper bound of h0.
Proposition 5.1. The Ricci potential h0 of ω2P is uniformly bounded from above on G. In particular, eh0 is uniformly bounded on G.
Proof. As in [36,Sections 3.2 and 4.3], the proof is based on a direct computation of asymptotic behavior of h0 near every point of ∂(2P+). Recall that
J(x)=∏α∈Φ+sinh2α(x),x∈aandπ(y)=∏α∈Φ+α2(y),y∈2P. |
Since the Ricci potential of h0 is also K×K-invariant, by (5.1) and (3.9),
h0=−logdet(ψ2P,ij)−ψ2P+logJ(x)−log∏α∈Φ+α2(∇ψ2P)=logdet(u2P,ij)−yiu2P,i+u2P+logJ(∇u2P)−logπ(y). | (5.2) |
Here in the second line we take the Legendre transformation
u2P(y(x))=yi(x)xi−ψ2P(x)andy(x)=∇ψ2P(x). |
Note that
u2P,i=12d0∑A=1(−uiA)(1+loglA),u2P,ij=12d0∑A=1uiAujAlA |
and
logJ(t)=2∑α∈Φ+logsinh(t). |
Thus we have
h0=−d0∑A=1loglA+12d0∑A=1(uiAyi)loglA+2∑α∈Φ+logsinh(−12d0∑A=1α(uA)loglA)−2∑α∈Φ+logα(y)+O(1). | (5.3) |
By (5.3), h0 is locally bounded in the interior of 2P+. Thus we need to prove that h0 is bounded above near each y0∈∂(2P+). There will be three cases as follows.
Case-1. y0∈∂(2P+) and is away from any Weyl wall (see Figure 1).
Note that
logsinh(t)={t+O(1),t→+∞,logt+O(1),t→0+. | (5.4) |
Then we get as y→y0,
∑α∈Φ+logsinh(−12d0∑A=1α(uA)loglA)=−∑Aρ(uA)loglA+O(1). |
By (5.3), it follows that
h0=−∑{A|lA(y0)=0}(1−12yiuiA+2ρiuiA)loglA(y)+O(1). |
However, by Lemma 3.1, we have
h0=−12∑{A|lA(y0)=0}lA(y)loglA(y)+O(1). |
Hence h0 is bounded near y0.
Case-2. y0 lies on some Weyl walls but away from any facet of 2P (see Figure 2).
In this case it is direct to see that h0 is bounded near y0 since
logdet(u2P,ij),yiu2P,i,J(∇u2P)π(y) |
are all bounded.
Case-3. y0 lies on the intersection of ∂(2P) with some Weyl walls. In this case, by (3.1), we rewrite (5.3) as
h0=2d0∑A=1ρA(uA)loglA+2∑α∈Φ+logsinh(−12d0∑A=1α(uA)loglA)−2∑α∈Φ+logα(y)+O(1)=∑α∈Φ+[d0∑A=1|α(uA)|loglA+2logsinh(−12d0∑A=1α(uA)loglA)−2logα(y)]+O(1),y→y0. |
Here we used a fact that
2ρA(uA)=∑α∈Φ+|α(uA)|. |
Set
Iα(y)=d0∑A=1|α(uA)|loglA+2logsinh(−12d0∑A=1α(uA)loglA)−2logα(y) |
for each α∈Φ+. Then
h0(y)=∑α∈Φ+Iα(y)+O(1),y→y0. | (5.5) |
Note that each Iα(y) involves only one root α. Thus, without loss of generality, we may assume that y0 lies on only one Weyl wall.
Assume that y0∈∂(2P)∩Wα0 for some simple Weyl wall Wα0, α0∈Φ+ and it is away from other Weyl walls. Now we estimate each Iα(y) in (5.5). When β≠α0, it is easy to see that
β(y)→cβ>0,asy→y0. |
Then, by (5.4), we have
logsinh(−12d0∑A=1β(uA)loglA)=−12∑{A|lA(y0)=0}β(uA)loglA+O(1),∀β≠α0. |
Note that y0∈{β(y)>0}. Thus any facet FA passing through y0 lies in {β(y)>0} or is orthogonal to Wβ. Since 2P is convex and sβ-invariant, where sβ is the reflection with respect to Wβ, these facets must satisfy
β(uA)≥0. |
Hence, for any β≠α0, we get
Iβ(y)=d0∑A=1|β(uA)|loglA−2d0∑A=1|β(uA)|loglA−2logβ(y)=O(1),asy→y0. | (5.6) |
It remains to estimate the second term in Iα0(y),
logsinh(−12∑Aα0(uA)loglA). | (5.7) |
We first consider a simple case that y0 lies on the intersection of Wα0 with at most two facets of 2P. Then there will be two subcases: y0∈Wα0∩F1 or y0∈Wα0∩F1∩F2, where F1,F2 are two facets of P.
Case-3.1. y0∈Wα0∩F1 is away from other facets of 2P. Then F1 is orthogonal to Wα0 (see Figure 3).
It follows that lA(y0)≠0 for any A≠1. Thus
⟨α0,y⟩=o(lA(y)),y→y0,A≠1. | (5.8) |
Let {F1,...,Fd1} be all facets of P such that α0(uA)≥0,A=1,...,d1. Let sα0 be the reflection with respect to Wα0. Then by sα0-invariance of P, for each A′∉{1,...,d1} there is some A∈{1,...,d1} such that
lA′=lA+2α0(uA)⟨α0,y⟩|α0|2. |
It follows that
α0(∇u2P)=−12d0∑A=1α0(uA)loglA=12d1∑A=2α0(uA)log(1+2α0(uA)⟨α0,y⟩|α0|2lA(y)). |
Thus, by (5.8) and the fact that α0(u1)=0, we obtain
logsinh(−12d0∑A=1α0(uA)loglA)=logsinhd1∑A=2α0(uA)log(1+2α0(uA)⟨α0,y⟩|α0|2lA(y))=log⟨α0,y⟩+O(1). |
Hence
Iα0(y)=O(1),as y→y0. |
Together with (5.6), we see that h0 is bounded near y0.
Case-3.2. y0∈Wα0∩F1∩F2 and is away from other facets of 2P (see Figure 4).
By the W-invariance of 2P, it must hold F1=sα0(F2). We may assume that F2⊆¯a+ and then
l1=l2+2α0(u2)⟨α0,y⟩|α0|2. |
As y→y0 we have
α0(y),l1(y),l2(y)→0,lA(y)↛0,∀A≠1,2. |
It follows that
d0∑A=1|α0(uA)|loglA=α0(u2)(logl1+logl2)+O(1). | (5.9) |
Then the second term in Iα0(y) becomes
logsinh(−12d0∑A=1α0(uA)loglA)=logsinh12[α0(u2)log(1+2α0(u2)⟨α0,y⟩|α0|2l2(y))+d1∑A≠2,α0(uA)>0α0(uA)log(1+2α0(uA)⟨α0,y⟩|α0|2lA(y))]. |
We will settle it down according to the different rate of α0(y)l2(y) below.
Case-3.2.1. α0(y)=o(l2(y)). Then
logsinhα0(∇u2P)=logα0(y)−logl2(y). | (5.10) |
Note that sα0(u1)=u2∈¯a+, we have
∑A=1,2|α0(uA)|loglA=α0(u2)(logl1+logl2). |
Using the above relation, (5.9) and (5.10), we get
Iα0(y)=α0(u2)logl1+(α0(u2)−2)logl2+O(1)=2(α0(u2)−1)logl2+O(1). | (5.11) |
Here we used l1=l2(1+o(1)) in the last equality.
Note that by our assumption α0(u2)>0. Then
α0(u2)≥1, |
since α0(u2)∈Z. Hence, as l1(y),l2(y)→0+, by (5.5), (5.6) and (5.11), we see that h0 is bounded from above in this case.
Case-3.2.2. c≤α0(y)l2(y)≤C for some constants C,c>0. Then
logα0(y)=logl2+O(1),logsinhα0(∇u2P)=O(1) |
and the right hand side of (5.5) becomes
α0(u2)(logl1+logl2)−2logα0(y)+O(1)=2(α0(u2)−1)logl2+O(1). | (5.12) |
Again h0 is also bounded from above.
Case-3.2.3. α0(y)l2(y)→+∞. Then
logsinhα0(∇u2P)=12α0(u2)(logα0(y)−logl2(y)),l1(y)=α0(y)(1+o(1)) |
and the right hand side of (5.5) becomes
α0(u2)(logl1+logl2)+α0(u2)(logα0(y)−logl2(y))−2logα0(y)+O(1)=α0(u2)logl1+[α0(u2)−2]logα0(y)+O(1)=2(α0(u2)−1)logα0(y)+O(1). | (5.13) |
Hence h0 is bounded from above as in Case-3.2.1.
Next we consider the case that there are facets F1...,Fs(s≥3) such that
y0∈Wα0∩F1∩...∩Fs |
and it is away from any other facet of 2P. We only need to control the term (5.7) as above. If F1,...,Fs are all orthogonal to Wα0 as in Case-3.1, we see that h0(y) is uniformly bounded. Otherwise, for any y nearby y0 there is a facet F=Fi′ for some i′∈{1,...,s} such that
li′(y)=min{li(y)|i=1,...,s such that α0(ui)≠0}. |
As y→y0, up to passing to a subsequence, we can fix this i′. Clearly, y0∈Wα0∩F1∩F2 as in Case-3.2, where F2=F⊆¯a+ and F1=sα0(F) for the reflection sα0. Hence by following the argument in Case-3.2, we can also prove that h0(y) is uniformly bounded from above. Therefore, the proposition is true in Case-3. The proof of our proposition is completed.
Remark 5.2. We note that h0 is always uniformly bounded in Case-1, Case-2 and Case-3.1. Furthermore, if rank(G)=2, there are at most two facets F1,F2 intersecting at a same point y0 of Wα0 as in Cases-3.2.1–3.2.3, thus, by the asymptotic expressions of h0 in (5.11), (5.12) and (5.13), respectively, we see that h0 is uniformly bounded if and only if the following relation holds,
α0(u2)=1. | (5.14) |
In other words, in Cases-3.2.1–3.2.3,
limy→y0h0=−∞, |
if (5.14) does not hold.
Remark 5.3. (5.14) always holds when M is a smooth G-compactification. This is because the Guillemin metric can be extended to a global one on M and so h0 is uniformly bounded (cf. [5,Proposition 3.2]). We also note that in this case (5.14) can come from Lemma 6.1 in [39] directly.
By Theorem 4.2, we see that for any u∈E1K×K(2P), its Legendre function
ψu(x)=supy∈2P{⟨x,y⟩−u(y)}≤v2P(x) |
corresponds to a K×K-invariant weak Kähler potential ϕu=ψu−ψ0 which belongs to E1K×K(M,−KM). Here we can choose ψ0 to be the Legendre function ψ2P of Guillemin function u2P as in (3.12). As we know, e−ϕu∈Lp(ω0) for any p≥0. Thus ∫a+e−ψuJ(x)dx is well-defined.
We introduce the following functional on E1K×K(2P) by
D(u)=L(u)+F(u), |
where
L(u)=1V∫2P+uπdy−u(4ρ) |
and
F(u)=−log(∫a+e−ψuJ(x)dx)+u(4ρ). |
It is easy to see that on a smooth Fano compactification of G,
L(uϕ)+uϕ(4ρ)=−1(n+1)Vn∑k=0∫Mϕωkϕ∧ωn−k0 |
and D(uϕ) is just the Ding functional F(ϕ). We note that a similar functional on such Fano manifolds has been studied for Mabuchi solitons in [37,Section 4]). Hence, for convenience, we call D(⋅) the reduced Ding functional on a Q-Fano compactifications of G.
In this section, we will use the variation method to prove Theorem 1.2 by verifying the properness of D(⋅). We assume that the associated polytope P is fine so that the Ricci potential h0 is uniformly bounded above by Proposition 5.1.
In this subsection, we establish a properness criterion for D(uϕ), namely,
Proposition 6.1. Let M be a Q-Fano compactification of G. Suppose that the associated polytope P is fine and satisfies (1.2). Then there are constants δ and Cδ such that
D(u)≥δ∫2P+uπ(y)dy+Cδ,u∈E1K×K(2P). | (6.1) |
The proof follows the line of argument in [37]. We note that uϕ satisfies the normalized condition u≥u(O)=0. Then we have the following estimate for the linear term L(⋅) as in [36,Proposition 4.2].
Lemma 6.2. Under the assumption (1.2), there exists a constant λ>0 such that
L(u)≥λ∫2P+uπ(y)dy,∀u∈E1K×K(2P). | (6.2) |
For the non-linear term F(⋅), we can also get an analogy of [37,Lemma 4.8] as follows.
Lemma 6.3. For any ϕ∈E1K×K(M,−KM), let
˜ψϕ:=ψϕ−4ρixi,x∈a+. |
Then
F(uϕ)=−log(∫a+e−(˜ψϕ−infa+˜ψϕ)∏α∈Φ+(1−e−2αixi2)2dx). | (6.3) |
Consequently, for any c>0,
F(uϕ)≥F(uϕ1+c)−n⋅log(1+c). | (6.4) |
Let ϕ0,ϕ1∈E1K×K(M,−KM) and u0,u1 be two Legendre functions of ψ0+ϕ0 and ψ0+ϕ1, respectively. Let ut (t∈[0,1]) be a linear path connecting u0 to u1 as in (4.14). Then by Theorem 4.2, the corresponding Legendre functions ψt of ut give a path in E1K×K(M,−KM). The following lemma shows that F(ψt) is convex in t.
Lemma 6.4. Let
ˆF(t)=−log∫a+e−ψtJ(x)dx,t∈[0,1]. |
Then ˆF(t) is convex in t and so is F(ψt).
Proof. By definition, we have
ψt(tx1+(1−t)x0)=supy{⟨y,tx1+(1−t)x0⟩−(tu1(y)+(1−t)u0(y))}≤tsupy{⟨y,x1⟩−u1(y)}+(1−t)supy{⟨y,x0⟩−u0(y))}≤tψ1(x1)+(1−t)ψ0(x0),∀x0,x1∈a. | (6.5) |
On the other hand,
logJ(tx1+(1−t)x0)≥tlogJ(x1)+(1−t)logJ(x0),∀x0,x1∈a+. |
Combining these two inequalities, we get
(e−ψtJ)(tx1+(1−t)x0)≥(e−ψ1J)t(x1)(e−ψ0J)1−t(x0),∀x0,x1∈a+. |
Hence, by applying the Prekopa-Leindler inequality to three functions e−ψtJ,e−ψ1J and e−ψ0J (cf. [27,Theorem 7.1]), we prove
−log∫a+e−ψtJ(x)dx≤−tlog∫a+e−ψ1J(x)dx−(1−t)log∫a+e−ψ0J(x)dx. |
This means that ˆF(t) is convex.
Proof of Proposition 6.1. By Proposition 5.1,
A(y)=V∫a+e−ψ0J(x)dxeh0(∇u0(y)) |
is bounded, where y(x)=∇ψ0(x). Then the functional
DA(u)=L0A(u)+F(u), |
is well-defined on E1K×K(2P), where
L0A(u)=1V∫2P+uA(y)π(y)dy−u(4ρ). |
It is easy to see that u0 is a critical point of DA(⋅). On the other hand, by Lemma 6.4, F(⋅) is convex along any path in E1K×K(M,−KM) determined by their Legendre functions as in (4.14). Note that L0A(⋅) is convex in E1K×K(2P). Hence
DA(u)≥DA(u0),∀u∈E1K×K(2P). |
Now together with Lemma 6.2 and Lemma 6.3, we can apply arguments in the proof of [37,Proposition 4.9] to proving that there is a constant C>0 such that for any u∈E1K×K(2P),
D(u)≥Cλ1+C∫2P+uπ(y)dy+DA(u0)−nlog(1+C). |
Therefore, we get (6.1).
Write E1K×K(2P) as
E1K×K(2P)=⋃κ≥0E1K×K(2P;κ), |
where
E1K×K(2P;κ)={u∈E1K×K(2P)|∫2P+uπdy≤κ}. |
By [36,Lemma 6.1] and Fatou's lemma, it is easy to see that any sequence {un}⊆E1K×K(2P;κ) has a subsequence which converges locally uniformly to some u∞∈E1K×K(2P;κ). Thus each E1K×K(2P;κ), and so E1K×K(2P) is complete. Moreover, we have
Proposition 6.5. The reduced Ding functional D(⋅) is lower semi-continuous on the space E1K×K(2P). Namely, for any sequence {un}⊆E1K×K(2P), which converges locally uniformly to some u∞, we have u∞∈E1K×K(2P) and it holds
D(u∞)≤lim infn→∞D(un). | (6.6) |
Proof. By Fatou's lemma, we have
∫2P+u∞πdy≤lim infn→+∞∫2P+unπdy<+∞. | (6.7) |
Then u∞∈E1K×K(2P) and
L(u∞)≤lim infn→+∞L(un). |
It remains to estimate F(u∞). Note that u∞ is finite everywhere in Int(2P) by the locally uniform convergence and its Legendre function ψ∞≤v2P. Thus, for any ϵ0∈(0,1) there is a constant Mϵ0>0 such that (cf. [17,Lemma 2.3]),
ψ∞(x)≥(1−ϵ0)v2P(x)−Mϵ0,∀x∈a. | (6.8) |
On the other hand, the Legendre function ψn of un also converges locally uniformly to ψ∞. Then
∂ψn→∂ψ∞ |
almost everywhere. Since
ψn(O)=ψ∞(O)=0,∀n∈N+, |
we have
ψn(x)≥(1−ϵ0)v2P(x)−Mϵ0,∀x∈a | (6.9) |
as long as n≫1. Note that
0≤J(x)≤e4ρ(x),∀x∈a+. |
By choosing an ϵ0 such that 4ρ∈(1−ϵ0)Int(2P), we get
∫a+eMϵ0−(1−ϵ0)v2P(x)J(x)dx<+∞. |
Hence, combining this with (6.8) and (6.9) and using Fatou's lemma, we derive
−log(∫a+e−ψ∞J(x)dx)≤lim infn→+∞[−log(∫a+e−ψnJ(x)dx)]. |
Therefore, we have proved (6.6) by (6.7).
Now we prove the sufficient part of Theorem 1.2. Suppose that (1.2) holds. Then by Propositions 6.1 and 6.5, there is a minimizing sequence {un} of D(⋅) on E1K×K(2P), which converges locally uniformly to some u⋆∈E1K×K(2P) such that
D(u⋆)≤limu∈E1K×K(2P)D(u). | (6.10) |
Let ψ⋆ be the Legendre function of u⋆. Then by Theorem 4.2, we have
ϕ⋆=ψ⋆−ψ0∈E1K×K(M,−KM). |
We need to show that ϕ⋆ satisfies the Kähler-Einstein equation (2.1).
Proposition 6.6. ϕ⋆ satisfies the Kähler-Einstein equation (2.1).
Proof. Let {ut}t∈[0,1]⊆E1K×K(2P) be a family of convex functions with u0=u⋆ and ψt the corresponding Legendre functions of ut. Then by Part (2) in Lemma 4.10,
˙ψ0=−˙u0,almosteverywhere. |
Note that
∫a+e−ψ⋆J(x)dx=V, |
Thus by (4.11) in Lemma 4.8, we get
ddt|t=0D(ut)=1V∫2P+˙u0πdy+∫a+˙ψ0e−ψ⋆J(x)dxV=1V∫a+˙ψ0[e−ψ⋆J(x)−MAR;π(ψ⋆)]dx. | (6.11) |
For any continuous, compactly supported W-invariant function η∈C0(a), we consider a family of functions u⋆+tη. In general, it may not be convex for t≠0 since u⋆ is just weakly convex. In the following, we use a trick to modify the function D(ut) as in [9,Section 2.6]. Define a family of W-invariant functions by
ˆψt=supϕ∈E1K×K(M,−KM){ψϕ|ψϕ≤ψ⋆+tη}. |
Then it is easy to see that the Legendre function ˆut of ˆψt satisfies
|ˆut−u0|≤C,∀|t|≪1. |
By Theorem 4.2, we see that (ˆψt−ψ0)∈E1K×K(M,−KM). Without loss of generality, we may assume that ˆψt satisfies (4.1).
Let
˜D(t)=L(ˆut)+F(ˆut). |
Then
˜D(0)=D(u⋆) | (6.12) |
and
˜D(t)≥D(u⋆). | (6.13) |
Claim 6.7. L(ˆut)+ˆut(4ρ) is differentiable for t. Moreover,
ddt|t=0(L(ˆut)+ˆut(4ρ))=−1V∫Mηωnϕ⋆. | (6.14) |
To prove this claim, we let a convex function g(t)=ˆut(p) for each fixed p∈2P. Then it has left and right derivatives g′−(t;p),g′+(t;p), respectively. Moreover, they are monotone and g′−(t;p)≤g′+(t;p). Thus, g′−,g′+∈L∞loc. It follows that
ddt|t=τ±∫2P+ˆuτπdy=limτ′→0±1τ′∫2P+(ˆuτ+τ′−ˆuτ)πdy |
and by the Lebesgue monotone convergence theorem,
ddt|t=τ±∫2P+ˆuτπdy=∫2P+g′±(τ;p)πdy. |
Recall that g′−(t;p)=g′+(t;p) holds almost everywhere. Thus we see that
L(ˆut)+u(4ρ)=1V∫2P+ˆutπdy |
is differentiable.
Note that
uˆψt=uψ⋆+tη, |
where uψ⋆+tη is the Legendre function of ψ⋆+tη. It follows from Part (2) in Lemma 4.10 that
˙ˆψ0=−˙u0=η,almosteverywhere. |
Hence by Lemma 4.5 (or Remark 4.11), we get
ddt|t=0(L(ˆut)+ˆut(4ρ))=1V∫2P+˙ˆu0πdy=−1V∫2Pηπdy=−1V∫a+ηMAR;π(ψ0)dx=−1V∫Mηωnϕ⋆, |
where ϕ⋆=ψ∗−ψ0. The claim is proved.
Similar with Claim 6.7, we have
ddt|t=0(F(ˆut)−ˆut(4ρ))=1V∫a+ηe−ψ⋆J(x)dx=∫Gηe−ϕ⋆+h0ωn0. | (6.15) |
Thus, by (6.12)–(6.15), we derive
0=ddt|t=0˜D(t)=1V∫Gη[e−ϕ⋆+h0ωn0−ωnϕ⋆]dx. | (6.16) |
As a consequence,
ωnϕ∗=e−ψ⋆+h0ωn0,inG. |
Therefore, by Lemma 4.5 and KAK-integration formula, we prove that ϕ⋆ satisfies (2.1) on G.
Next we show that ωϕ⋆ can be extended to a singular Kähler-Einstein metric on M. Choose an ϵ0 such that 4ρ∈Int(2(1−ϵ0)P). Since u⋆ is locally uniformly bounded on 2P, there is a constant C⋆>0 such that
ψ⋆≥(1−ϵ0)v2P−C⋆. |
Thus
e−ψ⋆(x)J(x) |
is bounded on a+. Also π(∂ψ⋆) is bounded. Therefore, by (4.11), for any ϵ>0, we can find a neighborhood Uϵ of M∖G such that
|∫Uϵ(ωnϕ⋆−eh0−ψ⋆ωn0)|<ϵ. |
This implies that ϕ⋆ can be extended to be a global solution of (2.1) on M. The proposition is proved.
In this section, we will construct Q-Fano compactifications of SO4(C) as examples and in particular, we will prove Theorem 1.3. Note that in this case rank(G)=2. Thus we can use Theorem 1.2 to verify whether there exists a Kähler-Einstein metric on a Q-Fano SO4(C)-compactification by computing the barycenter of their associated polytopes P+. For convenience, we will work with P+ instead of 2P+ throughout this section. It is easy to see that the condition (1.2) is equivalent to
bar(P+)∈2ρ+Ξ. | (7.1) |
Let
R(t)=(cost−sintsintcost). |
Then we can choose a maximal torus of SO4(C) in GL4(C) as follows,
TC={(R(z1)OOR(z2))|z1,z2∈C}. |
Recall M the lattice of SO4(C)-weights. Denote the basis of N=HomZ(M,Z) by E1,E2 which generates the actions of R(z1) and R(z2). Thus we have the two positive roots in M,
α1=(1,−1),α2=(1,1). |
Also we get
a∗+={(x,y)|−x<y<x},2ρ=(2,0) |
and
2ρ+Ξ={(x,y)|−x+2<y<x−2}. | (7.2) |
In this subsection, we use Lemma 3.1 to exhaust all polytopes associated to the Gorenstein Fano compactifications. Here by Gorenstein, we mean that K−1Mreg can be extended to a holomorphic line bundle on M. In this case, the whole polytope P is a lattice polytope. Also, since 2ρ=(2,0), each outer edge
of P+ must lies on some line
An edge of P+ is called an outer one if it does not lie in any Weyl wall, cf. [36].
lp,q(x,y)=(1+2p)−(px+qy)=0 | (7.3) |
for some coprime pair (p,q). Assume that lp,q≥0 on P. By the convexity and W-invariance of P, (p,q) must satisfy
p≥|q|≥0. |
Let us start at the outer edge F1 of P+ which intersects the Weyl wall
W1={x−y=0}. |
There are two cases: Case-1. F1 is orthogonal to W_1 ; Case-2. F_1 is not orthogonal to W_1 .
Case-1. F_1 is orthogonal to W_1 . Then F_1 lies on
\{(x, y)|\; l_{1, 1}(x, y) = 3-x-y = 0\}. |
Consider the vertex A_1 = (x_1, 3-x_1) of P_+ on this edge and suppose that the other edge F_2 at this point lies on
\{(x, y)|\; l_{p_2, q_2}(x, y) = 0\}. |
Thus
\begin{align} 2p_2+1 = x_1p_2+(3-x_1)q_2, \end{align} | (7.4) |
and by the convexity of P ,
p_2 > q_2\geq0. |
We will have two subcases according to the possible choice A_1 = (2, 1) or (3, 0) .
Case-1.1. A_1 = (2, 1) . Then by (7.4),
2p_2+1 = 2p_2+q_2. |
Thus q_2 = 1 and p_2\geq2 .
On the other hand, l_{p_2, q_2} must pass another lattice point A_2 = (x_2, y_2) as the other endpoint of F_2 . It is direct to see that there are only two possible choices p_2 = 2, 4 and three choices of A_2 = (5, -5) , (3, -1) and (3, -3) .
Case-1.1.1. A_2 = (5, -5) which lies on the other Weyl wall W_2 = \{x+y = 0\} . There can not be any other outer edges of P_+ , and P_+ is given by the first case in Figure 5 (we denote it by P_+^{(1)} ). By Theorem 1.2 (or equivalently (7.1)), this compactification admits no Kähler-Einstein metric.
Case-1.1.2. A_2 = (3, -1) . Then we exhaust the third edge F_3 which lies on
l_{p_3, q_3} = 2p_3+1-p_3x-q_3y, |
so that
\begin{align*} 2p_3+1& = 3p_3-q_3, \\ p_3& > 2q_3\geq0. \end{align*} |
Hence the only possible choice is p_3 = 1, q_3 = 0 and the other endpoint of F_3 is A_3 = (3, -3) . Then P_+ is given by the second case in Figure 5. Again, this compactification admits no Kähler-Einstein metric.
Case-1.1.3. A_2 = (3, -3) which lies on the other Weyl wall W_2 = \{x+y = 0\} . There can not be any other outer edges of P_+ , and P_+ is given by the third one in Figure 5. By Theorem 1.2, this compactification admits no Kähler-Einstein metric.
Case-1.2. A_1 = (3, 0) . By the same exhausting progress as in Case-1.1. There are two possible polytopes P_+ , Case-1.2.1 and Case-1.2.2 (see the first two cases of Figure 6).
Case-1.2.1. This compactification admits no Kähler-Einstein metric.
Case-1.2.2. This compactification admits a Kähler-Einstein metric.
Case-2. F_1 is not orthogonal to W_1 . Then its intersection A_1 = (x_1, x_1) with W_1 is a vertex of P . We see that F_1 lies on l_{p_1, q_1} and
\begin{align*} 2p_1+1& = (p_1+q_1)x_1, \\ p_1& > q_1\geq0, \\ x_1& = 2+\frac{1-2q_1}{p_1+q_1}\in\mathbb N_+. \end{align*} |
So the only choice is
p_1 = 1, q_1 = 0 |
and A_1 = (3, 3) . The only new polytope P_+ is given by the last one of Figure 6, which admits a Kähler-Einstein metric.
It is known that Case-1.1.2, Case-1.2.1 and Case-2 are the only smooth {\rm SO}_4(\mathbb C) -compactifications as shown in [41]. We summarize results of this subsection in Table 1.
\rm Cases. | Edges, except Weyl walls | Volume | KE? | Smoothness |
Case-1.1.1 | 3-x-y=0; 5-2x-y=0 | \frac{411}4 | No | Singular |
Case-1.1.2 | 3-x-y=0; 5-2x-y=0; 3-x=0 | \frac{10751}{180} | No | Smooth |
Case-1.1.3 | 3-x-y=0; 9-4x-y=0 | \frac{16349}{972} | No | Singular |
Case-1.2.1 | 3-x-y=0; 3-x=0 | \frac{1701}{20} | No | Smooth |
Case-1.2.2 | 3-x-y=0; 3-x+y=0 | \frac{81}2 | Yes | Singular |
Case-2 | 3-x=0 | \frac{648}5 | Yes | Smooth |
In general, for a fixed integer m > 0 , it will be hard to give a classification of all \mathbb Q -Fano compactifications such that -mK_X is Cartier. This is because when m is sufficiently divisible, there will be too many repeated polytopes according to Lemma 3.1. In the following, we give a way to exhaust all \mathbb Q -Fano polytopes according to the intersection point of \partial P_+ with x -axis.
We will adopt the notations from the previous subsection. We consider the intersection of P_+ with the positive part of the x -axis, namely (x_0, 0) . Then
x_0\, = \, 2+\frac1{p_0} |
for some p_0\in\mathbb N_+ , and there is an edge which lies on some \{l_{p_0, q_0} = 0\} . Without loss of generality, we may also assume that \{l_{p_0, q_0} = 0\}\cap\{y > 0\}\not = \emptyset . Thus by symmetry, it suffices to consider the case
\begin{align*} p_0\, \geq\, q_0\, \geq\, 0. \end{align*} |
Indeed, by the prime condition, q_0\not = 0, \pm p_0 if p_0\not = 1 . Hence, we may assume
\begin{align} p_0\, > \, q_0\, > \, 0, p_0\, \geq\, 2. \end{align} | (7.5) |
We associate this number p_0 to determine each \mathbb Q -Fano polytope P (and hence \mathbb Q -Fano compactifications of {\rm SO}_4(\mathbb C) ). By the convexity, other edges determined by l_{p, q} must satisfy (see Figure 7 below)
p\, \leq\, p_0, |
since we assume that
\begin{align} P_+\subseteq(\{l_{p_0, q_0}\geq0\}\cap\mathfrak a_+). \end{align} | (7.6) |
Thus, once p_0 is fixed, there are only finitely possible \mathbb Q -Fano compactifications of {\rm SO}_4(\mathbb C) associated to it. In Table 2, we list all possible \mathbb Q -Fano compactifications with p_0\leq2 , and test the existence of Kähler-Einstein metrics on these compactifications. In Appendix 1 we figure out the associated polytopes P of nine non-smooth examples in Table 2.
\rm No. | p_0 | (p, q) of edges, except Weyl walls | Volume | KE? | Smoothness/Multiple |
(1) | 1 | (1, 0) | \frac{648}{5} | Yes | Smooth |
(2) | (1, 0), (1, 1) | \frac{1701}{20} | No | Smooth | |
(3) | (1, -1), (1, 1) | \frac{81}{2} | Yes | Multiple=1 | |
(4) | 2 | (2, 1) | \frac{25000}{243} | No | Multiple=3 |
(5) | (2, 1), (1, 1) | \frac{411}{4} | No | Multiple=1 | |
(6) | (1, 0), (2, 1) | \frac{72728}{1215} | No | Multiple=3 | |
(7) | (2, 1), (1, -1) | \frac{947}{36} | No | Multiple=3 | |
(8) | (2, -1), (2, 1) | \frac{165625}{7776} | No | Multiple=6 | |
(9) | (2, 1), (1, 0), (1, 1) | \frac{10751}{180} | No | Smooth | |
(10) | (2, 1), (1, -1), (1, 1) | \frac{12721}{486} | No | Multiple=1 | |
(11) | (2, 1), (2, -1), (1, 1) | \frac{164609}{7776} | No | Multiple=6 | |
(12) | (2, 1), (2, -1), (1, 1), (1, -1) | \frac{6059}{288} | No | Multiple=6 |
Proof. We introduce some notations for convenience: For any domain \Omega\subset\overline{\mathfrak a_+^*} , define
\begin{aligned} \text{Vol}(\Omega)&: = \int_\Omega\pi dx\wedge dy, \\\bar x(\Omega)&: = \frac1{V(\Omega)}\int_\Omega x\pi dx\wedge dy, \\\bar y(\Omega)&: = \frac1{V(\Omega)}\int_\Omega y\pi dx\wedge dy, \end{aligned} |
and
\bar c(\Omega): = \bar x+\bar y. |
By Theorem 1.2 and (7.2), we have \bar c(P_+) > 2 whenever the \mathbb Q -Fano compactification of {\rm SO}_4(\mathbb C) admits a Kähler-Einstein metric.
Recall the numbers p_0, q_0 introduced in Section 7.2. Consider the line segment I_t cut by P_+ on \{y = x-2t\} for t\geq0 (see Figure 8 below). Set
\bar C(t): = \frac{\int_{I_t}\bar c(y)\pi(y)ds}{\int_{I_t}\pi(y)ds} \; \text{and}\; l(t): = {\int_{I_t}\pi(y)ds}, |
where ds is the standard Lebesgue measure on I_t . Then \bar C(t), l(t) are the mean value of \bar c(\cdot) and length of I_t against the weight \pi , respectively. Also, since P_+ is bounded, there is some 0\leq t_0 < +\infty so that I_t is non-empty only on [0, t_0] . Thus we get
\begin{align} \bar c(P_+)& = \frac{\int_0^{t_0}l(t)\bar C(t)dt}{\int_0^{t_0}l(t)dt}. \end{align} | (7.7) |
On the other hand, for each 0\leq t\leq t_0 , the line segment I_{t} = \{(x_0, -x_0)+s(1, 1)|\; 0\leq s\leq S_t\} , where (x_0, -x_0)\in P_+ and (x_0, -x_0)+S_t(1, 1)\in\partial P , satisfies
\bar C(t) = \frac32S_t\leq \frac32S_0\leq\frac{6p_0+3}{2p_0+2q_0}. |
Here in the first equality we use the relation (7.5), and the second follows from the fact that the endpoint of I_0 can not exceed the intersection point
l_{p_0, q_0}\cap\{t(1, 1)|t > 0\} = \left(\frac{2p_0+1}{p_0+q_0}, \frac{2p_0+1}{p_0+q_0}\right). |
Thus by (7.7), we get
\bar c(P_+)\leq\frac{6p_0+3}{2p_0+2q_0}. |
By the fact \bar c(P_+) > 2 , we derive from the above upper bound of \bar c(P_+) ,
\begin{align} q_0 < \frac12p_0+\frac34. \end{align} | (7.8) |
Then
\begin{align} \text{Vol}(P_+)&\leq\text{Vol}(\{l_{p_0, q_0}\geq0, x\geq y\geq-x\})\\ & = \frac{8(1 + 2 p_0)^6}{45 (p_0^2 - q_0^2)^3}\\ &\leq\frac{8(1 + 2 p_0)^6}{45 (p_0^2 - ((1/2) p_0 + (3/4))^2)^3}. \end{align} | (7.9) |
It turns that for p_0\geq9 ,
\text{Vol}(P_+)\leq\frac{224755712}{4100625}. |
However,
\text{Vol}(P^{(4)}_+) = \frac{1701}{20} > \text{Vol}(P^{(2)}_+) = \frac{10751}{180} > \frac{224755712}{4100625}, |
where \text{Vol}(P^{(2)}_+) and \text{Vol}(P^{(4)}_+) are volumes of polytopes in Case-1.1.2 and Case-1.2.1, respectively. Hence, there is no desired Kähler-Einstein polytope with its volume equals to {\rm Vol}(P^{(2)}_+) or {\rm Vol}(P^{(4)}_+) when p_0\geq9 .
Since q_0\in\mathbb N , we can improve (7.9) to
\begin{aligned}\text{Vol}(P_+) &\leq\frac{8(1 + 2 p_0)^6}{45 (p_0^2 - [(1/2) p_0 + (3/4)]^2)^3}.\end{aligned} |
Here [x] = \max_{n\in\mathbb Z}\{n\leq x\} . By the above estimation, when p_0 = 4, 6, 7, 8 , we have
\begin{align} \text{Vol}(P^{(4)}_+)\, > \, \text{Vol}(P^{(2)}_+)\, > \, \text{Vol}(P_+). \end{align} | (7.10) |
As a consequence, they are not Kähler-Einstein polytopes. Hence, it remains to deal with the cases when p_0 = 3, 5 . In these two cases, we shall rule out polytopes that may not satisfy (7.10).
When p_0 = 5 , there are three possible choices of q_0 , i.e., q_0 = 1, 2, 3 by (7.8). It is easy to see that (7.10) still holds for the first two cases by the second relation in (7.9). Thus we only need to consider all possible polytopes when q_0 = 3 . In this case, \{l_{5, 3} = 0\} is an edge of P_+ .
Case-7.3.1. P_+ has only one outer face which lies on \{l_{5, 3} = 0\} . Then
\text{Vol}(P_+)\, = \, \frac{1771561}{23040}. |
Case-7.3.2. P_+ has two outer edges. Assume that the second one lies on \{l_{p_1, q_1} = 0\} . Then
|q_1|\leq p_1\leq4\text{ or }p_1 = 5, q_1 = -3. |
By a direct computation, we see that (7.10) holds except the following two subcases:
Case-7.3.2.1. p_1 = 4, q_1 = 3 ,
\text{Vol}(P_+)\, = \, \frac{383478671}{5000940}. |
Case-7.3.2.2. p_1 = 2, q_1 = 1 ,
\text{Vol}(P_+)\, = \, \frac{567779}{7680}. |
Case-7.3.3. P_+ has three outer edges. Then P_+ is obtained by cutting one of polytopes in Case-7.3.2 with adding new edge \{l_{p_2, q_2} = 0\} . In fact we only need to consider P_+ obtained by cutting Case-7.3.2.1 and Case-7.3.2.2 above, since it obviously satisfies (7.10) in the other cases. By our construction, we can assume that |q_2|\leq p_2\leq p_1 . The only possible P which does not satisfy (7.10) is the case that p_1 = 4, q_1 = 3 and p_2 = 2, q_2 = 1 . However,
\text{Vol}(P_+)\, = \, \frac{92167583}{1250235}. |
Case-7.3.4. P_+ has four outer edges. We only need to consider P_+ which is obtained by cutting Case-7.3.3 with adding new edge \{l_{p_3, q_3} = 0\} with |q_3|\leq p_3\leq 2 . One can show that all of these possible P_+ satisfy (7.10). Thus we do not need to consider more polytopes with more than four outer edges in case of p_0 = 5 . Hence we conclude that for all polytopes P with p_0 = 5 ,
\text{Vol}(P_+)\neq {\rm Vol}(P^{(2)}_+)\; {\rm or}\; {\rm Vol}(P^{(4)}_+). |
Theorem 1.3 is true when p_0 = 5 .
The case p_0 = 3 can be ruled out in the same way. We only list the exceptional polytopes such that the volumes of P_+ do not satisfy (7.10):
Case-7.3.1'. P_+ has only one outer face \{l_{3, 2} = 0\} . Then
\text{Vol}(P_+)\, = \, \frac{941192}{5625}. |
Case-7.3.2'. P_+ has two outer face \{l_{3, 2} = 0\} and \{l_{2, 1} = 0\} . Then
\text{Vol}(P_+)\, = \, \frac{177064}{1875}. |
In summary, when p_0\ge 3 , the volume of P_+ is not equal to either {\rm Vol}(P^{(2)}_+) or {\rm Vol}(P^{(3)}_+) . Finally by exhausting all possible compactifications for p_0 = 1, 2 (see Table-2), we finish the proof of Theorem 1.3.
Remark 7.1. If P_+ is further symmetric under the reflection with respect to the x -axis, it is easy to see its barycenter is (\bar x(P_+), 0) and
\bar x(P_+)\leq\bar x(\{-x\leq y\leq x, 0\leq x\leq(2+\frac1{p_0})\}) = \frac67(2+\frac1{p_0}). |
Thus a Kähler-Einstein polytope of this type must satisfy
p_0\leq3. |
Yan Li was partially supported by NSFC Grant 12101043 and Beijing Institute of Technology Research Fund Program for Young Scholars. Xiaohua Zhu was partially supported by BJSF Grants Z180004. Both Gang Tian and Xiaohua Zhu were partially supported by National Key R & D Program of China 2020YFA0712800.
The authors declare no conflict of interest.
In this appendix we list all polytopes P_+ of non-smooth \mathbb Q -Fano {\rm SO}_4(\mathbb C) -compactifications with p_0\leq2 , namely, (3)–(8) and (10)–(12) labeled as in Table 2 (see Figure 9 below).
In this appendix, we show that the fine condition in Theorem 1.2 can be dropped by a recent result of Li in [32,Theorem 1.2]. Namely, we can generalize Theorem 1.1 to a \mathbb Q -Fano G -compactification M .
Let \mathbb G be the image of G\times G embedded in {\rm Aut}_0(M) , the identity component of {\rm Aut}(M) . Then the image \mathbb T' of Z(G)\times\{e\} is a subtorus of \mathbb T = Z(\mathbb G) . Let \mathbf{M}^{\rm NA}(\cdot) and \mathbf{J}^{\rm NA}_\mathbb T(\cdot) be the non-Archimedean Mabuchi K-energy and J -functional defined in [13,32], respectively. By [32,Theorem 1.2], it suffices to check that M is \mathbb G -uniformly K-stable under the assumption (1.2). More precisely, by [36,Proposition 4.2] (an analogous version of Lemma 6.2), we prove
Theorem B.1. Suppose that (1.2) holds. Then there is a constant c_0 > 0 such that for any \mathbb G -equivariant normal test configuration (\mathcal X, \mathcal L) of (M, -K_M) , it holds
\begin{align} \mathbf{M}^{\rm NA}(\mathcal X, \mathcal L)\geq c_0\cdot\mathbf{J}^{\rm NA}_\mathbb T(\mathcal X, \mathcal L). \end{align} | (B.1) |
Consequently, Proposition 6.1 holds and so M admits a (singular) Kähler-Einstein metric.
Proof. We need to compute the two functionals \mathbf{M}^{\rm NA}(\cdot) and \mathbf{J}^{\rm NA}_\mathbb T(\cdot) . Note that these two functionals can be computed via the \mathbb C^* -weights on (\mathcal X, \mathcal L) (cf. [13,40]). Moreover, the first one is also same with the CM-weight in [40,44,45] when the central fiber is reduced. In our case for a general \mathbb G -equivariant test configuration, it can be normalized with a reduced central fiber via a base change as follows.
Recall that the \mathbb G -equivariant normal test configurations are in one-one correspondence with W -invariant, convex, piecewise linear functions with rational coefficients on P (cf. [5,Section 2.4]). In particular, when M is a toric manifold with torus action \mathbb T , \mathbb T -equivariant normal test configurations are same with toric degenerations. Thus there is such a function f associated to \pi:(\mathcal X, \mathcal L)\to\mathbb {C} . By a base change z\to z^d on \mathbb C for sufficiently divisible d\in\mathbb N_+ , the normalization
(\mathcal X^{(d)}, \mathcal L^{(d)}): = (\mathcal X, \mathcal L)_{z\to z^d}^{\text{normalization}} |
has a reduced central fibre (cf. [13,Proposition 7.16]). In fact, from the proof of [35,Theorem 4.1], (\mathcal X^{(d)}, \mathcal L^{(d)}) is still a \mathbb G -equivariant normal test configuration associated to df .
By [13,Proposition 2.8], we have
\begin{align} \mathbf{M}^{\rm NA}(\mathcal X^{(d)}, \mathcal L^{(d)}) = {\rm Fut}(\mathcal X^{(d)}, \mathcal L^{(d)}). \end{align} | (B.2) |
Note that \mathbf{M}^{\rm NA}(\cdot) is linear under the base change. It follows
\begin{align*} \mathbf{M}^{\rm NA}(\mathcal X, \mathcal L) = \frac1d{\rm Fut}(\mathcal X^{(d)}, \mathcal L^{(d)}). \end{align*} |
On the other hand, by (3.11) and (3.13) in [36],
{\rm Fut}(\mathcal X^{(d)}, \mathcal L^{(d)}) = \frac d{V}\int_{2P_{+}}\langle y-4\rho, \nabla f\rangle\pi dy. |
This formula was proved by Donaldson for f with integral coefficients on a toric manifold [24,Proposition 4.2.1], but the arguments in his proof do not generalize to general cases.|| Thus
|| In fact, by using (B.3) and [13,40,44,45], Proposition 4.2.1 in [24] is equivalent to that the central fiber of (\mathcal X, \mathcal L) is reduced. Clearly, this condition on the central fiber does not hold in general cases.
\begin{align} \mathbf{M}^{\rm NA}(\mathcal X, \mathcal L) = \frac1{V}\int_{2P_{+}}\langle y-4\rho, \nabla f\rangle\pi dy \end{align} | (B.3) |
Also, by an analogous argument for any toric degeneration on a toric manifold in [29], we can get
\begin{align} \mathbf{J}^{\rm NA}(\mathcal X, \mathcal L) = \frac1{V}\int_{2P_+}(f-\min\limits_{2P_+} f)\pi dy, \end{align} | (B.4) |
and for a twisted test configuration it holds
\begin{align} \mathbf{J}^{\rm NA}(\mathcal X_\xi, \mathcal L_\xi) = \frac1{V}\int_{2P_+}(f-\xi(y)-\min\limits_{2P_+}\{f-\xi(y)\})\pi dy, \end{align} | (B.5) |
where \xi lies in the Lie algebra {\rm Lie}(\mathbb T') of \mathbb T' [29,32].
By [36,Proposition 4.2], we see
\begin{align} &\int_{2P_+}\langle y-4\rho, \nabla u\rangle\pi dy\\ &\geq c_0\int_{\partial(2P_+)}u \langle y, \nu\rangle \pi d\sigma_0, \; \forall\; \text{convex, W -invariant u satisfying (4.2).} \end{align} | (B.6) |
Since u is convex and W -invariant,
\nabla u(y) \in\overline{\mathfrak a_+}, \; \forall y\in P_+. |
It implies that u(ty) is non-decreasing for t\geq0 . Thus by (4.2),
\begin{align*} \int_{2P_+}u\pi dy = &\int_0^1t^{n-1}\left(\int_{\partial (2P_+)}u(ty)\langle y, \nu\rangle \pi d\sigma_0\right) dt\\ \leq&\frac1n\int_{\partial (2P_+)}u(y)\langle y, \nu\rangle \pi d\sigma_0. \end{align*} |
Combining with (B.6), we get
\begin{align} &\int_{2P_+}\langle y-4\rho, \nabla u\rangle\pi dy\\ &\geq c_0\int_{2P_+}u\pi dy, \; \forall\; \text{convex, W -invariant u satisfying (4.2).} \end{align} | (B.7) |
For the function f in (B.3), by the W -invariance, there is always an \xi\in{\rm Lie}(\mathbb T') such that
f_\xi: = f-\xi(y)-\min\{f-\xi(y)\} |
satisfies (4.2). Thus applying (B.7) to f_{\xi} together with (B.3) and (B.5), we obtain
\mathbf{M}^{\rm NA}(\mathcal X, \mathcal L)\geq \frac{c_0}V\int_{2P_+}f_\xi\pi dy = c_0\mathbf{J}^{\rm NA}(\mathcal X_\xi, \mathcal L_\xi)\geq c_0\mathbf{J}^{\rm NA}_{\mathbb T}(\mathcal X, \mathcal L). |
Hence (B.1) holds.
[1] |
M. Abreu, Kähler metrics on toric orbifolds, J. Differential Geom., 58 (2001), 151–187. https://doi.org/10.4310/jdg/1090348285 doi: 10.4310/jdg/1090348285
![]() |
[2] | A. D. Aleksandrov, Additive set-functions in abstract spaces, Matematicheski\check{i} Sbornik, 50 (1940), 307–348. |
[3] |
V. A. Alexeev, M. Brion, Stable reductive varieties I: Affine varieties, Invent. Math., 157 (2004), 227–274. https://doi.org/10.1007/s00222-003-0347-y doi: 10.1007/s00222-003-0347-y
![]() |
[4] |
V. A. Alexeev, M. Brion, Stable reductive varieties II: Projective case, Adv. Math., 184 (2004), 382–408. https://doi.org/10.1016/S0001-8708(03)00164-6 doi: 10.1016/S0001-8708(03)00164-6
![]() |
[5] |
V. A. Alexeev, L. V. Katzarkov, On K-stability of reductive varieties, GAFA, Geom. Funct. Anal., 15 (2005), 297–310. https://doi.org/10.1007/s00039-005-0507-x doi: 10.1007/s00039-005-0507-x
![]() |
[6] |
H. Azad, J. J. Loeb, Plurisubharmonic functions and Kählerian metrics on complexification of symmetric spaces, Indagat. Math., 3 (1992), 385–390. https://doi.org/10.1016/0019-3577(92)90017-F doi: 10.1016/0019-3577(92)90017-F
![]() |
[7] |
R. Bamler, Convergence of Ricci flows with bounded scalar curvature, Ann. Math., 188 (2018), 753–831. https://doi.org/10.4007/annals.2018.188.3.2 doi: 10.4007/annals.2018.188.3.2
![]() |
[8] |
R. Berman, K-stability of \mathbb Q-Fano varieties admitting Kähler-Einstein metrics, Invent. Math., 203 (2015), 973–1025. https://doi.org/10.1007/s00222-015-0607-7 doi: 10.1007/s00222-015-0607-7
![]() |
[9] |
R. Berman, B. Berndtsson, Real Monge-Ampère equations and Kähler-Ricci solitons on toric log Fano varieties, Annales de la Faculté des sciences de Toulouse : Mathématiques, Série 6, 22 (2013), 649–711. https://doi.org/10.5802/afst.1386 doi: 10.5802/afst.1386
![]() |
[10] |
R. Berman, S. Boucksom, P. Eyssidieux, V. Guedj, A. Zeriahi, Kähler-Einstein metrics and the Kähler-Ricci flow on log Fano varieties, J. Reine Angew. Math., 751 (2019), 27–89. https://doi.org/10.1515/crelle-2016-0033 doi: 10.1515/crelle-2016-0033
![]() |
[11] |
R. Berman, S. Boucksom, M. Jonsson, A variational approach to the Yau-Tian-Donaldson conjecture, J. Amer. Math. Soc., 34 (2021), 605–652. https://doi.org/10.1090/jams/964 doi: 10.1090/jams/964
![]() |
[12] |
B. Berndtsson, A Brunn-Minkowski type inequality for Fano manifolds and some uniqueness theorems in Kähler geometry, Invent. Math., 200 (2015), 149–200. https://doi.org/10.1007/s00222-014-0532-1 doi: 10.1007/s00222-014-0532-1
![]() |
[13] |
S. Boucksom, T. Hisamoto, M. Jonsson, Uniform K-stability and asymptotics of energy functionals in Kähler geometry, J. Eur. Math. Soc., 21 (2019), 2905–2944. https://doi.org/10.4171/JEMS/894 doi: 10.4171/JEMS/894
![]() |
[14] |
M. Brion, Groupe de Picard et nombres caractéristiques des variétés sphériques, Duke. Math. J., 58 (1989), 397–424. https://doi.org/10.1215/S0012-7094-89-05818-3 doi: 10.1215/S0012-7094-89-05818-3
![]() |
[15] | M. Brion, Curves and divisors in spherical varieties, In: Algebraic groups and Lie groups, Cambridge Univ. Press, 1997, 21–34. |
[16] | X.-X. Chen, B. Wang, Space of Ricci flows (II)–Part B: Weak compactness of the flows. J. Differential Geom., 116 (2020), 1–123. https://doi.org/10.4310/jdg/1599271253 |
[17] |
D. Coman, V. Guedj, S. Sahin, A. Zeriahi, Toric pluripotential theory, Ann. Pol. Math., 123 (2019), 215–242. https://doi.org/10.4064/ap180409-3-7 doi: 10.4064/ap180409-3-7
![]() |
[18] |
T. Darvas, The Mabuchi geometry of finite energy classes, Adv. Math., 285 (2015), 182–219. https://doi.org/10.1016/j.aim.2015.08.005 doi: 10.1016/j.aim.2015.08.005
![]() |
[19] |
T. Darvas, Y. Rubinstein, Tian's properness conjectures and Finsler geometry of the space of Kähler metrics, J. Amer. Math. Soc., 30 (2017), 347–387. https://doi.org/10.1090/jams/873 doi: 10.1090/jams/873
![]() |
[20] | T. Delcroix. Kähler-Einstein metrics on group compactifications, Geom. Funct. Anal., 27 (2017), 78–129. https://doi.org/10.1007/s00039-017-0394-y |
[21] | T. Delcroix, K-Stability of Fano spherical varieties, Ann Sci Éc Norm Supér., 53 (2020), 615–662. https://doi.org/10.24033/asens.2430 |
[22] |
W.-Y. Ding, Remarks on the existence problem of positive Kähler-Einstein metrics, Math. Ann., 282 (1988), 463–471. https://doi.org/10.1007/BF01460045 doi: 10.1007/BF01460045
![]() |
[23] |
W.-Y. Ding, G. Tian, Kähler-Einstein metrics and the generalized Futaki invariants, Invent. Math., 110 (1992), 315–335. https://doi.org/10.1007/BF01231335 doi: 10.1007/BF01231335
![]() |
[24] |
S. Donaldson, Scalar curvature and stability of toric varieties, J. Differential Geom., 62 (2002), 289–349. https://doi.org/10.4310/jdg/1090950195 doi: 10.4310/jdg/1090950195
![]() |
[25] | S. Donaldson, Interior estimates for solutions of Abreu's equation, Collect. Math., 56 (2005), 103–142. |
[26] |
G. Gagliardi, J. Hofscheier, Gorenstein spherical Fano varieties, Geom. Dedicata, 178 (2015), 111–133. https://doi.org/10.1007/s10711-015-0047-y doi: 10.1007/s10711-015-0047-y
![]() |
[27] |
R. J. Gardner, The Brunn-Minkowski inequality, Bull. Amer. Math. Soc., 39 (2002), 355–405. https://doi.org/10.1090/S0273-0979-02-00941-2 doi: 10.1090/S0273-0979-02-00941-2
![]() |
[28] | S. Helgason, Differential geometry, Lie groups, and symmetric spaces, New York-London: Academic Press, Inc., 1978. http://dx.doi.org/10.1090/gsm/034 |
[29] | T. Hisamoto, Stability and coercivity for toric polarizations, arXiv: 1610.07998v3. |
[30] | A. W. Knapp, Representation theory of semisimple groups, Princeton, NJ: Princeton Univ. Press, 1986. http://dx.doi.org/10.1515/9781400883974 |
[31] | A. W. Knapp, Lie groups beyond an introduction, Boston: Birkhäuser, 2002. http://dx.doi.org/10.1007/978-1-4757-2453-0 |
[32] |
C. Li, G-uniform stability and Kähler-Einstein metrics on Fano varieties, Invent. Math., 227 (2022), 661–744. http://dx.doi.org/10.1007/s00222-021-01075-9 doi: 10.1007/s00222-021-01075-9
![]() |
[33] |
C. Li, G. Tian, F. Wang, On Yau-Tian-Donaldson conjecture for singular Fano varieties, Commun. Pure Appl. Math., 74 (2021), 1748–1800. http://dx.doi.org/10.1002/cpa.21936 doi: 10.1002/cpa.21936
![]() |
[34] | C. Li, G. Tian, F. Wang, The uniform version of Yau-Tian-Donaldson conjecture for singular Fano varieties, Peking Math. J., in press. http://dx.doi.org/10.1007/s42543-021-00039-5 |
[35] | Y. Li, Z.-Y. Li, Equivariant \mathbb R-test configurations and semistable limits of \mathbb Q-Fano group compactifications, arXiv: 2103.06439. |
[36] |
Y. Li, B. Zhou, X.-H. Zhu, K-energy on polarized compactifications of Lie groups, J. Funct. Anal., 275 (2018), 1023–1072. http://dx.doi.org/10.1016/j.jfa.2018.04.009 doi: 10.1016/j.jfa.2018.04.009
![]() |
[37] |
Y. Li, B. Zhou, Mabuchi metrics and properness of modified Ding functional, Pac. J. Math., 302 (2019), 659–692. http://dx.doi.org/10.2140/pjm.2019.302.659 doi: 10.2140/pjm.2019.302.659
![]() |
[38] | Y. Li, G. Tian, X.-H. Zhu, Singular limits of Kähler-Ricci flow on Fano G-manifolds, arXiv: 1807.09167. |
[39] |
Y. Li, X.-H. Zhu, Tian's \alpha_{m, k}^{\hat K}-invariants on group compactifications, Math. Z., 298 (2021), 231–259. https://doi.org/10.1007/s00209-020-02591-9 doi: 10.1007/s00209-020-02591-9
![]() |
[40] |
S. Paul, Hyperdiscriminant polytopes, Chow polytopes, and Mabuchi energy asymptotics, Ann. Math., 175 (2012), 255–296. http://dx.doi.org/10.4007/annals.2012.175.1.7 doi: 10.4007/annals.2012.175.1.7
![]() |
[41] |
A. Ruzzi, Fano symmetric varieties with low rank, Publ. Res. Inst. Math. Sci., 48 (2012), 235–278. http://dx.doi.org/10.2977/PRIMS/69 doi: 10.2977/PRIMS/69
![]() |
[42] |
Y.-L. Shi, X.-H. Zhu, Kähler-Ricci solitons on toric Fano orbifolds, Math. Z., 271 (2012), 1241–1251. http://dx.doi.org/10.1007/s00209-011-0913-8 doi: 10.1007/s00209-011-0913-8
![]() |
[43] |
G. Tian, On Kähler-Einstein metrics on certain Kähler manifolds with C_1(M)>0, Invent. Math., 89 (1987), 225–246. https://doi.org/10.1007/BF01389077 doi: 10.1007/BF01389077
![]() |
[44] |
G. Tian, Kähler-Einstein metrics with positive scalar curvature, Invent. Math., 130 (1997), 1–37. https://doi.org/10.1007/s002220050176 doi: 10.1007/s002220050176
![]() |
[45] | G. Tian, K-stability implies CM-stability, In: Geometry, analysis and probability, Cham: Birkhäuser, 2017,245-261. http://dx.doi.org/10.1007/978-3-319-49638-2_11 |
[46] |
G. Tian, K-stability and Kähler-Einstein metrics, Commun. Pure Appl. Math., 68 (2015), 1085–1156. https://doi.org/10.1002/cpa.21578 doi: 10.1002/cpa.21578
![]() |
[47] |
D. A. Timashëv, Equivariant compactification of reductive groups, Mat. Sb., 194 (2003), 119–146. https://doi.org/10.4213/sm731 doi: 10.4213/sm731
![]() |
[48] |
F. Wang, X.-H. Zhu, Tian's partial C^0-estimate implies Hamilton-Tian's conjecture, Adv. Math., 381 (2021), 107619. http://dx.doi.org/10.1016/j.aim.2021.107619 doi: 10.1016/j.aim.2021.107619
![]() |
[49] |
B. Zhou, X.-H. Zhu, Minimizing weak solutions for Calabi's extremal metrics on toric manifolds, Calc. Var., 32 (2008), 191–217. http://dx.doi.org/10.1007/s00526-007-0136-3 doi: 10.1007/s00526-007-0136-3
![]() |
1. | Yan Li, Zhenye Li, Finiteness of ℚ-Fano Compactifications of Semisimple Groups with Kähler–Einstein Metrics, 2022, 2022, 1073-7928, 11776, 10.1093/imrn/rnab046 | |
2. | Jae-Hyouk Lee, Kyeong-Dong Park, Sungmin Yoo, K-stability of Gorenstein Fano group compactifications with rank two, 2022, 33, 0129-167X, 10.1142/S0129167X22500835 | |
3. | Yuji Sano, Weight Polytopes and Energy Functionals of Toric Varieties, 2023, 2096-6075, 10.1007/s42543-023-00079-z | |
4. | Julie Clutterbuck, Jiakun Liu, Preface to the Special Issue: Nonlinear PDEs and geometric analysis – Dedicated to Neil Trudinger on the occasion of his 80th birthday, 2023, 5, 2640-3501, 1, 10.3934/mine.2023095 | |
5. | Feng Wang, Xiaohua Zhu, Kähler–Ricci Flow on {\textbf{G}}-Spherical Fano Manifolds, 2024, 2096-6075, 10.1007/s42543-024-00088-6 | |
6. | Sungmin Yoo, 2025, Chapter 16, 978-981-96-0446-3, 177, 10.1007/978-981-96-0447-0_16 |
\rm Cases. | Edges, except Weyl walls | Volume | KE? | Smoothness |
Case-1.1.1 | 3-x-y=0; 5-2x-y=0 | \frac{411}4 | No | Singular |
Case-1.1.2 | 3-x-y=0; 5-2x-y=0; 3-x=0 | \frac{10751}{180} | No | Smooth |
Case-1.1.3 | 3-x-y=0; 9-4x-y=0 | \frac{16349}{972} | No | Singular |
Case-1.2.1 | 3-x-y=0; 3-x=0 | \frac{1701}{20} | No | Smooth |
Case-1.2.2 | 3-x-y=0; 3-x+y=0 | \frac{81}2 | Yes | Singular |
Case-2 | 3-x=0 | \frac{648}5 | Yes | Smooth |
\rm No. | p_0 | (p, q) of edges, except Weyl walls | Volume | KE? | Smoothness/Multiple |
(1) | 1 | (1, 0) | \frac{648}{5} | Yes | Smooth |
(2) | (1, 0), (1, 1) | \frac{1701}{20} | No | Smooth | |
(3) | (1, -1), (1, 1) | \frac{81}{2} | Yes | Multiple=1 | |
(4) | 2 | (2, 1) | \frac{25000}{243} | No | Multiple=3 |
(5) | (2, 1), (1, 1) | \frac{411}{4} | No | Multiple=1 | |
(6) | (1, 0), (2, 1) | \frac{72728}{1215} | No | Multiple=3 | |
(7) | (2, 1), (1, -1) | \frac{947}{36} | No | Multiple=3 | |
(8) | (2, -1), (2, 1) | \frac{165625}{7776} | No | Multiple=6 | |
(9) | (2, 1), (1, 0), (1, 1) | \frac{10751}{180} | No | Smooth | |
(10) | (2, 1), (1, -1), (1, 1) | \frac{12721}{486} | No | Multiple=1 | |
(11) | (2, 1), (2, -1), (1, 1) | \frac{164609}{7776} | No | Multiple=6 | |
(12) | (2, 1), (2, -1), (1, 1), (1, -1) | \frac{6059}{288} | No | Multiple=6 |
\rm Cases. | Edges, except Weyl walls | Volume | KE? | Smoothness |
Case-1.1.1 | 3-x-y=0; 5-2x-y=0 | \frac{411}4 | No | Singular |
Case-1.1.2 | 3-x-y=0; 5-2x-y=0; 3-x=0 | \frac{10751}{180} | No | Smooth |
Case-1.1.3 | 3-x-y=0; 9-4x-y=0 | \frac{16349}{972} | No | Singular |
Case-1.2.1 | 3-x-y=0; 3-x=0 | \frac{1701}{20} | No | Smooth |
Case-1.2.2 | 3-x-y=0; 3-x+y=0 | \frac{81}2 | Yes | Singular |
Case-2 | 3-x=0 | \frac{648}5 | Yes | Smooth |
\rm No. | p_0 | (p, q) of edges, except Weyl walls | Volume | KE? | Smoothness/Multiple |
(1) | 1 | (1, 0) | \frac{648}{5} | Yes | Smooth |
(2) | (1, 0), (1, 1) | \frac{1701}{20} | No | Smooth | |
(3) | (1, -1), (1, 1) | \frac{81}{2} | Yes | Multiple=1 | |
(4) | 2 | (2, 1) | \frac{25000}{243} | No | Multiple=3 |
(5) | (2, 1), (1, 1) | \frac{411}{4} | No | Multiple=1 | |
(6) | (1, 0), (2, 1) | \frac{72728}{1215} | No | Multiple=3 | |
(7) | (2, 1), (1, -1) | \frac{947}{36} | No | Multiple=3 | |
(8) | (2, -1), (2, 1) | \frac{165625}{7776} | No | Multiple=6 | |
(9) | (2, 1), (1, 0), (1, 1) | \frac{10751}{180} | No | Smooth | |
(10) | (2, 1), (1, -1), (1, 1) | \frac{12721}{486} | No | Multiple=1 | |
(11) | (2, 1), (2, -1), (1, 1) | \frac{164609}{7776} | No | Multiple=6 | |
(12) | (2, 1), (2, -1), (1, 1), (1, -1) | \frac{6059}{288} | No | Multiple=6 |