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

On the K-stability of blow-ups of projective bundles

Daniel Mallory
Abstract.

We investigate the K-stability of certain blow-ups of 1\mathbb{P}^{1}-bundles over a Fano variety VV, where the 1\mathbb{P}^{1}-bundle is the projective compactification of a line bundle LL proportional to KV-K_{V} and the center of the blow-up is the image along a positive section of a divisor BB also proportional to LL. When VV and BB are smooth, we show that, for B2LB\sim_{\mathbb{Q}}2L, the K-semistability and K-polystability of the blow-up is equivalent to the K-semistability and K-polystability of the log Fano pair (V,aB)(V,aB) for some coefficient aa explicitly computed. We also show that, for BlLB\sim_{\mathbb{Q}}lL, l2l\neq 2, the blow-up is K-unstable.

1. Introduction

Originally introduced in [Tia97, Don02] to provide a criterion for the existence of Kähler-Einstein metrics on Fano manifolds, K-stability has recently been a major area of research over the last decade. In the field of complex geometry, the work of Chen-Donaldson-Sun [CDS15] and Tian [Tia15] proved the Yau-Tian-Donaldson conjecture, showing that a Kähler-Einstein metric exists on a given Fano manifold precisely if and only the manifold is K-polystable. Recently, the algebraic theory of K-stability has seen remarkable progress, culminating in the discovery that K-stability provides a setting in which to construct moduli stacks of K-semistable Fano varieties with corresponding good moduli spaces of K-polystable Fano varieties, called K-moduli (see [Xu21, LXZ22, Xu24]).

Since K-stability provides the framework for a theory of moduli spaces for Fano varieties, a central aspect of research in the field of K-stability is developing methods to determine when a Fano variety is indeed K-stable (or K-polystable, or K-semistable, etc.). Key results towards this end include: Tian’s criterion on α\alpha-invariants of Fano varieties, providing a sufficient condition to show K-polystability ([Tia87, OS12]); the Fujita-Li criterion ([Fuj16, Li17]), relating the Futaki invariants of Donaldson’s algebraic reformulation of K-stability to β\beta-invariants consisting of birational data of the variety (or, more specifically, that of divisors on birational models of the variety); the development of stability thresholds ([BJ20]), relating β\beta-invariants to the so-called δ\delta-invariant of a variety, thus relating K-stability to basis-type divisors, log canonical thresholds, and log canonical places of complements, which in turn led to an “inverse of adjunction”-type theorem by Abban-Zhuang ([AZ22]); and the equivalence of K-poly/semistability to a GG-equivariant setting of the same (with the additional restriction of GG reductive for K-polystability) ([DS16, LX20, LWX21, Zhu21]).

Much of the current literature on determining the K-stability of explicit Fano varieties are in low dimensions (i.e. dimensions 22 and 33). In dimension 2, the K-stability of smooth del Pezzo surfaces is well-investigated (see [Tia90, Che08, PW18]); as is that of singular del Pezzo surfaces (see [MM93, OSS16] for example). In dimension 3, the question of K-stability has been systematically approached via the Mori-Mukai classification of Fano threefolds into 105 deformation families. For example, in [ACC+23] they determine for which of the 105 families is a general member K-polystable. The wall-crossing phenomenon for certain log Fano pairs of dimension at most 33 have been studied (see e.g. [ADL24, ADL23, Zha24, Zha23]). In higher dimensions, there is much less known about the K-stability of explicit Fano varieties. Much of what is known is related to hypersurfaces of projective space (see e.g. [Fuj19a, AZ22, AZ23]).

In this paper, we construct classes of K-polystable Fano varieties from K-polystable log Fano pairs of dimension one less, thus leading to many new examples of K-polystable Fano varieties in higher dimensions. Specifically, we relate the K-stability of Fano varieties that are blow-ups of certain 1\mathbb{P}^{1}-bundles over Fano varieties to the log K-stability of the base of the 1\mathbb{P}^{1}-bundle structure. More explicitly, let VV be a Fano variety of dimension n1n-1, \mathcal{L} an ample line bundle on VV such that rKVr\mathcal{L}\sim_{\mathbb{Q}}-K_{V} for some r>1r>1, and BB an effective divisor on VV such that B2B\sim_{\mathbb{Q}}2\mathcal{L}. Let Y=BlBV(𝒪V)Y=\mathrm{Bl}_{B_{\infty}}\mathbb{P}_{V}(\mathcal{L}\oplus\mathcal{O}_{V}), where BB_{\infty} is the image of BB under some positive section of the 1\mathbb{P}^{1}-bundle structure of V(𝒪V)V\mathbb{P}_{V}(\mathcal{L}\oplus\mathcal{O}_{V})\to V.

This construction of YY, although it may seem rather artificial, appears with some frequency among collections of Fano varieties. For example, smooth del Pezzo surfaces of degree 6 are of this form, as are the smooth members of the families №3.9, №3.19, and №4.2 in the Mori-Mukai classification of Fano threefolds. See Section 5 for more examples of Fano varieties arising from this construction.

Theorem 1.1.

Let VV, BB be as above. Further, let VV and BB be smooth. Then the variety YY as constructed above is K-semistable (resp. K-polystable) if and only if (V,aB)(V,aB) is K-semistable (resp. K-polystable), where

a=a(n,r):=rn+1(r1)n+1(n+1)(r1)n2(n+1)(rn(r1)n)a=a(n,r):=\frac{r^{n+1}-(r-1)^{n+1}-(n+1)(r-1)^{n}}{2(n+1)(r^{n}-(r-1)^{n})}

We note the additional hypothesis of VV and BB being smooth. We believe that this statement holds in more generality, and this will be the focus of a future work that will also explore the implications on the K-moduli of the relevant Fano varieties that such a more general statement would imply.

We note that similar results regarding the K-stability of such varieties were obtained in [CGF+23]; specifically [CGF+23, Theorem 1.10], which gives a numerical criterion for the K-polystability of smooth Casagrande-Druel varieties. They further conjecture that such varieties are K-polystable if and only if the base space and the related double cover are both K-polystable. In the smooth case, we confirm their conjecture as a consequence of Theorem 1.1.

Corollary 1.2.

[CGF+23, Conjecture 1.16] Let V,BV,B be as above. Further, let VV and BB be smooth. Let WVW\to V be the double cover of VV ramified over BB. Suppose that both WW and VV are K-polystable. Then YY is also K-polystable.

Our approach to Theorem 1.1 is as follows: we first reduce the statement to that of 𝕋\mathbb{T}-equivariant K-stability with the standard 𝕋\mathbb{T}-action induced by the 1\mathbb{P}^{1}-bundle structure where 𝕋=𝔾m\mathbb{T}=\mathbb{G}_{m}. Then for the reverse direction of the statement on K-semistability, we use [AZ22, Theorem 3.3] to bound the β\beta-invariants of most 𝕋\mathbb{T}-equivariant divisors over YY (specifically, those with “vertical” centers on YY) and directly compute the β\beta-invariants of the rest (those with “horizontal” centers). For the forward direction, we construct an explicit destabilizing divisor on YY given a destabilizing divisor on (V,aB)(V,aB). For the K-polystability case of the statement, we directly show that, given the K-polystability of (V,aB)(V,aB) (resp. YY), a divisor EE over YY (resp. (V,aB)(V,aB)) with β\beta-invariant 0 must induce a product test configuration.

Theorem 1.1 reduces questions of the K-stability of certain families of Fano varieties to similar questions set in one dimension smaller, in exchange for the added complexity of now dealing with the K-stability of log pairs. However, this added complexity is often already of research interest; as seen in several examples, the K-stability of pairs is worked out in the investigation of the wall-crossing phenomenon of moduli spaces and, from some perspectives, is a main focus of investigating such wall-crossing phenomena. We direct readers towards [ADL24, Zho23] for more on the wall-crossing phenomenon for K-stability.

We also note that, in the construction of YY, whose geometry we explore in Section 3, the choice of BB being \mathbb{Q}-linearly equivalent to 22\mathcal{L} is necessary for YY to be K-semistable. Analysis of the generalization of the construction when BlB\sim_{\mathbb{Q}}l\mathcal{L} for l2l\neq 2 leads to the following:

Theorem 1.3.

Let YY be constructed as above with VV and BB smooth and BlB\sim_{\mathbb{Q}}l\mathcal{L} for 0l<r+1,l20\leq l<r+1,l\neq 2. Then YY is K-unstable. Furthermore, either the strict transform of the image of the positive section containing BB_{\infty}, denoted as V¯\overline{V}_{\infty}, or the strict transform of the zero section, V0V_{0}, is a destabilizing divisor for YY.

We comment that the restriction on ll of 0l<r+10\leq l<r+1 is precisely the range such that YY is also Fano, with the interpretation that for l=0l=0, YY is simply the 1\mathbb{P}^{1}-bundle V(𝒪V)\mathbb{P}_{V}(\mathcal{L}\oplus\mathcal{O}_{V}). The case l=0l=0 was previously known, see [ZZ22, Theorem 1.3]. From Theorem 1.3 we provide a simple construction of a K-unstable Fano family in each dimension. Specifically, by Theorem 1.3, The blow-up of n\mathbb{P}^{n} along a codimension 22 linear subvariety is K-unstable; see Example 6.4.

Finally, during the preparation for our manuscript, we learned from Linsheng Wang that a different proof of Theorem 1.1 can be obtained from combining our Lemma 3.5 and [Wan24, Theorem 1.1].

1.1. Acknowledgements

I would like to thank Yuchen Liu for suggesting the problem and for many helpful conversations regarding it.

The author was partially supported by NSF Grant DMS-2148266 and NSF CAREER Grant DMS-2237139.

2. Preliminaries

First we review several definitions and theorems regarding the K-stability of Fano varieties. We work over \mathbb{C} for the entirety. A log Fano pair is a klt projective pair (X,D)(X,D) such that (KX+D)-(K_{X}+D) is ample. We recall the definitions of β\beta-invariants and product-type divisors; these contain all of the necessary data to define the K-semistability and K-polystability of log Fano pairs.

Definition 2.1 (β\beta-invariant, [Fuj19b], [Li17]).

For (X,D)(X,D) a log Fano pair and EE some divisor over XX (that is to say, EE is a prime divisor on some birational model π:YX\pi:Y\to X of XX), we define the β\beta-invariant βX,D(E)\beta_{X,D}(E) as follows: Let AX,D(E)A_{X,D}(E) denote the log discrepancy of EE with respect to the log pair (X,D)(X,D), and let

SX,D(L;E)=1vol(L)0vol(π(L)tE)𝑑tS_{X,D}(L;E)=\frac{1}{\operatorname{vol}(L)}\int_{0}^{\infty}\operatorname{vol}(\pi^{*}(L)-tE)dt

for LL a big \mathbb{Q}-Cartier divisor on XX where vol(L)=limkdimH0(X,Lk)kn/n!\operatorname{vol}(L)=\lim_{k\to\infty}\frac{\dim H^{0}(X,L^{\otimes k})}{k^{n}/n!} denotes the volume of LL. When L=(KX+D)L=-(K_{X}+D), we will usually omit the line bundle from the notation and simply write SX,D(E)S_{X,D}(E), which is often called the expected order of vanishing of EE with respect to the pair (X,D)(X,D). Then we define βX,D(E):=AX,D(E)SX,D(E)\beta_{X,D}(E):=A_{X,D}(E)-S_{X,D}(E).

Since we often will pass between divisors and divisorial valuations, we wish to mention the definition of the β\beta-invariant of a valuation vv. By a valuation on XX, we mean a valuation v:(X)v:\mathbb{C}(X)^{*}\to\mathbb{R} that is trivial on \mathbb{C}. Since XX is projective, given a valuation vv on XX there will exist a unique point ηX\eta\in X such that v0v\geq 0 on 𝒪X,η\mathcal{O}_{X,\eta} and v>0v>0 on the maximal ideal mX,η𝒪X,ηm_{X,\eta}\subset\mathcal{O}_{X,\eta} called the center of vv on XX. We will denote the closure of the center η\eta of a valuation vv as cX(v)c_{X}(v). For a divisor EE over XX, we will write cX(E)c_{X}(E) for the center of the valuation ordE\operatorname{ord}_{E}.

We will define βX,D(v):=AX,D(v)SX,D(v)\beta_{X,D}(v):=A_{X,D}(v)-S_{X,D}(v), where AX,D(v)A_{X,D}(v) is the log discrepancy of the valuation (see [JM12, Proposition 5.1] and [BdFFU15, Theorem 3.1]), while

SX,D(L;v)=1vol(L)0vol(Ltv)𝑑tS_{X,D}(L;v)=\frac{1}{\operatorname{vol}(L)}\int_{0}^{\infty}\operatorname{vol}(L-tv)dt

where vol(Ltv)=limkdim{sH0(X,Lk)|v(s)kt}kn/n!\operatorname{vol}(L-tv)=\lim_{k\to\infty}\frac{\dim\{s\in H^{0}(X,L^{\otimes k})|v(s)\geq kt\}}{k^{n}/n!}. For EE a divisor over XX, we have that SX,D(L;ordD)=SX,D(L;E)S_{X,D}(L;\operatorname{ord}_{D})=S_{X,D}(L;E).

Definition 2.2 (test configurations).

For a triple (X,D,L)(X,D,L) of a log Fano pair (X,D)(X,D) with XX dimension nn and LL an ample line bundle on XX with Lm(KX+D)L\sim-m(K_{X}+D) for some mm\in\mathbb{N}, a test configuration (𝒳,𝒟,)(\mathcal{X},\mathcal{D},\mathcal{L}) is a triple consisting of:

  • a normal variety 𝒳\mathcal{X};

  • an effective \mathbb{Q}-divisor 𝒟\mathcal{D} on 𝒳\mathcal{X};

  • a flat projective morphism π:(𝒳,𝒟)𝔸1\pi:(\mathcal{X},\mathcal{D})\to\mathbb{A}^{1} such that \mathcal{L} is a π\pi-ample line bundle;

  • a 𝔾m\mathbb{G}_{m}-action on the triple (𝒳,𝒟,)(\mathcal{X},\mathcal{D},\mathcal{L}) such that π\pi is equivariant with respect to the standard multiplicative action of 𝔾m\mathbb{G}_{m} on 𝔸1\mathbb{A}^{1};

  • a 𝔾m\mathbb{G}_{m}-equivariant isomorphism ν:(𝒳𝒳0,𝒟|𝒳𝒳0,|𝒳𝒳0)(X,D,L)×(𝔸1{0})\nu:(\mathcal{X}\setminus\mathcal{X}_{0},\mathcal{D}|_{\mathcal{X}\setminus\mathcal{X}_{0}},\mathcal{L}|_{\mathcal{X}\setminus\mathcal{X}_{0}})\to(X,D,L)\times(\mathbb{A}^{1}\setminus\{0\}).

We say a test configuration is a product test configuration if the above isomorphism ν\nu extends to an isomorphism (𝒳,𝒟,)(X,D,L)×𝔸1(\mathcal{X},\mathcal{D},\mathcal{L})\cong(X,D,L)\times\mathbb{A}^{1}. The restriction of the valuation ord𝒳0\operatorname{ord}_{\mathcal{X}_{0}} to (X)(𝒳)(X)(t)\mathbb{C}(X)\subset\mathbb{C}(\mathcal{X})\cong\mathbb{C}(X)(t) yields a divisorial valuation on XX; we will call the divisor EE such that ordE=ord𝒳0|(X)\operatorname{ord}_{E}=\operatorname{ord}_{\mathcal{X}_{0}}|_{\mathbb{C}(X)} the divisor associated to the test configuration (𝒳,𝒟,)(\mathcal{X},\mathcal{D},\mathcal{L}).

Definition 2.3 (product-type divisors).

Let (X,D)(X,D) be a log Fano pair and let EE be a divisor over XX. Let

Rees(E):=mp{sH0(mKX)|ordE(s)p}\text{Rees}(E):=\bigoplus_{m\in\mathbb{N}}\bigoplus_{p\in\mathbb{Z}}\{s\in H^{0}(-mK_{X})|\operatorname{ord}_{E}(s)\geq p\}

be the Rees algebra of EE. Suppose Rees(E)\text{Rees}(E) is finitely generated (in such case, we call EE a dreamy divisor). We define the test configuration associated to EE to be

TC(E):=ProjX(Rees(E))TC(E):=\text{Proj}_{X}\left(\text{Rees}(E)\right)

We say EE is a product-type divisor over XX if TC(E)TC(E) is a product test configuration.

Definition 2.4.

(K-stability, [Fuj19b], [Li17], [LX20, Theorem E], [LWX21, Theorem 1.4], [Zhu21, Theorem 1.1]) Let (X,D)(X,D) be a log Fano pair. Suppose we have an action of an algebraic torus 𝕋=𝔾mr\mathbb{T}=\mathbb{G}_{m}^{r} on (X,D)(X,D). Then, we say that:

  • (X,D)(X,D) is K-semistable, if, for all 𝕋\mathbb{T}-invariant divisors EE over XX, βX,D(E)0\beta_{X,D}(E)\geq 0.

  • Then (X,D)(X,D) is K-polystable, if XX is K-semistable and, for all 𝕋\mathbb{T}-invariant divisors EE over XX, βX,D(E)=0\beta_{X,D}(E)=0 implies that EE is a product-type divisor.

  • (X,D)(X,D) is K-unstable if it is not K-semistable.

Definition 2.5 (filtrations).

For a finite dimensional vector space VV, we will say a filtration (or \mathbb{R}-filtration) \mathcal{F} on VV is a collection of subspaces λVV\mathcal{F}^{\lambda}V\subseteq V for λ\lambda\in\mathbb{R} such that λ0V=V\mathcal{F}^{\lambda_{0}}V=V for some λ0\lambda_{0}\in\mathbb{R}, λ1V=0\mathcal{F}^{\lambda_{1}}V=0 for some λ1\lambda_{1}\in\mathbb{R}, λλ\lambda\geq\lambda^{\prime} implies λVλV\mathcal{F}^{\lambda^{\prime}}V\subseteq\mathcal{F}^{\lambda}V, and that the collection is left-continuous, which is to say that λϵV=λ\mathcal{F}^{\lambda-\epsilon}V=\mathcal{F}^{\lambda} for all λ\lambda and sufficiently small 0<ϵ0<\epsilon.

A valuation vv on VV has a naturally associated filtration vV\mathcal{F}_{v}V where vλV={xV|v(x)λ}\mathcal{F}_{v}^{\lambda}V=\{x\in V|v(x)\geq\lambda\}. In particular, given a linear system VH0(X,L)V\subseteq H^{0}(X,L) and a divisor DD on XX, there is a naturally associated filtration on VV given by the filtration associated to ordD\operatorname{ord}_{D}.

Given a filtration \mathcal{F} on a vector space VV, we say a basis {s0,,sm}V\{s_{0},\ldots,s_{m}\}\subset V of VV is compatible with \mathcal{F} if, for all λ\lambda\in\mathbb{R}, λV\mathcal{F}^{\lambda}V is spanned by some subset of the sis_{i}.

Next we recall details of δ\delta-invariants, as related to both β\beta-invariants and basis-type divisors, to the end of recalling [AZ22, Theorem 3.3], which will be a key lemma in the proof of our main result.

Definition 2.6 (δ\delta-invariants, [BJ20]).

For a log Fano pair (X,D)(X,D), a subvariety ZXZ\subseteq X, and a big \mathbb{Q}-Cartier divisor LL on XX, we define the δ\delta-invariant of the triple (X,D,L)(X,D,L) along ZZ as

δZ(X,D,L)=infE,ZcX(E)AX,D(E)SX,D(L;E)\delta_{Z}(X,D,L)=\inf_{E,Z\subseteq c_{X}(E)}\frac{A_{X,D}(E)}{S_{X,D}(L;E)}

where the infimum is taken over divisors EE over XX whose centers on XX contain ZZ. As above, when L=(KX+D)L=-(K_{X}+D), we will often omit LL from the notation. Likewise, when Z=XZ=X, we will omit the subscript. From Definition 2.4, we see that (X,D)(X,D) is K-semistable precisely when δ(X,D)1\delta(X,D)\geq 1.

There is an alternate approach to determining δ\delta-invariants due to [BJ20]; towards that end, we recall basis-type divisors.

Definition 2.7.

(basis-type divisors, [FO18, Definition 0.1]) Let LL be a big \mathbb{Q}-Cartier \mathbb{Q}-divisor on (X,D)(X,D), and VV some linear series VH0(X,mL)V\subset H^{0}(X,mL). For a global section sVs\in V, let {s=0}\{s=0\} denote the divisor Div(s)+mL\operatorname{Div}(s)+mL. Then, a basis-sum divisor of VV is a divisor Δ\Delta of the form

Δ=i=0Nm{si=0}\Delta=\sum_{i=0}^{N_{m}}\{s_{i}=0\}

where Nm=h0(X,mL)N_{m}=h^{0}(X,mL) and {s0,sNm}\{s_{0},\ldots s_{N_{m}}\} form a basis of VV. We define a divisor Δ\Delta^{\prime} to be an mm-basis type \mathbb{Q}-divisor of the line bundle LL if Δ=1mNmΔ\Delta^{\prime}=\frac{1}{mN_{m}}\Delta for Δ\Delta a basis-sum divisor of H0(X,mL)H^{0}(X,mL).

Given a valuation vv on XX, we say that an mm-basis type \mathbb{Q}-divisor Δ\Delta is compatible with vv if the basis {s0,sNm}\{s_{0},\ldots s_{N_{m}}\} is compatible with the filtration on VV associated to vv. If v=ordDv=\operatorname{ord}_{D} for a divisor DD over XX, then we will say Δ\Delta is compatible with DD.

Let δZ,m(X,D,L)=sup{t0|(X,D+tΔ) is log canonical near Z}\delta_{Z,m}(X,D,L)=\sup\{t\geq 0|(X,D+t\Delta)\text{ is log canonical near }Z\} where the supremum is over all mm-basis type \mathbb{Q}-divisors Δ\Delta of LL.

By [BJ20, Theorem 4.4], we have that δZ(X,D,L)=limmδZ,m(X,D,L)\delta_{Z}(X,D,L)=\lim_{m\to\infty}\delta_{Z,m}(X,D,L), thus allowing us to approach K-stability via basis-type divisors.

We also can approximate SX,DS_{X,D}. Let

Sm(L;v)=supDv(D)S_{m}(L;v)=\sup_{D}v(D)

where the supremum is taken over mm-basis type \mathbb{Q}-divisors of LL. Then, by [BJ20, Corollary 3.6], we have that SX,D(L;E)=limmSm(L;E)S_{X,D}(L;E)=\lim_{m\to\infty}S_{m}(L;E).

Definition 2.8.

(multi-graded linear systems, [AZ22, Definition 2.11]) Let L1,LkL_{1},\ldots L_{k} be big \mathbb{Q}-Cartier \mathbb{Q}-divisors on (X,D)(X,D). A multi-graded linear system VV_{\vec{\bullet}} graded by k\mathbb{N}^{k} on (X,D)(X,D) consists of a set of subspaces

VuH0(X,u1L1+ukLk)V_{\vec{u}}\subset H^{0}(X,u_{1}L_{1}+\ldots u_{k}L_{k})

for all uk\vec{u}\in\mathbb{N}^{k} such that V0=V_{\vec{0}}=\mathbb{C}, and, for any u1,u2\vec{u}_{1},\vec{u}_{2}\in\mathbb{N}, Vu1Vu2Vu1+u2V_{\vec{u_{1}}}\cdot V_{\vec{u_{2}}}\subset V_{\vec{u_{1}}+\vec{u_{2}}}. Denote

Nm=u{m}×rh0(X,Vu)N_{m}=\sum_{\vec{u}\in\{m\}\times\mathbb{N}^{r}}h^{0}(X,V_{\vec{u}})

We can now analogously define an mm-basis type \mathbb{Q}-divisor of a Nk+1N^{k+1}-graded linear system VV_{\vec{\bullet}} to be a divisor Δ\Delta of the form

Δ=1mNmu{m}×rΔu\Delta=\frac{1}{mN_{m}}\sum_{\vec{u}\in\{m\}\times\mathbb{N}^{r}}\Delta_{\vec{u}}

where Δu\Delta_{\vec{u}} is a basis-sum divisor of VuV_{\vec{u}}. We also analogously define Sm(V;E)S_{m}(V_{\vec{\bullet}};E), SX,D(V;E)S_{X,D}(V_{\vec{\bullet}};E), δm(X,D,V)\delta_{m}(X,D,V_{\vec{\bullet}}) and δ(X,D,V)\delta(X,D,V_{\vec{\bullet}}), and note that, for a big line bundle LL, we have that

δ(X,D,W)=δ(X,D,L)\delta(X,D,W_{\bullet})=\delta(X,D,L)

for the \mathbb{N}-graded linear system WW_{\bullet} with Wm=H0(X,Lm)W_{m}=H^{0}(X,L^{\otimes m}).

Definition 2.9.

(refinements of linear systems, [AZ22, Example 2.1]) Let (X,D)(X,D) be a log Fano pair, VV_{\vec{\bullet}} an k\mathbb{N}^{k}-graded linear system on XX for some big \mathbb{Q}-Cartier \mathbb{Q}-divisors L1,LkL_{1},\ldots L_{k}, and EY𝜋XE\subset Y\xrightarrow{\pi}X a divisor over XX. We define the refinement of VV by EE to be the k+1\mathbb{N}^{k+1}-graded linear series WW_{\vec{\bullet}} associated to the big \mathbb{Q}-Cartier \mathbb{Q}-divisors L1|E,Lk|E,E|EL_{1}|_{E},\ldots L_{k}|_{E},-E|_{E}, with

Wu,j\displaystyle W_{\vec{u},j} =Im(jVuH0(X,π(u1L1++ukLk)jE)\displaystyle=\text{Im}\left(\mathcal{F}_{j}V_{\vec{u}}\to H^{0}(X,\pi^{*}(u_{1}L_{1}+\ldots+u_{k}L_{k})-jE)\right.
𝜌H0(E,π(u1L1++ukLk)|EjE|E))\displaystyle\left.\xrightarrow{\rho}H^{0}(E,\pi^{*}(u_{1}L_{1}+\ldots+u_{k}L_{k})|_{E}-jE|_{E})\right)

where the map ρ\rho is the restriction map on global sections.

The invariant data of a multi-graded linear system has a close relation with that of its refinement, as is seen from  [AZ22, Theorem 3.3]; this theorem will in turn be a key lemma in the argument for the semistable version of Theorem 1.1.

Theorem 2.10.

[AZ22, Theorem 3.3] Let (X,D)(X,D) be a log pair, VV_{\vec{\bullet}} a multi-graded linear system on XX, EY𝜋XE\subset Y\xrightarrow{\pi}X a primitive divisor over XX, and ZXZ\subset X a subvariety. Let Z0Z_{0} denote an irreducible component of ZcX(E)Z\cap c_{X}(E), let DYD_{Y} denote the strict transform of DD on YY, and let DFD_{F} denote the different (that is, the divisor on EE such that (KY+DY+E)|E=KE+DE(K_{Y}+D_{Y}+E)|_{E}=K_{E}+D_{E}. Let WW_{\vec{\bullet}} denote the refinement of VV_{\vec{\bullet}} by EE. Then, if ZcX(E)Z\subset c_{X}(E), we have the following inequality of δ\delta-invariants

δZ(X,D,V)min{AX,D(E)SX,D(V),infZδZ(E,DE,W)}\delta_{Z}(X,D,V_{\vec{\bullet}})\geq\min\left\{\frac{A_{X,D}(E)}{S_{X,D}(V_{\vec{\bullet}})},\inf_{Z^{\prime}}\delta_{Z^{\prime}}(E,D_{E},W_{\vec{\bullet}})\right\}

and, if ZcX(E)Z\not\subset c_{X}(E),

δZ(X,D,V)infZδZ(E,DE,W)\delta_{Z}(X,D,V_{\vec{\bullet}})\geq\inf_{Z^{\prime}}\delta_{Z^{\prime}}(E,D_{E},W_{\vec{\bullet}})

with both infima over subvarieties ZYZ^{\prime}\subset Y such that π(Z)=Z0\pi(Z^{\prime})=Z_{0}.

Due to the natural torus action on 1\mathbb{P}^{1}-bundles, we also recall the equivalence between K-stability and the notion of 𝕋\mathbb{T}-equivariant K-stability.

Theorem 2.11.

[LX20, Theorem E], [LWX21, Theorem 1.4], [Zhu21, Theorem 1.1] Let (X,D)(X,D) be a log Fano pair with a 𝕋\mathbb{T}-action for 𝕋𝔾mr\mathbb{T}\cong\mathbb{G}_{m}^{r} a torus group. Then the K-semistability (resp. K-polystability) of (X,D)(X,D) is equivalent to the 𝕋\mathbb{T}-equivariant K-semistability (resp 𝕋\mathbb{T}-equivariant K-polystability) of (X,D)(X,D).

3. Blow-ups of Projective Compactifications of Proportional Line Bundles

Let VV be a smooth, (n1)(n-1)-dimensional Fano variety. We construct an nn-dimensional Fano variety from the data of VV, a choice of an ample line bundle \mathcal{L} on VV such that rKVr\mathcal{L}\sim_{\mathbb{Q}}-K_{V} for some proportionality constant r>1r>1, and a divisor BB on VV such that BB is smooth and BlB\sim_{\mathbb{Q}}l\mathcal{L} for some proportionality constant 0l<r+10\leq l<r+1.

Firstly we consider XX, the projective compactification of the total space of the line bundle \mathcal{L}. XX can be geometrically realized as the projectivization (𝒪V)\mathbb{P}(\mathcal{L}\oplus\mathcal{O}_{V}) of the total space of the vector bundle 𝒪V\mathcal{L}\oplus\mathcal{O}_{V}. The morphism from the vector bundle structure 𝒪VV\mathcal{L}\oplus\mathcal{O}_{V}\to V induces a morphism ϕ:XV\phi:X\to V, giving XX the structure of a 1\mathbb{P}^{1}-bundle over VV. With this structure, we denote the image of the zero section as V0V_{0}. We have the following formula for the anti-canonical divisor KX-K_{X} on XX in relation to that of VV:

KX=2H+ϕ(KVdet(𝒪V))=2H+r1rϕ(KV)-K_{X}=2H+\phi^{*}\left(-K_{V}-\text{det}(\mathcal{L}\oplus\mathcal{O}_{V})\right)=2H+\frac{r-1}{r}\phi^{*}(-K_{V})

where HH denotes the relative hyperplane section from the 1\mathbb{P}^{1}-bundle structure of XX. Thus, we see that XX is also Fano exactly when r>1r>1. We also compute the anti-canonical volume of XX, that is, the top intersection power of KX-K_{X}, as a function of r,n,r,n, and the anti-canonical volume of VV:

Lemma 3.1.

XX is a Fano variety. Moreover, the anti-canonical volume of XX is

vol(X):=(KX)n=(r+1)n(r1)nrn1vol(V).\operatorname{vol}(X):=(-K_{X})^{n}=\frac{(r+1)^{n}-(r-1)^{n}}{r^{n-1}}\operatorname{vol}(V).
Proof.

The anti-canonical divisor of XX is

KX=2H+ϕ(KV)det(𝒪V)-K_{X}=2H+\phi^{*}(-K_{V})-\text{det}(\mathcal{O}_{V}\oplus\mathcal{L})

where HH is the relative hyperplane section. Since 1rKV\mathcal{L}\sim\frac{-1}{r}K_{V}, the above simplifies to

KX=2H+r1rϕ(KV)-K_{X}=2H+\frac{r-1}{r}\phi^{*}(-K_{V})

We also have that H=V0+ϕ()H=V_{0}+\phi^{*}(\mathcal{L}) and thus

H2\displaystyle H^{2} =V0H+ϕ()H\displaystyle=V_{0}\cdot H+\phi^{*}(\mathcal{L})\cdot H
=ϕ()H\displaystyle=\phi^{*}(\mathcal{L})\cdot H

and, more generally,

Hk=ϕ()k1HH^{k}=\phi^{*}(\mathcal{L})^{k-1}\cdot H

From this, we calculate (KX)n(-K_{X})^{n}

(KX)n\displaystyle(-K_{X})^{n} =[2H+r1rϕ(KV)]n\displaystyle=\left[2H+\frac{r-1}{r}\phi^{*}(-K_{V})\right]^{n}
=k=0n(nk)2nk(r1r)kHnkϕ(KV)k\displaystyle=\sum_{k=0}^{n}\binom{n}{k}2^{n-k}(\frac{r-1}{r})^{k}H^{n-k}\cdot\phi^{*}(-K_{V})^{k}
=k=0n1(nk)2nk(r1r)kϕ()nk1Hϕ(KV)k\displaystyle=\sum_{k=0}^{n-1}\binom{n}{k}2^{n-k}(\frac{r-1}{r})^{k}\phi^{*}(\mathcal{L})^{n-k-1}\cdot H\cdot\phi^{*}(-K_{V})^{k}
=k=0n1(nk)2nk(r1r)k(1r)nk1ϕ(KV)nk1H(KV)k\displaystyle=\sum_{k=0}^{n-1}\binom{n}{k}2^{n-k}(\frac{r-1}{r})^{k}(\frac{1}{r})^{n-k-1}\phi^{*}(-K_{V})^{n-k-1}\cdot H\cdot(-K_{V})^{k}
=k=0n1(nk)2nk(r1)krn1Hϕ(KV)n1\displaystyle=\sum_{k=0}^{n-1}\binom{n}{k}2^{n-k}\frac{(r-1)^{k}}{r^{n-1}}H\cdot\phi^{*}(-K_{V})^{n-1}
=(2+r1)n(r1)nrn1vol(V)\displaystyle=\frac{(2+r-1)^{n}-(r-1)^{n}}{r^{n-1}}\operatorname{vol}(V)

In particular, we see that KX-K_{X} has strictly positive top intersection power.

Now, we consider the intersection of KX-K_{X} with effective classes of curves on XX. In particular, since KX-K_{X} is invariant under the 𝕋\mathbb{T}-action of the 1\mathbb{P}^{1}-bundle, we consider intersection products of KX-K_{X} with effective classes of curves fixed by the 𝕋\mathbb{T}-action. Since 𝕋\mathbb{T} is 11 dimensional, a 𝕋\mathbb{T}-fixed curve is either the closure of a single orbit or a curve that is fixed point-wise. The closure of a single orbit would be a fibre under the 1\mathbb{P}^{1}-bundle structure; a curve that is fixed point-wise would be contained in the fixed locus of the 𝕋\mathbb{T}-action, which is V0VV_{0}\sqcup V_{\infty}. Thus, we concern ourselves with the following classes of curves: curves c0V0c_{0}\subset V_{0}, curves cVc_{\infty}\subset V_{\infty}, and fibres ff of the 1\mathbb{P}^{1}-bundle structure. We have the following intersection products:

KXc0\displaystyle-K_{X}\cdot c_{0} =(2V+r1rϕ(KV))c0=r1rϕ(KV)c0>0\displaystyle=(2V_{\infty}+\frac{r-1}{r}\phi^{*}(-K_{V}))\cdot c_{0}=\frac{r-1}{r}\phi^{*}(-K_{V})\cdot c_{0}>0
KXc\displaystyle-K_{X}\cdot c_{\infty} =(2V0+r+1rϕ(KV))c=r+1rϕ(KV)c>0\displaystyle=(2V_{0}+\frac{r+1}{r}\phi^{*}(-K_{V}))\cdot c_{\infty}=\frac{r+1}{r}\phi^{*}(-K_{V})\cdot c_{\infty}>0
KXf\displaystyle-K_{X}\cdot f =(2H+r1rϕ(KV))f=2Hf>0\displaystyle=(2H+\frac{r-1}{r}\phi^{*}(-K_{V}))\cdot f=2H\cdot f>0

Thus, we see KX-K_{X} is strictly nef, and thus the computation of the top intersection power of KX-K_{X} shows that it is big. Thus, by the Basepoint-free Theorem (see [KM98, Theorem 3.3]), KX-K_{X} is ample. ∎

With XX Fano, it is natural to question whether XX is K-semistable. In this case, a quick computation of a specific β\beta-invariant shows that such XX is always K-unstable. This was previously shown in [ZZ22]; however, we include the computation of the β\beta-invariant here as it is indicative of the general procedure we will apply in Section 6 to show various blow-ups of XX are also K-unstable.

Proposition 3.2.

XX is KK-unstable, with β(V0)<0\beta(V_{0})<0.

Proof.
βX(V0)=AX(V0)SX(V0)=1SX(V0)\displaystyle\beta_{X}(V_{0})=A_{X}(V_{0})-S_{X}(V_{0})=1-S_{X}(V_{0})

Thus it suffices to show that SX(V0)>1S_{X}(V_{0})>1. We note that V0V_{0} has pseudo-effective threshold τ=2\tau=2, and that for 0t20\leq t\leq 2, KXtV0-K_{X}-tV_{0} is nef, thus:

SX(V0)\displaystyle S_{X}(V_{0}) =1vol(X)02(KXtV0)n𝑑t\displaystyle=\frac{1}{\operatorname{vol}(X)}\int_{0}^{2}(-K_{X}-tV_{0})^{n}dt
=1vol(X)02vol(V)rn1((r+1)n(r+t1)n)𝑑t\displaystyle=\frac{1}{\operatorname{vol}(X)}\int_{0}^{2}\frac{\operatorname{vol}(V)}{r^{n-1}}\left((r+1)^{n}-(r+t-1)^{n}\right)dt
=vol(V)rn1vol(X)(2(r+1)n(r+1)n+1(r1)n+1n+1)\displaystyle=\frac{\operatorname{vol}(V)}{r^{n-1}\operatorname{vol}(X)}\left(2(r+1)^{n}-\frac{(r+1)^{n+1}-(r-1)^{n+1}}{n+1}\right)
=1(r+1)n(r1)n(2(r+1)n(r+1)n+1(r1)n+1n+1)\displaystyle=\frac{1}{(r+1)^{n}-(r-1)^{n}}\left(2(r+1)^{n}-\frac{(r+1)^{n+1}-(r-1)^{n+1}}{n+1}\right)
=1(n+1)((r+1)n(r1)n)(f(r1)f(r+1))+1\displaystyle=\frac{1}{(n+1)((r+1)^{n}-(r-1)^{n})}\left(f(r-1)-f(r+1)\right)+1

Where f(x)=xn+1(n+1)xnf(x)=x^{n+1}-(n+1)x^{n} which is strictly decreasing on the range 0<x<n0<x<n, thus showing the first term is positive and thus SX(V0)>1S_{X}(V_{0})>1 as desired. ∎

Due to the 1\mathbb{P}^{1}-bundle structure of ϕ:XV\phi:X\to V, we have a natural 𝕋\mathbb{T}-action on XX where 𝕋𝔾m\mathbb{T}\cong\mathbb{G}_{m} acts fibre-wise, and on each fiber of ϕ\phi, 𝕋\mathbb{T} acts with the standard 𝔾m\mathbb{G}_{m}-action on 1\mathbb{P}^{1}. Given this action, XX has two divisors that are point-wise fixed by 𝕋\mathbb{T}: V0V_{0} and the image of a positive section s:VXs_{\infty}:V\to X of the 1\mathbb{P}^{1}-bundle structure, which we will denote as VV_{\infty} (hereafter also referred to as the infinity section).

We continue with the construction. We label the the image of the divisor BB under ss_{\infty} as BB_{\infty}. This image is a codimension 22 subvariety of XX isomorphic to BB since s\mathrm{s}_{\infty} is an embedding. We denote the blow-up of XX along BB_{\infty} as π:Y=BlBXX\pi:Y=\mathrm{Bl}_{B_{\infty}}X\to X, and the exceptional divisor as EE. For ease of notation, we also denote the strict transform of the pullback of BB along ϕ\phi as F:=ϕ1(B)F:=\phi^{-1}_{*}(B) and the strict transform of VV_{\infty} as V¯:=ϕ1(V)\overline{V}_{\infty}:=\phi^{-1}_{*}(V_{\infty}). By abuse of notation, we will denote the pullbacks π(H)\pi^{*}(H) and π(V0)\pi^{*}(V_{0}) as HH and V0V_{0} respectively. We note that YVY\to V has the structure of a conic bundle, and the fibers are reducible conics precisely over BVB\subset V. We note that, since BB_{\infty} is fixed under the 𝕋\mathbb{T}-action on XX, said action lifts to an action on YY, with the locus of fixed points of the 𝕋\mathbb{T}-action is precisely V¯V0(EF)\overline{V}_{\infty}\sqcup V_{0}\sqcup(E\cap F).

Lemma 3.3.

YY is a Fano variety for 0l<r+10\leq l<r+1 (where we interpret l=0l=0 to mean Y=XY=X). Moreover, the anti-canonical volume of YY is

vol(Y)={(rn(r+1l)nl1+rn(r1)n)vol(V)rn1,if l1(nrn1+rn(r1)n)vol(V)rn1,if l=1\operatorname{vol}(Y)=\begin{cases}\left(\frac{r^{n}-(r+1-l)^{n}}{l-1}+r^{n}-(r-1)^{n}\right)\frac{\operatorname{vol}(V)}{r^{n-1}},&\text{if }l\neq 1\\ \left(nr^{n-1}+r^{n}-(r-1)^{n}\right)\frac{\operatorname{vol}(V)}{r^{n-1}},&\text{if }l=1\end{cases}

In particular, when l=2l=2, the anti-canonical volume of YY is

vol(Y)=2(rn(r1)n)rn1vol(V).\operatorname{vol}(Y)=\frac{2(r^{n}-(r-1)^{n})}{r^{n-1}}\operatorname{vol}(V).
Proof.
(KY)n\displaystyle(-K_{Y})^{n} =(2πH+r1rπϕ(KV)E)n\displaystyle=(2\pi^{*}H+\frac{r-1}{r}\pi^{*}\phi^{*}(-K_{V})-E)^{n}
=(πH+r1rπϕ(KV)+V¯)n\displaystyle=(\pi^{*}H+\frac{r-1}{r}\pi^{*}\phi^{*}(-K_{V})+\overline{V}_{\infty})^{n}
=(πV0+πϕ(KV)+V¯)n\displaystyle=(\pi^{*}V_{0}+\pi^{*}\phi^{*}(-K_{V})+\overline{V}_{\infty})^{n}
=(πV0+πϕ(KV))n+(πϕ(KV)+V¯)n(πϕ(KV))n\displaystyle=(\pi^{*}V_{0}+\pi^{*}\phi^{*}(-K_{V}))^{n}+(\pi^{*}\phi^{*}(-K_{V})+\overline{V}_{\infty})^{n}-(\pi^{*}\phi^{*}(-K_{V}))^{n}
=(V0+πϕ(KV))n+(πϕ(KV)+V¯)n\displaystyle=(V_{0}+\pi^{*}\phi^{*}(-K_{V}))^{n}+(\pi^{*}\phi^{*}(-K_{V})+\overline{V}_{\infty})^{n}

with V0,V¯V_{0},\overline{V}_{\infty} as denote above. Now, computing monomial products, we have

V0kπϕ(KV)nk\displaystyle V_{0}^{k}\cdot\pi^{*}\phi^{*}(-K_{V})^{n-k} =V0k1V0πϕ(KV)nk\displaystyle=V_{0}^{k-1}\cdot V_{0}\cdot\pi^{*}\phi^{*}(-K_{V})^{n-k}
=(H1rπϕ(KV))k1V0πϕ(KV)nk\displaystyle=(H-\frac{1}{r}\pi^{*}\phi^{*}(-K_{V}))^{k-1}\cdot V_{0}\cdot\pi^{*}\phi^{*}(-K_{V})^{n-k}
=(1rπϕ(KV))k1V0πϕ(KV)nk\displaystyle=(-\frac{1}{r}\pi^{*}\phi^{*}(-K_{V}))^{k-1}\cdot V_{0}\cdot\pi^{*}\phi^{*}(-K_{V})^{n-k}
=(1r)k1vol(V)\displaystyle=(\frac{-1}{r})^{k-1}\operatorname{vol}(V)

and, similarly,

V¯kπϕ(KV)nk\displaystyle\overline{V}_{\infty}^{k}\cdot\pi^{*}\phi^{*}(-K_{V})^{n-k} =V¯k1V¯πϕ(KV)nk\displaystyle=\overline{V}_{\infty}^{k-1}\cdot\overline{V}_{\infty}\cdot\pi^{*}\phi^{*}(-K_{V})^{n-k}
=(V0+1rπϕ(KV)E)k1V¯πϕ(KV)nk\displaystyle=(V_{0}+\frac{1}{r}\pi^{*}\phi^{*}(-K_{V})-E)^{k-1}\cdot\overline{V}_{\infty}\cdot\pi^{*}\phi^{*}(-K_{V})^{n-k}
=(1rπϕ(KV)E)k1V¯πϕ(KV)nk\displaystyle=(\frac{1}{r}\pi^{*}\phi^{*}(-K_{V})-E)^{k-1}\cdot\overline{V}_{\infty}\cdot\pi^{*}\phi^{*}(-K_{V})^{n-k}
=(1lr)k1vol(V)\displaystyle=(\frac{1-l}{r})^{k-1}\operatorname{vol}(V)

Substituting yields

(KY)n\displaystyle(-K_{Y})^{n} =i=1n(ni)V¯iπϕ(KV)ni+j=1n(nj)V0jπϕ(KV)nj\displaystyle=\sum_{i=1}^{n}\binom{n}{i}\overline{V}_{\infty}^{i}\pi^{*}\phi^{*}(-K_{V})^{n-i}+\sum_{j=1}^{n}\binom{n}{j}V_{0}^{j}\pi^{*}\phi^{*}(-K_{V})^{n-j}
=i=1n(ni)(1lr)i1vol(V)+j=1n(nj)(1r)j1vol(V)\displaystyle=\sum_{i=1}^{n}\binom{n}{i}(\frac{1-l}{r})^{i-1}\operatorname{vol}(V)+\sum_{j=1}^{n}\binom{n}{j}(\frac{-1}{r})^{j-1}\operatorname{vol}(V)
={(rn(r+1l)nl1+rn(r1)n)vol(V)rn1,if l1(nrn1+rn(r1)n)vol(V)rn1,if l=1=\begin{cases}\left(\frac{r^{n}-(r+1-l)^{n}}{l-1}+r^{n}-(r-1)^{n}\right)\frac{\operatorname{vol}(V)}{r^{n-1}},&\text{if }l\neq 1\\ \left(nr^{n-1}+r^{n}-(r-1)^{n}\right)\frac{\operatorname{vol}(V)}{r^{n-1}},&\text{if }l=1\end{cases}

Note that, under our assumptions on rr and VV, the top intersection power of KY-K_{Y} is positive when 0l<r+10\leq l<r+1.

Now, we consider the intersection of KY-K_{Y} with effective classes of curves on YY. In particular, since KY-K_{Y} is invariant under the 𝕋\mathbb{T}-action, we consider intersection products of KY-K_{Y} with effective classes of curves fixed by the 𝕋\mathbb{T}-action. A curve fixed by the 𝕋\mathbb{T}-action will either be the closure of a single orbit or a curve that is fixed point-wise. A curve that is the closure of a single orbit would either be the strict transform of a fibre from the 1\mathbb{P}^{1}-bundle structure XVX\to V or a fibre of the 1\mathbb{P}^{1}-bundle EBE\to B_{\infty}. A curve that is fixed point-wise would be contained in the fixed locus of the 𝕋\mathbb{T}-action, which is V¯V0(EF)\overline{V}_{\infty}\sqcup V_{0}\sqcup(E\cap F). Thus, we consider intersection products of KY-K_{Y} with curves c0V0c_{0}\in V_{0}, curves cV¯c_{\infty}\in\overline{V}_{\infty}, fibres eEe\in E, fibres ff of the conic bundle structure over VBV\setminus B, and fibres that lie in FF in the class fef-e. We have the following intersection products:

KYc0\displaystyle-K_{Y}\cdot c_{0} =(πH+r1rπϕ(KV)+V¯)c0=r1rπϕ(KV)c0>0\displaystyle=(\pi^{*}H+\frac{r-1}{r}\pi^{*}\phi^{*}(-K_{V})+\overline{V}_{\infty})\cdot c_{0}=\frac{r-1}{r}\pi^{*}\phi^{*}(-K_{V})\cdot c_{0}>0
KYc\displaystyle-K_{Y}\cdot c_{\infty} =(2πV0+r+1rπϕ(KV)E)c=r1rπϕ(KV)c>0\displaystyle=(2\pi^{*}V_{0}+\frac{r+1}{r}\pi^{*}\phi^{*}(-K_{V})-E)\cdot c_{\infty}=\frac{r-1}{r}\pi^{*}\phi^{*}(-K_{V})\cdot c_{\infty}>0
KYf\displaystyle-K_{Y}\cdot f =(2H+r1rπϕ(KV)E)f=2Hf=2>0\displaystyle=(2H+\frac{r-1}{r}\pi^{*}\phi^{*}(-K_{V})-E)\cdot f=2H\cdot f=2>0
KYe\displaystyle-K_{Y}\cdot e =(2H+r1rπϕ(KV)E)e=Ee=1>0\displaystyle=(2H+\frac{r-1}{r}\pi^{*}\phi^{*}(-K_{V})-E)\cdot e=-E\cdot e=1>0
KY(fe)\displaystyle-K_{Y}\cdot(f-e) =(2H+r1rπϕ(KV)E)(fe)=2HE(fe)=1>0\displaystyle=(2H+\frac{r-1}{r}\pi^{*}\phi^{*}(-K_{V})-E)\cdot(f-e)=2H-E\cdot(f-e)=1>0

Thus, we see KY-K_{Y} is strictly nef, and thus the computation of the top intersection power of KY-K_{Y} shows that it is big, and thus ample. ∎

For the next few sections, until Section 6, we will work with this construction under the additional assumption that l=2l=2. We note that there exists an alternate construction of YY when l=2l=2 related to Fano double covers, see [CGF+23, Section 2]. One property of such YY that we use frequently is the existence of an involution ι:YY\iota:Y\to Y such that ι(V¯)=V0\iota(\overline{V}_{\infty})=V_{0}, ι(E)=F\iota(E)=F, and ι\iota respects the conic bundle structure of YY over VV. The existence of such an involution is readily apparent from the point of view of [CGF+23]. From this involution we see that YY permits a second contraction morphism YXY\to X distinct from π\pi: the blow-down of FF, i.e. the composition πι\pi\circ\iota of the involution and the blow-up morphism.

With the construction of YY from the triple (V,B,)(V,B,\mathcal{L}), we now compute certain key β\beta-invariants. In particular, given the induced 𝕋\mathbb{T}-action on YY, we can consider the test configurations induced by the coweights of the 𝕋\mathbb{T}-action. Since the torus action on YY is that of a 11-dimensional algebraic torus, the co-character lattice NN is isomorphic to \mathbb{Z}, with 11 the identity co-character.

Definition 3.4 (Futaki character).

We define the Futaki character on YY to be Fut|N:Fut|_{N}:\mathbb{Z}\to\mathbb{R} where

Fut|N(i)=βY(wti)Fut|_{N}(i)=\beta_{Y}(\mathrm{wt}_{i})

with wti\mathrm{wt}_{i} the valuation induced by the coweight ii\in\mathbb{Z},

Lemma 3.5.

On YY, we have Fut|N=0\operatorname{Fut}|_{N}=0. In particular, β(V¯)=β(V0)=0\beta(\overline{V}_{\infty})=\beta(V_{0})=0.

Proof.

The test configuration induced by ξ\xi has ordt\operatorname{ord}_{t} as its associated valuation under the isomorphism (Y)(V)(t)\mathbb{C}(Y)\cong\mathbb{C}(V)(t) induced by the 𝕋\mathbb{T}-action. This valuation is divisorial with associated divisor V0V_{0}, so Fut(ξ)=βY(V0)\operatorname{Fut}(\xi)=\beta_{Y}(V_{0}). Similarly, the test configuration induced by ξ-\xi has associated divisor V¯\overline{V}_{\infty}. Any other coweight is a multiple of either of these, and Fut(bξ)=bFut(ξ)\operatorname{Fut}(b\xi)=b\operatorname{Fut}(\xi), it suffices to show βY(V0)=βY(V¯)=0\beta_{Y}(V_{0})=\beta_{Y}(\overline{V}_{\infty})=0. We compute both β\beta-invariants simultaneously. As both V0V_{0} and V¯\overline{V}_{\infty} are prime divisors on YY, each has log discrepancy of 11. Thus it remains to show that SY(V0)=SY(V¯)=1S_{Y}(V_{0})=S_{Y}(\overline{V}_{\infty})=1. Let P(t),N(t)P_{\infty}(t),N_{\infty}(t) (resp. P0(t),N0(t)P_{0}(t),N_{0}(t)) be the positive and negative parts of the Zariski decomposition of KYtV¯-K_{Y}-t\overline{V}_{\infty} (resp. KYtV0-K_{Y}-tV_{0}). Then

P(t)={KYtV¯=(1t)V¯+πϕ(KV)+V0,if 0t1(2t)H+r1rπϕ(KV)=r+1trπϕ(KV)+(2t)V0,if 1t2P_{\infty}(t)=\begin{cases}-K_{Y}-t\overline{V}_{\infty}=(1-t)\overline{V}_{\infty}+\pi^{*}\phi^{*}(-K_{V})+V_{0},&\text{if }0\leq t\leq 1\\ (2-t)H+\frac{r-1}{r}\pi^{*}\phi^{*}(-K_{V})=\frac{r+1-t}{r}\pi^{*}\phi^{*}(-K_{V})+(2-t)V_{0},&\text{if }1\leq t\leq 2\end{cases}
N(t)={0,if 0t1(1t)E,if 1t2N_{\infty}(t)=\begin{cases}0,&\text{if }0\leq t\leq 1\\ (1-t)E,&\text{if }1\leq t\leq 2\end{cases}
P0(t)={KYtV0=V¯+πϕ(KV)+(1t)V0,if 0t1(2t)V¯+r+1trπϕ(KV),if 1t2P_{0}(t)=\begin{cases}-K_{Y}-tV_{0}=\overline{V}_{\infty}+\pi^{*}\phi^{*}(-K_{V})+(1-t)V_{0},&\text{if }0\leq t\leq 1\\ (2-t)\overline{V}_{\infty}+\frac{r+1-t}{r}\pi^{*}\phi^{*}(-K_{V}),&\text{if }1\leq t\leq 2\end{cases}
N0(t)={0,if 0t1(t1)F=(t1)(V¯+1rπϕ(KV)V0),if 1t2N_{0}(t)=\begin{cases}0,&\text{if }0\leq t\leq 1\\ (t-1)F=(t-1)(\overline{V}_{\infty}+\frac{1}{r}\pi^{*}\phi^{*}(-K_{V})-V_{0}),&\text{if }1\leq t\leq 2\end{cases}

This can be seen as follows: for the range 0t10\leq t\leq 1, KYtV¯-K_{Y}-t\overline{V}_{\infty} and KYtV0-K_{Y}-tV_{0} are both nef, thus the negative parts of their Zariski decompositions are trivial. For KYtV¯-K_{Y}-t\overline{V}_{\infty} with 1t21\leq t\leq 2, we observe that PP_{\infty} is the pullback along the contraction π\pi of a nef class on XX and NN_{\infty} is supported on the exceptional locus of the same contraction. Thus by [Oka16, Proposition 2.13], we see that this is the Zariski decomposition of KYtV¯-K_{Y}-t\overline{V}_{\infty}. A similar argument shows that P0+N0P_{0}+N_{0} is the Zariski decomposition of KYtV0-K_{Y}-tV_{0}, using the contraction π:YX\pi^{\prime}:Y\to X that contracts FF (this contraction is πι\pi\circ\iota).

We observe that ι(P)=P0\iota(P_{\infty})=P_{0} and ι(N)=N0\iota(N_{\infty})=N_{0}. From this and the computations that V0V¯=0V_{0}\cdot\overline{V}_{\infty}=0 and V0kπϕ(KV)nk=V¯kπϕ(KV)nkV_{0}^{k}\cdot\pi^{*}\phi^{*}(-K_{V})^{n-k}=\overline{V}_{\infty}^{k}\cdot\pi^{*}\phi^{*}(-K_{V})^{n-k}, we conclude that SY(V0)=SY(V¯)S_{Y}(V_{0})=S_{Y}(\overline{V}_{\infty}), and thus we need only compute one of them (this can also be more directly seen from ι(V0)=V¯\iota(V_{0})=\overline{V}_{\infty}).

SY(V¯)\displaystyle S_{Y}(\overline{V}_{\infty}) =1vol(Y)0vol(KYtV¯)𝑑t\displaystyle=\frac{1}{\operatorname{vol}(Y)}\int_{0}^{\infty}\operatorname{vol}(-K_{Y}-t\overline{V}_{\infty})dt
=1vol(Y)02(P(t))n𝑑t\displaystyle=\frac{1}{\operatorname{vol}(Y)}\int_{0}^{2}\left(P_{\infty}(t)\right)^{n}dt
=1vol(Y)01((1t)V¯+πϕ(KV)+V0)n𝑑t\displaystyle=\frac{1}{\operatorname{vol}(Y)}\int_{0}^{1}((1-t)\overline{V}_{\infty}+\pi^{*}\phi^{*}(-K_{V})+V_{0})^{n}dt
+1vol(Y)12(r+1trπϕ(KV)+(2t)V0)n𝑑t\displaystyle+\frac{1}{\operatorname{vol}(Y)}\int_{1}^{2}(\frac{r+1-t}{r}\pi^{*}\phi^{*}(-K_{V})+(2-t)V_{0})^{n}dt
=1vol(Y)012rn(r1)n(r+t1)nrn1vol(V)𝑑t\displaystyle=\frac{1}{\operatorname{vol}(Y)}\int_{0}^{1}\frac{2r^{n}-(r-1)^{n}-(r+t-1)^{n}}{r^{n-1}}\operatorname{vol}(V)dt
+1vol(Y)12(r+t1)n(r1)nrn1vol(V)𝑑t\displaystyle+\frac{1}{\operatorname{vol}(Y)}\int_{1}^{2}\frac{(r+t-1)^{n}-(r-1)^{n}}{r^{n-1}}\operatorname{vol}(V)dt
=1vol(Y)((r1)n+1rn+1n+1+2rn(r1)n)vol(V)rn1\displaystyle=\frac{1}{\operatorname{vol}(Y)}\left(\frac{(r-1)^{n+1}-r^{n+1}}{n+1}+2r^{n}-(r-1)^{n}\right)\frac{\operatorname{vol}(V)}{r^{n-1}}
+1vol(Y)(rn+1(r1)n+1n+1(r1)n)vol(V)rn1\displaystyle+\frac{1}{\operatorname{vol}(Y)}\left(\frac{r^{n+1}-(r-1)^{n+1}}{n+1}-(r-1)^{n}\right)\frac{\operatorname{vol}(V)}{r^{n-1}}
=1\displaystyle=1

We note that there is a second, simpler proof. Since βY(bv)=bβY(v)\beta_{Y}(bv)=b\beta_{Y}(v). By the linearity of the Futaki character we have that βY(V0)=βY(wt1)=βY(wt1)=βY(wt1)=βY(V¯)\beta_{Y}(V_{0})=\beta_{Y}(\mathrm{wt}_{1})=-\beta_{Y}(-\mathrm{wt}_{1})=-\beta_{Y}(\mathrm{wt}_{-1})=-\beta_{Y}(\overline{V}_{\infty}), and that βY(V0)=βY(V¯)\beta_{Y}(V_{0})=\beta_{Y}(\overline{V}_{\infty}). ∎

4. Proof of Main Theorem

4.1. K-Semistability

In the light of Theorem 2.11 and the natural 𝕋\mathbb{T}-action on YY, we investigate the 𝕋\mathbb{T}-equivariant K-semistability of YY in relation to the K-semistability of (V,aB)(V,aB).

4.1.1. Reverse Implication

Definition 4.1.

(vertical and horizontal divisors [Che08, Definition 1.8]) Let DD be a 𝕋\mathbb{T}-invariant divisor over YY. DD is called a vertical divisor if the maximal 𝕋\mathbb{T}-orbit in DD has the same dimension as 𝕋\mathbb{T}, and is otherwise called a horizontal divisor.

Lemma 4.2.

For DD a vertical divisor over YY, there exists a vertical divisor DD^{\prime} over YY such that βY(D)=βY(D)\beta_{Y}(D)=\beta_{Y}(D^{\prime}) and cY(D)V¯c_{Y}(D^{\prime})\cap\overline{V}_{\infty} is nonempty.

Proof.

Let DD be a vertical divisor over YY. Then cY(D)c_{Y}(D) is a 𝕋\mathbb{T}-invariant subvariety of YY. Either the closed points of cY(D)c_{Y}(D) are all fixed by the 𝕋\mathbb{T}-action or cY(D)c_{Y}(D) consists of closures of orbits of points not fixed by 𝕋\mathbb{T}. In the later case, these closures of orbits are either fibres of EE thought of as a 1\mathbb{P}^{1}-bundle over BB_{\infty}, pullbacks of fibres of XVX\to V away from BVB\in V, or the strict transform of fibres of XVX\to V over BB. If cY(D)c_{Y}(D) contains any fibres from the first two cases, these fibres, and thus cY(D)c_{Y}(D), have nonempty intersection with V¯\overline{V}_{\infty}, so we set D:=DD^{\prime}:=D. If this is not the case, then cY(D)Fc_{Y}(D)\subset F. From here, we consider the divisorial valuation v:=ordDιv^{\prime}:=\operatorname{ord}_{D}\circ\iota and let DD^{\prime} be a divisor associated to the valuation vv^{\prime}. Then cY(D)=ι(cY(D))Ec_{Y}(D^{\prime})=\iota(c_{Y}(D))\subset E and cY(D)c_{Y}(D^{\prime}) consists of fibres of EBE\to B_{\infty}, so cY(D)V¯c_{Y}(D^{\prime})\cap\overline{V}_{\infty} is nonempty. Since ordD=ordDι\operatorname{ord}_{D^{\prime}}=\operatorname{ord}_{D}\circ\iota, βY(D)=βY(D)\beta_{Y}(D)=\beta_{Y}(D^{\prime}).

If all closed points ycY(D)y\in c_{Y}(D) are fixed by the 𝕋\mathbb{T}-action, then we have that cY(D)c_{Y}(D) is either contained in V¯\overline{V}_{\infty}, V0V_{0}, or EFE\cap F. If cY(D)V¯c_{Y}(D)\subset\overline{V}_{\infty}, then take D:=DD^{\prime}:=D. If cY(D)V0c_{Y}(D)\subset V_{0}, we consider the divisorial valuation v:=ordDιv^{\prime}:=\operatorname{ord}_{D}\circ\iota and let DD^{\prime} be a divisor associated to the valuation vv^{\prime}. Then cY(D)=ι(cY(D))c_{Y}(D^{\prime})=\iota(c_{Y}(D)) and cY(D)V¯=ι(cY(D)V0)c_{Y}(D^{\prime})\cap\overline{V}_{\infty}=\iota(c_{Y}(D)\cap V_{0})\neq\emptyset. Again, since ordD=ordDι\operatorname{ord}_{D^{\prime}}=\operatorname{ord}_{D}\circ\iota, βY(D)=βY(D)\beta_{Y}(D)=\beta_{Y}(D^{\prime}).

Suppose cY(D)Eι(E)c_{Y}(D)\subset E\cap\iota(E), and denote ordD\operatorname{ord}_{D} as vv. Let μ:=ordD|(V)\mu:=\operatorname{ord}_{D}|_{\mathbb{C}(V)} where the inclusion (V)(Y)\mathbb{C}(V)\subset\mathbb{C}(Y) is induced by the composition πϕ\pi\circ\phi.

Let μ\mu be a valuation on VV and cV(μ)UVc_{V}(\mu)\subset U\subset V be a Zariski-open neighborhood of cV(μ)c_{V}(\mu). Choose a trivialization of \mathcal{L} over UU. This trivialization gives a birational map XV×𝔸1X\dashrightarrow V\times\mathbb{A}^{1}, and composing by π\pi gives us a birational map YV×𝔸1Y\dashrightarrow V\times\mathbb{A}^{1}. This birational map induces an isomorphism on function fields, so we have (Y)(V×𝔸1)(V)(t)\mathbb{C}(Y)\cong\mathbb{C}(V\times\mathbb{A}^{1})\cong\mathbb{C}(V)(t).

We define the valuation vμ,bv_{\mu,b} on YY for μ\mu a valuation on VV and bb\in\mathbb{Z} as follows:

vμ,b(f)=mini{μ(fi)+bi}v_{\mu,b}(f)=\min_{i\in\mathbb{Z}}\{\mu(f_{i})+bi\}

where f=ifitif=\sum_{i}f_{i}t^{i} is the weight decomposition under the 𝔾m\mathbb{G}_{m} action via the above isomorphism of function fields. Moreover, every 𝕋\mathbb{T}-invariant valuation on YY is of this form for some (not necessarily unique) choice of μ,\mu, U,U, and bb. Indeed, suppose vv is a 𝕋\mathbb{T}-invariant valuation and let μ=vs0¯\mu=v\circ\overline{s_{0}} where s0¯:VY\overline{s_{0}}:V\to Y is the lifting of the zero section s0:VXs_{0}:V\to X. Choose UU such that \mathcal{L} has a trivialization over UU and cV(μ)Uc_{V}(\mu)\subset U. Let b=v(t)b=v(t) under the isomorphism (Y)(V)(t)\mathbb{C}(Y)\cong\mathbb{C}(V)(t) induced by the trivialization. Then, for f=fiti(Y)f=\sum f_{i}t^{i}\in\mathbb{C}(Y),

v(f)=mini{v(fiti)}=mini{v(fi)+bi}=vμ,b(f)\displaystyle v(f)=\min_{i\in\mathbb{Z}}\{v(f_{i}t^{i})\}=\min_{i\in\mathbb{Z}}\{v(f_{i})+bi\}=v_{\mu,b}(f)

We then have by [Li22, Proposition 3.12],

β(vb)=β(v)+bFut(ξ)=β(v)=β(μ),\beta(v_{b})=\beta(v)+b\operatorname{Fut}(\xi)=\beta(v)=\beta(\mu^{*}),

with μ=vμ,0\mu^{*}=v_{\mu,0}. The second equality follows from Lemma 3.5. Now, we note that cY(μ)=ϕ1(cV(μ))c_{Y}(\mu^{*})=\phi^{-1}_{*}(c_{V}(\mu)) and thus consists of 11-dimensional orbits of the 𝕋\mathbb{T}-action. If cY(μ)V¯c_{Y}(\mu^{*})\cap\overline{V}_{\infty} is nonempty, we take DD^{\prime} to be a divisor over YY such that DD^{\prime} is the associated divisor to the divisorial valuation μ\mu^{*}. If cY(μ)V¯c_{Y}(\mu^{*})\cap\overline{V}_{\infty} is empty, then cY(μ)Fc_{Y}(\mu^{*})\subset F, so we consider the valuation μ:=μι\mu^{*^{\prime}}:=\mu^{*}\circ\iota, and take DD^{\prime} to be a divisor associated to μ\mu^{*^{\prime}}. From there we see that βY(D)=βY(μ)=βY(D)\beta_{Y}(D^{\prime})=\beta_{Y}(\mu^{*})=\beta_{Y}(D). ∎

Lemma 4.3.

The only horizontal divisors over YY are V¯\overline{V}_{\infty} and V0V_{0}.

Proof.

Any horizontal divisor must solely consist of points fixed by the torus action since the maximal orbit contained in the divisor is strictly less than 1 and thus must be zero dimensional. Thus, the only horizontal 𝕋\mathbb{T}-invariant divisors on YY are V¯\overline{V}_{\infty} and V0V_{0}, as all fixed points are either contained with these two divisors or in the intersection EFE\cap F.

A nontrivial horizontal valuation vv when restricted to (Y)𝕋\mathbb{C}(Y)^{\mathbb{T}} will be the trivial valuation, and thus will be equal to wtb\mathrm{wt}_{b} for some b,b0b\in\mathbb{Z},b\neq 0. If b>0b>0, then the associated divisor will be V0V_{0}, and if b<0b<0 then the associated divisor will be V¯\overline{V}_{\infty}. ∎

In order to compare δ\delta-invariants using the techniques developed in [AZ22], we want to consider the refinement of the linear system TT_{\bullet} where Tm=H0(Y,mKY)T_{m}=H^{0}(Y,-mK_{Y}) by the divisor V¯\overline{V}_{\infty}, which we will denote as W,W_{\bullet,\bullet}. This is defined (see Definition 2.9) as

Wm,j:=Im(H0(Y,mKYjV¯)𝜌H0(V¯,(mKYjV¯)|V¯))W_{m,j}:=\text{Im}\left(H^{0}(Y,-mK_{Y}-j\overline{V}_{\infty})\xrightarrow{\rho}H^{0}(\overline{V}_{\infty},(-mK_{Y}-j\overline{V}_{\infty})|_{\overline{V}_{\infty}})\right)

where the map ρ\rho is induced by the inclusion of V¯\overline{V}_{\infty} into YY.

Lemma 4.4.

The refinement of TT_{\bullet} by V¯\overline{V}_{\infty} is

Wm,j={H0(V¯,(mKYjV¯)|V¯),if jm(jm)B+H0(V¯,P(m,j)|V¯),if m<j2m0,if j>2mW_{m,j}=\begin{cases}H^{0}(\overline{V}_{\infty},(-mK_{Y}-j\overline{V}_{\infty})|_{\overline{V}_{\infty}}),&\text{if }j\leq m\\ (j-m)B+H^{0}(\overline{V}_{\infty},P(m,j)|_{\overline{V}_{\infty}}),&\text{if }m<j\leq 2m\\ 0,&\text{if }j>2m\end{cases}

The movable part of Wm,jW_{m,j}, denoted as Mm,jM_{m,j}, is

Mm,j={H0(V,mKV(mj)),if jmH0(V,mKV+(mj)),if m<j2m0,if j>2mM_{m,j}=\begin{cases}H^{0}(V,-mK_{V}-(m-j)\mathcal{L}),&\text{if }j\leq m\\ H^{0}(V,-mK_{V}+(m-j)\mathcal{L}),&\text{if }m<j\leq 2m\\ 0,&\text{if }j>2m\end{cases}

and the fixed part of Wm,jW_{m,j}, denote as Fm,jF_{m,j}, is

Fm,j={0,if jm(jm)B,if n<j2m0,elseF_{m,j}=\begin{cases}0,&\text{if }j\leq m\\ (j-m)B,&\text{if }n<j\leq 2m\\ 0,&else\end{cases}
Proof.

For j>2mj>2m, mKYjV¯-mK_{Y}-j\overline{V}_{\infty} is no longer psuedo-effective, and thus the image of ρ\rho is 0.

Denoting the positive (resp. negative) part of the Zariski decomposition of the divisor mKYjV¯-mK_{Y}-j\overline{V}_{\infty} with P(m,j)P(m,j) (resp. N(m,j)N(m,j)), we have

P(m,j)={mKYjV¯,if jmmKYjπ(V),if m<j2mP(m,j)=\begin{cases}-mK_{Y}-j\overline{V}_{\infty},&\text{if }j\leq m\\ -mK_{Y}-j\pi^{*}(V_{\infty}),&\text{if }m<j\leq 2m\end{cases}

and

N(m,j)={0,if jm(jm)E,if m<j2mN(m,j)=\begin{cases}0,&\text{if }j\leq m\\ (j-m)E,&\text{if }m<j\leq 2m\end{cases}

Again, we can see that the above is the Zariski decomposition as follows: for jmj\leq m, the divisor is nef, and thus has no negative part. For m<j2mm<j\leq 2m, P(m,j)P(m,j) is the pullback of a nef divisor along the contraction YXY\to X and N(m,j)N(m,j) is supported on the contracted locus.

To compute W,W_{\bullet,\bullet} we use the following long exact sequence from the restriction ρ\rho:

H0(Y,mKY(j+1)V¯)H0(Y,mKYjV¯)𝜌\displaystyle\dots\rightarrow H^{0}(Y,-mK_{Y}-(j+1)\overline{V}_{\infty})\rightarrow H^{0}(Y,-mK_{Y}-j\overline{V}_{\infty})\xrightarrow{\rho}
H0(V¯,(mKYjV¯)|V¯)H1(Y,mKY(j+1)V¯)\displaystyle H^{0}(\overline{V}_{\infty},(-mK_{Y}-j\overline{V}_{\infty})|_{\overline{V}_{\infty}})\rightarrow H^{1}(Y,-mK_{Y}-(j+1)\overline{V}_{\infty})\rightarrow\dots

Thus, we see that the map on global sections induced by restriction is surjective when H1(Y,mKY(j+1)V¯)H^{1}(Y,-mK_{Y}-(j+1)\overline{V}_{\infty}) vanishes, which occurs when jmj\leq m by the Kawamata-Viehweg Vanishing Theorem. More specifically, when jmj\leq m, t:=j+1m+11t:=\frac{j+1}{m+1}\leq 1 and KYtV¯-K_{Y}-t\overline{V}_{\infty} is both big and nef, and thus so is (m+1)(KYtV¯)=(m+1)KY(j+1)V¯(m+1)(-K_{Y}-t\overline{V}_{\infty})=-(m+1)K_{Y}-(j+1)\overline{V}_{\infty}, so by Kawamata-Viehweg Vanishing Theorem, Hi(Y,mKY(j+1)V¯)=0H^{i}(Y,-mK_{Y}-(j+1)\overline{V}_{\infty})=0 for i>0i>0.

Now we consider when m<j2mm<j\leq 2m. In this case, the restriction of the negative part of the Zariski decomposition gives us the fixed part of the refinement, which is (jm)EV¯=(jm)B(j-m)E_{\overline{V}_{\infty}}=(j-m)B. For the movable part, we see that P(m,j)P(m,j) is again big and nef, so H1(Y,P(m,j))=0H^{1}(Y,P(m,j))=0. Thus, by the analogous long exact sequence from the restriction of P(m,j)P(m,j) to V¯\overline{V}_{\infty}, we see that the restriction map is surjective, implying that the free part of the refinement is the space of global sections of the restriction of P(m,j)P(m,j). ∎

Note that, since rKYr\mathcal{L}\sim_{\mathbb{Q}}-K_{Y}, we have

mKY|V¯jV¯|V¯\displaystyle-mK_{Y}|_{\overline{V}_{\infty}}-j\overline{V}_{\infty}|_{\overline{V}_{\infty}} =m(KY+V¯)|V¯(jm)V¯|V¯\displaystyle=-m(K_{Y}+\overline{V}_{\infty})|_{\overline{V}_{\infty}}-(j-m)\overline{V}_{\infty}|_{\overline{V}_{\infty}}
=mKV¯(jm)(VE)|V¯\displaystyle=-mK_{\overline{V}_{\infty}}-(j-m)(V_{\infty}-E)|_{\overline{V}_{\infty}}
=mKV¯(jm)(B)=mrL(jm)()\displaystyle=-mK_{\overline{V}_{\infty}}-(j-m)(\mathcal{L}-B)=mrL-(j-m)(-\mathcal{L})
=(mrm+j)\displaystyle=(mr-m+j)\mathcal{L}

When jmj\leq m, this is then the free part of Wm,jW_{m,j}. When m<j2mm<j\leq 2m, we see that the fixed part Fm,jF_{m,j} of Wm,jW_{m,j} is (jm)B=2(jm)(j-m)B=2(j-m)\mathcal{L}, and thus the free part is (mrm+j)2(jm)=(mr+mj)(mr-m+j)\mathcal{L}-2(j-m)\mathcal{L}=(mr+m-j)\mathcal{L}.

Let Nm,jN_{m,j} denote the dimension of H0(V,Wm,j)H^{0}(V,W_{m,j}), and Nm=jNm,jN_{m}=\sum_{j}N_{m,j}. Let

a(n,r):=rn+1(r1)n+1(n+1)(r1)n2(n+1)(rn(r1)n)a(n,r):=\frac{r^{n+1}-(r-1)^{n+1}-(n+1)(r-1)^{n}}{2(n+1)(r^{n}-(r-1)^{n})}

and let amBa_{m}B be the fixed part of an mm-basis type \mathbb{Q}-divisor of W,W_{\bullet,\bullet}. Then from the above description of W,W_{\bullet,\bullet}, we have that

am=1mNmj=02mNm,jam,ja_{m}=\frac{1}{mN_{m}}\sum_{j=0}^{2m}N_{m,j}a_{m,j}
Lemma 4.5.
limmam(n,r)=a(n,r)\lim_{m\to\infty}a_{m}(n,r)=a(n,r)
Proof.

We note that, asymptotically,

Nm\displaystyle N_{m} =j=0m(mrm+j)n1vol()(n1)!\displaystyle=\sum_{j=0}^{m}\frac{(mr-m+j)^{n-1}\text{vol}(\mathcal{L})}{(n-1)!}
+j=m+12m(mr+mj)n1vol()(n1)!+O(mn2)\displaystyle+\sum_{j=m+1}^{2m}\frac{(mr+m-j)^{n-1}\text{vol}(\mathcal{L})}{(n-1)!}+O(m^{n-2})
=mnvol()(n1)!(j=0m1m(r1+jm)n1\displaystyle=\frac{m^{n}\text{vol}(\mathcal{L})}{(n-1)!}\left(\sum_{j=0}^{m}\frac{1}{m}(r-1+\frac{j}{m})^{n-1}\right.
+j=m+12m1m(r+1jm)n1+(n1)!mnvol()O(mn2))\displaystyle+\left.\sum_{j=m+1}^{2m}\frac{1}{m}(r+1-\frac{j}{m})^{n-1}+\frac{(n-1)!}{m^{n}\text{vol}(\mathcal{L})}O(m^{n-2})\right)

Thus, we have

j=02mNm,jam,j\displaystyle\sum_{j=0}^{2m}N_{m,j}a_{m,j} =vol()(n1)!(m+12m(mr+mj)n1(jm)+O(mn2))\displaystyle=\frac{\text{vol}(\mathcal{L})}{(n-1)!}\left(\sum_{m+1}^{2m}(mr+m-j)^{n-1}(j-m)+O(m^{n-2})\right)
=mn+1vol()(n1)!(m+12m1m(r+1jm)n1(jm1)+1mn+1O(mn2))\displaystyle=\frac{m^{n+1}\text{vol}(\mathcal{L})}{(n-1)!}\left(\sum_{m+1}^{2m}\frac{1}{m}(r+1-\frac{j}{m})^{n-1}(\frac{j}{m}-1)+\frac{1}{m^{n}+1}O(m^{n-2})\right)

Thus we compute the limit as follows:

limmam\displaystyle\lim_{m\to\infty}a_{m} =limm1mNmj=02mNm,jam,j\displaystyle=\lim_{m\to\infty}\frac{1}{mN_{m}}\sum_{j=0}^{2m}N_{m,j}a_{m,j}
=limmj=0m1m(r1+jm)n1+j=m+12m1m(r+1jm)n1m+12m1m(r+1jm)n1(jm1)\displaystyle=\lim_{m\to\infty}\frac{\sum_{j=0}^{m}\frac{1}{m}(r-1+\frac{j}{m})^{n-1}+\sum_{j=m+1}^{2m}\frac{1}{m}(r+1-\frac{j}{m})^{n-1}}{\sum_{m+1}^{2m}\frac{1}{m}(r+1-\frac{j}{m})^{n-1}(\frac{j}{m}-1)}
=12(r+1t)n1(t1)𝑑t212(r+1t)n1𝑑t\displaystyle=\frac{\int_{1}^{2}(r+1-t)^{n-1}(t-1)dt}{2\int_{1}^{2}(r+1-t)^{n-1}dt}
=12[(r1)n(r1)n+1rn+1n+1rn(r1)n]\displaystyle=\frac{1}{2}\left[\frac{-(r-1)^{n}-\frac{(r-1)^{n+1}-r^{n+1}}{n+1}}{r^{n}-(r-1)^{n}}\right]
=a\displaystyle=a

In the terminology of [AZ22], a(n,r)a(n,r) is the coefficient of BB in the asymptotic fixed part F(W,)F(W_{\bullet,\bullet}) of W,W_{\bullet,\bullet}.

Lemma 4.6.

Let DD be a divisor over YY and ZcY(D)YZ\subset c_{Y}(D)\subset Y be a subvariety of YY contained in the center of DD on YY. If δZ(Y)1\delta_{Z}(Y)\geq 1, then βY(D)0\beta_{Y}(D)\geq 0.

Proof.

Suppose δZ(Y)1\delta_{Z}(Y)\geq 1. Then, by definition of the local δ\delta-invariant, AY(D)SY(D)1\frac{A_{Y}(D^{\prime})}{S_{Y}(D^{\prime})}\geq 1 for all divisors DD^{\prime} over YY whose center on YY contains ZZ. This includes DD. Thus, by rearranging terms, βY(D)=AY(D)SY(D)0\beta_{Y}(D)=A_{Y}(D)-S_{Y}(D)\geq 0. ∎

Lemma 4.7.

Let ZZ be a subvariety of YY such that ZV¯Z\subset\overline{V}_{\infty}. Suppose (V,aB)(V,aB) is K-semistable. Then δZ(Y)1\delta_{Z}(Y)\geq 1.

Proof.

By [AZ22, Theorem 3.3], we have that

δZ(Y)min{AY(V¯)SY(V¯),δZ(V¯,W,)}\delta_{Z}(Y)\geq\text{min}\left\{\frac{A_{Y}(\overline{V}_{\infty})}{S_{Y}(\overline{V}_{\infty})},\delta_{Z}(\overline{V}_{\infty},W_{\bullet,\bullet})\right\}

Where W,W_{\bullet,\bullet} is the refinement of the graded linear series associated to KY-K_{Y} by V¯\overline{V}_{\infty}. The first value in the minimum is 11 by Lemma 3.5. Thus, it remains to show that δZ(V¯,W,)1\delta_{Z}(\overline{V}_{\infty},W_{\bullet,\bullet})\geq 1.

By the assumption that the pair (V,aB)(V,aB) is K-semistable, we have that δ(V,aB)1\delta(V,aB)\geq 1, which in particular means that, for a choice of ϵ>0\epsilon>0, δM(V,aB)>1ϵ\delta_{M}(V,aB)>1-\epsilon for all MM sufficiently large. So, for any MM-basis type \mathbb{Q} divisor ΔM\Delta_{M} of KVaB-K_{V}-aB, we have that (V,(1ϵ)ΔM+aB)(V,(1-\epsilon)\Delta_{M}+aB) is lc.

Since KVaB=(r2a)-K_{V}-aB=(r-2a)\mathcal{L}, for any such ΔM\Delta_{M}, (r2a)ΔM(r-2a)\Delta_{M} is an M(r2a)M(r-2a)-basis type \mathbb{Q}-divisor of \mathcal{L}, and thus we have δM(r2a)(V,aB;)>(r2a)(1ϵ)\delta_{M(r-2a)}(V,aB;\mathcal{L})>(r-2a)(1-\epsilon).

Now, let DmD_{m} be an mm-basis type divisor of W,W_{\bullet,\bullet}, then

Dm=1mNmj=02mDm,jD_{m}=\frac{1}{mN_{m}}\sum_{j=0}^{2m}D_{m,j}

where each Dm,jD_{m,j} is a basis sum divisor of Wm,jW_{m,j}. By our computation of Wm,jW_{m,j}, we see that Dm,j=Δmr|mj|+am,jNm,jBD_{m,j}=\Delta_{mr-|m-j|}+a_{m,j}N_{m,j}B, where Δmr|mj|\Delta_{mr-|m-j|} is a basis sum divisor of H0((mr|mj|))H^{0}((mr-|m-j|)\mathcal{L}) and am,j=jma_{m,j}=j-m for 2mj>m2m\geq j>m and 0 otherwise. In light of the previous discussion, we can choose mm such that mr|mj|mr-|m-j| is sufficiently large and hence the pair

(V,(r2a)(1ϵ)1(mr|mj|)Nm,jΔmr|mj|+aB)\left(V,(r-2a)(1-\epsilon)\frac{1}{(mr-|m-j|)N_{m,j}}\Delta_{mr-|m-j|}+aB\right)

is log canonical for all 0j2m0\leq j\leq 2m. Letting wm,j=(mr|mj|)Nm,jm(r2am)Nmw_{m,j}=\frac{(mr-|m-j|)N_{m,j}}{m(r-2a_{m})N_{m}}. Then we see that j=02mwm,j=1\sum_{j=0}^{2m}w_{m,j}=1, so by the convexity of log canonicity, we have that the pair

(V,(r2a)(1ϵ)j=02mwm,j(1(mr|mj|)Nm,jΔmr|mj|+aB))\displaystyle\left(V,(r-2a)(1-\epsilon)\sum_{j=0}^{2m}w_{m,j}\left(\frac{1}{(mr-|m-j|)N_{m,j}}\Delta_{mr-|m-j|}+aB\right)\right)
=(V,(1ϵ)1mNm(j=02mΔmr|mj|+am,jNm,jB))\displaystyle=\left(V,(1-\epsilon)\frac{1}{mN_{m}}\left(\sum_{j=0}^{2m}\Delta_{mr-|m-j|}+a_{m,j}N_{m,j}B\right)\right)
=(V,(1ϵ)Dm)\displaystyle=\left(V,(1-\epsilon)D_{m}\right)

is log canonical. Thus, we see that δZ(V¯,W,)1\delta_{Z}(\overline{V}_{\infty},W_{\bullet,\bullet})\geq 1 as desired. ∎

Proposition 4.8.

If the pair (V,aB)(V,aB) is K-semistable, then YY is K-semistable.

Proof.

We will show that the K-semistability of the pair (V,aB)(V,aB) implies the 𝕋\mathbb{T}-equivariant K-semistability of YY, which then implies the K-semistability of YY by Theorem 2.11. Suppose DD is a 𝕋\mathbb{T}-invariant divisor over YY. We proceed to show βY(D)0\beta_{Y}(D)\geq 0. If DD is a horizontal divisor, then βY(D)=0\beta_{Y}(D)=0 by Lemma 4.3 and Lemma 3.5. Thus, we suppose DD is vertical. Then, by Lemma 4.2, we may assume cY(D)V¯c_{Y}(D)\cap\overline{V}_{\infty} is nonempty and thus contains some subvariety ZZ. By Theorem 4.7, δY(Z)1\delta_{Y}(Z)\geq 1, and thus βY(D)0\beta_{Y}(D)\geq 0 by Lemma 4.6. ∎

4.1.2. Forward Implication

First we proceed with a 𝔾m\mathbb{G}_{m}-equivariant strengthening of [AZ22, Lemma 3.1] and [AZ22, Proposition 3.2].

Lemma 4.9.

Let HH be a finite dimensional vector space over \mathbb{C} with an action of 𝔾m\mathbb{G}_{m} and two 𝔾m\mathbb{G}_{m}-equivariant filtrations ,𝒢\mathcal{F},\mathcal{G}. Then there exists a 𝔾m\mathbb{G}_{m}-invariant basis of HH compatible with both \mathcal{F} and 𝒢\mathcal{G}.

Proof.

The proof mostly follows that of the analogous statement in [AZ22]. Consider the induced filtrations 𝒢i\mathcal{G}_{i} on GriHGr_{i}^{\mathcal{F}}H for each ii. The equivariance of \mathcal{F} implies that the 𝔾m\mathbb{G}_{m} action on HH restricts to an action on each GriHGr_{i}^{\mathcal{F}}H; the equivariance of 𝒢\mathcal{G} implies the equivariance of each 𝒢i\mathcal{G}_{i} and thus the aforementioned action of 𝔾m\mathbb{G}_{m} restricts to an action on Grj𝒢iGriHGr_{j}^{\mathcal{G}_{i}}Gr_{i}^{\mathcal{F}}H for all i,ji,j. For each ii and jj, take a basis Bi,jB_{i,j} of 𝔾m\mathbb{G}_{m}-eigenvectors of Grj𝒢iGriHGr_{j}^{\mathcal{G}_{i}}Gr_{i}^{\mathcal{F}}H; for each ii lift these bases to a basis BiB_{i} of GriHGr_{i}^{\mathcal{F}}H. Each of these bases will consist of 𝔾m\mathbb{G}_{m}-eigenvectors for the action on GriHGr_{i}^{\mathcal{F}}H; these then lift to a \mathcal{F}-compatible 𝔾m\mathbb{G}_{m}-invariant basis BB of HH. Since Grj𝒢iGriH=GrijGrj𝒢HGr_{j}^{\mathcal{G}_{i}}Gr_{i}^{\mathcal{F}}H=Gr_{i}^{\mathcal{F}_{j}}Gr_{j}^{\mathcal{G}}H for all i,ji,j, we see that for each jj, iBi,j\bigcup_{i}B_{i,j} forms a basis of Grj𝒢HGr_{j}^{\mathcal{G}}H and that BB is the lift of these bases and is thus 𝒢\mathcal{G}-compatible. ∎

Lemma 4.10.

For a linear system HH_{\bullet}, a filtration \mathcal{F} on HH_{\bullet}, and a valuation vv on HH_{\bullet}, we have

S(H;v)=S𝔾m(H;;v)S(H_{\bullet};v)=S_{\mathbb{G}_{m}}(H_{\bullet};\mathcal{F};v)
Proof.

We have that Sm(H;v)S𝔾m,m(H;v)S_{m}(H_{\bullet};v)\geq S_{\mathbb{G}_{m},m}(H_{\bullet};v) by definition. Let DmD_{m} be a m-basis type \mathbb{Q}-divisor of HH_{\bullet} that is 𝔾m\mathbb{G}_{m}-invariant and compatible with \mathcal{F} and v\mathcal{F}_{v}; such a DmD_{m} exists by Lemma 4.9. Then Sm(H;v)=v(Dm)S𝔾m,m(H;v)S_{m}(H_{\bullet};v)=v(D_{m})\leq S_{\mathbb{G}_{m},m}(H_{\bullet};v). Taking the limit of each side as mm approaches infinity yields the lemma. ∎

Lemma 4.11.

Fix mm such that h0(Y,mKY)0h^{0}(Y,-mK_{Y})\neq 0. There is a one-to-one correspondence between 𝕋\mathbb{T}-invariant mm-basis type \mathbb{Q}-divisors of KY-K_{Y} and mm-basis type \mathbb{Q}-divisors of W,W_{\bullet,\bullet}.

Proof.

Since there is a clear one-to-one correspondence between mm-basis type \mathbb{Q} divisors and bases of global sections modulo scaling each basis element, this lemma follows from an analogous correspondence between bases of Rm:=H0(mKY)R_{m}:=H^{0}(-mK_{Y}) invariant under the 𝕋\mathbb{T}-action and sets of bases for H0(Wm,j)H^{0}(W_{m,j}) for 0j2m0\leq j\leq 2m. We can decompose Rm=Rm,jR_{m}=\oplus R_{m,j} into the direct sum of eigenspaces under the 𝕋\mathbb{T}-action, where 𝕋\mathbb{T} acts on sRm,js\in R_{m,j} by σ(λ)s=λjs\sigma(\lambda)s=\lambda^{j}s; a 𝕋\mathbb{T}-invariant basis on RmR_{m} is exactly a union of bases on each Rm,jR_{m,j}. In particular, each Rm,jR_{m,j} is isomorphic to the jj-th graded piece of the weight filtration on RmR_{m}, which is exactly the filtration induced by wt1=ordV¯\mathrm{wt}_{-1}=\operatorname{ord}_{\overline{V}_{\infty}}. Thus, by construction, Rm,jH0(Wm,j)R_{m,j}\cong H^{0}(W_{m,j}), and the correspondence between bases follows. ∎

With lemmata in hand, we begin our argument for the forward implication of the main theorem by constructing divisors over YY whose β\beta-invariants are dictated by divisors over VV.

Convention 4.12.

Let DVD_{V} be a divisor over VV. Then DVD_{V} defines a divisor DYD_{Y} over YY as follows:

Let WVW\to V be a model on which DVD_{V} is a divisor. We consider the birational model XW:=W×VX𝜂XX_{W}:=W\times_{V}X\xrightarrow{\eta}X of XX, and the divisor DX:=ϕW(DV)D_{X}:={\phi_{W}}^{*}(D_{V}). DXD_{X} defines a divisorial valuation on XX, which we pullback along the blow-up map π\pi to a divisorial valuation ordDX\operatorname{ord}_{D_{X}} on YY. We then pull back ordDX\operatorname{ord}_{D_{X}} along ι\iota to a divisorial valuation ordDY\operatorname{ord}_{D_{Y}} achieved by some divisor DYD_{Y} over YY.

Lemma 4.13.

Under the inclusion (V)(Y)\mathbb{C}(V)\subset\mathbb{C}(Y) of function fields induced by the composition πϕ\pi\circ\phi, for a divisor DVD_{V} over VV and DYD_{Y} as in Convention 4.12, we have the identification ordDY|(V)=ordDV\operatorname{ord}_{D_{Y}}|_{\mathbb{C}(V)}=\operatorname{ord}_{D_{V}}.

Proof.

It is clear that ordDX|(V)=ordDV\operatorname{ord}_{D_{X}}|_{\mathbb{C}(V)}=\operatorname{ord}_{D_{V}} via the inclusion of function fields induced by ϕ\phi. Now, ordDY=ordDXι\operatorname{ord}_{D_{Y}}=\operatorname{ord}_{D_{X}}\circ\iota by construction, so ordDY=ordDX\operatorname{ord}_{D_{Y}}=\operatorname{ord}_{D_{X}} on meromorphic functions that are fixed by composition with ι\iota. Since ι\iota fixes fibres of the conic bundle structure of πϕ:YV\pi\circ\phi:Y\to V, it in particular fixes (V)(Y)\mathbb{C}(V)\subset\mathbb{C}(Y), and the lemma follows. ∎

Lemma 4.14.

Let DVD_{V} be a divisor over VV and define DYD_{Y} as in Convention 4.12. Then AY(DY)=AV(DV)A_{Y}(D_{Y})=A_{V}(D_{V}).

Proof.

First we note that AY(DY)=AY(ordDX)A_{Y}(D_{Y})=A_{Y}(\operatorname{ord}_{D_{X}}) since ordDXι=ordDY\operatorname{ord}_{D_{X}}\circ\iota=\operatorname{ord}_{D_{Y}}. Then, we compute AY(DX)A_{Y}(D_{X}) as follows:

AY(DX)\displaystyle A_{Y}(D_{X}) =ordDX(K/Y)+1\displaystyle=\operatorname{ord}_{D_{X}}(K_{-/Y})+1
=ordDX(K/X)ordDX(E)+1\displaystyle=\operatorname{ord}_{D_{X}}(K_{-/X})-\operatorname{ord}_{D_{X}}(E)+1
=AX(DX)ordDX(E)\displaystyle=A_{X}(D_{X})-\operatorname{ord}_{D_{X}}(E)
=AX(DX)\displaystyle=A_{X}(D_{X})
=AV(DV)\displaystyle=A_{V}(D_{V})

where the notation K/ZK_{-/Z} for Z=X,YZ=X,Y denotes the relative canonical divisor of an appropriate choice of resolution of ZZ, and the last equality follows from the canonical bundle formula for projective bundles. In particular, KXWη(KX)=ϕ(KWfKV)K_{X_{W}}-\eta^{*}(K_{X})=\phi^{*}(K_{W}-f^{*}K_{V}). ∎

Proposition 4.15.

For DVD_{V} a divisor over VV and DYD_{Y} as in Convention 4.12, βY(DY)=βV,aB(DV)\beta_{Y}(D_{Y})=\beta_{V,aB}(D_{V}).

Proof.

Using Lemma 4.9, let ΔmY\Delta_{m}^{Y} be a 𝕋\mathbb{T}-invariant mm-basis type \mathbb{Q}-divisor of KY-K_{Y} compatible with DYD_{Y} and V¯\overline{V}_{\infty}. Then ΔmY\Delta_{m}^{Y} decomposes 𝕋\mathbb{T}-equivariantly as below, with Supp(Γ)\operatorname{Supp}(\Gamma) consisting of 𝕋\mathbb{T}-invariant divisors distinct from V¯\overline{V}_{\infty} (see [AZ22, Section 3.1]).

ΔmY=Γ+Sm(KY,V¯)V¯\Delta_{m}^{Y}=\Gamma+S_{m}(-K_{Y},\overline{V}_{\infty})\cdot\overline{V}_{\infty}

In particular, Γ|V¯\Gamma|_{\overline{V}_{\infty}} is the mm-basis type \mathbb{Q}-divisor of W,W_{\bullet,\bullet} corresponding to ΔmY\Delta_{m}^{Y} under Lemma 4.11, i.e. Γ|V¯=1mNmj(Δm,j+am,jNm,jB)\Gamma|_{\overline{V}_{\infty}}=\frac{1}{mN_{m}}\sum_{j}(\Delta_{m,j}+a_{m,j}N_{m,j}B), where Δm,j\Delta_{m,j} is a basis-sum divisor of the movable part of the refinement Mm,jM_{m,j}. Since ΔmY\Delta_{m}^{Y} is compatible with DYD_{Y}, each Δm,j\Delta_{m,j} is compatible with the restriction of ordDY\operatorname{ord}_{D_{Y}} to (V)\mathbb{C}(V), which in particular means that each Δm,j\Delta_{m,j} is compatible with DVD_{V}. Since the irreducible components of Supp(Γ)\operatorname{Supp}(\Gamma) are invariant under the 𝕋\mathbb{T} action, we see that

Γ=1mNM(jΓj+am,jNm,jE)\Gamma=\frac{1}{mN_{M}}\left(\sum_{j}\Gamma_{j}+a_{m,j}N_{m,j}E\right)

where Γj|V¯=Δm,j\Gamma_{j}|_{\overline{V}_{\infty}}=\Delta_{m,j} and each Γj\Gamma_{j} is strictly supported on the 𝕋\mathbb{T}-invariant divisors on YY. We compute ordDY(Γj)\operatorname{ord}_{D_{Y}}(\Gamma_{j}) by decomposing Γj\Gamma_{j} into its parts supported on specific 𝕋\mathbb{T}-invariant divisors. Let DD denote all divisors on VV (other than BB) such that Γ\Gamma has support on π1(ϕD){\pi}_{*}^{-1}(\phi^{*}D). We note that ordDY(V¯),ordDY(V0)\operatorname{ord}_{D_{Y}}(\overline{V}_{\infty}),\operatorname{ord}_{D_{Y}}(V_{0}) and ordDY(F)\operatorname{ord}_{D_{Y}}(F) are all 0: for a choice any one of these three divisors we can find an open neighborhood UU of cY(DY)c_{Y}(D_{Y}) that has empty intersection with the chosen divisor. Thus, on UU, the chosen divisor is defined locally by a nonvanishing holomorphic function ff, in particular ff has vanishing order 0 at cY(DY)c_{Y}(D_{Y}). Similarly, we compute ordDY(E)=ordπ1(DX)(F)=ordDX(ϕ(B))=ordV(B)\operatorname{ord}_{D_{Y}}(E)=\operatorname{ord}_{{\pi}_{*}^{-1}(D_{X})}(F)=\operatorname{ord}_{D_{X}}(\phi^{*}(B))=\operatorname{ord}_{V}(B) and ordDY(π1(ϕD))=ordDV(D)\operatorname{ord}_{D_{Y}}({\pi}_{*}^{-1}(\phi^{*}D))=\operatorname{ord}_{D_{V}}(D). Thus, we have the following computation:

ordDY(Γj)\displaystyle\operatorname{ord}_{D_{Y}}(\Gamma_{j}) =coeffV0(Γj)ordDY(V0)+coeffV¯(Γj)ordDY(V¯)\displaystyle=\mathrm{coeff}_{V_{0}}(\Gamma_{j})\cdot\operatorname{ord}_{D_{Y}}(V_{0})+\mathrm{coeff}_{\overline{V}_{\infty}}(\Gamma_{j})\cdot\operatorname{ord}_{D_{Y}}(\overline{V}_{\infty})
+coeffE(Γj)ordDY(E)+coeffF(Γj)ordDY(F)\displaystyle+\mathrm{coeff}_{E}(\Gamma_{j})\cdot\operatorname{ord}_{D_{Y}}(E)+\mathrm{coeff}_{F}(\Gamma_{j})\cdot\operatorname{ord}_{D_{Y}}(F)
+coeffπ1(ϕD)(Γj)ordDY(π1(ϕD))\displaystyle+\mathrm{coeff}_{{\pi}_{*}^{-1}(\phi^{*}D)}(\Gamma_{j})\cdot\operatorname{ord}_{D_{Y}}({\pi}_{*}^{-1}(\phi^{*}D))
=coeffE(Γj)ordDY(E)+coeffπ1(ϕD)(Γj)ordDY(π1(ϕD))\displaystyle=\mathrm{coeff}_{E}(\Gamma_{j})\cdot\operatorname{ord}_{D_{Y}}(E)+\mathrm{coeff}_{{\pi}_{*}^{-1}(\phi^{*}D)}(\Gamma_{j})\cdot\operatorname{ord}_{D_{Y}}({\pi}_{*}^{-1}(\phi^{*}D))
=coeffB(Δm,j)ordDV(B)+coeffD(Δm,j)ordDV(D)\displaystyle=\mathrm{coeff}_{B}(\Delta_{m,j})\cdot\operatorname{ord}_{D_{V}}(B)+\mathrm{coeff}_{D}(\Delta_{m,j})\operatorname{ord}_{D_{V}}(D)
=ordDV(Δm,j)\displaystyle=\operatorname{ord}_{D_{V}}(\Delta_{m,j})

As an aside, we note that coeffV¯(Γj)=0\mathrm{coeff}_{\overline{V}_{\infty}}(\Gamma_{j})=0 from the decomposition of ΔmY\Delta_{m}^{Y}.

With this, we have the following computation of Sm(DY)S_{m}(D_{Y}):

Sm(DY)=ordDYΔmY\displaystyle S_{m}(D_{Y})=\operatorname{ord}_{D_{Y}}\Delta_{m}^{Y} =ordDY(Γ+Sm(KY,V¯)V¯)\displaystyle=\operatorname{ord}_{D_{Y}}\left(\Gamma+S_{m}(-K_{Y},\overline{V}_{\infty})\cdot\overline{V}_{\infty}\right)
=1mNmordDY(jΓj+am,jNm,jE)\displaystyle=\frac{1}{mN_{m}}\operatorname{ord}_{D_{Y}}\left(\sum_{j}\Gamma_{j}+a_{m,j}N_{m,j}E\right)
+Sm(KY,V¯)ordDY(V¯)\displaystyle+S_{m}(-K_{Y},\overline{V}_{\infty})\cdot\operatorname{ord}_{D_{Y}}(\overline{V}_{\infty})
=1mNmordDY(jΓj+am,jNm,jE)\displaystyle=\frac{1}{mN_{m}}\operatorname{ord}_{D_{Y}}\left(\sum_{j}\Gamma_{j}+a_{m,j}N_{m,j}E\right)
=1mNm(jordDY(Γj)+am,jNm,jordDY(E))\displaystyle=\frac{1}{mN_{m}}\left(\sum_{j}\operatorname{ord}_{D_{Y}}(\Gamma_{j})+a_{m,j}N_{m,j}\operatorname{ord}_{D_{Y}}(E)\right)
=1mNm(jordDV(Δm,j)+am,jNm,jordDV(B))\displaystyle=\frac{1}{mN_{m}}\left(\sum_{j}\operatorname{ord}_{D_{V}}(\Delta_{m,j})+a_{m,j}N_{m,j}\operatorname{ord}_{D_{V}}(B)\right)
=1mNm(jmNm,jSmr|mj|(DV)+am,jNm,jordDVB)\displaystyle=\frac{1}{mN_{m}}\left(\sum_{j}mN_{m,j}S_{mr-|m-j|}(D_{V})+a_{m,j}N_{m,j}\operatorname{ord}_{D_{V}}B\right)

Taking the limit as mm tends to \infty, we are left with:

S(DY)\displaystyle S(D_{Y}) =S(DV)+aordDVB\displaystyle=S(D_{V})+a\operatorname{ord}_{D_{V}}B

Also by Lemma 4.14 we have AV(DV)=AY(DY)A_{V}(D_{V})=A_{Y}(D_{Y}), so we have

βV,aB(DV)\displaystyle\beta_{V,aB}(D_{V}) =AV,aB(DV)S(DV)\displaystyle=A_{V,aB}(D_{V})-S(D_{V})
=AV(DV)aordDV(B)S(DV)\displaystyle=A_{V}(D_{V})-a\operatorname{ord}_{D_{V}}(B)-S(D_{V})
=AY(DY)S(DY)\displaystyle=A_{Y}(D_{Y})-S(D_{Y})
=βY(DY)\displaystyle=\beta_{Y}(D_{Y})

Proposition 4.16.

If (V,aB)(V,aB) is K-unstable, then so is YY.

Proof.

Assume that (V,aB)(V,aB) is K-unstable. Let DVD_{V} be a destabilizing divisor of the pair (V,aB)(V,aB), that is a divisor on a birational model WW of VV such that βV,aB(DV)<0\beta_{V,aB}(D_{V})<0. Then by Proposition 4.15 βY(DY)<0\beta_{Y}(D_{Y})<0. ∎

4.2. K-Polystability

By Proposition 4.16 and Proposition 4.8, if either of (V,aB)(V,aB) or YY is not K-semistable, then neither are K-semistable. Since K-polystability implies K-semistability, we assume, for this section, that both of YY and the pair (V,aB)(V,aB) are K-semistable.

4.2.1. Forward Implication

Proposition 4.17.

Suppose YY is K-polystable. Then the log Fano pair (V,aB)(V,aB) is K-polystable.

Proof.

Suppose YY is K-polystable. Let DVD_{V} be a divisor over the pair (V,aB)(V,aB) such that βV,aB(D)=0\beta_{V,aB}(D)=0. To show (V,aB)(V,aB) is K-polystable it suffices to show that DVD_{V} is a product-type divisor. We know that βY(DY)=βV,aB(DV)=0\beta_{Y}(D_{Y})=\beta_{V,aB}(D_{V})=0 with DYD_{Y} as in Convention 4.12, so DYD_{Y} is a product-type divisor. Denote this test configuration as 𝒴Y×𝔸1\mathcal{Y}\cong Y\times\mathbb{A}^{1}, and denote the 𝔾m\mathbb{G}_{m}-action of the test configuration as σDY:𝔾m×𝒴𝒴\sigma_{D_{Y}}:\mathbb{G}_{m}\times\mathcal{Y}\to\mathcal{Y}. Further, since DYD_{Y} is invariant with respect to the 𝕋\mathbb{T}-action on YY, 𝒴\mathcal{Y} is a 𝕋\mathbb{T}-equivariant test configuration.

Consider the closure of the orbit orb¯σD(V¯×{1})𝒴\overline{\text{orb}}_{\sigma_{D}}(\overline{V}_{\infty}\times\{1\})\subset\mathcal{Y} under the action of σDY\sigma_{D_{Y}} on 𝒴\mathcal{Y}. Let 𝒱\mathcal{V} denote this closure. Since 𝒴\mathcal{Y} is a 𝕋\mathbb{T}-invariant test configuration, the 𝔾m\mathbb{G}_{m}-action on YY induced by σDY\sigma_{D_{Y}} commutes with the 𝕋\mathbb{T}-action σ\sigma. We will denote this restriction as σDY0\sigma_{D_{Y}}^{0}. For t𝕋,s𝔾m,t\in\mathbb{T},s\in\mathbb{G}_{m}, and yV¯Yy\in\overline{V}_{\infty}\subset Y, we have

σ(t)σDY0(s)(y)\displaystyle\sigma(t)\sigma_{D_{Y}}^{0}(s)(y) =σDY0(s)σ(t)(y)\displaystyle=\sigma_{D_{Y}}^{0}(s)\sigma(t)(y)
=σDY0(s)(y)\displaystyle=\sigma_{D_{Y}}^{0}(s)(y)

So σDY0(s)(y)\sigma_{D_{Y}}^{0}(s)(y) is in the fixed locus of σ\sigma, which is V¯V0(EF)\overline{V}_{\infty}\sqcup V_{0}\sqcup(E\cap F). However, since σDY0\sigma_{D_{Y}}^{0} has connected orbits, we see that σDY0(s)(y)V¯\sigma_{D_{Y}}^{0}(s)(y)\in\overline{V}_{\infty}. Thus, σDY0\sigma_{D_{Y}}^{0} maps points in V¯\overline{V}_{\infty} to points in V¯\overline{V}_{\infty}, which implies that σDY\sigma_{D_{Y}} maps points in V¯×𝔸1𝒴\overline{V}_{\infty}\times\mathbb{A}^{1}\subset\mathcal{Y} to points in V¯×𝔸1\overline{V}_{\infty}\times\mathbb{A}^{1}. From this we see that 𝒱V×𝔸1\mathcal{V}\cong V\times\mathbb{A}^{1} and that 𝒱\mathcal{V} with the restriction of the 𝔾m\mathbb{G}_{m}-action σDY\sigma_{D_{Y}} to 𝒱\mathcal{V} forms a product test configuration of VV.

The valuation associated to 𝒱\mathcal{V} is, by definition, is ord𝒱0|(V)\operatorname{ord}_{\mathcal{V}_{0}}|_{\mathbb{C}(V)}. We see that the 𝕋\mathbb{T}-action σ\sigma extends to 𝒴\mathcal{Y}, and induces a weight decomposition on (𝒴)\mathbb{C}(\mathcal{Y}) (similar to that on (Y)\mathbb{C}(Y)) such that (𝒴)𝕋(𝒱)\mathbb{C}(\mathcal{Y})^{\mathbb{T}}\cong\mathbb{C}(\mathcal{V}). From this, we see the ord𝒴0|(𝒱)=ord𝒱0\operatorname{ord}_{\mathcal{Y}_{0}}|_{\mathbb{C}(\mathcal{V})}=\operatorname{ord}_{\mathcal{V}_{0}}. Since ordDY=bord𝒴0|(𝒴)\operatorname{ord}_{D_{Y}}=b\operatorname{ord}_{\mathcal{Y}_{0}}|_{\mathbb{C}(\mathcal{Y})} for some b>0b\in\mathbb{Z}_{>0}, we have that bord𝒱0|(V)=bord𝒴0|(V)=ordDY|(V)=ordDVb\operatorname{ord}_{\mathcal{V}_{0}}|_{\mathbb{C}(V)}=b\operatorname{ord}_{\mathcal{Y}_{0}}|_{\mathbb{C}(V)}=\operatorname{ord}_{D_{Y}}|_{\mathbb{C}(V)}=\operatorname{ord}_{D_{V}}. Thus, we see that the test configuration associated to DVD_{V} is isomorphic to 𝒱\mathcal{V} and is thus a product test configuration. ∎

4.2.2. Reverse Implication

Proposition 4.18.

Suppose the log pair (V,aB)(V,aB) is K-polystable. Then YY is K-polystable.

Proof.

Suppose that the pair (V,aB)(V,aB) is K-polystable; we seek to show the same for the construction YY. In fact, due to 2.11, it suffices to show that YY is 𝕋\mathbb{T}-equivariantly K-polystable.

Suppose DD is a 𝕋\mathbb{T}-invariant divisor over YY such that βY(D)=0\beta_{Y}(D)=0; we seek to show that the test configuration induced by DD is a product test configuration. If DD is horizontal, then D=V0D=V_{0} or V¯\overline{V}_{\infty}; both of these induce product test configurations. Indeed, ordV0=wtξ\operatorname{ord}_{V_{0}}=\mathrm{wt}_{\xi} and ordV¯=wtξ\operatorname{ord}_{\overline{V}_{\infty}}=\mathrm{wt}_{-\xi}, so the test configurations induced by DD is the product test configuration given by either the 𝕋\mathbb{T}-action on YY or its inverse.

Now, assume DD is vertical. Since the underlying space of two test configurations are isomorphic if one of the associated valuations is achieved by twisting the other by a suitable cocharacter, using Lemma 4.2 we may assume cY(D)V¯c_{Y}(D)\cap\overline{V}_{\infty}\neq\emptyset.

Let μ:=ordD|(V)\mu:=\operatorname{ord}_{D}|_{\mathbb{C}(V)}, in which case we see via [Li22] that ordD=vμ,nξ\operatorname{ord}_{D}=v_{\mu,n\xi} for some nn\in\mathbb{N}, and denote the divisor associated to μ\mu as DμD_{\mu}. Then βV,aB(Dμ)=βY(Dμ,Y)\beta_{V,aB}(D_{\mu})=\beta_{Y}(D_{\mu,Y}), with Dμ,YD_{\mu,Y} as in Convention 4.12 with respect to DμD_{\mu}. Since ordDμ,Y=vμ,0\operatorname{ord}_{D_{\mu,Y}}=v_{\mu,0}, we see that Dμ,YD_{\mu,Y} and DD differ by a twist and thus βV,aB(Dμ)=βY(Dμ,Y)=βY(D)=0\beta_{V,aB}(D_{\mu})=\beta_{Y}(D_{\mu,Y})=\beta_{Y}(D)=0. Thus, by our assumption that the log pair (V,aB)(V,aB) is K-polystable, the test configuration associated to DμD_{\mu} is a product test configuration.

In what follows, we use a linearization on \mathcal{L} of the 𝔾m\mathbb{G}_{m} action σμ\sigma_{\mu} on VV induced by the product test configuration associated to DμD_{\mu}. Such a linearization exists by [KKLV89, Proposition 2.4] and its antecedent remark.

Denote the test configuration associated to μ\mu as 𝒱μV×𝔸1\mathcal{V}_{\mu}\cong V\times\mathbb{A}^{1}. Lifting \mathcal{L} along the first projection to pr2()=:𝕃\text{pr}_{2}^{*}(\mathcal{L})=:\mathbb{L}, we can then consider the variety 𝒳:=𝒱μ(𝕃𝒪𝒱μ)\mathcal{X}:=\mathbb{P}_{\mathcal{V}_{\mu}}(\mathbb{L}\oplus\mathcal{O}_{\mathcal{V}_{\mu}}) (which in fact is the test configuration of XX associated to some 𝕋\mathbb{T}-invariant lift of the valuation μ\mu). The aforementioned 𝔾m\mathbb{G}_{m}-linearization of σμ\sigma_{\mu} on \mathcal{L} lifts to a linearization on 𝕃𝒪V×A1\mathbb{L}\oplus\mathcal{O}_{V\times A^{1}}, thus we have an induced 𝔾m\mathbb{G}_{m} action on 𝒳\mathcal{X} that fixes a positive section of the 1\mathbb{P}^{1}-bundle 𝒳V×𝔸1\mathcal{X}\to V\times\mathbb{A}^{1}. Thus, we may define the variety 𝒴\mathcal{Y} to be the blow-up of 𝒳\mathcal{X} along the image of B×𝔸1B\times\mathbb{A}^{1} along the 𝕋\mathbb{T}-invariant positive section. Thus, the 𝔾m\mathbb{G}_{m} action lifts to 𝒴\mathcal{Y}, which is in fact a product test configuration for YY.

By construction we have (𝒴)(Y)(s)(V)(t,s)\mathbb{C}(\mathcal{Y})\cong\mathbb{C}(Y)(s)\cong\mathbb{C}(V)(t,s), and the valuation associated to the test configuration 𝒴\mathcal{Y} is the restriction w:=ords=0|(Y)w:=\operatorname{ord}_{s=0}|_{\mathbb{C}(Y)}. The further restriction of ww to (V)\mathbb{C}(V) is in fact μ\mu, so w=vμ,kξw=v_{\mu,k\xi} for some kk\in\mathbb{N}, and thus ww is equivalent to some twist of ordD\operatorname{ord}_{D}, so their associated test configurations are isomorphic as varieties by [Li22, Example 3.7], and thus the test configuration associated to DD is a product test configuration. ∎

Proof of Theorem 1.1.

By combining Proposition 4.16 and Proposition 4.8, we have that (V,aB)(V,aB) is K-semistable if and only if YY is K-semistable. Similarly, Proposition 4.17 and Proposition 4.18 together show that (V,aB)(V,aB) is K-polystable if and only if YY is K-polystable. ∎

With Theorem 1.1 established, we now prove Corollary 1.2 through an application of interpolation of K-stability.

Proof of Corollary 1.2.

By [LZ22, Zhu21], the K-polystability of WW is equivalent to that of the pair (V,12B)(V,\frac{1}{2}B). A quick calculation shows that 0<a(n,r)<120<a(n,r)<\frac{1}{2} for all n,rn,r, so by interpolation of K-stability (see e.g. [ADL24, Proposition 2.13]), we have that (V,aB)(V,aB) is K-polystable, so by Theorem 1.1, YY is K-polystable. ∎

5. Examples

Example 5.1.

The Fano family №3.93.9 (previously known, see [ACC+23]): Letting V=2V=\mathbb{P}^{2}, r=32r=\frac{3}{2}, and BB be a smooth quartic curve, then \mathcal{L} is 𝒪(2)\mathcal{O}(2), XX is (𝒪(2)𝒪)\mathbb{P}(\mathcal{O}(2)\oplus\mathcal{O}), and we have then that by Theorem 1.1 YY is a member of the Fano family №3.93.9 with K-poly/semistability equivalent to that of the pair (V,952B)(V,\frac{9}{52}B). By [ADL24], we see that such pairs are K-poly/semistable exactly when the quartic plane curve BB is poly/semistable in the GIT sense, both of which are implied in this case by the smoothness of BB. This provides another proof that all smooth members of family №3.93.9 are K-polystable, as previously shown in [ACC+23].

Example 5.2.

The Fano family №3.193.19 (previously known, see [ACC+23]): Letting V=2V=\mathbb{P}^{2}, r=3r=3, and BB be a smooth conic, we have then that =𝒪V(1)\mathcal{L}=\mathcal{O}_{V}(1), XX is 3\mathbb{P}^{3} blown up at a point, and YY is a member of the Fano family №3.193.19 with K-poly/semistability equivalent to that of the pair (V,33152B)(V,\frac{33}{152}B) by Theorem 1.1. By [LS14], we see that such a pair K-polystable, thus providing another proof that the unique smooth Fano variety in family №3.193.19 is K-polystable, originally shown in [IS17].

Example 5.3.

The Fano family №4.24.2 (previously known, see [ACC+23]): Let V=1×1V=\mathbb{P}^{1}\times\mathbb{P}^{1}, r=2r=2, and BB be a smooth curve of bidegree (2,2)(2,2). Then, =𝒪V(1,1)\mathcal{L}=\mathcal{O}_{V}(1,1), X=(𝒪(1,1)𝒪)X=\mathbb{P}(\mathcal{O}(1,1)\oplus\mathcal{O}), and YY is a smooth member of the Fano family №4.24.2 and, by the above theorem, the K-polystability of YY is equivalent to that of the pair (V,1156B)(V,\frac{11}{56}B). The K-polystability of this pair follows from the interpolation of K-stability (see [ADL24, Proposition 2.13] and [ADL23, Theorem 2.10]) since VV is K-polystable and (V,B)(V,B) is a plt log Calabi-Yau pair. Thus, by Theorem 1.1, such YY is K-polystable, giving a new proof of this result previously shown in [ACC+23].

Example 5.4.

New examples from blow-ups related to quartic surfaces in 3\mathbb{P}^{3} and higher dimensional analogs: Let V=3V=\mathbb{P}^{3}, r=2r=2, and BB a smooth quartic surface in 3\mathbb{P}^{3}. Then, =𝒪V(2)\mathcal{L}=\mathcal{O}_{V}(2), and YY is the blow-up of the cone over the second Veronese embedding of 3\mathbb{P}^{3} in 9\mathbb{P}^{9} along the cone point and a quartic surface in the base. In this case, (V,aB)=(3,1375B)(V,aB)=(\mathbb{P}^{3},\frac{13}{75}B). By [ADL23], such pairs are K-poly/semistable exactly when they are poly/semistable in the GIT sense. As BB is smooth, it is GIT polystable, thus all smooth fourfolds constructed in this manner are K-polystable by Theorem 1.1.

We generalize this to higher dimensions: let V=n1V=\mathbb{P}^{n-1} for nn even, r=2r=2, and BB a smooth degree nn hypersurface in VV. Then =𝒪V(n2)\mathcal{L}=\mathcal{O}_{V}(\frac{n}{2}), and YY is the blow-up of the cone over the embedding ϕ||:VN\phi_{|\mathcal{L}|}:V\hookrightarrow\mathbb{P}^{N} along the cone point and the inclusion of BB in the base of the cone. By [ADL24, Theorem 1.4], there exists some c1c_{1} such that (V,cB)(V,cB) is K-semi/polystable if and only if BB is semi/polystable in the GIT sense for all c<c1c<c_{1}. If a=a(n,2)<c1a=a(n,2)<c_{1}, then BB smooth implies BB is polystable in the GIT sense and thus (V,aB)(V,aB) is K-polystable, implying by Theorem 1.1 that YY is K-polystable. If c1ac_{1}\leq a, then again since BB is smooth, (V,cB)(V,cB) is K-polystable for some c<ac<a. Then, since (V,B)(V,B) is a plt log Calabi-Yau pair, (V,(1ϵ)B)(V,(1-\epsilon)B) is K-polystable for some sufficiently small ϵ\epsilon (see [ADL23, Theorem 2.10]), and thus by interpolation of K-stability (see [ADL24, Proposition 2.13]), we again have that (V,aB)(V,aB) is K-polystable, implying the K-polystability of YY by Theorem 1.1. This improves [CGF+23, Theorem 1.9].

Example 5.5.

New K-unstable example: Let V=Blp3V=\mathrm{Bl}_{p}\mathbb{P}^{3}, r=2r=2, and BB a smooth member of |KV||-K_{V}|. Then, =π(2H)E\mathcal{L}=\pi^{*}(2H)-E where π:V3\pi:V\to\mathbb{P}^{3} is the blow-up map and EE is the exceptional divisor of the blow-up, X=V(𝒪V)X=\mathbb{P}_{V}(\mathcal{L}\oplus\mathcal{O}_{V}), and YY is BlBX\mathrm{Bl}_{B_{\infty}}X. As in the previous example, a=1375a=\frac{13}{75}. A straightforward computation shows that the pair (V,aB)(V,aB) is K-unstable with βV,aB(π(H))<0\beta_{V,aB}(\pi^{*}(H))<0, where π(H)\pi^{*}(H) is the pullback of the hyperplane section of 3\mathbb{P}^{3} along the blow-up. Thus, by Theorem 1.1, YY is K-unstable.

6. Other Blow-ups of Projective Compactifications of Proportional Line Bundles

In this section, we see that, when replacing the assumption that l=2l=2 in the construction of YY with l2l\neq 2 (where BlB\sim_{\mathbb{Q}}l\mathcal{L}), the resulting Fano varieties YY are always K-unstable.

Theorem 6.1 (Theorem 1.3).

Let YY be constructed as above with VV and BB smooth and BlB\sim_{\mathbb{Q}}l\mathcal{L} for 0<l<r+1,l20<l<r+1,l\neq 2. Then YY is K-unstable. Furthermore, either the strict transform of the image of the positive section containing BB_{\infty}, denoted as V¯\overline{V}_{\infty}, or the strict transform of the zero section, V0V_{0}, is a destabilizing divisor for YY.

Lemma 6.2.
βY(V0)+βY(V¯)=0\displaystyle\beta_{Y}(V_{0})+\beta_{Y}(\overline{V}_{\infty})=0
Proof.

As V0,V¯V_{0},\overline{V}_{\infty} are both divisors on YY, AY(V0)=AY(V¯)=1A_{Y}(V_{0})=A_{Y}(\overline{V}_{\infty})=1, so βY(V0)+βY(V¯)=2(SY(V0)+SY(V¯))\beta_{Y}(V_{0})+\beta_{Y}(\overline{V}_{\infty})=2-(S_{Y}(V_{0})+S_{Y}(\overline{V}_{\infty})). It remains to show the sum SY(V0)+SY(V¯)S_{Y}(V_{0})+S_{Y}(\overline{V}_{\infty}) is equal to 22.

The Zariski decompositions of KYtV¯,KYtV0-K_{Y}-t\overline{V}_{\infty},-K_{Y}-tV_{0} on YY are as follows:

P(t)={KYtV¯=(1t)V¯+πϕ(KV)+V0,if 0t1(2t)H+r1rπϕ(KV)=r+1trπϕ(KV)+(2t)V0,if 1t2P_{\infty}(t)=\begin{cases}-K_{Y}-t\overline{V}_{\infty}=(1-t)\overline{V}_{\infty}+\pi^{*}\phi^{*}(-K_{V})+V_{0},&\text{if }0\leq t\leq 1\\ (2-t)H+\frac{r-1}{r}\pi^{*}\phi^{*}(-K_{V})=\frac{r+1-t}{r}\pi^{*}\phi^{*}(-K_{V})+(2-t)V_{0},&\text{if }1\leq t\leq 2\end{cases}
N(t)={0,if 0t1(1t)E,if 1t2N_{\infty}(t)=\begin{cases}0,&\text{if }0\leq t\leq 1\\ (1-t)E,&\text{if }1\leq t\leq 2\end{cases}
P0(t)={KYtV0=V¯+πϕ(KV)+(1t)V0,if 0t1(2t)V¯+r(t1)(l1)rπϕ(KV),if 1t2P_{0}(t)=\begin{cases}-K_{Y}-tV_{0}=\overline{V}_{\infty}+\pi^{*}\phi^{*}(-K_{V})+(1-t)V_{0},&\text{if }0\leq t\leq 1\\ (2-t)\overline{V}_{\infty}+\frac{r-(t-1)(l-1)}{r}\pi^{*}\phi^{*}(-K_{V}),&\text{if }1\leq t\leq 2\end{cases}
N0(t)={0,if 0t1(t1)F=(t1)(V¯+l1rπϕ(KV)V0),if 1t2N_{0}(t)=\begin{cases}0,&\text{if }0\leq t\leq 1\\ (t-1)F=(t-1)(\overline{V}_{\infty}+\frac{l-1}{r}\pi^{*}\phi^{*}(-K_{V})-V_{0}),&\text{if }1\leq t\leq 2\end{cases}

This can be seen as follows: for the range 0t10\leq t\leq 1, KYtV¯-K_{Y}-t\overline{V}_{\infty} and KYtV0-K_{Y}-tV_{0} are both nef, thus the negative parts of their Zariski decompositions are trivial. For KYtV¯-K_{Y}-t\overline{V}_{\infty} with 1t21\leq t\leq 2, we observe that PP_{\infty} is the pullback along the contraction π\pi of a nef class on XX and NN_{\infty} is supported on the exceptional locus of the same contraction. Thus by [Oka16, Proposition 2.13], we see that this is the Zariski decomposition of KYtV¯-K_{Y}-t\overline{V}_{\infty}. For l1l\neq 1, a similar argument shows that P0+N0P_{0}+N_{0} is the Zariski decomposition of KYtV0-K_{Y}-tV_{0}, using the contraction π:YX:=V(𝒪V)\pi^{\prime}:Y\to X^{\prime}:=\mathbb{P}_{V}(\mathcal{L}^{\prime}\oplus\mathcal{O}_{V}) (for some line bundle \mathcal{L}^{\prime} such that lKVl^{\prime}\mathcal{L}^{\prime}\sim_{\mathbb{Q}}-K_{V} where (l1)(l1)=1(l^{\prime}-1)(l-1)=1) that contracts FF. For l=1l=1, we have the same contraction map π:YX:=V(𝒪V𝒪V)\pi^{\prime}:Y\to X^{\prime}:=\mathbb{P}_{V}(\mathcal{O}_{V}\oplus\mathcal{O}_{V}) but now for =𝒪V\mathcal{L}^{\prime}=\mathcal{O}_{V}.

We have the following intersection products on YY, for k>0k>0:

V0kπϕ(KV)nk\displaystyle V_{0}^{k}\cdot\pi^{*}\phi^{*}(-K_{V})^{n-k} =V0k1V0πϕ(KV)nk\displaystyle=V_{0}^{k-1}\cdot V_{0}\cdot\pi^{*}\phi^{*}(-K_{V})^{n-k}
=(H1rπϕ(KV))k1V0πϕ(KV)nk\displaystyle=(H-\frac{1}{r}\pi^{*}\phi^{*}(-K_{V}))^{k-1}\cdot V_{0}\cdot\pi^{*}\phi^{*}(-K_{V})^{n-k}
=(1rπϕ(KV))k1V0πϕ(KV)nk\displaystyle=(-\frac{1}{r}\pi^{*}\phi^{*}(-K_{V}))^{k-1}\cdot V_{0}\cdot\pi^{*}\phi^{*}(-K_{V})^{n-k}
=(1r)k1vol(V)\displaystyle=(\frac{-1}{r})^{k-1}\operatorname{vol}(V)
V¯kπϕ(KV)nk\displaystyle\overline{V}_{\infty}^{k}\cdot\pi^{*}\phi^{*}(-K_{V})^{n-k} =V¯k1V¯πϕ(KV)nk\displaystyle=\overline{V}_{\infty}^{k-1}\cdot\overline{V}_{\infty}\cdot\pi^{*}\phi^{*}(-K_{V})^{n-k}
=(V0+1rπϕ(KV)E)k1V¯πϕ(KV)nk\displaystyle=(V_{0}+\frac{1}{r}\pi^{*}\phi^{*}(-K_{V})-E)^{k-1}\cdot\overline{V}_{\infty}\cdot\pi^{*}\phi^{*}(-K_{V})^{n-k}
=(1rπϕ(KV)E)k1V¯πϕ(KV)nk\displaystyle=(\frac{1}{r}\pi^{*}\phi^{*}(-K_{V})-E)^{k-1}\cdot\overline{V}_{\infty}\cdot\pi^{*}\phi^{*}(-K_{V})^{n-k}
=(1lr)k1vol(V)\displaystyle=(\frac{1-l}{r})^{k-1}\operatorname{vol}(V)

Let us assume l1l\neq 1, then we have for vol(Y)SY(V¯)\operatorname{vol}(Y)S_{Y}(\overline{V}_{\infty}):

vol(Y)SY(V¯)\displaystyle\operatorname{vol}(Y)S_{Y}(\overline{V}_{\infty}) =01((1t)V¯+πϕ(KV)+V0)n𝑑t\displaystyle=\int_{0}^{1}\left((1-t)\overline{V}_{\infty}+\pi^{*}\phi^{*}(-K_{V})+V_{0}\right)^{n}dt
+12(r+1trπϕ(KV)+(2t)V0)n𝑑t\displaystyle+\int_{1}^{2}\left(\frac{r+1-t}{r}\pi^{*}\phi^{*}(-K_{V})+(2-t)V_{0}\right)^{n}dt
=vol(V)rn1(0111l((1t)(1l)+r)nrn)dt\displaystyle=\frac{\operatorname{vol}(V)}{r^{n-1}}(\int_{0}^{1}\frac{1}{1-l}\left((1-t)(1-l)+r)^{n}-r^{n}\right)dt
+01rn(r1)ndt12(r1)n(r+1t)ndt)\displaystyle+\int_{0}^{1}r^{n}-(r-1)^{n}dt-\int_{1}^{2}(r-1)^{n}-(r+1-t)^{n}dt)
=vol(V)rn101((1t)(1l)+r)nrn1l+rn2(r1)n+(r+t1)ndt\displaystyle=\frac{\operatorname{vol}(V)}{r^{n-1}}\int_{0}^{1}\frac{((1-t)(1-l)+r)^{n}-r^{n}}{1-l}+r^{n}-2(r-1)^{n}+(r+t-1)^{n}dt

Similarly, for vol(Y)SY(V¯)\operatorname{vol}(Y)S_{Y}(\overline{V}_{\infty}):

vol(Y)SY(V0)\displaystyle\operatorname{vol}(Y)S_{Y}(V_{0}) =01(V¯+πϕ(KV)+(1t)V0)n𝑑t\displaystyle=\int_{0}^{1}\left(\overline{V}_{\infty}+\pi^{*}\phi^{*}(-K_{V})+(1-t)V_{0}\right)^{n}dt
+12((2t)V¯+r(t1)(l1)rπϕ(KV))n𝑑t\displaystyle+\int_{1}^{2}\left((2-t)\overline{V}_{\infty}+\frac{r-(t-1)(l-1)}{r}\pi^{*}\phi^{*}(-K_{V})\right)^{n}dt
=vol(V)rn1(0111l((r+1l)nrn)dt+01rn(r+t1)ndt\displaystyle=\frac{\operatorname{vol}(V)}{r^{n-1}}(\int_{0}^{1}\frac{1}{1-l}\left((r+1-l)^{n}-r^{n}\right)dt+\int_{0}^{1}r^{n}-(r+t-1)^{n}dt
+1211l((r+1l)n(r(t1)(l1))n)dt)\displaystyle+\int_{1}^{2}\frac{1}{1-l}\left((r+1-l)^{n}-(r-(t-1)(l-1))^{n}\right)dt)
=vol(V)rn101(r+1l)nrn1l+rn(r+t1)n\displaystyle=\frac{\operatorname{vol}(V)}{r^{n-1}}\int_{0}^{1}\frac{(r+1-l)^{n}-r^{n}}{1-l}+r^{n}-(r+t-1)^{n}
+(r+1l)n(r+(t1)(l1))n1ldt\displaystyle+\frac{(r+1-l)^{n}-(r+(t-1)(l-1))^{n}}{1-l}dt

Summing these two terms we get precisely 2vol(Y)2\operatorname{vol}(Y), and so β(V0)+β(V¯)=0\beta(V_{0})+\beta(\overline{V}_{\infty})=0.

Now, for l=1l=1, we have for vol(Y)SY(V¯)\operatorname{vol}(Y)S_{Y}(\overline{V}_{\infty}):

vol(Y)SY(V¯)\displaystyle\operatorname{vol}(Y)S_{Y}(\overline{V}_{\infty}) =01((1t)V¯+πϕ(KV)+V0)n𝑑t\displaystyle=\int_{0}^{1}\left((1-t)\overline{V}_{\infty}+\pi^{*}\phi^{*}(-K_{V})+V_{0}\right)^{n}dt
+12(r+1trπϕ(KV)+(2t)V0)n𝑑t\displaystyle+\int_{1}^{2}\left(\frac{r+1-t}{r}\pi^{*}\phi^{*}(-K_{V})+(2-t)V_{0}\right)^{n}dt
=vol(V)rn1(01n(1t)rn1+rn(r1)ndt\displaystyle=\frac{\operatorname{vol}(V)}{r^{n-1}}(\int_{0}^{1}n(1-t)r^{n-1}+r^{n}-(r-1)^{n}dt
12(r1)n(r+1t)ndt)\displaystyle-\int_{1}^{2}(r-1)^{n}-(r+1-t)^{n}dt)
=vol(V)rn101n(1t)rn1+rn2(r1)n+(r+t1)ndt\displaystyle=\frac{\operatorname{vol}(V)}{r^{n-1}}\int_{0}^{1}n(1-t)r^{n-1}+r^{n}-2(r-1)^{n}+(r+t-1)^{n}dt

and similarly for vol(Y)SY(V0)\operatorname{vol}(Y)S_{Y}(V_{0}):

vol(Y)SY(V0)\displaystyle\operatorname{vol}(Y)S_{Y}(V_{0}) =01(V¯+πϕ(KV)+(1t)V0)n𝑑t\displaystyle=\int_{0}^{1}\left(\overline{V}_{\infty}+\pi^{*}\phi^{*}(-K_{V})+(1-t)V_{0}\right)^{n}dt
+12((2t)V¯+r(t1)(l1)rπϕ(KV))n𝑑t\displaystyle+\int_{1}^{2}\left((2-t)\overline{V}_{\infty}+\frac{r-(t-1)(l-1)}{r}\pi^{*}\phi^{*}(-K_{V})\right)^{n}dt
=vol(V)rn1(01nrn1+rn(r+t1)ndt+12n(2t)rn1𝑑t)\displaystyle=\frac{\operatorname{vol}(V)}{r^{n-1}}(\int_{0}^{1}nr^{n-1}+r^{n}-(r+t-1)^{n}dt+\int_{1}^{2}n(2-t)r^{n-1}dt)

So for the l=1l=1 case, summing these two terms we again get 2vol(Y)2\operatorname{vol}(Y), and so β(V0)+β(V¯)=0\beta(V_{0})+\beta(\overline{V}_{\infty})=0. ∎

Thus, to finish the prove of Theorem 1.3, we simply must show that βY(V¯)0\beta_{Y}(\overline{V}_{\infty})\neq 0 for l2l\neq 2.

Proof.

In fact, we show that rn1vol(Y)vol(V)βY(V¯)0\frac{r^{n-1}\operatorname{vol}(Y)}{\operatorname{vol}(V)}\beta_{Y}(\overline{V}_{\infty})\neq 0.

vol(Y)SY(V¯)\displaystyle\operatorname{vol}(Y)S_{Y}(\overline{V}_{\infty}) =01((1t)V¯+πϕ(KV)+V0)n𝑑t\displaystyle=\int_{0}^{1}\left((1-t)\overline{V}_{\infty}+\pi^{*}\phi^{*}(-K_{V})+V_{0}\right)^{n}dt
+12(r+1trπϕ(KV)+(2t)V0)n𝑑t\displaystyle+\int_{1}^{2}\left(\frac{r+1-t}{r}\pi^{*}\phi^{*}(-K_{V})+(2-t)V_{0}\right)^{n}dt
=vol(V)rn1(0111l((1t)(1l)+r)nrn)dt\displaystyle=\frac{\operatorname{vol}(V)}{r^{n-1}}(\int_{0}^{1}\frac{1}{1-l}\left((1-t)(1-l)+r)^{n}-r^{n}\right)dt
+01rn(r1)ndt+12(r1)n(r+1t)ndt)\displaystyle+\int_{0}^{1}r^{n}-(r-1)^{n}dt+\int_{1}^{2}(r-1)^{n}-(r+1-t)^{n}dt)
=01((1t)(1l)+r)nrn1l+rn2(r1)n+(r+t1)ndt\displaystyle=\int_{0}^{1}\frac{((1-t)(1-l)+r)^{n}-r^{n}}{1-l}+r^{n}-2(r-1)^{n}+(r+t-1)^{n}dt
rn1vol(Y)vol(V)βY(V¯)\displaystyle\frac{r^{n-1}\operatorname{vol}(Y)}{\operatorname{vol}(V)}\beta_{Y}(\overline{V}_{\infty}) =rn1vol(Y)vol(V)rn1vol(Y)vol(V)SY(V¯)\displaystyle=\frac{r^{n-1}\operatorname{vol}(Y)}{\operatorname{vol}(V)}-\frac{r^{n-1}\operatorname{vol}(Y)}{\operatorname{vol}(V)}S_{Y}(\overline{V}_{\infty})
=(rn(r+1l)nl1+rn(r1)n)\displaystyle=\left(\frac{r^{n}-(r+1-l)^{n}}{l-1}+r^{n}-(r-1)^{n}\right)
01((1t)(1l)+r)nrn1l+rn2(r1)n+(r+t1)ndt\displaystyle-\int_{0}^{1}\frac{((1-t)(1-l)+r)^{n}-r^{n}}{1-l}+r^{n}-2(r-1)^{n}+(r+t-1)^{n}dt
=01(rn(r+1l)nl1+rn(r1)n)\displaystyle=\int_{0}^{1}\left(\frac{r^{n}-(r+1-l)^{n}}{l-1}+r^{n}-(r-1)^{n}\right)
((1t)(1l)+r)nrn1lrn+2(r1)n(r+t1)ndt\displaystyle-\frac{((1-t)(1-l)+r)^{n}-r^{n}}{1-l}-r^{n}+2(r-1)^{n}-(r+t-1)^{n}dt
=01(r(t1)(1l))n(r+1l)nl1+(r1)n(r+t1)ndt\displaystyle=\int_{0}^{1}\frac{(r-(t-1)(1-l))^{n}-(r+1-l)^{n}}{l-1}+(r-1)^{n}-(r+t-1)^{n}dt

We will show the integrand is strictly positive (resp. negative) when r>2r>2 (resp. r<2r<2) for 0<t<10<t<1. Let f(x)=(rx)nf(x)=(r-x)^{n}, w=1tw=1-t and y=l1y=l-1. Then the integrand is equal to

(1w)(f(1)f(w)1wf(wy)f(y)wyy)(1-w)\left(\frac{f(1)-f(w)}{1-w}-\frac{f(wy)-f(y)}{wy-y}\right)

Then, since ff is strictly convex on the interval (0,r)(0,r), we have, for l>2l>2,

f(1)f(w)1w<f(1)f(wx)1wx<f(wy)f(y)wyy\frac{f(1)-f(w)}{1-w}<\frac{f(1)-f(wx)}{1-wx}<\frac{f(wy)-f(y)}{wy-y}

with the inequalities reversing for l<2l<2. Thus the integrand is strictly positive (resp. negative), so βY(V¯)0\beta_{Y}(\overline{V}_{\infty})\neq 0 as desired. ∎

6.1. Examples of Other Blow-ups

Example 6.3.

The Fano family №3.143.14: Letting VV be 2\mathbb{P}^{2}, r=3r=3, and BB a smooth planar cubic curve, so that l=3l=3, we have that X=Blp3X=\mathrm{Bl}_{p}\mathbb{P}^{3} is the blow-up of 3\mathbb{P}^{3} at a point and YY is a smooth member of the family of Fano threefolds №3.143.14, and all smooth members of said family are obtained in this way. Thus, Theorem 1.3 recovers the K-unstability of members of this family, originally due to Fujita [Fuj16, Theorem 1.4].

Example 6.4.

K-unstable in each dimension: Let VV be n1\mathbb{P}^{n-1}, r=nr=n, and BB a degree dd hypersurface in VV, for d<nd<n. Then X=BlpnX=\mathrm{Bl}_{p}\mathbb{P}^{n} the blow up of n\mathbb{P}^{n} at a point and YY is the blow-up of projective nn-space along a codimension 22 subvariety contained in the pullback of a hyperplane in XX that doesn’t contain pp. Such YY is K-unstable by Theorem 1.3. Thus, we have for every nn several examples of a K-unstable Fano variety of dimension nn.

References

  • [ACC+23] Carolina Araujo, Ana-Maria Castravet, Ivan Cheltsov, Kento Fujita, Anne-Sophie Kaloghiros, Jesus Martinez-Garcia, Constantin Shramov, Hendrik Süß, and Nivedita Viswanathan. The Calabi problem for Fano threefolds, volume 485 of London Mathematical Society Lecture Note Series. Cambridge University Press, Cambridge, 2023.
  • [ADL23] Kenneth Ascher, Kristen DeVleming, and Yuchen Liu. K-stability and birational models of moduli of quartic k3 surfaces. Inventiones mathematicae, 232:471–552, 2023.
  • [ADL24] Kenneth Ascher, Kristin DeVleming, and Yuchen Liu. Wall crossing for K-moduli spaces of plane curves. Proc. Lond. Math. Soc. (3), 128(6):Paper No. e12615, 113, 2024.
  • [AZ22] Hamid Abban and Ziquan Zhuang. K-stability of Fano varieties via admissible flags. Forum Math. Pi, 10:Paper No. e15, 43, 2022.
  • [AZ23] Hamid Abban and Ziquan Zhuang. Seshadri constants and KK-stability of Fano manifolds. Duke Math. J., 172(6):1109–1144, 2023.
  • [BdFFU15] S. Boucksom, T. de Fernex, C. Favre, and S. Urbinati. Valuation spaces and multiplier ideals on singular varieties. In Recent advances in algebraic geometry, volume 417 of London Math. Soc. Lecture Note Ser., pages 29–51. Cambridge Univ. Press, Cambridge, 2015.
  • [BJ20] Harold Blum and Mattias Jonsson. Thresholds, valuations, and K-stability. Adv. Math., 365:107062, 57, 2020.
  • [CDS15] Xiuxiong Chen, Simon Donaldson, and Song Sun. Kähler-Einstein metrics on Fano manifolds. I, II, III. J. Amer. Math. Soc., 28(1):235–278, 2015.
  • [CGF+23] Ivan Cheltsov, Tiago Duarte Guerreiro, Kento Fujita, Igor Krylov, and Jesus Martinez-Garcia. K-stability of Casagrande-Druel varieties. J. Reine Angew. Math., to appear, 2023. arXiv:2309.12522.
  • [Che08] Ivan Cheltsov. Log canonical thresholds of del Pezzo surfaces. Geom. Funct. Anal., 18(4):1118–1144, 2008.
  • [Don02] S. K. Donaldson. Scalar curvature and stability of toric varieties. J. Differential Geom., 62(2):289–349, 2002.
  • [DS16] Ved Datar and Gábor Székelyhidi. Kähler-Einstein metrics along the smooth continuity method. Geom. Funct. Anal., 26(4):975–1010, 2016.
  • [FO18] Kento Fujita and Yuji Odaka. On the K-stability of Fano varieties and anticanonical divisors. Tohoku Math. J. (2), 70(4):511–521, 2018.
  • [Fuj16] Kento Fujita. On KK-stability and the volume functions of \mathbb{Q}-Fano varieties. Proc. Lond. Math. Soc. (3), 113(5):541–582, 2016.
  • [Fuj19a] Kento Fujita. K-stability of Fano manifolds with not small alpha invariants. J. Inst. Math. Jussieu, 18(3):519–530, 2019.
  • [Fuj19b] Kento Fujita. A valuative criterion for uniform K-stability of \mathbb{Q}-Fano varieties. J. Reine Angew. Math., 751:309–338, 2019.
  • [IS17] Nathan Ilten and Hendrik Süß. K-stability for Fano manifolds with torus action of complexity 1. Duke Mathematical Journal, 166(1), January 2017.
  • [JM12] Mattias Jonsson and Mircea Mustaţă. Valuations and asymptotic invariants for sequences of ideals. Ann. Inst. Fourier (Grenoble), 62(6):2145–2209, 2012.
  • [KKLV89] Friedrich Knop, Hanspeter Kraft, Domingo Luna, and Thierry Vust. Local properties of algebraic group actions. In Hanspeter Kraft, Peter Slodowy, and Tonny A. Springer, editors, Algebraische Transformationsgruppen und Invariantentheorie Algebraic Transformation Groups and Invariant Theory, pages 63–75. Birkhäuser Basel, Basel, 1989.
  • [KM98] János Kollár and Shigefumi Mori. Birational geometry of algebraic varieties, volume 134 of Cambridge Tracts in Mathematics. Cambridge University Press, Cambridge, 1998. With the collaboration of C. H. Clemens and A. Corti, Translated from the 1998 Japanese original.
  • [Li17] Chi Li. K-semistability is equivariant volume minimization. Duke Math. J., 166(16):3147–3218, 2017.
  • [Li22] Chi Li. GG-uniform stability and Kähler-Einstein metrics on Fano varieties. Invent. Math., 227(2):661–744, 2022.
  • [LS14] Chi Li and Song Sun. Conical Kähler-Einstein metrics revisited. Comm. Math. Phys., 331(3):927–973, 2014.
  • [LWX21] Chi Li, Xiaowei Wang, and Chenyang Xu. Algebraicity of the metric tangent cones and equivariant K-stability. J. Amer. Math. Soc., 34(4):1175–1214, 2021.
  • [LX20] Chi Li and Chenyang Xu. Stability of valuations and Kollár components. J. Eur. Math. Soc. (JEMS), 22(8):2573–2627, 2020.
  • [LXZ22] Yuchen Liu, Chenyang Xu, and Ziquan Zhuang. Finite generation for valuations computing stability thresholds and applications to K-stability. Ann. of Math. (2), 196(2):507–566, 2022.
  • [LZ22] Yuchen Liu and Ziwen Zhu. Equivariant KK-stability under finite group action. Internat. J. Math., 33(1):Paper No. 2250007, 21, 2022.
  • [MM93] Toshiki Mabuchi and Shigeru Mukai. Stability and Einstein-Kähler metric of a quartic del Pezzo surface. In Einstein metrics and Yang-Mills connections (Sanda, 1990), volume 145 of Lecture Notes in Pure and Appl. Math., pages 133–160. Dekker, New York, 1993.
  • [Oka16] Shinnosuke Okawa. On images of Mori dream spaces. Math. Ann., 364(3-4):1315–1342, 2016.
  • [OS12] Yuji Odaka and Yuji Sano. Alpha invariant and K-stability of \mathbb{Q}-Fano varieties. Adv. Math., 229(5):2818–2834, 2012.
  • [OSS16] Yuji Odaka, Cristiano Spotti, and Song Sun. Compact moduli spaces of del Pezzo surfaces and Kähler-Einstein metrics. J. Differential Geom., 102(1):127–172, 2016.
  • [PW18] Jihun Park and Joonyeong Won. K-stability of smooth del Pezzo surfaces. Math. Ann., 372(3-4):1239–1276, 2018.
  • [Tia87] Gang Tian. On Kähler-Einstein metrics on certain Kähler manifolds with C1(M)>0C_{1}(M)>0. Invent. Math., 89(2):225–246, 1987.
  • [Tia90] Gang Tian. On Calabi’s conjecture for complex surfaces with positive first Chern class. Invent. Math., 101(1):101–172, 1990.
  • [Tia97] Gang Tian. Kähler-Einstein metrics with positive scalar curvature. Invent. Math., 130(1):1–37, 1997.
  • [Tia15] Gang Tian. K-stability and Kähler-Einstein metrics. Comm. Pure Appl. Math., 68(7):1085–1156, 2015.
  • [Wan24] Linsheng Wang. A valuative criterion of K-polystability. 2024. arXiv:2406.06176.
  • [Xu21] Chenyang Xu. K-stability of Fano varieties: an algebro-geometric approach. EMS Surv. Math. Sci., 8(1-2):265–354, 2021.
  • [Xu24] Chenyang Xu. K-stability of Fano varieties. 2024. https://web.math.princeton.edu/~chenyang/Kstabilitybook.pdf.
  • [Zha23] Junyan Zhao. Compactifications of moduli of del Pezzo surfaces via line arrangement and K-stability. Canadian Journal of Mathematics, 76(6):2115–2135, November 2023.
  • [Zha24] Junyan Zhao. Moduli of genus six curves and K-stability. Trans. Amer. Math. Soc. Ser. B, 11:863–900, 2024.
  • [Zho23] Chuyu Zhou. On wall-crossing for K-stability. Adv. Math., 413:Paper No. 108857, 26, 2023.
  • [Zhu21] Ziquan Zhuang. Optimal destabilizing centers and equivariant K-stability. Inventiones mathematicae, 226(1):195–223, April 2021.
  • [ZZ22] Kewei Zhang and Chuyu Zhou. Delta invariants of projective bundles and projective cones of Fano type. Math. Z., 300(1):179–207, 2022.