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

Anticanonically balanced metrics and the Hilbert–Mumford criterion for the δm\delta_{m}-invariant of Fujita–Odaka

Yoshinori Hashimoto Department of Mathematics, Osaka Metropolitan University, 3-3-138, Sugimoto, Sumiyoshi-ku, Osaka, 558-8585, Japan. [email protected]
Abstract.

We prove that the stability condition for Fano manifolds defined by Saito–Takahashi, given in terms of the sum of the Ding invariant and the Chow weight, is equivalent to the existence of anticanonically balanced metrics. Combined with the result by Rubinstein–Tian–Zhang, we obtain the following algebro-geometric corollary: the δm\delta_{m}-invariant of Fujita–Odaka satisfies δm>1\delta_{m}>1 if and only if the Fano manifold is stable in the sense of Saito–Takahashi, establishing a Hilbert–Mumford type criterion for δm>1\delta_{m}>1. We also extend this result to the Kähler–Ricci gg-solitons and the coupled Kähler–Einstein metrics, and as a by-product we obtain a formula for the asymptotic slope of the coupled Ding functional in terms of multiple test configurations.

1. Introduction

1.1. Background

Canonical metrics on Fano manifolds have been a focus of intensive studies in recent years, particularly in relation to the stability notions in Geometric Invariant Theory (GIT) [MFK]. There are many important results, but we mention here only the ones that are directly related to the results in this paper. The existence of Kähler–Einstein metrics on a Fano manifold with no nontrivial holomorphic vector fields is equivalent to the stability condition called the uniform Ding stability [BBJ, Theorem A]. This stability condition is given in terms of objects called the test configurations, which is a certain class of degenerations of Fano manifolds, and plays the role of the \mathbb{C}^{*}-action in the Hilbert–Mumford criterion of stability [MFK]. The uniform Ding stability can be characterised by an invariant called the δ\delta-invariant introduced by Fujita–Odaka [fo18], in the sense that it is equivalent to δ>1\delta>1 by [bj20, Fujita2019, fo18]. In a recent paper, K. Zhang [Zhang21] directly proved that δ>1\delta>1 implies the existence of Kähler–Einstein metrics, without using the uniform Ding stability itself and based instead on a finite dimensional approximation (often called quantisation) which is briefly reviewed below; indeed, all the objects mentioned above have finite dimensional approximations in an appropriate sense.

Donaldson [donproj1] proved a foundational theorem that a Kähler metric of constant scalar curvature (such as Kähler–Einstein metrics) can be approximated by a sequence of Fubini–Study metrics called the balanced metrics, as long as the automorphism group is discrete. Combined with the theorem due to H. Luo [Luo] and S. Zhang [zhang96] (see also Phong–Sturm [PS03]), which proves that a balanced metric exists if and only if the manifold is Chow stable, Donaldson’s theorem implies that a Kähler manifold admitting a constant scalar curvature Kähler metric is asymptotically Chow stable [donproj1, Corollary 4]. The asymptotic Chow stability can be regarded as a finite dimensional approximation of the KK-stability, which has a well-understood relationship to the Ding stability, in the sense that the Donaldson–Futaki invariant can be written as the limit of a sequence of Chow weights.

For Fano manifolds, we have a variant of balanced metrics that is called the anticanonically balanced metrics, which was introduced by Donaldson [donnum09, §2.2.2]. The analogue of Donaldson’s theorem [donproj1] above was established by Berman–Witt Nyström [BWN14], who proved that the existence of Kähler–Einstein metrics (resp. Kähler–Ricci gg-solitons) implies that we can find a sequence of anticanonically balanced metrics that converges to the Kähler–Einstein metric (resp. the Kähler–Ricci gg-soliton), in the sense of currents; see also [tak15] for the case of solitons. The convergence can in fact be improved to the smooth convergence by [tak19, Theorem 1.3 with N=1N=1] and [Ioos20, Ioos21]. From the point of view of stability, Saito–Takahashi [st19] defined a notion of stability that can be regarded as an anticanonical version of the Chow stability, and proved that it is implied by the existence of anticanonically balanced metrics.

The δ\delta-invariant of Fujita–Odaka also has a finite dimensional approximation called the δm\delta_{m}-invariant, which originally appeared as an intermediate object for the definition of the δ\delta-invariant. Recently, Rubinstein–Tian–Zhang [rtz] proved that the δm\delta_{m}-invariant characterises the existence of anticanonically balanced metrics. The theorem of K. Zhang [Zhang21] mentioned above relies on this result.

While many foundational works so far established a deep understanding of the uniform Ding stability, the δ\delta-invariant, and the Kähler–Einstein metrics, the relationship between the finite dimensional analogues of these concepts does not seem to be complete in the sense that no GIT stability condition, given in terms of a Hilbert–Mumford type criterion involving test configurations, has been proved to fully characterise δm>1\delta_{m}>1 or the anticanonically balanced metrics. This is the question that is addressed in this paper.

1.2. Statement of the results

The main result of this paper is the following, which can be regraded as the anticanonical version of the theorem by H. Luo [Luo] and S. Zhang [zhang96], and establishes the required correspondence between the anticanonically balanced metrics and the GIT stability.

Theorem 1.1.

Let mm\in\mathbb{N} be large enough such that mKX-mK_{X} is very ample. A Fano manifold (X,KX)(X,-K_{X}) admits an anticanonically balanced metric at level mm, which is unique up to Aut0(X)\mathrm{Aut}_{0}(X), if and only if it satisfies the following stability condition: for any very ample test configuration (𝒳,)(\mathcal{X},\mathcal{L}) for (X,KX)(X,-K_{X}) of exponent mm we have Ding(𝒳,)+Chowm(𝒳,)0\mathrm{Ding}(\mathcal{X},\mathcal{L})+\mathrm{Chow}_{m}(\mathcal{X},\mathcal{L})\geq 0, with equality if and only if (𝒳,)(\mathcal{X},\mathcal{L}) is product.

One direction of the above result, i.e. the existence of anticanonically balanced metrics implying the stated stability condition, was proved by Saito–Takahashi [st19, Theorem 1.2]. Thus the main point of the above theorem is the stability implying the anticanonically balanced metrics, although the proof that we give in this paper easily establishes both directions.

The main application of the above theorem, combined with Rubinstein–Tian–Zhang [rtz, Theorem 2.3], is the following result.

Corollary 1.2.

Suppose that a Fano manifold XX has no nontrivial holomorphic vector fields. For mm\in\mathbb{N} large enough such that mKX-mK_{X} is very ample, the δm\delta_{m}-invariant of Fujita–Odaka satisfies δm>1\delta_{m}>1 if and only if Ding(𝒳,)+Chowm(𝒳,)>0\mathrm{Ding}(\mathcal{X},\mathcal{L})+\mathrm{Chow}_{m}(\mathcal{X},\mathcal{L})>0 for any nontrivial very ample test configuration (𝒳,)(\mathcal{X},\mathcal{L}) for (X,KX)(X,-K_{X}) of exponent mm.

While the above is a result in algebraic geometry, it seems that no purely algebro-geometric proof is known at the moment of writing this paper; the proof that we give in §5.1 relies on Theorem 1.1 and [rtz, Theorem 2.3] which concern the anticanonically balanced metrics in differential geometry. Note also that, in the above corollary, the triviality of test configurations is defined in a way that is not commonly used recently; see Remark 3.9.

For the definitions of the terminologies that arise in the above results, the reader is referred to §2.2, §3.2, and §A.1.

We also extend these results to the Kähler–Ricci gg-solitons (Theorem 5.6 and Corollary 5.8) and the coupled Kähler–Einstein metrics (Theorem 5.11 and Corollary 5.12). While the main argument of the proof carries over almost word by word for both cases, it turns out that we need to strengthen the stability condition for the coupled Kähler–Einstein case proposed by Hultgren–Witt Nyström [HWN19]. For this purpose, we first define the notion of a test configuration generated by the \mathbb{C}^{*}-actions of multiple test configurations (Definition 3.15). We then define a strengthened version of the coupled Ding invariant (Definition 3.16), and prove that it naturally arises as the asymptotic slope of the coupled Ding functional (Theorem 5.9). It seems natural to expect that this result can be applied to give further results for the coupled Kähler–Einstein metrics, as pointed out at the end of §3.4, but we shall only consider its implications to the anticanonically balanced metrics in this paper. The case of Kähler–Ricci gg-solitons is also treated similarly, but the analogue of Corollary 1.2 is weaker than the full characterisation of δmg>1\delta^{g}_{m}>1 (see Corollary 5.8 and §A.2) because of nontrivial holomorphic vector fields, which is analogous to the well-known phenomenon that varieties with nontrivial holomorphic vector fields are never (uniformly) Ding stable.

Remark 1.3.

We point out how Theorem 1.1 relates to some known results for toric Fano manifolds. Yotsutani [Yotsu2017] proved that the asymptotic Chow semistability implies the Ding polystability for toric Fano manifolds, where we note that there are various results (e.g. [Ono2011, Yotsu2016]) available for the Chow semistability of toric Fano manifolds. Combined with Theorem 1.1, we immediately find that asymptotically Chow semistable toric Fano manifolds admit infinitely many anticanonically balanced metrics, by noting that the Chow weight must be zero for product test configurations under such hypotheses (cf. [st19, Proposition 5.4] or the proof of Lemma 5.5); this is not a new result, however, as it also follows from [BWN14, Theorem 1.7] and [WZ04, Theorem 1.1] since the higher Futaki invariants [Futaki04] vanish by the Chow semistability [st19, Theorem 5.5].

It should also be possible to consider the twisted version of above results as in [rtz, §4.3], but we decide not to treat such cases since Proposition 5.3, which is the key estimate in the proof of Theorem 1.1, seems more complicated with the twist term.

1.3. Organisation of the paper

We start by reviewing the differential-geometric preliminaries in §2. While most of the materials in §2 are well-known, Lemma 2.18 for the coupled Ding functional seems to be a new result that plays an important role later. Algebro-geometric preliminaries are recalled in §3; again this is mostly a review of well-known results, but in §3.4 we define the notion of a test configuration generated by the \mathbb{C}^{*}-actions of multiple test configurations, and define the strengthened version of the coupled Ding invariant (Definition 3.16) which seems to be new. The relationship between these analytic and algebraic concepts are reviewed in §4, which is a summary of many well-established foundational results.

The proof of the main results is given in §5; we note that, almost as a by-product of the proof for the coupled version, we obtain a formula (Theorem 5.9) for the asymptotic slope of the coupled Ding functional along a coupled psh ray that does not seem to have been considered before and seems appropriate for the study of the coupled Kähler–Einstein metrics. While the invariants δ\delta and δm\delta_{m} are important objects that appear in Corollary 1.2, we only review them in Appendix A since the main body of this paper does not quite depend on these invariants.


Acknowledgements. The author thanks Kento Fujita, Tomoyuki Hisamoto, Eiji Inoue, Yuji Odaka, and Ryosuke Takahashi for helpful discussions. This work is partially supported by JSPS KAKENHI (Grant-in-Aid for Early-Career Scientists), Grant Number JP19K14524.

2. Differential-geometric preliminaries

2.1. Kähler–Einstein metrics and Ding functional

Let (X,KX)(X,-K_{X}) be a Fano manifold of complex dimension nn, polarised by the anticanonical bundle. Throughout this paper, we use the additive notation for the tensor product of line bundles and write Vol(X)\mathrm{Vol}(X) for Xc1(KX)n\int_{X}c_{1}(-K_{X})^{n}. We work with the very ample line bundle mKX-mK_{X}, by choosing mm\in\mathbb{N} to be sufficiently large; we shall further assume that mm is sufficiently divisible for the coupled Kähler–Einstein case in §2.4. We fix a reference hermitian metric h0h_{0} on mKX-mK_{X}, rather than KX-K_{X}, and write ω0c1(KX)\omega_{0}\in c_{1}(-K_{X}) for the associated Kähler metric re-scaled by 1/m1/m. This reference metric is assumed to satisfy various extra hypotheses according to the situation under consideration; we further assume h0h_{0} to be a Fubini–Study metric defined in §2.2 for convenience, and moreover to be TT-invariant when we consider the Kähler–Ricci gg-solitons in §2.3. The coupled Kähler–Einstein case §2.4 is slightly more complicated as there will be an auxiliary reference metric (16), but the one in (18) is the relevant one which we denote by h0h_{0}. In any case, without loss of generality we may assume that h0h_{0} satisfies all these required properties and fix the notation once and for all to streamline the notational convention.

We define

:={ϕC(X,)ω0+1¯ϕ/2π>0}\mathcal{H}:=\{\phi\in C^{\infty}(X,\mathbb{R})\mid\omega_{0}+\sqrt{-1}\partial\bar{\partial}\phi/2\pi>0\}

for the space of smooth Kähler metrics in c1(KX)c_{1}(-K_{X}). Note that a hermitian metric emϕh0e^{-m\phi}h_{0} on mKX-mK_{X}, with ϕ\phi\in\mathcal{H}, corresponds to the Kähler metric ωϕ:=ω0+1¯ϕ/2π\omega_{\phi}:=\omega_{0}+\sqrt{-1}\partial\bar{\partial}\phi/2\pi.

An elementary yet important observation is that a hermitian metric on KX-K_{X} naturally defines a volume form on XX; more precisely, a hermitian metric on KX-K_{X} is a smooth positive section of omCX((KX)(KX)¯,)\mathcal{H}om_{C^{\infty}_{X}}((-K_{X})\otimes\overline{(-K_{X})},\mathbb{C}), which corresponds to the one of KXKX¯K_{X}\otimes\overline{K_{X}}, i.e. a volume form. To clarify the notation, we shall write dμϕd\mu_{\phi} for the volume form on XX defined by the hermitian metric emϕh0e^{-m\phi}h_{0} on mKX-mK_{X} with ϕ\phi\in\mathcal{H}; note that we have

dμϕ=eϕdμ0.d\mu_{\phi}=e^{-\phi}d\mu_{0}. (1)

It is convenient to re-scale h0h_{0} by a nonzero constant if necessary so that X𝑑μ0=1\int_{X}d\mu_{0}=1.

Another notational convention that we use is

Xf𝑑μϕ:=(X𝑑μϕ)1Xf𝑑μϕ\fint_{X}fd\mu_{\phi}:=\left(\int_{X}d\mu_{\phi}\right)^{-1}\int_{X}fd\mu_{\phi}

for an integrable function ff on XX, noting that this is just a re-scaling so as to have X𝑑μϕ=1\fint_{X}d\mu_{\phi}=1.

We recall the following definition.

Definition 2.1.

The Ding functional 𝒟:\mathscr{D}:\mathcal{H}\to\mathbb{R} is defined by

𝒟(ϕ):=(ϕ)(ϕ),\mathscr{D}(\phi):=\mathscr{L}(\phi)-\mathscr{E}(\phi),

where ,:\mathscr{L},\mathscr{E}:\mathcal{H}\to\mathbb{R} are defined respectively as

(ϕ):=logX𝑑μϕ,\mathscr{L}(\phi):=-\log\int_{X}d\mu_{\phi},

and

(ϕ):=1(n+1)Vol(X)j=0nXϕω0njωϕj.\mathscr{E}(\phi):=\frac{1}{(n+1)\mathrm{Vol}(X)}\sum_{j=0}^{n}\int_{X}\phi\omega_{0}^{n-j}\wedge\omega^{j}_{\phi}.

A straightforward computation reveals that a critical point of 𝒟\mathscr{D} is precisely the metric that satisfies ωϕn=dμϕ\omega^{n}_{\phi}=d\mu_{\phi}, which is equivalent to the Kähler–Einstein equation

Ric(ωϕ)=ωϕ.\mathrm{Ric}(\omega_{\phi})=\omega_{\phi}.

An important result due to Berndtsson [Ber09a, ber15] is that \mathscr{L} is convex along psh (plurisubharmonic) rays in \mathcal{H}, which we recall later (Theorem 4.2) in the form that is necessary for the proof of our main results. Note that Berndtsson’s convexity implies that the critical point, i.e. the Kähler–Einstein metric, is unique [ber15, Theorems 1.2 and 5.1], giving an alternative proof of the result originally obtained by Bando–Mabuchi [BanMab] that can be generalised to more singular situations.

Note finally that the convention X𝑑μ0=1\int_{X}d\mu_{0}=1 implies (0)=(0)=𝒟(0)=0\mathscr{L}(0)=\mathscr{E}(0)=\mathscr{D}(0)=0.

2.2. Anticanonically balanced metrics

An important subset of \mathcal{H} is the Fubini–Study metrics given in terms of the Kodaira embedding

ι:X(H0(X,mKX))\iota:X\hookrightarrow\mathbb{P}(H^{0}(X,-mK_{X})^{\vee})

defined as follows. Suppose that we have a positive definite hermitian form HH on H0(X,mKX)H^{0}(X,-mK_{X}). This naturally induces a hermitian metric on the hyperplane bundle 𝒪(1)\mathcal{O}(1) over (H0(X,mKX))\mathbb{P}(H^{0}(X,-mK_{X})^{\vee}), and hence on mKX=ι𝒪(1)-mK_{X}=\iota^{*}\mathcal{O}(1) by pullback. By taking the mm-th root, we get a hermitian metric on KX-K_{X} and write FS(H)\mathrm{FS}(H) for the corresponding Kähler potential in \mathcal{H}. Noting that we may write H=eAeAH=e^{A^{*}}e^{A} for some A𝔤𝔩(H0(X,mKX))A\in\mathfrak{gl}(H^{0}(X,-mK_{X})), the set m\mathcal{B}_{m} of positive definite hermitian forms on H0(X,mKX)H^{0}(X,-mK_{X}) can be identified with the right coset space

m=U(Nm)\GL(Nm,),\mathcal{B}_{m}=U(N_{m})\backslash GL(N_{m},\mathbb{C}), (2)

where Nm:=dimH0(X,mKX)N_{m}:=\dim_{\mathbb{C}}H^{0}(X,-mK_{X}). Thus, the construction as above defines the Fubini–Study map [donproj2, §2]

FS:m,\mathrm{FS}:\mathcal{B}_{m}\to\mathcal{H},

which is injective [Lempert2021, Theorem 1.1]. We write m:=FS(m)\mathcal{H}_{m}:=\mathrm{FS}(\mathcal{B}_{m}) for its image, and the elements of m\mathcal{H}_{m} are called the Fubini–Study metrics. We also note that we have the Hilbert map [donnum09, §2.2],

Hilbμ:m,\mathrm{Hilb}_{\mu}:\mathcal{H}\to\mathcal{B}_{m},

defined by associating ϕ\phi\in\mathcal{H} to the (re-scaled) L2L^{2}-inner product NmX𝑑μϕXemϕh0(,)𝑑μϕ\frac{N_{m}}{\int_{X}d\mu_{\phi}}\int_{X}e^{-m\phi}h_{0}(\cdot,\cdot)d\mu_{\phi} on H0(X,mKX)H^{0}(X,-mK_{X}).

We can write down the Fubini–Study metric more explicitly as follows. We choose a reference hermitian form H0mH_{0}\in\mathcal{B}_{m} once and for all, and we define the reference metric h0h_{0} on mKX-mK_{X} to be the Fubini–Study metric defined by H0H_{0}. More precisely, writing h~0\tilde{h}_{0} for the hermitian metric on the hyperplane bundle over (H0(X,mKX))\mathbb{P}(H^{0}(X,-mK_{X})^{\vee}) defined by H0H_{0}, we write

h0:=ιh~0.h_{0}:=\iota^{*}\tilde{h}_{0}. (3)

With this reference metric chosen, the Fubini–Study metric FS(H)m\mathrm{FS}(H)\in\mathcal{H}_{m} defined by HmH\in\mathcal{B}_{m} can be written as

FS(H)=1mlogi=1NmsiHh02,\mathrm{FS}(H)=\frac{1}{m}\log\sum_{i=1}^{N_{m}}\left\|s^{H}_{i}\right\|^{2}_{h_{0}}\in\mathcal{H}, (4)

where {siH}i=1Nm\{s^{H}_{i}\}_{i=1}^{N_{m}} is any HH-orthonormal basis for H0(X,mKX)H^{0}(X,-mK_{X}). Note that FS(H)\mathrm{FS}(H) can be characterised as the unique element in \mathcal{H} such that hH:=exp(mFS(H))h0h_{H}:=\exp(-m\mathrm{FS}(H))h_{0} satisfies i=1NmsiHhH2=1\sum_{i=1}^{N_{m}}\left\|s^{H}_{i}\right\|^{2}_{h_{H}}=1 over XX.

Remark 2.2.

In what follows, {si}i=1Nm\{s_{i}\}_{i=1}^{N_{m}} stands for a fixed H0H_{0}-orthonormal basis for H0(X,mKX)H^{0}(X,-mK_{X}), which we shall refer to as the reference basis, and H0H_{0} will often be regarded as the identity matrix with respect to it. For each A𝔤𝔩(H0(X,mKX))A\in\mathfrak{gl}(H^{0}(X,-mK_{X})), we write AijA_{ij} for the (i,j)(i,j)-th entry of the matrix representing AA with respect to the reference basis, unless otherwise stated. The hermitian conjugate AA^{*} of AA will henceforth be assumed to be with respect to H0H_{0}; we write AA^{\dagger} for the hermitian conjugate with respect to another hermitian form specified in the context.

We observe that m=U(Nm)\GL(Nm,)\mathcal{B}_{m}=U(N_{m})\backslash GL(N_{m},\mathbb{C}) is a Riemannian symmetric space with respect to the natural bi-invariant metric. An important object is a one-parameter family in m\mathcal{H}_{m} defined by the geodesic in m\mathcal{B}_{m} with respect to the bi-invariant metric, defined more explicitly as follows.

Definition 2.3.

The Bergman geodesic ray emanating from H0H_{0} is a family {Ht}t0m\{H_{t}\}_{t\geq 0}\subset\mathcal{B}_{m} of positive definite hermitian forms defined by

Ht:=etAetA,H_{t}:=e^{-tA^{*}}e^{-tA}, (5)

for an H0H_{0}-hermitian endomorphism A𝔤𝔩(H0(X,mKX))A\in\mathfrak{gl}(H^{0}(X,-mK_{X})).

By abuse of terminology, the resulting family {FS(Ht)}t0m\{\mathrm{FS}(H_{t})\}_{t\geq 0}\subset\mathcal{H}_{m} emanating from h0h_{0} is also called the Bergman geodesic ray. The formula (4) immediately yields

FS(Ht)=1mlogi=1NmetAsih02,\mathrm{FS}(H_{t})=\frac{1}{m}\log\sum_{i=1}^{N_{m}}\left\|e^{tA}s_{i}\right\|^{2}_{h_{0}}, (6)

where {si}i=1Nm\{s_{i}\}_{i=1}^{N_{m}} is any H0H_{0}-orthonormal basis for H0(X,mKX)H^{0}(X,-mK_{X}), which we may of course take to be the reference basis, by noting that {etAsi}i=1Nm\{e^{tA}s_{i}\}_{i=1}^{N_{m}} is an HtH_{t}-orthonormal basis.

We recall the following functional defined by Berman–Witt Nyström [BWN14, §4.2.2], which “quantises” the Ding functional.

Definition 2.4.

The quantised Ding functional 𝒟m:m\mathscr{D}_{m}:\mathcal{B}_{m}\to\mathbb{R} is defined as

𝒟m(H):=(FS(H))m(H)\mathscr{D}_{m}(H):=\mathscr{L}(\mathrm{FS}(H))-\mathscr{E}_{m}(H)

where

m(H):=1mNmlogdet(HH01).\mathscr{E}_{m}(H):=-\frac{1}{mN_{m}}\log\det(HH^{-1}_{0}).

The computation of the asymptotic slope of 𝒟m\mathscr{D}_{m} along Bergman geodesic rays will be of crucial importance in what follows, and hence we record an elementary result for the derivative of 𝒟m\mathscr{D}_{m}.

Lemma 2.5.

(cf. [BBGZ, (7.3) and (7.5)]) Suppose that {Ht}t0m\{H_{t}\}_{t\geq 0}\subset\mathcal{B}_{m} is a Bergman geodesic ray (5). Then we have

ddt(FS(Ht))=1mi,j=1Nm(Aij+Aij)Xh0(etAsi,etAsj)l=1NmetAslh02𝑑μFS(Ht),\frac{d}{dt}\mathscr{L}(\mathrm{FS}(H_{t}))=\frac{1}{m}\sum_{i,j=1}^{N_{m}}(A_{ij}+A^{*}_{ij})\fint_{X}\frac{h_{0}(e^{tA}s_{i},e^{tA}s_{j})}{\sum_{l=1}^{N_{m}}\|e^{tA}s_{l}\|^{2}_{h_{0}}}d\mu_{\mathrm{FS}(H_{t})},

and

ddtm(Ht)=tr(A+A)mNm.\frac{d}{dt}\mathscr{E}_{m}(H_{t})=\frac{\mathrm{tr}(A^{*}+A)}{mN_{m}}.
Proof.

The first formula is obvious from

ddt(FS(Ht))=(X𝑑μFS(Ht))1X(ddtFS(Ht))𝑑μFS(Ht)\frac{d}{dt}\mathscr{L}(\mathrm{FS}(H_{t}))=\left(\int_{X}d\mu_{\mathrm{FS}(H_{t})}\right)^{-1}\int_{X}\left(\frac{d}{dt}\mathrm{FS}(H_{t})\right)d\mu_{\mathrm{FS}(H_{t})}

and the equation (6). The second is also immediate from the definition. ∎

Note that the above lemma immediately implies

d2dt2m(Ht)=0.\frac{d^{2}}{dt^{2}}\mathscr{E}_{m}(H_{t})=0. (7)

Recall that the Bergman function ρm(FS(H))C(X,)\rho_{m}(\mathrm{FS}(H))\in C^{\infty}(X,\mathbb{R}) (also called the distortion function in [BBGZ, §7.3]) is defined as follows: writing {σi}i=1Nm\{\sigma_{i}\}_{i=1}^{N_{m}} for an orthonormal basis for H0(X,mKX)H^{0}(X,-mK_{X}) with respect to Vol(X)NmHilbμFS(H)\frac{\mathrm{Vol}(X)}{N_{m}}\mathrm{Hilb}_{\mu}\circ\mathrm{FS}(H), we define

ρm(FS(H)):=i=1NmσihH2,\rho_{m}(\mathrm{FS}(H)):=\sum_{i=1}^{N_{m}}\|\sigma_{i}\|^{2}_{h_{H}}, (8)

where hH:=exp(mFS(H))h0h_{H}:=\exp(-m\mathrm{FS}(H))h_{0}. The following well-known result is contained in [BBGZ, §7.3], and also stated slightly implicitly in [st19, tak19].

Proposition 2.6.

The following are equivalent for HmH\in\mathcal{B}_{m}.

  1. (i)

    There exists an H0H_{0}-hermitian A𝔤𝔩(H0(X,mKX))A\in\mathfrak{gl}(H^{0}(X,-mK_{X})) such that H=eAeAH=e^{-A^{*}}e^{-A} and that for all i,j=1,,Nmi,j=1,\dots,N_{m} we have

    Xh0(eAsi,eAsj)l=1NmeAslh02𝑑μFS(H)1Nmδij=0,\fint_{X}\frac{h_{0}(e^{A}s_{i},e^{A}s_{j})}{\sum_{l=1}^{N_{m}}\|e^{A}s_{l}\|^{2}_{h_{0}}}d\mu_{\mathrm{FS}(H)}-\frac{1}{N_{m}}\delta_{ij}=0,

    where δij\delta_{ij} is the Kronecker delta.

  2. (ii)

    The Bergman function ρm(FS(H))\rho_{m}(\mathrm{FS}(H)) is constant over XX.

  3. (iii)

    HilbμFS(H)=H\mathrm{Hilb}_{\mu}\circ\mathrm{FS}(H)=H.

  4. (iv)

    HmH\in\mathcal{B}_{m} is a critical point of 𝒟m\mathscr{D}_{m}.

Proof.

The proof of (i)(iv)\text{(i)}\Leftrightarrow\text{(iv)} is an obvious consequence of Lemma 2.5, which gives

ddt𝒟m(Ht)=1mi,j=1Nm(Aij+Aij)(Xh0(etAsi,etAsj)l=1NmetAslh02𝑑μFS(Ht)1Nmδij).\frac{d}{dt}\mathscr{D}_{m}(H_{t})=\frac{1}{m}\sum_{i,j=1}^{N_{m}}(A_{ij}+A^{*}_{ij})\left(\fint_{X}\frac{h_{0}(e^{tA}s_{i},e^{tA}s_{j})}{\sum_{l=1}^{N_{m}}\|e^{tA}s_{l}\|^{2}_{h_{0}}}d\mu_{\mathrm{FS}(H_{t})}-\frac{1}{N_{m}}\delta_{ij}\right).

(i)(iii)\text{(i)}\Leftrightarrow\text{(iii)} immediately follows from (4) and the definition of Hilbμ\mathrm{Hilb}_{\mu}, and (ii)(iii)\text{(ii)}\Leftrightarrow\text{(iii)} is obvious from (4) and (8). ∎

Definition 2.7.

A Fubini–Study metric FS(H)m\mathrm{FS}(H)\in\mathcal{H}_{m} is said to be anticanonically balanced at level mm if it satisfies one of the equivalent conditions in Proposition 2.6.

The item (i) of Proposition 2.6 can be regarded as the “zero of the moment map” condition, and 𝒟m\mathscr{D}_{m} can be regarded as the Kempf–Ness type functional. It is thus natural to expect that the existence of the anticanonically balanced metric can be characterised by a Hilbert–Mumford type criterion in Geometric Invariant Theory. The appropriate criterion, as it turns out, is the FF-stability defined by Saito–Takahashi [st19] that we review in Definition 3.8. For more details on the anticanonically balanced metrics, the reader is referred to [BBGZ, §7].

Remark 2.8.

Note that the functionals 𝒟\mathscr{D} and 𝒟m\mathscr{D}_{m} are translation invariant, in the sense that they satisfy 𝒟(ϕ)=𝒟(ϕ+c)\mathscr{D}(\phi)=\mathscr{D}(\phi+c) and 𝒟m(H)=𝒟m(ecH)\mathscr{D}_{m}(H)=\mathscr{D}_{m}(e^{c}H) for any cc\in\mathbb{R}, which follows immediately from their definitions.

We finally note that Berndtsson’s convexity (recalled later in Theorem 4.2) and (7) immediately imply that a critical point of 𝒟m\mathscr{D}_{m} must be the global minimum over m\mathcal{B}_{m}.

2.3. Kähler–Ricci gg-solitons and balanced metrics

We start with the preliminary materials on the automorphism group of Fano manifolds. We write Aut0(X)\mathrm{Aut}_{0}(X) for the identity component of the automorphism group of XX. First observe that Aut0(X)=Aut0(X,KX)\mathrm{Aut}_{0}(X)=\mathrm{Aut}_{0}(X,-K_{X}) is a linear algebraic group, since the action Aut0(X)X\mathrm{Aut}_{0}(X)\curvearrowright X naturally lifts to the anticanonical bundle, which is ample since XX is Fano. This in turn implies that we have the action Aut0(X)(H0(X,mKX))\mathrm{Aut}_{0}(X)\curvearrowright\mathbb{P}(H^{0}(X,-mK_{X})^{\vee}) which preserves the image ι(X)(H0(X,mKX))\iota(X)\subset\mathbb{P}(H^{0}(X,-mK_{X})^{\vee}) by the equivariant embedding theorem [Kambayashi] (see also [CG, §5.1]), for any mm\in\mathbb{N} such that mKX-mK_{X} is very ample. More precisely, for mm large enough there exists a faithful rational representation

θ:Aut0(X)GL(H0(X,mKX)),\theta:\mathrm{Aut}_{0}(X)\hookrightarrow GL(H^{0}(X,-mK_{X})), (9)

which is unique up to the choice of the linearisation (i.e. an overall constant multiple by \mathbb{C}^{*}), such that it is equivariant with respect to ι\iota, i.e. for any xXx\in X and aAut0(X)a\in\mathrm{Aut}_{0}(X) we have

ι(ax)=θ(a)ι(x).\iota(a\cdot x)=\theta(a)\cdot\iota(x).

We observe that an element of GL(H0(X,mKX))GL(H^{0}(X,-mK_{X})) which preserves ι(X)(H0(X,mKX))\iota(X)\subset\mathbb{P}(H^{0}(X,-mK_{X})^{\vee}) must be contained in the image of θ\theta above.

Lemma 2.9.

The group action Aut0(X)m\mathrm{Aut}_{0}(X)\curvearrowright\mathcal{B}_{m} given by the representation (9) as

aH:=θ(a1)Hθ(a1),a\cdot H:=\theta(a^{-1})^{\dagger}H\theta(a^{-1}), (10)

where θ(a1)\theta(a^{-1})^{\dagger} is the hermitian conjugate of θ(a1)\theta(a^{-1}) with respect to HH, is an isometry with respect to the bi-invariant metric on m\mathcal{B}_{m} which is consistent with the natural action Aut0(X)\mathrm{Aut}_{0}(X)\curvearrowright\mathcal{H}.

Proof.

Recalling the isomorphism m=U(Nm)\GL(Nm,)\mathcal{B}_{m}=U(N_{m})\backslash GL(N_{m},\mathbb{C}) in (2) given by H=eBeBH=e^{B^{\dagger}}e^{B}, (10) is exactly the action on m\mathcal{B}_{m} given by the right multiplication eBeBθ(a1)e^{B}\mapsto e^{B}\cdot\theta(a^{-1}) and hence is an isometry with respect to the bi-invariant metric. It is consistent with the natural action Aut0(X)\mathrm{Aut}_{0}(X)\curvearrowright\mathcal{H} since aFS(H)=FS(aH)a^{*}\mathrm{FS}(H)=\mathrm{FS}(a\cdot H) by [yhextremal, Lemma 8].∎

In what follows, we shall mostly regard Aut0(X)\mathrm{Aut}_{0}(X) as a closed subgroup of GL(H0(X,mKX))GL(H^{0}(X,-mK_{X})) by (9), and suppress θ\theta in the notation. Likewise, its Lie algebra 𝔞𝔲𝔱(X)\mathfrak{aut}(X) will be regarded as a Lie subalgebra of 𝔤𝔩(H0(X,mKX))\mathfrak{gl}(H^{0}(X,-mK_{X})). We define a maximal compact subgroup K0:=Aut0(X)U(Nm)K_{0}:=\mathrm{Aut}_{0}(X)\cap U(N_{m}) of Aut0(X)\mathrm{Aut}_{0}(X), where the unitarity is defined with respect to the reference hermitian form H0mH_{0}\in\mathcal{B}_{m}. Note that K0K_{0} is the isometry group of the reference Kähler metric ω0\omega_{0} by Lemma 2.9 and [Lempert2021, Theorem 1.1]. We also define a reductive subgroup Aut0(X)r:=K0\mathrm{Aut}_{0}(X)_{r}:=K_{0}^{\mathbb{C}} of Aut0(X)\mathrm{Aut}_{0}(X) by its complexification, and note that Aut0(X)\mathrm{Aut}_{0}(X) can be written as a semidirect product of its unipotent radical and Aut0(X)r\mathrm{Aut}_{0}(X)_{r} by the Jordan–Chevalley decomposition. We write 𝔞𝔲𝔱(X)r\mathfrak{aut}(X)_{r} for the Lie algebra of Aut0(X)r\mathrm{Aut}_{0}(X)_{r}, and note that if A𝔞𝔲𝔱(X)A\in\mathfrak{aut}(X) is H0H_{0}-hermitian then it must be contained in 𝔞𝔲𝔱(X)r\mathfrak{aut}(X)_{r}.

For any connected subgroup KK of K0K_{0}, we define

mK:={HmH commutes with all elements in K.},\mathcal{B}_{m}^{K}:=\{H\in\mathcal{B}_{m}\mid\text{$H$ commutes with all elements in $K$.}\},

where we identified HmH\in\mathcal{B}_{m} with a hermitian endomorphism with respect to the reference basis. Note that we have uH=Hu\cdot H=H for all uKu\in K and HmKH\in\mathcal{B}_{m}^{K}, with the action uHu\cdot H given by (10), since uH=uHu=u1Huu\cdot H=u^{\dagger}Hu=u^{-1}Hu for all uKu\in K by KU(Nm)K\subset U(N_{m}), noting that uu^{\dagger} agrees with the H0H_{0}-hermitian conjugate uu^{*} since uu commutes with HH. Defining K\mathcal{H}^{K} for the set of Kähler potentials that are invariant under the KK-action and setting mK:=Km\mathcal{H}_{m}^{K}:=\mathcal{H}^{K}\cap\mathcal{H}_{m}, we thus get FS(mK)mK\mathrm{FS}(\mathcal{B}_{m}^{K})\subset\mathcal{H}^{K}_{m} by Lemma 2.9. Just as in (2), we can write mK\mathcal{B}^{K}_{m} as a right coset space

mK=C(K)\C(K)\mathcal{B}_{m}^{K}=C(K)\backslash C(K)^{\mathbb{C}}

where C(K)C(K) is the centraliser of KK in U(Nm)U(N_{m}) and C(K)C(K)^{\mathbb{C}} is its complexification. In particular, mK\mathcal{B}^{K}_{m} is a Riemannian symmetric space with respect to the bi-invariant metric.

We now define the Kähler–Ricci gg-solitons, following [BWN14, §2.2] (see also [rtz, §6.2]). Let TT be a (compact) real torus in Aut0(X)\mathrm{Aut}_{0}(X), and suppose that (X,KX)(X,-K_{X}) admits a Hamiltonian (with respect to ω0\omega_{0}) TT-action which is also holomorphic; note that this necessarily implies that TT is a subgroup of K0K_{0}. We write TT^{\mathbb{C}} for the complexification of TT, Aut0(X,T)\mathrm{Aut}_{0}(X,T^{\mathbb{C}}) for the elements in Aut0(X)\mathrm{Aut}_{0}(X) that commute with TT^{\mathbb{C}}, and 𝔞𝔲𝔱(X,T)\mathfrak{aut}(X,T^{\mathbb{C}}) for its Lie algebra. Just as in Lemma 2.9, we have a natural action of Aut0(X,T)\mathrm{Aut}_{0}(X,T^{\mathbb{C}}) on mT\mathcal{B}_{m}^{T} which is an isometry with respect to the bi-invariant metric and consistent with the natural action on T\mathcal{H}^{T}. As before, we define its reductive subgroup Aut0(X,T)r\mathrm{Aut}_{0}(X,T^{\mathbb{C}})_{r} by the complexification of Aut0(X,T)U(Nm)\mathrm{Aut}_{0}(X,T^{\mathbb{C}})\cap U(N_{m}), whose Lie algebra is denoted by 𝔞𝔲𝔱(X,T)r\mathfrak{aut}(X,T^{\mathbb{C}})_{r}.

The action T(X,KX)T\curvearrowright(X,-K_{X}), with the Kähler form ωϕ\omega_{\phi} (ϕT\phi\in\mathcal{H}^{T}), defines a moment map

mϕ:X𝔱dimT,m_{\phi}:X\to\mathfrak{t}^{\vee}\cong\mathbb{R}^{\dim T},

and its image P:=mϕ(X)P:=m_{\phi}(X) is a compact convex polytope in dimT\mathbb{R}^{\dim T} known as the Delzant polytope. The Duistermaat–Heckman measure ν\nu is the measure on dimT\mathbb{R}^{\dim T}, supported on PP, defined by

ν:=(mϕ)(1Vol(X)ωϕnn!)\nu:=(m_{\phi})_{*}\left(\frac{1}{\mathrm{Vol}(X)}\frac{\omega^{n}_{\phi}}{n!}\right)

which is known to be absolutely continuous and independent of ϕ\phi [DuiHec82].

Definition 2.10.

For a TT-invariant Kähler potential ϕT\phi\in\mathcal{H}^{T} and a smooth function g:P>0g:P\to\mathbb{R}_{>0}, the gg-Monge–Ampère measure MAg(ϕ)\mathrm{MA}_{g}(\phi) is a smooth volume form on XX defined by

MAg(ϕ):=g(mϕ)Vol(X)ωϕnn!.\mathrm{MA}_{g}(\phi):=\frac{g(m_{\phi})}{\mathrm{Vol}(X)}\frac{\omega^{n}_{\phi}}{n!}.

We can define MAg(ϕ)\mathrm{MA}_{g}(\phi) for a more general singular potential ϕ\phi as in [BWN14, §2.2] but we only consider the smooth case. Note that the definition of ν\nu means that we have the volume normalisation

XMAg(ϕ)=Pgν.\int_{X}\mathrm{MA}_{g}(\phi)=\int_{P}g\nu.
Definition 2.11.

Let mϕ:X𝔱m_{\phi}:X\to\mathfrak{t}^{\vee} be a moment map for the torus action with respect to the Kähler form ωϕ\omega_{\phi}, and g:P>0g:P\to\mathbb{R}_{>0} be a smooth function. We say that ϕT\phi\in\mathcal{H}^{T} is a Kähler–Ricci gg-soliton if it satisfies

Ric(ωϕ)=ωϕ+1¯logg(mϕ).\mathrm{Ric}(\omega_{\phi})=\omega_{\phi}+\sqrt{-1}\partial\bar{\partial}\log g(m_{\phi}).

An equivalent way of writing down the above equation is

MAg(ϕ)Pgν=eϕ+fϕdμϕX𝑑μϕ,\frac{\mathrm{MA}_{g}(\phi)}{\int_{P}g\nu}=e^{-\phi+f_{\phi}}\frac{d\mu_{\phi}}{\int_{X}d\mu_{\phi}},

where fϕf_{\phi} is the Ricci potential, which is the unique (TT-invariant) smooth function on XX which satisfies

Ric(ωϕ)=ωϕ+1¯fϕandXefϕ𝑑μϕ=1.\mathrm{Ric}(\omega_{\phi})=\omega_{\phi}+\sqrt{-1}\partial\bar{\partial}f_{\phi}\quad\text{and}\quad\fint_{X}e^{f_{\phi}}d\mu_{\phi}=1.

The above description in terms of the gg-Monge–Ampère measure can be easily extended to allow for more singular solutions, as explained in [BWN14, §2.2].

Remark 2.12.

The Kähler–Ricci gg-soliton reduces to the usual Kähler–Ricci soliton for an appropriate choice of gg (see [BWN14, §2.3], [Hisamotomab, (2.35)]) when we take mϕm_{\phi} to be the Hamiltonian function for the real part of the soliton vector field (i.e. the holomorphy potential of the soliton vector field); recall that the soliton vector field is a holomorphic vector field uniquely determined by the volume minimisation, as proved by Tian–Zhu [TianZhu2002, Lemma 2.2]. We may assume that TT^{\mathbb{C}} contains this soliton vector field, as it provides the most interesting examples, although it is not necessary for the proof of the results in this paper. Note also that the holomorphy potential is a real function as long as the Kähler metrics under consideration are TT-invariant. The Kähler–Ricci gg-solitons also include the Mabuchi soliton, which was defined by Mabuchi [Mab2001] and seems to attract renewed attention after the work of Y. Yao [Yaomab], by taking gg to be as in [Hisamotomab, (2.32)].

Remark 2.13.

Han–Li [HanLi20] (and Hisamoto [Hisamotomab] for the case of Mabuchi solitons) proved that the existence of Kähler–Ricci gg-solitons is equivalent to a version of uniform stability.

Following [BWN14, §2.6], we define a functional 𝒟g:T\mathscr{D}^{g}:\mathcal{H}^{T}\to\mathbb{R} by

𝒟g(ϕ):=(ϕ)g(ϕ),\mathscr{D}^{g}(\phi):=\mathscr{L}(\phi)-\mathscr{E}^{g}(\phi),

where g\mathscr{E}^{g} is defined by its first variation as

δg|ϕ(ψ):=XψMAg(ϕ)=Xψg(mϕ)Vol(X)ωϕnn!.\delta\mathscr{E}^{g}|_{\phi}(\psi):=\int_{X}\psi\mathrm{MA}_{g}(\phi)=\int_{X}\psi\frac{g(m_{\phi})}{\mathrm{Vol}(X)}\frac{\omega^{n}_{\phi}}{n!}. (11)

That g\mathscr{E}^{g} is well-defined is proved by Berman–Witt Nyström [BWN14, Lemma 2.14], where the case of the Kähler–Ricci solitons was originally proved by X. Zhu [Zhu2000, Lemma 3.1]. It then follows that ϕT\phi\in\mathcal{H}^{T} is a critical point of 𝒟g\mathscr{D}^{g} if and only if it is a Kähler–Ricci gg-soliton. See [BWN14, §2.4] for more details.

The above functional can be “quantised”, as proposed by [BWN14, §4]. First we write

Pm:={λHom(T,)λ appears as a weight of TH0(X,mKX)}.P_{m}:=\{\lambda\in\mathrm{Hom}(T^{\mathbb{C}},\mathbb{C}^{*})\mid\lambda\text{ appears as a weight of }T^{\mathbb{C}}\curvearrowright H^{0}(X,-mK_{X})\}.

It is well-known that 1mPmP=mϕ(X)\frac{1}{m}P_{m}\subset P=m_{\phi}(X), as in [Lah19, Lemma 13]. We have the weight decomposition

H0(X,mKX)=λPmRm,λH^{0}(X,-mK_{X})=\bigoplus_{\lambda\in P_{m}}R_{m,\lambda} (12)

where TT^{\mathbb{C}} acts with weight λPm\lambda\in P_{m} on Rm,λR_{m,\lambda}. We also write

Nm,λ:=dimRm,λN_{m,\lambda}:=\dim_{\mathbb{C}}R_{m,\lambda}

and

gm¯:=1NmλPmg(λ/m)Nm,λ.\overline{g_{m}}:=\frac{1}{N_{m}}\sum_{\lambda\in P_{m}}g(\lambda/m)N_{m,\lambda}.
Definition 2.14.

The quantised Ding functional for the Kähler–Ricci gg-soliton is a map 𝒟mg:mT\mathscr{D}^{g}_{m}:\mathcal{B}_{m}^{T}\to\mathbb{R} defined by

𝒟mg(H):=(FS(H))mg(H),\mathscr{D}^{g}_{m}(H):=\mathscr{L}(\mathrm{FS}(H))-\mathscr{E}^{g}_{m}(H),

where mg:mT\mathscr{E}_{m}^{g}:\mathcal{B}_{m}^{T}\to\mathbb{R} is defined by

mg(H):=1mNmgm¯λPmg(λ/m)logdet(HH01|Rm,λ).\mathscr{E}^{g}_{m}(H):=-\frac{1}{mN_{m}\overline{g_{m}}}\sum_{\lambda\in P_{m}}g(\lambda/m)\log\det\left(HH_{0}^{-1}|_{R_{m,\lambda}}\right).

An anticanonically gg-balanced metric at level mm is a critical point of 𝒟mg\mathscr{D}^{g}_{m}.

Note that, for the Bergman geodesic ray {Ht}t0mT\{H_{t}\}_{t\geq 0}\subset\mathcal{B}^{T}_{m} as defined in (5), we have

ddtmg(Ht)=1mNmgm¯λPmg(λ/m)tr((A+A)|Rm,λ).\frac{d}{dt}\mathscr{E}^{g}_{m}(H_{t})=\frac{1}{mN_{m}\overline{g_{m}}}\sum_{\lambda\in P_{m}}g(\lambda/m)\mathrm{tr}\left((A+A^{*})|_{R_{m,\lambda}}\right). (13)

Just as we saw in Proposition 2.6, we can characterise the critical point of 𝒟mg\mathscr{D}^{g}_{m} by the “zero of the moment map” condition. Suppose that we write {sα(λ)}α=1Nm,λ\left\{s_{\alpha}^{(\lambda)}\right\}_{\alpha=1}^{N_{m,\lambda}} for an H0H_{0}-orthonormal basis for each Rm,λR_{m,\lambda}, noting that the weight subspaces are orthogonal to each other with respect to H0H_{0} which is TT-invariant. Then H=eAeAmTH=e^{-A^{*}}e^{-A}\in\mathcal{B}^{T}_{m} is a critical point of 𝒟mg\mathscr{D}^{g}_{m} if and only if

Xh0(eAsα(λ),eAsβ(λ))λPmγ=1Nm,λeAsγ(λ)h02𝑑μFS(H)g(λ/m)mNmgm¯δαβ=0,\fint_{X}\frac{h_{0}\left(e^{A}s^{(\lambda)}_{\alpha},e^{A}s^{(\lambda)}_{\beta}\right)}{\sum_{\lambda^{\prime}\in P_{m}}\sum_{\gamma=1}^{N_{m,\lambda^{\prime}}}\left\|e^{A}s^{(\lambda^{\prime})}_{\gamma}\right\|^{2}_{h_{0}}}d\mu_{\mathrm{FS}(H)}-\frac{g(\lambda/m)}{mN_{m}\overline{g_{m}}}\delta_{\alpha\beta}=0,

for all α,β=1,,Nm,λ\alpha,\beta=1,\dots,N_{m,\lambda} and all λPm\lambda\in P_{m}; note that AA preserves the weight decomposition (12) as HmTH\in\mathcal{B}^{T}_{m}. We can also characterise the above as the metric for which the gg-Bergman function [BWN14, §4.2.1] is constant, by arguing exactly as in Proposition 2.6.

Just as we pointed out for 𝒟m\mathscr{D}_{m} in §2.2, a critical point of 𝒟mg\mathscr{D}^{g}_{m} is necessarily the global minimum over mT\mathcal{B}^{T}_{m}; in fact it attains the global minimum over m\mathcal{B}_{m}, as pointed out in [rtz, Proof of Theorem 2.11].

2.4. Coupled Kähler–Einstein metrics and balanced metrics

Suppose that we have a kk-tuple of ample \mathbb{Q}-line bundles L1,,LkL_{1},\dots,L_{k} over a Fano manifold XX such that

KX=L1++Lk.-K_{X}=L_{1}+\cdots+L_{k}.

We also write (X,KX;L1,,Lk)(X,-K_{X};L_{1},\dots,L_{k}) for the above data.

Coupled Kähler–Einstein metrics were introduced by Hultgren–Witt Nyström [HWN19] as a generalisation of the Kähler–Einstein metrics which is compatible with the above “decomposition” of KX-K_{X}, and have been actively studied recently. We recall here some basic results established in [HWN19], and also its relationship to the geometric quantisation as given by Takahashi [tak19]; in fact we will complement it by adding some new materials that were not discussed in [tak19], which turns out to be important later.

We choose a positive integer mm to be sufficiently large and divisible so that each of mKX-mK_{X}, mL1,,mLkmL_{1},\dots,mL_{k} is a very ample line bundle and that the natural multiplication map

H0(X,mL1)H0(X,mLk)H0(X,mL1++mLk)=H0(X,mKX)H^{0}(X,mL_{1})\otimes\cdots\otimes H^{0}(X,mL_{k})\to H^{0}(X,mL_{1}+\cdots+mL_{k})=H^{0}(X,-mK_{X})

is surjective (see e.g. [LazarsfeldI, Example 1.2.22]). An elementary observation, which follows from taking the dual of the above, is that we have a sequence of embeddings

X(H0(X,mKX))(H0(X,mL1)H0(X,mLk))X\hookrightarrow\mathbb{P}(H^{0}(X,-mK_{X})^{\vee})\hookrightarrow\mathbb{P}(H^{0}(X,mL_{1})^{\vee}\otimes\cdots\otimes H^{0}(X,mL_{k})^{\vee}) (14)

which turns out to be important. We write ιcoupled\iota_{\mathrm{coupled}} for the composition of the two embeddings above. Note that the hyperplane bundle over (H0(X,mL1)H0(X,mLk))\mathbb{P}(H^{0}(X,mL_{1})^{\vee}\otimes\cdots\otimes H^{0}(X,mL_{k})^{\vee}) is pulled back by ιcoupled\iota_{\mathrm{coupled}} to mKX-mK_{X} since the second embedding is linear.

On the other hand, since mL1,,mLkmL_{1},\dots,mL_{k} are very ample, for each i=1,,ki=1,\dots,k we have the Kodaira embedding ιi:X(H0(X,mLi))\iota_{i}:X\hookrightarrow\mathbb{P}(H^{0}(X,mL_{i})^{\vee}). For each ii we fix a reference hermitian form Hi,0H_{i,0} for H0(X,mLi)H^{0}(X,mL_{i}), and define h~i,0\tilde{h}_{i,0} to be the Fubini–Study metric on (H0(X,mLi))\mathbb{P}(H^{0}(X,mL_{i})^{\vee}) with respect to Hi,0H_{i,0}, which is pulled back to a hermitian metric hi,0:=ιih~i,0h_{i,0}:=\iota_{i}^{*}\tilde{h}_{i,0} on mLimL_{i} by ιi\iota_{i}, entirely analogously to (3) in §2.2. We write θic1(Li)\theta_{i}\in c_{1}(L_{i}) for the associated Kähler metric, scaled by 1/m1/m to be in c1(Li)c_{1}(L_{i}). As in (4), for a general positive definite hermitian form HiH_{i} on H0(X,mLi)H^{0}(X,mL_{i}) we define

FSi(Hi):=1mlogj=1Ni,msjHihi,02,\mathrm{FS}_{i}(H_{i}):=\frac{1}{m}\log\sum_{j=1}^{N_{i,m}}\left\|s^{H_{i}}_{j}\right\|^{2}_{h_{i,0}}, (15)

where {sjHi}j=1Ni,m\{s^{H_{i}}_{j}\}_{j=1}^{N_{i,m}} is any HiH_{i}-orthonormal basis for H0(X,mLi)H^{0}(X,mL_{i}) and we wrote

Ni,m:=dimH0(X,mLi).N_{i,m}:=\dim_{\mathbb{C}}H^{0}(X,mL_{i}).

We pick an auxiliary reference hermitian metric on mKX-mK_{X} to be

h0:=h1,0hk,0,h^{\prime}_{0}:=h_{1,0}\otimes\cdots\otimes h_{k,0}, (16)

with the associated volume form dμ0d\mu^{\prime}_{0} which we may assume has unit volume over XX by re-scaling; a slightly subtle point is that, while this is a metric on mKX-mK_{X} that is naturally determined by the reference hermitian metrics on mL1,,mLkmL_{1},\dots,mL_{k}, we need another reference metric h0h_{0} on KX-K_{X}, defined later in (18).

With the reference metrics chosen as above, we define the space of “coupled” Kähler potentials defined as

𝓗:=1××k,\bm{\mathcal{H}}:=\mathcal{H}_{1}\times\cdots\times\mathcal{H}_{k},

where for each i=1,,ki=1,\dots,k we define

i:={ϕC(X,)θi+1¯ϕ/2π>0}.\mathcal{H}_{i}:=\{\phi\in C^{\infty}(X,\mathbb{R})\mid\theta_{i}+\sqrt{-1}\partial\bar{\partial}\phi/2\pi>0\}.
Definition 2.15.

Suppose KX=L1++Lk-K_{X}=L_{1}+\cdots+L_{k}. A kk-tuple of Kähler metrics ω1,,ωk\omega_{1},\dots,\omega_{k}, ωic1(Li)\omega_{i}\in c_{1}(L_{i}) for i=1,,ki=1,\dots,k, is said to be coupled Kähler–Einstein if

Ric(ω1)==Ric(ωk)=i=1kωi.\mathrm{Ric}(\omega_{1})=\cdots=\mathrm{Ric}(\omega_{k})=\sum_{i=1}^{k}\omega_{i}.

Hultgren–Witt Nyström [HWN19, Proposition 2.8] proved (see also Remark 2.17 below) that the above metric is precisely the critical point of the following functional.

Definition 2.16.

The coupled Ding functional is a map 𝒟coupled:𝓗\mathscr{D}^{\mathrm{coupled}}:\bm{\mathcal{H}}\to\mathbb{R} defined by

𝒟coupled(ϕ1,,ϕk):=coupled(ϕ1,,ϕk)i=1k(ϕi),\mathscr{D}^{\mathrm{coupled}}(\phi_{1},\dots,\phi_{k}):=\mathscr{L}^{\mathrm{coupled}}(\phi_{1},\dots,\phi_{k})-\sum_{i=1}^{k}\mathscr{E}(\phi_{i}),

where

coupled(ϕ1,,ϕk):=logXexp(i=1kϕi)𝑑μ0.\mathscr{L}^{\mathrm{coupled}}(\phi_{1},\dots,\phi_{k}):=-\log\int_{X}\exp\left(-\sum_{i=1}^{k}\phi_{i}\right)d\mu^{\prime}_{0}.

Note that Pingali [Pin18] reduced the existence of coupled Kähler–Einstein metrics to a priori C0C^{0}-estimates.

Remark 2.17.

The choice of the reference metric is in fact a slightly subtle point, as it is used to show that the Euler–Lagrange equation for 𝒟coupled\mathscr{D}^{\mathrm{coupled}} is exactly the coupled Kähler–Einstein equation [HWN19, Lemma 2.1 and Proposition 2.8]. In [HWN19, §2.1] the reference hermitian metric on KX-K_{X} is chosen to be the volume form ηn\eta^{n} (normalised to have unit volume over XX) given by a Kähler metric η\eta satisfying

Ric(η)=i=1kθi,\mathrm{Ric}(\eta)=\sum_{i=1}^{k}\theta_{i},

which exists by Yau’s theorem [Yau78]. This is in fact the same as our choice of the reference (16) since ηn=dμ0\eta^{n}=d\mu^{\prime}_{0}; note that the Kähler metric associated to h0h^{\prime}_{0}, re-scaled to be in c1(KX)c_{1}(-K_{X}), satisfies

1m1¯logh0=1m1¯logh1,0hk,0=i=1kθi,-\frac{1}{m}\sqrt{-1}\partial\bar{\partial}\log h^{\prime}_{0}=-\frac{1}{m}\sqrt{-1}\partial\bar{\partial}\log h_{1,0}\otimes\cdots\otimes h_{k,0}=\sum_{i=1}^{k}\theta_{i},

which immediately implies ηn=dμ0\eta^{n}=d\mu^{\prime}_{0} since Ric(η)=i=1kθi\mathrm{Ric}(\eta)=\sum_{i=1}^{k}\theta_{i} and both volume forms have unit volume over XX.

Just as we did in §2.2, for each i=1,,ki=1,\dots,k we define the right coset space

i,m:=U(Ni,m)\GL(Ni,m,),\mathcal{B}_{i,m}:=U(N_{i,m})\backslash GL(N_{i,m},\mathbb{C}),

which is the set of positive definite hermitian forms on H0(X,mLi)H^{0}(X,mL_{i}). We further set

𝓑m:=1,m××k,m,\bm{\mathcal{B}}_{m}:=\mathcal{B}_{1,m}\times\cdots\times\mathcal{B}_{k,m},

which is a Riemannian symmetric space with the bi-invariant metric, and define 𝓗m𝓗\bm{\mathcal{H}}_{m}\subset\bm{\mathcal{H}} by

𝓗m:=FS1(1,m)××FSk(k,m).\bm{\mathcal{H}}_{m}:=\mathrm{FS}_{1}(\mathcal{B}_{1,m})\times\cdots\times\mathrm{FS}_{k}(\mathcal{B}_{k,m}).

So far we mostly considered kk-tuples of Fubini–Study metrics associated to the very ample line bundles mL1,,mLkmL_{1},\dots,mL_{k}. On the other hand, we have the embedding

ιcoupled:X(H0(X,mL1)H0(X,mLk))\iota_{\mathrm{coupled}}:X\hookrightarrow\mathbb{P}(H^{0}(X,mL_{1})^{\vee}\otimes\cdots\otimes H^{0}(X,mL_{k})^{\vee})

given by (14). Recalling that we fixed a reference hermitian form Hi,0i,mH_{i,0}\in\mathcal{B}_{i,m} for each i=1,,ki=1,\dots,k, we can define a reference hermitian form

𝑯0:=H1,0Hk,0\bm{H}_{0}:=H_{1,0}\otimes\cdots\otimes H_{k,0}

on H0(X,mL1)H0(X,mLk)H^{0}(X,mL_{1})\otimes\cdots\otimes H^{0}(X,mL_{k}). Just as in Remark 2.2, we fix a reference basis for H0(X,mL1)H0(X,mLk)H^{0}(X,mL_{1})\otimes\cdots\otimes H^{0}(X,mL_{k}) as an 𝑯0\bm{H}_{0}-orthonormal basis which we write as {𝒔𝒋}𝒋\{\bm{s}_{\bm{j}}\}_{\bm{j}}; the multi-index 𝒋\bm{j} here is such that

𝒋:=(j1,,jk),ji{1,,Ni,m} for each i=1,,k,\bm{j}:=(j_{1},\dots,j_{k}),\quad j_{i}\in\{1,\dots,N_{i,m}\}\text{ for each }i=1,\dots,k,

and

𝒔𝒋:=sj1sjk,\bm{s}_{\bm{j}}:=s_{j_{1}}\otimes\cdots\otimes s_{j_{k}}, (17)

where {sji}ji=1Ni,m\{s_{j_{i}}\}_{j_{i}=1}^{N_{i,m}} is an Hi,0H_{i,0}-orthonormal basis for H0(X,mLi)H^{0}(X,mL_{i}). Recalling also that the hyperplane bundle is pulled back to mKX-mK_{X} by ιcoupled\iota_{\mathrm{coupled}}, we see that 𝑯0\bm{H}_{0} defines a hermitian metric h~0\tilde{h}_{0} on the hyperplane bundle over (H0(X,mL1)H0(X,mLk))\mathbb{P}(H^{0}(X,mL_{1})^{\vee}\otimes\cdots\otimes H^{0}(X,mL_{k})^{\vee}) which defines a hermitian metric

h0:=ιcoupledh~0h_{0}:=\iota^{*}_{\mathrm{coupled}}\tilde{h}_{0} (18)

on mKX-mK_{X}, with the associated volume form dμ0d\mu_{0}. We take this h0h_{0} to be the reference hermitian metric on mKX-mK_{X}. Writing ω0c1(KX)\omega_{0}\in c_{1}(-K_{X}) for the associated Kähler metric (scaled by 1/m1/m), we take ω0\omega_{0} to be the reference Kähler metric for the Kähler class c1(KX)c_{1}(-K_{X}), and write

={ϕC(X,)ω0+1¯ϕ/2π>0},\mathcal{H}=\{\phi\in C^{\infty}(X,\mathbb{R})\mid\omega_{0}+\sqrt{-1}\partial\bar{\partial}\phi/2\pi>0\},

continuing with the notation so far.

For a general element (H1,,Hk)𝓑𝒎(H_{1},\dots,H_{k})\in\bm{\mathcal{B}_{m}}, we define 𝑯:=H1Hk\bm{H}:=H_{1}\otimes\cdots\otimes H_{k} for the positive definite hermitian form on H0(X,mL1)H0(X,mLk)H^{0}(X,mL_{1})\otimes\cdots\otimes H^{0}(X,mL_{k}). We also write {𝒔𝒋𝑯}𝒋\left\{\bm{s}^{\bm{H}}_{\bm{j}}\right\}_{\bm{j}} for the associated orthonormal basis, where

𝒔𝒋𝑯:=sj1H1sjkHk\bm{s}^{\bm{H}}_{\bm{j}}:=s^{H_{1}}_{j_{1}}\otimes\cdots\otimes s^{H_{k}}_{j_{k}}

and {sjiHi}ji=1Ni,m\left\{s^{H_{i}}_{j_{i}}\right\}_{j_{i}=1}^{N_{i,m}} is an HiH_{i}-orthonormal basis for H0(X,mLi)H^{0}(X,mL_{i}). The Fubini–Study metric defined by 𝑯\bm{H} and ιcoupled\iota_{\mathrm{coupled}} is given by

FS(𝑯)=1mlog(i=1kji=1Ni,m𝒔𝒋𝑯h02),\mathrm{FS}(\bm{H})=\frac{1}{m}\log\left(\sum_{i=1}^{k}\sum_{j_{i}=1}^{N_{i,m}}\left\|\bm{s}^{\bm{H}}_{\bm{j}}\right\|^{2}_{h_{0}}\right)\in\mathcal{H}, (19)

just as in the usual case (4).

Thus, we find that a kk-tuple of hermitian forms (H1,,Hk)𝓑𝒎(H_{1},\dots,H_{k})\in\bm{\mathcal{B}_{m}} gives rise to a kk-tuple of Kähler potentials (FS1(H1),,FSk(Hk))𝓗m(\mathrm{FS}_{1}(H_{1}),\dots,\mathrm{FS}_{k}(H_{k}))\in\bm{\mathcal{H}}_{m} and also an additional Kähler potential FS(𝑯)\mathrm{FS}(\bm{H})\in\mathcal{H}. An intriguing relationship satisfied by these metrics is the following lemma, which turns out to be useful later.

Lemma 2.18.

With the notation as above, we have

exp(i=1kFSi(Hi))dμ0=dμFS(𝑯).\exp\left(-{\sum_{i=1}^{k}\mathrm{FS}_{i}(H_{i})}\right)d\mu^{\prime}_{0}=d\mu_{\mathrm{FS}(\bm{H})}.

In particular,

coupled(FS1(H1),,FSk(Hk))=(FS(𝑯)).\mathscr{L}^{\mathrm{coupled}}(\mathrm{FS}_{1}(H_{1}),\dots,\mathrm{FS}_{k}(H_{k}))=\mathscr{L}(\mathrm{FS}(\bm{H})).
Proof.

First note that there exists φC(X,)\varphi\in C^{\infty}(X,\mathbb{R}) such that eφh0=h0e^{-\varphi}h_{0}=h^{\prime}_{0}, or equivalently eφ/mdμ0=dμ0e^{-\varphi/m}d\mu_{0}=d\mu^{\prime}_{0} in terms of the volume form. We then get the claim by the following straightforward computation that follows from the definitions (15) and (19):

exp(i=1kFSi(Hi))dμ0\displaystyle\exp\left(-\sum_{i=1}^{k}\mathrm{FS}_{i}(H_{i})\right)d\mu^{\prime}_{0} =i=1k(j=1Ni,msjiHihi,02)1/mdμ0\displaystyle=\prod_{i=1}^{k}\left(\sum_{j=1}^{N_{i,m}}\left\|s^{H_{i}}_{j_{i}}\right\|^{2}_{h_{i,0}}\right)^{-1/m}d\mu^{\prime}_{0}
=(j1=1N1,mjk=1Nk,msj1HisjkHkh1,0hk,02)1/meφ/mdμ0\displaystyle=\left(\sum_{j_{1}=1}^{N_{1,m}}\cdots\sum_{j_{k}=1}^{N_{k,m}}\left\|s^{H_{i}}_{j_{1}}\otimes\cdots\otimes s^{H_{k}}_{j_{k}}\right\|^{2}_{h_{1,0}\otimes\cdots\otimes h_{k,0}}\right)^{-1/m}e^{-\varphi/m}d\mu_{0}
=(i=1kji=1Ni,m𝒔𝒋𝑯h02)1/mdμ0\displaystyle=\left(\sum_{i=1}^{k}\sum_{j_{i}=1}^{N_{i,m}}\left\|\bm{s}^{\bm{H}}_{\bm{j}}\right\|^{2}_{h_{0}}\right)^{-1/m}d\mu_{0}
=exp(FS(𝑯))dμ0=dμFS(𝑯),\displaystyle=\exp\left(-\mathrm{FS}(\bm{H})\right)d\mu_{0}=d\mu_{\mathrm{FS}(\bm{H})},

where in the third line we used eφh0=eφh1,0hk,0=h0e^{\varphi}h^{\prime}_{0}=e^{\varphi}h_{1,0}\otimes\cdots\otimes h_{k,0}=h_{0} and the last equality is the definitional (1). ∎

We can now define the following “quantised” functional.

Definition 2.19.

The quantised coupled Ding functional 𝒟mcoupled:𝓑m\mathscr{D}^{\mathrm{coupled}}_{m}:\bm{\mathcal{B}}_{m}\to\mathbb{R} is defined for (H1,,Hk)𝓑m(H_{1},\dots,H_{k})\in\bm{\mathcal{B}}_{m} as

𝒟mcoupled(H1,,Hk):=(FS(𝑯))i=1ki,m(Hi),\mathscr{D}^{\mathrm{coupled}}_{m}(H_{1},\dots,H_{k}):=\mathscr{L}(\mathrm{FS}(\bm{H}))-\sum_{i=1}^{k}\mathscr{E}_{i,m}(H_{i}),

where

i,m(Hi):=1mNi,mlogdet(HiHi,01).\mathscr{E}_{i,m}(H_{i}):=-\frac{1}{mN_{i,m}}\log\det(H_{i}\cdot H^{-1}_{i,0}).

A coupled anticanonically balanced metric at level mm is a critical point of 𝒟mcoupled\mathscr{D}^{\mathrm{coupled}}_{m}.

The original definition given by Takahashi [tak19, §3.1.2] was

𝒟mcoupled(H1,,Hk):=coupled(FS1(H1),,FSk(Hk))i=1ki,m(Hi),\mathscr{D}^{\mathrm{coupled}}_{m}(H_{1},\dots,H_{k}):=\mathscr{L}^{\mathrm{coupled}}(\mathrm{FS}_{1}(H_{1}),\dots,\mathrm{FS}_{k}(H_{k}))-\sum_{i=1}^{k}\mathscr{E}_{i,m}(H_{i}),

but Lemma 2.18 ensures that it is exactly the same as the one above. We can also write down the “zero of the moment map” definition for the coupled anticanonically balanced metric, but for the detailed discussions we refer the reader to [tak19, Proof of Proposition 3.4 and also Definition 3.1]; the condition for the Bergman function, analogously to Proposition 2.6, can be established equally easily (cf. [tak19, §4.1]).

Again as pointed out for 𝒟m\mathscr{D}_{m} in §2.2, a critical point of 𝒟mcoupled\mathscr{D}^{\mathrm{coupled}}_{m} is necessarily the global minimum over 𝓑m\bm{\mathcal{B}}_{m}.

3. Algebro-geometric preliminaries

3.1. Test configurations

We recall the basics of test configurations that will be used later.

Definition 3.1.

A very ample test configuration (𝒳,)(\mathcal{X},\mathcal{L}) of exponent mm for a Fano manifold (X,KX)(X,-K_{X}) is a scheme 𝒳\mathcal{X} endowed with a flat projective morphism π:𝒳\pi:\mathcal{X}\to\mathbb{C}, which is \mathbb{C}^{*}-equivariant with respect to the natural \mathbb{C}^{*}-action on \mathbb{C}, with a relatively very ample Cartier divisor \mathcal{L} to which the action 𝒳\mathbb{C}^{*}\curvearrowright\mathcal{X} linearises, such that π1(1)X\pi^{-1}(1)\cong X and |π1(1)mKX\mathcal{L}|_{\pi^{-1}(1)}\cong-mK_{X}. The preimage of 00\in\mathbb{C}, written 𝒳0:=π1(0)\mathcal{X}_{0}:=\pi^{-1}(0), is called the central fibre.

We say that (𝒳,)(\mathcal{X},\mathcal{L}) is product if 𝒳\mathcal{X} is isomorphic to X×X\times\mathbb{C}, and trivial if it is \mathbb{C}^{*}-equivariantly isomorphic to X×X\times\mathbb{C} (i.e. 𝒳X×\mathcal{X}\stackrel{{\scriptstyle\sim}}{{\to}}X\times\mathbb{C} and \mathbb{C}^{*} acts trivially on XX).

Remark 3.2.

It is not common in the literature to assume \mathcal{L} to be relatively very ample, and often important to consider the case when \mathcal{L} is merely semiample on 𝒳\mathcal{X}. In this paper, however, we restrict to very ample test configurations unless otherwise stated, since all test configurations of interest arise as a subscheme of a fixed projective space. The notion of the trivial test configuration as stated above is also rather uncommon recently, but the one above turns out to be the appropriate definition for this paper; see Remark 3.9.

The reason for considering very ample test configurations is that we can write down its “matrix generator”, which is defined as the generator of the one-parameter subgroup associated to the \mathbb{C}^{*}-action; throughout the text, we decree that a one-parameter subgroup is always algebraic, i.e. a morphism between algebraic groups. More precisely, we have the following result by Ross–Thomas [rt07] which is important later in connecting differential-geometric and algebro-geometric arguments.

Proposition 3.3.

(see [rt07, Proposition 3.7] and also [bhj1, §2.3]) Let (𝒳,)(\mathcal{X},\mathcal{L}) be a very ample test configuration of exponent mm. Then there exists a one-parameter subgroup GL(H0(X,mKX))\mathbb{C}^{*}\to GL(H^{0}(X,-mK_{X})), with the generator A𝔤𝔩(H0(X,mKX))A\in\mathfrak{gl}(H^{0}(X,-mK_{X})), such that (𝒳,)(\mathcal{X},\mathcal{L}) can be realised as the Zariski closure of the orbit of ι(X)(H0(X,mKX))\iota(X)\subset\mathbb{P}(H^{0}(X,-mK_{X})^{\vee}) under the \mathbb{C}^{*}-action given by AA; in other words 𝒳\mathcal{X} is isomorphic to

𝒳A:={(τAι(x),τ)xX}τ¯(H0(X,mKX))×\mathcal{X}_{A}:=\overline{\{(\tau^{A}\cdot\iota(x),\tau)\mid x\in X\}_{\tau\in\mathbb{C}^{*}}}\subset\mathbb{P}(H^{0}(X,-mK_{X})^{\vee})\times\mathbb{C} (20)

where the bar denotes the Zariski closure, π:𝒳A\pi:\mathcal{X}_{A}\to\mathbb{C} is the second projection, and the polarisation (denoted by A\mathcal{L}_{A}) is given by the restriction of the hyperplane bundle.

Conversely, the Zariski closure 𝒳A\mathcal{X}_{A} as above, with the polarisation A\mathcal{L}_{A}, defines a very ample test configuration of exponent mm if A𝔤𝔩(H0(X,mKX))A\in\mathfrak{gl}(H^{0}(X,-mK_{X})) is diagonalisable with integral eigenvalues.

Note that the endomorphism A𝔤𝔩(H0(X,mKX))A\in\mathfrak{gl}(H^{0}(X,-mK_{X})) in (20) is exactly the generator of the \mathbb{C}^{*}-action on H0(𝒳0,|𝒳0)H^{0}(\mathcal{X}_{0},\mathcal{L}|_{\mathcal{X}_{0}}), as we can see from the proof of [rt07, Proposition 3.7] (see also [bhj1, Proposition 1.3 and §2.3]). Note also that AA is required to be diagonalisable with integral eigenvalues so that the one-parameter subgroup ττAGL(H0(X,mKX))\mathbb{C}^{*}\ni\tau\mapsto\tau^{A}\in GL(H^{0}(X,-mK_{X})^{\vee}) is a morphism of algebraic groups (rather than complex Lie groups), which is necessary for us to stay in the category of varieties and schemes.

Finally, we note that we can compactify a test configuration to form a family over 1\mathbb{P}^{1} (see also [bhj1, §2.2]).

Definition 3.4.

Let (𝒳,)(\mathcal{X},\mathcal{L}) be a very ample test configuration for a Fano manifold (X,KX)(X,-K_{X}). The compactification 𝒳¯\bar{\mathcal{X}} of 𝒳\mathcal{X} is defined by gluing together 𝒳\mathcal{X} and X×(1{0})X\times(\mathbb{P}^{1}\setminus\{0\}) along their respective open subsets 𝒳𝒳0\mathcal{X}\setminus\mathcal{X}_{0} and X×({0})X\times(\mathbb{C}\setminus\{0\}), identified by the canonical \mathbb{C}^{*}-equivariant isomorphism 𝒳𝒳0X×({0})\mathcal{X}\setminus\mathcal{X}_{0}\stackrel{{\scriptstyle\sim}}{{\to}}X\times(\mathbb{C}\setminus\{0\}). The line bundle ¯\bar{\mathcal{L}} is a natural line bundle over 𝒳¯\bar{\mathcal{X}} constructed from \mathcal{L} and the procedure for the compactification as above.

3.2. Ding invariant and Chow weight

We collect some definitions that are standard in the literature; see [Fujita18, Definition 3.1], [Fujita2019, Definition 2.3], and also [Berman16, §3] for more details on the following.

Definition 3.5.

Let (𝒳,)(\mathcal{X},\mathcal{L}) be a very ample test configuration for (X,KX)(X,-K_{X}) of exponent mm, and ν:𝒳ν𝒳\nu:\mathcal{X}^{\nu}\to\mathcal{X} be its normalisation with ν:=ν\mathcal{L}^{\nu}:=\nu^{*}\mathcal{L}, whose compactification over 1\mathbb{P}^{1} is written as (𝒳¯ν,¯ν)(\bar{\mathcal{X}}^{\nu},\bar{\mathcal{L}}^{\nu}). Let D(𝒳ν,ν)D_{(\mathcal{X}^{\nu},\mathcal{L}^{\nu})} be a \mathbb{Q}-divisor on 𝒳ν\mathcal{X}^{\nu}, whose support is contained in 𝒳0ν:=(νπ)1(0)\mathcal{X}^{\nu}_{0}:=(\nu\circ\pi)^{-1}(0), such that m(K𝒳¯ν/1+D(𝒳ν,ν))-m(K_{\bar{\mathcal{X}}^{\nu}/\mathbb{P}^{1}}+D_{(\mathcal{X}^{\nu},\mathcal{L}^{\nu})}) is a Cartier divisor corresponding to ¯ν\bar{\mathcal{L}}^{\nu}; it is well-known that such a \mathbb{Q}-divisor D(𝒳ν,ν)D_{(\mathcal{X}^{\nu},\mathcal{L}^{\nu})} exists uniquely. The Ding invariant of (𝒳,)(\mathcal{X},\mathcal{L}) is a real number defined by

Ding(𝒳,):=(¯ν)n+1(n+1)mn+1Vol(X)1+lct(𝒳ν,D(𝒳ν,ν);𝒳0ν),\mathrm{Ding}(\mathcal{X},\mathcal{L}):=-\frac{(\bar{\mathcal{L}}^{\nu})^{n+1}}{(n+1)m^{n+1}\mathrm{Vol}(X)}-1+\mathrm{lct}(\mathcal{X}^{\nu},D_{(\mathcal{X}^{\nu},\mathcal{L}^{\nu})};\mathcal{X}^{\nu}_{0}),

where (¯ν)n+1(\bar{\mathcal{L}}^{\nu})^{n+1} stands for the intersection product over 𝒳¯ν\bar{\mathcal{X}}^{\nu}, and lct(𝒳ν,D(𝒳ν,ν);𝒳0ν)\mathrm{lct}(\mathcal{X}^{\nu},D_{(\mathcal{X}^{\nu},\mathcal{L}^{\nu})};\mathcal{X}^{\nu}_{0}) is the log canonical threshold of 𝒳0ν\mathcal{X}^{\nu}_{0} with respect to (𝒳ν,D(𝒳ν,ν))(\mathcal{X}^{\nu},D_{(\mathcal{X}^{\nu},\mathcal{L}^{\nu})}), as in Definition A.1, noting that 𝒳0ν\mathcal{X}_{0}^{\nu} is an effective Cartier divisor on 𝒳ν\mathcal{X}^{\nu} by the flatness of π\pi.

Note that there are several alternative definitions for the Ding invariant, including the version which is given in terms of the infimum over valuations; see [BBJ, bhj1, bhj2] for more details.

Definition 3.6.

Let (𝒳,)(\mathcal{X},\mathcal{L}) be a very ample test configuration for (X,KX)(X,-K_{X}) of exponent mm, which we realise as the Zariski closure of the orbit of ι(X)(H0(X,mKX))\iota(X)\subset\mathbb{P}(H^{0}(X,-mK_{X})^{\vee}) under the \mathbb{C}^{*}-action generated by A𝔤𝔩(H0(X,mKX))A\in\mathfrak{gl}(H^{0}(X,-mK_{X})) by Proposition 3.3. The Chow weight is a rational number defined by

Chowm(𝒳,):=(¯ν)n+1(n+1)mn+1Vol(X)tr(A+A)mh0(X,mKX).\mathrm{Chow}_{m}(\mathcal{X},\mathcal{L}):=\frac{(\bar{\mathcal{L}}^{\nu})^{n+1}}{(n+1)m^{n+1}\mathrm{Vol}(X)}-\frac{\mathrm{tr}(A+A^{*})}{mh^{0}(X,-mK_{X})}.

Ding polystability and Chow polystability can be defined by the nonnegativity of the Ding invariant and the Chow weight respectively, with equality if and only if the test configuration is product. We decide not to elaborate on the details since they will not be used in what follows.

Remark 3.7.

Both these invariants are tightly connected to the Donaldson–Futaki invariant [dontoric, §2.1]. Indeed, Berman [Berman16, §3.1 and Corollary 3.9] proved that the Ding invariant agrees with the Donaldson–Futaki invariant for special test configurations. It is also well-known (see [dontoric, rt07]) that the limit of the Chow weight is the Donaldson–Futaki invariant in the sense that

DF(𝒳,)=limkmkChowmk(𝒳,k),\mathrm{DF}(\mathcal{X},\mathcal{L})=\lim_{k\to\infty}mk\cdot\mathrm{Chow}_{mk}(\mathcal{X},k\mathcal{L}),

where we note that (𝒳,k)(\mathcal{X},k\mathcal{L}) is a very ample test configuration for (X,KX)(X,-K_{X}) of exponent mkmk, if (𝒳,)(\mathcal{X},\mathcal{L}) is of exponent mm.

We recall the following stability condition defined by Saito–Takahashi [st19, Definition 3.4].

Definition 3.8.

A Fano manifold (X,KX)(X,-K_{X}) is said to be FF-semistable at level mm if for any very ample test configuration (𝒳,)(\mathcal{X},\mathcal{L}) of exponent mm for (X,KX)(X,-K_{X}) we have

Ding(𝒳,)+Chowm(𝒳,)0.\mathrm{Ding}(\mathcal{X},\mathcal{L})+\mathrm{Chow}_{m}(\mathcal{X},\mathcal{L})\geq 0.

(X,KX)(X,-K_{X}) is FF-stable if it is FF-semistable and the equality holds if and only if (𝒳,)(\mathcal{X},\mathcal{L}) is trivial, and FF-polystable if it is FF-semistable with equality if and only if (𝒳,)(\mathcal{X},\mathcal{L}) is product.

Remark 3.9.

In the above, we recall that a test configuration (𝒳,)(\mathcal{X},\mathcal{L}) was defined to be trivial if it is \mathbb{C}^{*}-equivariantly isomorphic to (X,KX)×(X,-K_{X})\times\mathbb{C}; note that, in terms of matrix generators, this amounts to saying that (𝒳,)=(𝒳A,A)(\mathcal{X},\mathcal{L})=(\mathcal{X}_{A},\mathcal{L}_{A}) with AA being a constant multiple of the identity matrix.

Recall that this is not exactly the same as requiring JNA(𝒳,)=0J^{\mathrm{NA}}(\mathcal{X},\mathcal{L})=0 (or equivalently the minimum norm being zero), which is another widely used definition for the trivial test configurations, because of the phenomenon first observed by Li–Xu [lx14]; see [bhj1, dertwisted] for more details.

Definition 3.8 is slightly different from the one defined in [st19, Definition 3.4], where the Donaldson–Futaki invariant, instead of the Ding invariant, was used. Note also that the invariant Ding(𝒳,)+Chowm(𝒳,)\mathrm{Ding}(\mathcal{X},\mathcal{L})+\mathrm{Chow}_{m}(\mathcal{X},\mathcal{L}) agrees with the quantised Futaki invariant introduced by Berman–Witt Nyström [BWN14, §4.4] if (𝒳,)(\mathcal{X},\mathcal{L}) is a special test configuration; this is proved in [st19, Lemma 3.2], where we recall [Berman16] that the Ding invariant agrees with the Donaldson–Futaki invariant for the special test configurations (Remark 3.7). Finally, while the Ding invariant is more inherently defined for the non-Archimedean metrics [bhj1, §7.7], and hence its value unchanged under the normalisation, the Chow weight and the Donaldson–Futaki invariant are not. It is well-known, however, that these invariants decrease under the normalisation; see [rt07, Proposition 5.1] for the Chow weight and [bhj1, Proposition 3.15] for the Donaldson–Futaki invariant. This point is also mentioned in [st19, Lemma 3.5].

Remark 3.10.

All the invariants above are translation invariant, in the sense that they remain unchanged when the linearisation of the \mathbb{C}^{*}-action on \mathcal{L} is twisted by the character τc\tau^{c} for some cc\in\mathbb{Z}; note that the twist of the test configuration (𝒳,)(\mathcal{X},\mathcal{L}) by τc\tau^{c} is (𝒳,+c𝒳0)(\mathcal{X},\mathcal{L}+c\mathcal{X}_{0}) [bhj1, page 763]. In the formalism of Proposition 3.3, this amounts to saying that the invariants for (𝒳A,A)(\mathcal{X}_{A},\mathcal{L}_{A}) remain unchanged when we replace AA by A+cidNmA+c\cdot\mathrm{id}_{N_{m}}.

3.3. Kähler–Ricci gg-solitons case

Stability conditions for the Kähler–Ricci gg-solitons must be modified appropriately, as we describe below. Let TT^{\mathbb{C}} be an algebraic torus of automorphisms. We first need the definition below, following [sze2007, Definition 2.1].

Definition 3.11.

Let TT^{\mathbb{C}} be an algebraic torus contained in Aut0(X)\mathrm{Aut}_{0}(X). A very ample test configuration (𝒳,)(\mathcal{X},\mathcal{L}) for a Fano manifold (X,KX)(X,-K_{X}) is said to be TT^{\mathbb{C}}-equivariant if the action Tπ1(1)=(X,KX)T^{\mathbb{C}}\curvearrowright\pi^{-1}(1)=(X,-K_{X}) extends to (𝒳,)(\mathcal{X},\mathcal{L}) in such a way that it commutes with the \mathbb{C}^{*}-action of (𝒳,)(\mathcal{X},\mathcal{L}).

We observe the following, by noting that ι(X)\iota(X) is not contained in any proper linear subspace of (H0(X,mKX))\mathbb{P}(H^{0}(X,-mK_{X})^{\vee}).

Lemma 3.12.

Let (𝒳,)(\mathcal{X},\mathcal{L}) be a very ample test configuration for (X,KX)(X,-K_{X}) of exponent mm which we can write as (𝒳A,A)(\mathcal{X}_{A},\mathcal{L}_{A}) with some A𝔤𝔩(H0(X,mKX))A\in\mathfrak{gl}(H^{0}(X,-mK_{X})) as in Proposition 3.3. Suppose that we identify TAut0(X)T^{\mathbb{C}}\subset\mathrm{Aut}_{0}(X) with a subgroup of GL(H0(X,mKX))GL(H^{0}(X,-mK_{X})) by (9). Then, (𝒳,)(\mathcal{X},\mathcal{L}) is TT^{\mathbb{C}}-equivariant if and only if AA commutes with all elements in TT^{\mathbb{C}}.

The version of FF-stability for the Kähler–Ricci gg-solitons is as follows.

Definition 3.13.

Let (𝒳,)(\mathcal{X},\mathcal{L}) be a TT^{\mathbb{C}}-equivariant very ample test configuration for (X,KX)(X,-K_{X}) of exponent mm, which we realise as the Zariski closure of the \mathbb{C}^{*}-orbit of ι(X)(H0(X,mKX))\iota(X)\subset\mathbb{P}(H^{0}(X,-mK_{X})^{\vee}) generated by A𝔤𝔩(H0(X,mKX))A\in\mathfrak{gl}(H^{0}(X,-mK_{X})) that commutes with TT^{\mathbb{C}} as in the above lemma. We define an invariant

𝒟mg,NA(𝒳,):=1+lct(𝒳ν,D(𝒳ν,ν);𝒳0ν)1mNmgm¯λPmg(λ/m)tr((A+A)|Rm,λ).\mathscr{D}^{g,\mathrm{NA}}_{m}(\mathcal{X},\mathcal{L}):=-1+\mathrm{lct}(\mathcal{X}^{\nu},D_{(\mathcal{X}^{\nu},\mathcal{L}^{\nu})};\mathcal{X}^{\nu}_{0})-\frac{1}{mN_{m}\overline{g_{m}}}\sum_{\lambda\in P_{m}}g(\lambda/m)\mathrm{tr}\left((A+A^{*})|_{R_{m,\lambda}}\right).

We say that (X,KX)(X,-K_{X}) is FF-semistable in the sense of gg-solitons at level mm if for any TT^{\mathbb{C}}-equivariant very ample test configuration (𝒳,)(\mathcal{X},\mathcal{L}) of exponent mm for (X,KX)(X,-K_{X}) we have

𝒟mg,NA(𝒳,)0.\mathscr{D}^{g,\mathrm{NA}}_{m}(\mathcal{X},\mathcal{L})\geq 0.

FF-stability and FF-polystability in the sense of gg-solitons can be defined entirely analogously to Definition 3.8.

Note that the second term of 𝒟mg,NA\mathscr{D}^{g,\mathrm{NA}}_{m} is precisely the slope of mg-\mathscr{E}^{g}_{m} along Bergman geodesic rays by (13).

Remark 3.14.

The following point was communicated to the author by Ryosuke Takahashi. Han–Li [HanLi20, Definition 5.3] defined an invariant g,NA(𝒳,)\mathscr{E}^{g,\mathrm{NA}}(\mathcal{X},\mathcal{L}) for any very ample test configuration (more precisely they defined it for any positive non-Archimedean metric). Using this invariant and defining

Dingg(𝒳,):=1+lct(𝒳ν,D(𝒳ν,ν);𝒳0ν)g,NA(𝒳,),\mathrm{Ding}^{g}(\mathcal{X},\mathcal{L}):=-1+\mathrm{lct}(\mathcal{X}^{\nu},D_{(\mathcal{X}^{\nu},\mathcal{L}^{\nu})};\mathcal{X}^{\nu}_{0})-\mathscr{E}^{g,\mathrm{NA}}(\mathcal{X},\mathcal{L}),

and

Chowmg(𝒳,):=g,NA(𝒳,)1mNmgm¯λPmg(λ/m)tr((A+A)|Rm,λ),\mathrm{Chow}^{g}_{m}(\mathcal{X},\mathcal{L}):=\mathscr{E}^{g,\mathrm{NA}}(\mathcal{X},\mathcal{L})-\frac{1}{mN_{m}\overline{g_{m}}}\sum_{\lambda\in P_{m}}g(\lambda/m)\mathrm{tr}\left((A+A^{*})|_{R_{m,\lambda}}\right),

for a very ample test configuration (𝒳,)(\mathcal{X},\mathcal{L}) of exponent mm, just as we defined the Chow weight in Definition 3.6, we may write

𝒟mg,NA(𝒳,)=Dingg(𝒳,)+Chowmg(𝒳,)\mathscr{D}^{g,\mathrm{NA}}_{m}(\mathcal{X},\mathcal{L})=\mathrm{Ding}^{g}(\mathcal{X},\mathcal{L})+\mathrm{Chow}^{g}_{m}(\mathcal{X},\mathcal{L})

analogously to the strategy in [st19]. This could be a more natural invariant since mg\mathscr{E}^{g}_{m} is related to the “quantisation” of g\mathscr{E}^{g} (defined in (11)) by [BWN14, Proposition 4.5], and g,NA\mathscr{E}^{g,\mathrm{NA}} can be identified with the asymptotic slope of the functional g\mathscr{E}^{g} as

limt+g(FS(Ht))t=g,NA(𝒳A,A)\lim_{t\to+\infty}\frac{\mathscr{E}^{g}(\mathrm{FS(H_{t})})}{t}=\mathscr{E}^{g,\mathrm{NA}}(\mathcal{X}_{A},\mathcal{L}_{A})

for the Bergman geodesic ray, by [HanLi20, Proposition 5.8], analogously to the results in Theorem 4.3 which we review later.

3.4. Coupled Ding invariant

For the coupled Kähler–Einstein metrics, the relevant stability condition was introduced by Hultgren–Witt Nyström [HWN19, Definition 1.14]. It was adapted to the balanced metrics by Takahashi [tak19, Definition 3.10] who defined the coupled version of the FF-polystability. Takahashi’s definition, however, does not seem strong enough to be equivalent to the existence of coupled anticanonically balanced metrics, as suggested by the argument in §5.3. Thus we define a more stringent version of stability, which turns out to involve modifying the “coupled” test configurations defined by Hultgren–Witt Nyström [HWN19, Definition 1.11]; see Remark 3.18 for more details on the comparison with their original definition.

Definition 3.15.

Let (𝒳i,i)(\mathcal{X}_{i},\mathcal{L}_{i}) be a very ample test configuration for (X,Li)(X,L_{i}) of exponent mm, for i=1,,ki=1,\dots,k, defined as the Zariski closure of ιi(X)(H0(X,mLi))\iota_{i}(X)\subset\mathbb{P}(H^{0}(X,mL_{i})^{\vee}) under the one-parameter subgroup τAi\tau^{A_{i}} generated by Ai𝔤𝔩(H0(X,mLi))A_{i}\in\mathfrak{gl}(H^{0}(X,mL_{i})), as in Proposition 3.3. Recall also that we have the embedding

ιcoupled:X(H0(X,mL1)H0(X,mLk))=:\iota_{\mathrm{coupled}}:X\hookrightarrow\mathbb{P}(H^{0}(X,mL_{1})^{\vee}\otimes\cdots\otimes H^{0}(X,mL_{k})^{\vee})=:\mathbb{P}

by (14). We say that a very ample test configuration (𝒴,𝒴)(\mathcal{Y},\mathcal{L}_{\mathcal{Y}}) is generated by the \mathbb{C}^{*}-actions of (𝒳i,i)i=1k(\mathcal{X}_{i},\mathcal{L}_{i})_{i=1}^{k}, if 𝒴\mathcal{Y} is defined as the Zariski closure of ιcoupled(X)\iota_{\mathrm{coupled}}(X) in \mathbb{P}, with the reduced scheme structure, under the natural (dual) tensor product action of the one-parameter subgroup

τA1τAk\tau^{A_{1}}\otimes\cdots\otimes\tau^{A_{k}}

on H0(X,mL1)H0(X,mLk)H^{0}(X,mL_{1})^{\vee}\otimes\cdots\otimes H^{0}(X,mL_{k})^{\vee}, and 𝒴:=𝒪(1)|𝒴\mathcal{L}_{\mathcal{Y}}:=\mathcal{O}_{\mathbb{P}}(1)|_{\mathcal{Y}}. Note that (𝒴,𝒴)(\mathcal{Y},\mathcal{L}_{\mathcal{Y}}) is a very ample test configuration for (X,KX)(X,-K_{X}) of exponent mm since mKX=ιcoupled𝒪(1)-mK_{X}=\iota^{*}_{\mathrm{coupled}}\mathcal{O}_{\mathbb{P}}(1).

The author believes that there should be a better way of formulating the above test configuration (𝒴,𝒴)(\mathcal{Y},\mathcal{L}_{\mathcal{Y}}), hopefully in terms of non-Archimedean metrics. For example, it would be good if we can define what it means for a non-Archimedean metric ϕ𝒴NA\phi^{\mathrm{NA}}_{\mathcal{Y}} on KX-K_{X} to be generated by the \mathbb{C}^{*}-actions of non-Archimedean metrics ϕ1NA,,ϕkNA\phi^{\mathrm{NA}}_{1},\dots,\phi^{\mathrm{NA}}_{k} on L1,,LkL_{1},\dots,L_{k}. This does not seem obvious at all from the above definition, since a priori 𝒴\mathcal{Y} depends on a particular choice of very ample test configurations and even their exponents: 𝒴\mathcal{Y} may change when we change (𝒳i,i)(\mathcal{X}_{i},\mathcal{L}_{i}) to (𝒳i,mi)(\mathcal{X}_{i},m^{\prime}\mathcal{L}_{i}) for mm^{\prime}\in\mathbb{N} and for each i=1,,ki=1,\dots,k. Moreover, (𝒳1,1),,(𝒳k,k)(\mathcal{X}_{1},\mathcal{L}_{1}),\dots,(\mathcal{X}_{k},\mathcal{L}_{k}) all being product does not seem to necessarily imply that (𝒴,𝒴)(\mathcal{Y},\mathcal{L}_{\mathcal{Y}}) is product unless these test configurations are all generated by the same holomorphic vector field, while it is easy to see that (𝒴,𝒴)(\mathcal{Y},\mathcal{L}_{\mathcal{Y}}) is trivial if and only if (𝒳1,1),,(𝒳k,k)(\mathcal{X}_{1},\mathcal{L}_{1}),\dots,(\mathcal{X}_{k},\mathcal{L}_{k}) are all trivial (in the sense of Remark 3.9). Still, Definition 3.15 suffices for the purposes in this paper in which all test configurations arise as a subscheme of a fixed projective space.

Definition 3.16.

Suppose that (𝒳1,1),,(𝒳k,k)(\mathcal{X}_{1},\mathcal{L}_{1}),\dots,(\mathcal{X}_{k},\mathcal{L}_{k}) is a kk-tuple of very ample test configurations, respectively for (X,L1),,(X,Lk)(X,L_{1}),\dots,(X,L_{k}), each of exponent mm. Let (𝒴,𝒴)(\mathcal{Y},\mathcal{L}_{\mathcal{Y}}) be a very ample test configuration generated by the \mathbb{C}^{*}-actions of (𝒳i,i)i=1k(\mathcal{X}_{i},\mathcal{L}_{i})_{i=1}^{k}. Let ν:𝒴ν𝒴\nu:\mathcal{Y}^{\nu}\to\mathcal{Y} be the normalisation of 𝒴\mathcal{Y} and 𝒴ν:=ν𝒴\mathcal{L}_{\mathcal{Y}^{\nu}}:=\nu^{*}\mathcal{L}_{\mathcal{Y}}. We define the coupled Ding invariant by

Ding((𝒳i,i)i=1k):=i=1k(¯i)n+1(n+1)mn+1Xc1(Li)n1+lct(𝒴ν,D𝒴ν,𝒴ν;𝒴0ν),\mathrm{Ding}\left((\mathcal{X}_{i},\mathcal{L}_{i})_{i=1}^{k}\right):=-\sum_{i=1}^{k}\frac{(\bar{\mathcal{L}}_{i})^{n+1}}{(n+1)m^{n+1}\int_{X}c_{1}(L_{i})^{n}}-1+\mathrm{lct}(\mathcal{Y}^{\nu},D_{\mathcal{Y}^{\nu},\mathcal{L}_{\mathcal{Y}^{\nu}}};\mathcal{Y}^{\nu}_{0}),

where lct(𝒴ν,D(𝒴ν,𝒴ν);𝒴0ν)\mathrm{lct}(\mathcal{Y}^{\nu},D_{(\mathcal{Y}^{\nu},\mathcal{L}_{\mathcal{Y}^{\nu}})};\mathcal{Y}^{\nu}_{0}) is the log canonical threshold of 𝒴0ν\mathcal{Y}^{\nu}_{0} with respect to (𝒴ν,D(𝒴ν,𝒴ν))(\mathcal{Y}^{\nu},D_{(\mathcal{Y}^{\nu},\mathcal{L}_{\mathcal{Y}^{\nu}})}), which is defined exactly as in Definitions 3.5 and A.1.

It turns out that this invariant arises as the asymptotic slope of the coupled Ding functional (Theorem 5.9). It is natural to define a stability condition as follows.

Definition 3.17.

A Fano manifold (X,KX)(X,-K_{X}) with the decomposition KX=L1++Lk-K_{X}=L_{1}+\cdots+L_{k} is said to be coupled Ding stable if for any mm\in\mathbb{N} and any kk-tuple of very ample test configurations (𝒳1,1),,(𝒳k,k)(\mathcal{X}_{1},\mathcal{L}_{1}),\dots,(\mathcal{X}_{k},\mathcal{L}_{k}) respectively for (X,L1),,(X,Lk)(X,L_{1}),\dots,(X,L_{k}), each of exponent mm, we have

Ding((𝒳i,i)i=1k)0\mathrm{Ding}\left((\mathcal{X}_{i},\mathcal{L}_{i})_{i=1}^{k}\right)\geq 0

with equality if and only if (𝒳i,i)(\mathcal{X}_{i},\mathcal{L}_{i}) is trivial for each i=1,,ki=1,\dots,k.

The appropriate condition for the equality case could be JNA(𝒳i,i)=0J^{\mathrm{NA}}(\mathcal{X}_{i},\mathcal{L}_{i})=0 for all i=1,,ki=1,\dots,k (cf. Remark 3.9), or may even be JNA(𝒴,𝒴)=0J^{\mathrm{NA}}(\mathcal{Y},\mathcal{L}_{\mathcal{Y}})=0, but we decide not to discuss this stability condition any further as it will not be used in the rest of this paper. The one that we use in this paper is its FF-stability version, which will be stated later in Theorem 5.11 and Corollary 5.12.

Remark 3.18.

In the original definition [HWN19, Definition 1.10] by Hultgren–Witt Nyström (and also its FF-stability version of Takahashi [tak19, Definition 3.10]), they further assume 𝒴=𝒳1==𝒳k\mathcal{Y}=\mathcal{X}_{1}=\cdots=\mathcal{X}_{k} and 𝒴=i=1ki\mathcal{L}_{\mathcal{Y}}=\sum_{i=1}^{k}\mathcal{L}_{i} as a definition of the coupled test configurations, which may be too stringent (hence leading to a weaker stability condition) as suggested by the computation for the balanced metrics that we provide in §5.3. They also assume that 𝒴=𝒳1==𝒳k\mathcal{Y}=\mathcal{X}_{1}=\cdots=\mathcal{X}_{k} is a normal \mathbb{Q}-Gorenstein variety and consider the KK-stability as opposed to the Ding stability, but this seems to be a minor difference. While the author believes that the stability condition in Definition 3.17, or its appropriately modified version, is a useful one in studying the coupled Kähler–Einstein metrics, it is important to note that none of the results in [DelHul, FutZha2019, FutZha2020, Hultgren, HWN19, Nakamura, tak19] seems to be affected when we adopt Definition 3.17 as the relevant notion of stability. The works [DelHul, FutZha2019, FutZha2020, Hultgren, Nakamura], from the point of view of stability, essentially consider the case when (𝒳1,1),,(𝒳k,k)(\mathcal{X}_{1},\mathcal{L}_{1}),\dots,(\mathcal{X}_{k},\mathcal{L}_{k}) are all product and generated by the same holomorphic vector field, and hence we have 𝒴=𝒳1==𝒳k=X×\mathcal{Y}=\mathcal{X}_{1}=\cdots=\mathcal{X}_{k}=X\times\mathbb{C}; [DelHul, Hultgren] also consider multiple holomorphic vector fields but in this case the equation in question seems different. Definition 3.17 does not affect the result in [HWN19] (resp. [tak19]) which proves that the existence of coupled Kähler–Einstein metrics implies their weaker version of coupled KK-stability (resp. its FF-stability version [tak19, §3.2]). It also seems interesting to compare the above definition with the version for the constant scalar curvature Kähler metrics in [DatPin].

Given the above definition for the coupled Ding invariant, it is natural to consider the uniform coupled Ding stability, following the definition of [bhj1, dertwisted] (and also [szethesis]), and the GG-uniform coupled Ding stability when the automorphism group is nondiscrete [Hisamotosl, Li19G]. Moreover, given recent results for the Kähler–Einstein metrics and the Kähler–Ricci gg-solitons [BBJ, HanLi20, Hisamotomab, Li19G], it seems natural to expect that the GG-uniform coupled Ding stability is equivalent to the existence of coupled Kähler–Einstein metrics. We decide not to discuss these problems any further in this paper; related results may appear elsewhere.

4. Psh rays, convexity, and the slope formulae

We now recall several foundational results that connect the asymptotic slope of the functionals 𝒟,,\mathscr{D},\mathscr{L},\mathscr{E} to the algebro-geometric invariants defined above. We start by recalling the following facts that are surely well-known to the experts.

Lemma 4.1.

The following hold for the Bergman geodesic ray (6).

  1. (i)

    The Bergman geodesic ray is a psh (plurisubharmonic) ray of linear growth.

  2. (ii)

    Suppose that A𝔤𝔩(H0(X,mKX))A\in\mathfrak{gl}(H^{0}(X,-mK_{X})) in (6) is H0H_{0}-hermitian with integral eigenvalues. Writing λmin\lambda_{\mathrm{min}}\in\mathbb{Z} for the minimum eigenvalue of AA, the Bergman geodesic ray defined by A:=AλminidNmA^{\prime}:=A-\lambda_{\mathrm{min}}\cdot\mathrm{id}_{N_{m}} admits ϕANA\phi_{A^{\prime}}^{\mathrm{NA}} as non-Archimedean limit, where ϕANANA\phi_{A^{\prime}}^{\mathrm{NA}}\in\mathcal{H}^{\mathrm{NA}} is the positive non-Archimedean metric on KX-K_{X} represented by the very ample test configuration (𝒳A,A)(\mathcal{X}_{A^{\prime}},\mathcal{L}_{A^{\prime}}) of exponent mm.

In the above, a psh ray is as defined in [BBJ, Definition 1.3], which is also called a “subgeodesic” in [bhj2, §3.1] and many other papers. Recall also that a positive non-Archimedean metric on KX-K_{X} is a certain equivalence class of semiample test configurations for (X,KX)(X,-K_{X}) (see [bhj1, Definition 6.2]), and the notions of psh rays and non-Archimedean limits are as defined in [bhj2, Definition 3.1]. We refer the reader to [bhj1, bhj2] for more details on non-Archimedean metrics and non-Archimedean limits, and do not give a detailed review of them in this paper as it is rather technical. The only remark here is that the re-scaling AAλmaxidNmA\mapsto A-\lambda_{\mathrm{max}}\cdot\mathrm{id}_{N_{m}} in (ii) does not change the total space of the test configuration and only results in a difference in linearising the \mathbb{C}^{*}-action on the polarisation, as we saw in Remark 3.10.

Proof.

Ascertaining both claims is an easy exercise in checking the definitions given in [BBJ, Definition 1.3 and (4.1)] for the first item, and [bhj2, §3.1] for the second. Writing p1:X×Xp_{1}:X\times\mathbb{C}^{*}\to X for the natural projection, the required semipositivity for the first item is a consequence of the well-known inequality (p1ω0+XׯX×FS(Ht))n+10(p_{1}^{*}\omega_{0}+\partial_{X\times\mathbb{C}^{*}}\bar{\partial}_{X\times\mathbb{C}^{*}}\mathrm{FS}(H_{t}))^{n+1}\geq 0 which follows from e.g. [donsymm, Proposition 3] and [donproj2, page 351]. To see that it is of linear growth, we only need to note, from (6),

limt+1tsupXFS(Ht)=limt+1mlog(supXi=1NmetAsih02)1/tλmaxm\lim_{t\to+\infty}\frac{1}{t}\sup_{X}\mathrm{FS}(H_{t})=\lim_{t\to+\infty}\frac{1}{m}\log\left(\sup_{X}\sum_{i=1}^{N_{m}}\left\|e^{tA}s_{i}\right\|^{2}_{h_{0}}\right)^{1/t}\leq\frac{\lambda_{\mathrm{max}}}{m}

where λmax\lambda_{\mathrm{max}} is the largest modulus of the eigenvalues of AA.

For the second item, by replacing the reference basis {si}i=1Nm\{s_{i}\}_{i=1}^{N_{m}} by an H0H_{0}-unitarily equivalent basis if necessary, we may diagonalise A=diag(λ1,,λNm)A^{\prime}=\mathrm{diag}(\lambda^{\prime}_{1},\dots,\lambda^{\prime}_{N_{m}}), λ1λNm0\lambda^{\prime}_{1}\geq\cdots\geq\lambda^{\prime}_{N_{m}}\geq 0, and hence, for Ht:=et(A)etAH^{\prime}_{t}:=e^{-t(A^{\prime})^{*}}e^{-tA^{\prime}}, we have

FS(Ht)=1mlog(i=1Nmeλitsih02).\mathrm{FS}(H^{\prime}_{t})=\frac{1}{m}\log\left(\sum_{i=1}^{N_{m}}\left\|e^{\lambda^{\prime}_{i}t}s_{i}\right\|^{2}_{h_{0}}\right).

Then the metric exp(mFS(Ht))h0m\exp(-m\mathrm{FS}(H^{\prime}_{t}))h_{0}^{m} indeed extends to a smooth metric on A\mathcal{L}_{A^{\prime}} over \mathbb{C} as h~0|𝒳A\tilde{h}_{0}|_{\mathcal{X}_{A^{\prime}}}, where we recall our notational convention (3); note that this metric may be degenerate on the central fibre of 𝒳A\mathcal{X}_{A^{\prime}}. Thus we get the statement [bhj2, Definition 3.1] that we needed to prove.∎

We recall the convexity result due to Berndtsson [ber15] which plays a very important role in the proof, which we can apply to the Bergman geodesic rays by Lemma 4.1 (see also [BBGZ, Lemmas 6.5 and 7.2]).

Theorem 4.2.

(Berndtsson [ber15, Theorems 1.1 and 1.2]) \mathscr{L} is convex along the Bergman geodesic rays. More precisely, we have

d2dt2(FS(Ht))0\frac{d^{2}}{dt^{2}}\mathscr{L}(\mathrm{FS}(H_{t}))\geq 0

for the Bergman geodesic ray (6), and the equality holds only if there exists a holomorphic vector field on XX with the flow ψt\psi_{t} such that

ψt(¯FS(Ht))=¯FS(H0).\psi^{*}_{t}(\partial\bar{\partial}\mathrm{FS}(H_{t}))=\partial\bar{\partial}\mathrm{FS}(H_{0}).

We further recall the following foundational theorem, which is a collection of the results proved in [BBJ, Theorem 5.4], [Berman16, Proposition 3.8], [bhj1, Proposition 7.29], and [bhj2, Theorems 3.6 and 3.7], summarised for our purposes in this paper.

Theorem 4.3.

(see [Berman16, BBJ, bhj1, bhj2]) Suppose that an H0H_{0}-hermitian matrix A𝔤𝔩(H0(X,mKX))A\in\mathfrak{gl}(H^{0}(X,-mK_{X})) has rational eigenvalues, and choose cc\in\mathbb{N} so that cAcA has integral eigenvalues. Writing Ht:=etAetAH_{t}:=e^{-tA^{*}}e^{tA} for the Bergman geodesic ray and (𝒳cA,cA)(\mathcal{X}_{cA},\mathcal{L}_{cA}) for the test configuration associated to cAcA as in Proposition 3.3, we have

limt+ddt(FS(Ht))=limt+(FS(Ht))t=1c(1+lct(𝒳cAν,D(𝒳cAν,cAν);𝒳cA,0ν)),\lim_{t\to+\infty}\frac{d}{dt}\mathscr{L}(\mathrm{FS}(H_{t}))=\lim_{t\to+\infty}\frac{\mathscr{L}(\mathrm{FS}(H_{t}))}{t}=\frac{1}{c}\left(-1+\mathrm{lct}(\mathcal{X}_{cA}^{\nu},D_{(\mathcal{X}_{cA}^{\nu},\mathcal{L}_{cA}^{\nu})};\mathcal{X}_{cA,0}^{\nu})\right),

and

limt+ddt(FS(Ht))=limt+(FS(Ht))t=1c(¯cAν)n+1(n+1)mn+1Vol(X).\lim_{t\to+\infty}\frac{d}{dt}\mathscr{E}(\mathrm{FS}(H_{t}))=\lim_{t\to+\infty}\frac{\mathscr{E}(\mathrm{FS}(H_{t}))}{t}=\frac{1}{c}\frac{(\bar{\mathcal{L}}_{cA}^{\nu})^{n+1}}{(n+1)m^{n+1}\mathrm{Vol}(X)}.

In particular,

limt+ddt𝒟(FS(Ht))=limt+𝒟(FS(Ht))t=1cDing(𝒳cA,cA).\lim_{t\to+\infty}\frac{d}{dt}\mathscr{D}(\mathrm{FS}(H_{t}))=\lim_{t\to+\infty}\frac{\mathscr{D}(\mathrm{FS}(H_{t}))}{t}=\frac{1}{c}\mathrm{Ding}(\mathcal{X}_{cA},\mathcal{L}_{cA}).

The equality limt+ddt(FS(Ht))=limt+(FS(Ht))/t\lim_{t\to+\infty}\frac{d}{dt}\mathscr{L}(\mathrm{FS}(H_{t}))=\lim_{t\to+\infty}\mathscr{L}(\mathrm{FS}(H_{t}))/t is a direct consequence of Berndtsson’s convexity, and the corresponding statement for \mathscr{E} follows from the well-known fact that \mathscr{E} is convex along the Bergman geodesic rays (see [donproj2, Proposition 1]). We also observe

limt+(FS(Ht))t=limct+1c(FS(Hct))t=1c(1+lct(𝒳cAν,D(𝒳cAν,cAν);𝒳cA,0ν))\lim_{t\to+\infty}\frac{\mathscr{L}(\mathrm{FS}(H_{t}))}{t}=\lim_{ct\to+\infty}\frac{1}{c}\frac{\mathscr{L}(\mathrm{FS}(H_{ct}))}{t}=\frac{1}{c}\left(-1+\mathrm{lct}(\mathcal{X}_{cA}^{\nu},D_{(\mathcal{X}_{cA}^{\nu},\mathcal{L}_{cA}^{\nu})};\mathcal{X}_{cA,0}^{\nu})\right)

for cc\in\mathbb{N}, and similarly for \mathscr{E} and 𝒟\mathscr{D}. To get the same result, we can also consider the pullback by the base change ττc\mathbb{C}\ni\tau\mapsto\tau^{c}\in\mathbb{C} (cc\in\mathbb{N}) as in [bhj1, §6.3], combined with [bhj2, Proposition 3.5 and Theorem 3.7]. Note also that \mathscr{L} and \mathscr{E} are not translation invariant, but their difference 𝒟\mathscr{D} (and also 𝒟m\mathscr{D}_{m}) is (cf. Remark 2.8).

Remark 4.4.

For an H0H_{0}-hermitian A𝔤𝔩(H0(X,mKX))A\in\mathfrak{gl}(H^{0}(X,-mK_{X})), not necessarily of rational eigenvalues, we can show that the asymptotic slopes of the functionals ,\mathscr{L},\mathscr{E} are well-defined real numbers since the Bergman geodesic ray is a psh ray of linear growth by Lemma 4.1; see [BBJ, Theorem 5.4] for the proof for \mathscr{L}, and e.g. [BBJ, Proposition 4.1], [BDL17, Proposition 4.1] for \mathscr{E}. In particular, the asymptotic slopes for 𝒟,𝒟m\mathscr{D},\mathscr{D}_{m} are well-defined real numbers even when A𝔤𝔩(H0(X,mKX))A\in\mathfrak{gl}(H^{0}(X,-mK_{X})) does not have rational eigenvalues.

5. Proof of the main results

5.1. Proof of Theorem 1.1 and Corollary 1.2

We pick a basis for H0(X,mKX)H^{0}(X,-mK_{X})^{\vee} that is dual to the reference basis {si}i=1Nm\{s_{i}\}_{i=1}^{N_{m}} in Remark 2.2, and identify (H0(X,mKX))\mathbb{P}(H^{0}(X,-mK_{X})^{\vee}) with Nm1\mathbb{P}^{N_{m}-1} in what follows. We write [Z1::ZNm][Z_{1}:\cdots:Z_{N_{m}}] for the homogeneous coordinates for Nm1\mathbb{P}^{N_{m}-1}. The Fano manifold XX is considered to be embedded in Nm1\mathbb{P}^{N_{m}-1} by ι\iota, and we observe that the embedded variety ι(X)\iota(X) is not contained in any proper linear subspace of Nm1\mathbb{P}^{N_{m}-1}.

Recall first that for any A𝔤𝔩(H0(X,mKX))A\in\mathfrak{gl}(H^{0}(X,-mK_{X})), which will be assumed to be H0H_{0}-hermitian in what follows, and the embedded variety ι(X)Nm1\iota(X)\subset\mathbb{P}^{N_{m}-1}, the flat limit (or the limit in the Hilbert scheme)

𝒳0:=limt+etAι(X)Nm1\mathcal{X}_{0}:=\lim_{t\to+\infty}e^{-tA}\cdot\iota(X)\subset\mathbb{P}^{N_{m}-1} (21)

is a well-defined projective scheme in Nm1\mathbb{P}^{N_{m}-1} defined by, in a down-to-earth terminology, the homogeneous ideal generated by the limit of the defining homogeneous polynomials for ι(X)Nm1\iota(X)\subset\mathbb{P}^{N_{m}-1} (see e.g. [szebook, Definition 6.3]). We write 𝒳0,red\mathcal{X}_{0,\mathrm{red}} for the reduced part of 𝒳0\mathcal{X}_{0}, which, again in a down-to-earth terminology, is equal to 𝒳0\mathcal{X}_{0} as an algebraic set in Nm1\mathbb{P}^{N_{m}-1} but has the reduced scheme structure (note that 𝒳0\mathcal{X}_{0} can have multiple components). We also write 𝒳0,redreg\mathcal{X}^{\mathrm{reg}}_{0,\mathrm{red}} for the regular locus of the reduced part of 𝒳0\mathcal{X}_{0}.

Remark 5.1.

The family π:𝒳A\pi:\mathcal{X}_{A}\to\mathbb{C} defined as in (20) above is not a test configuration if AA has non-integral eigenvalues, but is still flat (see [szebook, §6.2] for more details). In fact we do not need the flatness later, and we only need dim𝒳0,redreg=n\dim_{\mathbb{C}}\mathcal{X}^{\mathrm{reg}}_{0,\mathrm{red}}=n for the later argument. This result is also a consequence of [hk18, proof of Lemma 3.15] which constructs a test configuration that has the flat limit (21) as its central fibre.

Recalling also the variable τ\tau in §3.1, we have a change of variables

τ=et,\tau=e^{-t}, (22)

so that the limit of τA\tau^{A} as τ0\tau\to 0 corresponds to t+t\to+\infty; a subtlety may be that τ\tau\in\mathbb{C}^{*} while tt\in\mathbb{R}, but this does not affect the description of the flat limit up to a difference in unitarily rotating the embedding inside Nm1\mathbb{P}^{N_{m}-1}, since the unitary group is compact (see also [donlb, Lemma 2]). This correspondence allows us to compare the asymptotic behaviour of the (differential-geometric) energy functionals and algebro-geometric invariants given in terms of the central fibre of test configurations.

For the H0H_{0}-hermitian endomorphism A𝔤𝔩(H0(X,mKX))A\in\mathfrak{gl}(H^{0}(X,-mK_{X})) we perform a unitary change of basis for H0(X,mKX)H^{0}(X,-mK_{X}), i.e. replace the reference basis {si}i=1Nm\{s_{i}\}_{i=1}^{N_{m}} (resp. homogeneous coordinates [Z1::ZNm][Z_{1}:\cdots:Z_{N_{m}}]) by an H0H_{0}-unitarily equivalent one which is still written as {si}i=1Nm\{s_{i}\}_{i=1}^{N_{m}} (resp. [Z1::ZNm][Z_{1}:\cdots:Z_{N_{m}}]) by abuse of notation, so that A=diag(λ1,,λNm)A=\mathrm{diag}(\lambda_{1},\dots,\lambda_{N_{m}}) with λ1λNm\lambda_{1}\geq\cdots\geq\lambda_{N_{m}}, with respect to this new basis. We first prove the following lemma.

Lemma 5.2.

Suppose that we write A:=AλNmidNmA^{\prime}:=A-\lambda_{N_{m}}\cdot\mathrm{id}_{N_{m}} and consider the flat family π:𝒳A\pi:\mathcal{X}_{A^{\prime}}\to\mathbb{C} as in (20), whose central fibre is the flat limit 𝒳0\mathcal{X}_{0} defined by AA^{\prime} as in (21). Writing 𝒪(1)\mathcal{O}(1) for the hyperplane bundle over Nm1\mathbb{P}^{N_{m}-1} which 𝒳0\mathcal{X}_{0} is embedded in, and h~0\tilde{h}_{0} for the Fubini–Study metric on 𝒪(1)\mathcal{O}(1) defined by the reference hermitian form H0mH_{0}\in\mathcal{B}_{m}, the fibrewise mm-th root h~01/m\tilde{h}_{0}^{1/m} defines a volume form on 𝒳0,redreg\mathcal{X}^{\mathrm{reg}}_{0,\mathrm{red}} which may be degenerate.

Proof.

Recalling the second item of Lemma 4.1, we find that h~0|𝒳A\tilde{h}_{0}|_{\mathcal{X}_{A^{\prime}}} is a smooth hermitian metric on A=𝒪(1)|𝒳A\mathcal{L}_{A^{\prime}}=\mathcal{O}(1)|_{\mathcal{X}_{A^{\prime}}} which is allowed to be degenerate on 𝒳0,red\mathcal{X}_{0,\mathrm{red}}. For any point p𝒳0,redregp\in\mathcal{X}^{\mathrm{reg}}_{0,\mathrm{red}} we pick an open set 𝒰𝒳A\mathcal{U}\subset\mathcal{X}_{A^{\prime}} (in the Euclidean topology) with (p,0)𝒰Nm1×(p,0)\in\mathcal{U}\subset\mathbb{P}^{N_{m}-1}\times\mathbb{C}.

We observe that there exists a holomorphic coordinate system of 𝒳0,redreg\mathcal{X}^{\mathrm{reg}}_{0,\mathrm{red}} around pp which can be perturbed to give a holomorphic coordinate system of a point in a nearby noncentral fibre, as follows. We first pick a set of homogeneous polynomials defining 𝒳0,redreg\mathcal{X}^{\mathrm{reg}}_{0,\mathrm{red}} near pp and choose smooth coordinates by the implicit function theorem. By the definition of the flat limit we can pick a point qq in π1(t)etAι(X)\pi^{-1}(t)\cong e^{-tA^{\prime}}\cdot\iota(X) such that t0t\neq 0 and qq is close to pp in Nm1×\mathbb{P}^{N_{m}-1}\times\mathbb{C} in the Euclidean topology, and we can choose tt to be sufficiently close to 0 so that the set of defining homogeneous polynomials remain nondegenerate to satisfy the hypothesis of the implicit function theorem. Then the parameter-dependent implicit function theorem (see e.g. [Gloeckner, Theorem 2.3]) implies that the holomorphic coordinates near qq thus chosen converge to the ones near pp at least continuously with respect to the Euclidean topology of Nm1×\mathbb{P}^{N_{m}-1}\times\mathbb{C}.

Writing 𝒱:=𝒰𝒳0,red\mathcal{V}:=\mathcal{U}\setminus\mathcal{X}_{0,\mathrm{red}}, we find that h~01/m|𝒱\tilde{h}^{1/m}_{0}|_{\mathcal{V}} defines a smooth family of 2n2n-forms on 𝒱\mathcal{V}, since at each point (q,et)𝒱(q,e^{-t})\in\mathcal{V} we have 𝒪(1)|(q,et)=mKetAι(X)|q\mathcal{O}(1)|_{(q,e^{-t})}=-mK_{e^{-tA^{\prime}}\cdot\iota(X)}|_{q} for the anticanonically embedded Fano manifold etAι(X)Nm1e^{-tA^{\prime}}\cdot\iota(X)\subset\mathbb{P}^{N_{m}-1}, by recalling the correspondence omCX((KX)(KX)¯,)=KXKX¯\mathcal{H}om_{C^{\infty}_{X}}((-K_{X})\otimes\overline{(-K_{X})},\mathbb{C})=K_{X}\otimes\overline{K_{X}} explained in §2.1. Since h~0|𝒳A\tilde{h}_{0}|_{\mathcal{X}_{A^{\prime}}} is a smooth hermitian metric over 𝒳A\mathcal{X}_{A^{\prime}} we may extend the positive 2n2n-form h~01/m|𝒱\tilde{h}^{1/m}_{0}|_{\mathcal{V}} to a semipositive 2n2n-form on 𝒰\mathcal{U} by continuity, noting that the resulting form on 𝒰𝒳0,redreg\mathcal{U}\cap\mathcal{X}^{\mathrm{reg}}_{0,\mathrm{red}} can be written explicitly in terms of the homogeneous coordinates [Z1::ZNm][Z_{1}:\cdots:Z_{N_{m}}] and the homogeneous ideal defining 𝒳0,redreg\mathcal{X}^{\mathrm{reg}}_{0,\mathrm{red}}, or equivalently the homogeneous coordinate system that can be perturbed to the one for the nearby point, as discussed above. Finally, we have dim𝒳0,redreg=n\dim_{\mathbb{C}}\mathcal{X}^{\mathrm{reg}}_{0,\mathrm{red}}=n by the flatness of π:𝒳A\pi:\mathcal{X}_{A^{\prime}}\to\mathbb{C}, and hence a semipositive 2n2n-form on 𝒳0,redreg\mathcal{X}^{\mathrm{reg}}_{0,\mathrm{red}} defines a volume form which may be degenerate. ∎

The key proposition, which we prove by adapting the strategy of [donlb, §4], is the following.

Proposition 5.3.

Let A𝔤𝔩(H0(X,mKX))A\in\mathfrak{gl}(H^{0}(X,-mK_{X})) be H0H_{0}-hermitian with eigenvalues λ1λNm\lambda_{1}\geq\cdots\geq\lambda_{N_{m}}. We have

limt+ddt(FS(Ht))=2mi=1NmλiCi(𝒳0),\lim_{t\to+\infty}\frac{d}{dt}\mathscr{L}(\mathrm{FS}(H_{t}))=\frac{2}{m}\sum_{i=1}^{N_{m}}\lambda_{i}C_{i}(\mathcal{X}_{0}),

for some real numbers C1(𝒳0),,CNm(𝒳0)C_{1}(\mathcal{X}_{0}),\dots,C_{N_{m}}(\mathcal{X}_{0}), with i=1NmCi(𝒳0)=1\sum_{i=1}^{N_{m}}C_{i}(\mathcal{X}_{0})=1, which depends only on the flat limit 𝒳0\mathcal{X}_{0} as defined in (21) and the H0H_{0}-orthonormal basis {si}i=1Nm\{s_{i}\}_{i=1}^{N_{m}} that diagonalises AA.

Recall that by Remark 4.4 we already know that the above asymptotic slope is a well-defined real number.

Proof.

We consider the embedded variety etAι(X)Nm1e^{-tA}\cdot\iota(X)\subset\mathbb{P}^{N_{m}-1} as before. As in the previous lemma, h~0|etAι(X)\tilde{h}_{0}|_{e^{-tA}\cdot\iota(X)} defines a hermitian metric on the line bundle mKetAι(X)-mK_{e^{-tA}\cdot\iota(X)} and hence a volume form h~01/m|etAι(X)\tilde{h}_{0}^{1/m}|_{e^{-tA}\cdot\iota(X)} on etAι(X)e^{-tA}\cdot\iota(X). With this understood, we can re-write Lemma 2.5 as

ddt(FS(Ht))\displaystyle\frac{d}{dt}\mathscr{L}(\mathrm{FS}(H_{t})) =2mi=1Nmλi(etAι(X)h~01/m)1etAι(X)Zih~02l=1NmZlh~02h~01/m\displaystyle=\frac{2}{m}\sum_{i=1}^{N_{m}}\lambda_{i}\left(\int_{e^{-tA}\cdot\iota(X)}\tilde{h}^{1/m}_{0}\right)^{-1}\int_{e^{-tA}\cdot\iota(X)}\frac{\|Z_{i}\|^{2}_{\tilde{h}_{0}}}{\sum_{l=1}^{N_{m}}\|Z_{l}\|^{2}_{\tilde{h}_{0}}}\tilde{h}^{1/m}_{0}
=2mi=1Nmλi(etAι(X)h~01/m)1etAι(X)Zih~02l=1NmZlh~02h~01/m,\displaystyle=\frac{2}{m}\sum_{i=1}^{N_{m}}\lambda_{i}\left(\int_{e^{-tA^{\prime}}\cdot\iota(X)}\tilde{h}^{1/m}_{0}\right)^{-1}\int_{e^{-tA^{\prime}}\cdot\iota(X)}\frac{\|Z_{i}\|^{2}_{\tilde{h}_{0}}}{\sum_{l=1}^{N_{m}}\|Z_{l}\|^{2}_{\tilde{h}_{0}}}\tilde{h}^{1/m}_{0},

where A:=AλminidNmA^{\prime}:=A-\lambda_{\mathrm{min}}\cdot\mathrm{id}_{N_{m}} for the least eigenvalue λmin\lambda_{\mathrm{min}}\in\mathbb{R} of AA. While the effect of re-scaling cancels out between the denominator and the numerator, it is chosen so that the hermitian metric exp(mFS(Ht))h0m=h~0|etAι(X)\exp(-m\mathrm{FS}(H^{\prime}_{t}))h^{m}_{0}=\tilde{h}_{0}|_{e^{-tA^{\prime}}\cdot\iota(X)} extends to a smooth metric over 𝒳A\mathcal{X}_{A^{\prime}} which may be degenerate along the central fibre, as we saw in the proof of Lemmas 4.1 and 5.2.

We take the limit of the domain of the integration, in the sense of the integral currents, as in [donlb, §4]. We write |𝒳0||\mathcal{X}_{0}| for the algebraic cycle defined by 𝒳0\mathcal{X}_{0}, which we may write as |𝒳0|=jνj|Vj||\mathcal{X}_{0}|=\sum_{j}\nu_{j}|V_{j}|, using a finite \mathbb{Z}-linear combination of algebraic cycles defined by the reduced and irreducible components of 𝒳0\mathcal{X}_{0} (νj>0\nu_{j}>0). By Lemma 5.2, h~01/m\tilde{h}_{0}^{1/m} defines a 2n2n-form (and hence a volume form) on 𝒳0,redreg\mathcal{X}^{\mathrm{reg}}_{0,\mathrm{red}}, which naturally defines an absolutely continuous measure on each reduced and irreducible component of 𝒳0\mathcal{X}_{0}, since h~0\tilde{h}_{0} is a smooth (in particular bounded) metric over 𝒳A\mathcal{X}_{A^{\prime}} and hence has no singular part that is supported on a Zariski closed subset. Thus, the definition of the flat limit (21) implies that we can write the limit t+t\to+\infty of the above integral as

limt+ddt(FS(Ht))=2mi=1Nmλi(jνjVj𝒳0,redregh~01/m)1(jνjVj𝒳0,redregZih~02l=1NmZlh~02h~01/m).\lim_{t\to+\infty}\frac{d}{dt}\mathscr{L}(\mathrm{FS}(H_{t}))=\frac{2}{m}\sum_{i=1}^{N_{m}}\lambda_{i}\left(\sum_{j}\nu_{j}\int_{V_{j}\cap\mathcal{X}^{\mathrm{reg}}_{0,\mathrm{red}}}\tilde{h}^{1/m}_{0}\right)^{-1}\left(\sum_{j}\nu_{j}\int_{V_{j}\cap\mathcal{X}^{\mathrm{reg}}_{0,\mathrm{red}}}\frac{\|Z_{i}\|^{2}_{\tilde{h}_{0}}}{\sum_{l=1}^{N_{m}}\|Z_{l}\|^{2}_{\tilde{h}_{0}}}\tilde{h}^{1/m}_{0}\right).

We claim that each summand on the right hand side is a well-defined real number which depends only on the flat limit 𝒳0\mathcal{X}_{0} and the homogeneous coordinates [Z0::ZNm][Z_{0}:\cdots:Z_{N_{m}}]. We first claim that the denominator is a well-defined nonzero real number. To prove this claim, pick the homogeneous coordinate ZminZ_{\mathrm{min}} corresponding to the eigenvector associated to the least eigenvalue λmin\lambda_{\mathrm{min}} of AA. Since ι(X)\iota(X) is not contained in any proper linear subspace of Nm1\mathbb{P}^{N_{m}-1}, there exists at least one point p1ι(X)p_{1}\in\iota(X) where ZminZ_{\mathrm{min}} is nonzero at p1p_{1}. Taking the limit limt+etAp1=:p0\lim_{t\to+\infty}e^{-tA^{\prime}}\cdot p_{1}=:p_{0}, which is a (well-defined) point in 𝒳0,red\mathcal{X}_{0,\mathrm{red}} that is fixed by the \mathbb{C}^{*}-action induced by AA^{\prime} (by the definition of the flat limit (21)), there exists at least one point p𝒳0,redregp\in\mathcal{X}^{\mathrm{reg}}_{0,\mathrm{red}} where ZminZ_{\mathrm{min}} is nonzero at pp by continuity. With the scaling of AA^{\prime} as above, we find that the volume form h~01/m|𝒰\tilde{h}_{0}^{1/m}|_{\mathcal{U}} on some Euclidean open set 𝒰\mathcal{U} in 𝒳0,redreg\mathcal{X}^{\mathrm{reg}}_{0,\mathrm{red}} containing pp, given by Lemma 5.2, is nondegenerate. Outside this Euclidean open set, h~01/m\tilde{h}_{0}^{1/m} remains bounded over the whole of 𝒳0,redreg\mathcal{X}^{\mathrm{reg}}_{0,\mathrm{red}} (an important point being that we allow it to be zero), again by the scaling that we chose for AA^{\prime}. By extending h~01/m\tilde{h}_{0}^{1/m} as an absolutely continuous measure over 𝒳0,red\mathcal{X}_{0,\mathrm{red}}, we thus find that the total volume over Vj𝒳0,redregV_{j}\cap\mathcal{X}^{\mathrm{reg}}_{0,\mathrm{red}} is finite for all jj and nonzero for some jj, since 𝒳0,red\mathcal{X}_{0,\mathrm{red}} is compact in the Euclidean topology. It is immediate from the discussions so far that each summand depends only on the flat limit 𝒳0\mathcal{X}_{0}, and not on the particular generator A𝔤𝔩(H0(X,mKX))A\in\mathfrak{gl}(H^{0}(X,-mK_{X})) which defines 𝒳0\mathcal{X}_{0} (see also Remark 3.10).

Thus we find that for each i=1,,Nmi=1,\dots,N_{m},

Ci(𝒳0):=(jνjVj𝒳0,redregh~01/m)1(jνjVj𝒳0,redregZih~02l=1NmZlh~02h~01/m)C_{i}(\mathcal{X}_{0}):=\left(\sum_{j}\nu_{j}\int_{V_{j}\cap\mathcal{X}^{\mathrm{reg}}_{0,\mathrm{red}}}\tilde{h}^{1/m}_{0}\right)^{-1}\left(\sum_{j}\nu_{j}\int_{V_{j}\cap\mathcal{X}^{\mathrm{reg}}_{0,\mathrm{red}}}\frac{\|Z_{i}\|^{2}_{\tilde{h}_{0}}}{\sum_{l=1}^{N_{m}}\|Z_{l}\|^{2}_{\tilde{h}_{0}}}\tilde{h}^{1/m}_{0}\right)

is a well-defined real number depending only on the flat limit 𝒳0\mathcal{X}_{0} and the homogeneous coordinates [Z1::ZNm][Z_{1}:\cdots:Z_{N_{m}}] which correspond to the reference basis {si}i=1Nm\{s_{i}\}_{i=1}^{N_{m}}, as required; it is obvious from the above that we have i=1NmCi(𝒳0)=1\sum_{i=1}^{N_{m}}C_{i}(\mathcal{X}_{0})=1. ∎

We now begin the proof of Theorem 1.1. Lemma 5.2 and Proposition 5.3 give us the necessary ingredients to pursue the strategy introduced by Keller and the author for the JJ-balanced metrics in [hk18, §3.2], while some additional arguments are necessary to deal with the nontrivial holomorphic vector fields.

Proof of Theorem 1.1.

Suppose that (X,KX)(X,-K_{X}) is FF-polystable at level mm. We first show that

limt+ddt𝒟m(Ht)0\lim_{t\to+\infty}\frac{d}{dt}\mathscr{D}_{m}(H_{t})\geq 0 (23)

for all Bergman geodesic rays {Ht}t0m\{H_{t}\}_{t\geq 0}\subset\mathcal{B}_{m}, Ht=etAetAH_{t}=e^{-tA^{*}}e^{-tA} as defined in (5), with equality if and only if A𝔞𝔲𝔱(X)rA\in\mathfrak{aut}(X)_{r}; recall that 𝒟m\mathscr{D}_{m} is convex along {Ht}t0\{H_{t}\}_{t\geq 0} by (7) and Theorem 4.2.

Suppose that AA is H0H_{0}-hermitian and has rational eigenvalues. Following the idea of Saito–Takahashi [st19], we write 𝒟m(Ht)=𝒟(FS(Ht))+(FS(Ht))m(Ht)\mathscr{D}_{m}(H_{t})=\mathscr{D}(\mathrm{FS}(H_{t}))+\mathscr{E}(\mathrm{FS}(H_{t}))-\mathscr{E}_{m}(H_{t}). Then Lemma 2.5, Definition 3.6, and Theorem 4.3 imply that, for some cc\in\mathbb{N} such that cAcA has integral eigenvalues, we have

limt+ddt𝒟m(Ht)=1c(Ding(𝒳cA,cA)+Chowm(𝒳cA,cA))\lim_{t\to+\infty}\frac{d}{dt}\mathscr{D}_{m}(H_{t})=\frac{1}{c}\left(\mathrm{Ding}(\mathcal{X}_{cA},\mathcal{L}_{cA})+\mathrm{Chow}_{m}(\mathcal{X}_{cA},\mathcal{L}_{cA})\right) (24)

where (𝒳cA,cA)(\mathcal{X}_{cA},\mathcal{L}_{cA}) is a test configuration as defined in Proposition 3.3. The above quantity is nonnegative, and zero if and only if (𝒳cA,cA)(\mathcal{X}_{cA},\mathcal{L}_{cA}) is product, by the FF-polystability.

For a general H0H_{0}-hermitian A𝔤𝔩(H0(X,mKX))A\in\mathfrak{gl}(H^{0}(X,-mK_{X})), by passing to a basis that is H0H_{0}-unitarily equivalent to the reference basis if necessary, we write

A=diag(λ1,,λNm)A=\mathrm{diag}(\lambda_{1},\dots,\lambda_{N_{m}})

with λ1λNm\lambda_{1}\geq\cdots\geq\lambda_{N_{m}}. We then have

limt+ddt𝒟m(Ht)=2mi=1Nmλi(Ci(𝒳0)1Nm),\lim_{t\to+\infty}\frac{d}{dt}\mathscr{D}_{m}(H_{t})=\frac{2}{m}\sum_{i=1}^{N_{m}}\lambda_{i}\left(C_{i}(\mathcal{X}_{0})-\frac{1}{N_{m}}\right),

by Proposition 5.3, where 𝒳0\mathcal{X}_{0} is the flat limit (21) defined by AA. We now take an approximation {Ap}p=1𝔤𝔩(H0(X,mKX))\{A_{p}\}_{p=1}^{\infty}\subset\mathfrak{gl}(H^{0}(X,-mK_{X})) of AA by H0H_{0}-hermitian matrices with rational eigenvalues, as in [hk18, Lemma 3.15], so that the following hold: ApAA_{p}\to A as pp\to\infty (say in the Hilbert–Schmidt norm), and the flat limit (21) defined by ApA_{p} is equal to 𝒳0\mathcal{X}_{0} (i.e. the one defined by AA) for all large enough pp. Writing Ht,p:=exp(tAp)exp(tAp)H_{t,p}:=\exp(-tA_{p}^{*})\exp(-tA_{p}), and λ1,p,,λNm,p\lambda_{1,p},\dots,\lambda_{N_{m},p} for the eigenvalues of ApA_{p}, again by Proposition 5.3 we find

limt+ddt𝒟m(Ht,p)=2mi=1Nmλi,p(Ci(𝒳0)1Nm),\lim_{t\to+\infty}\frac{d}{dt}\mathscr{D}_{m}(H_{t,p})=\frac{2}{m}\sum_{i=1}^{N_{m}}\lambda_{i,p}\left(C_{i}(\mathcal{X}_{0})-\frac{1}{N_{m}}\right),

where an important point is that the flat limit 𝒳0\mathcal{X}_{0} remains unchanged for ApA_{p} when pp is large enough. Since ApAA_{p}\to A as pp\to\infty, we get

limt+ddt𝒟m(Ht)=limplimt+ddt𝒟m(Ht,p)0.\lim_{t\to+\infty}\frac{d}{dt}\mathscr{D}_{m}(H_{t})=\lim_{p\to\infty}\lim_{t\to+\infty}\frac{d}{dt}\mathscr{D}_{m}(H_{t,p})\geq 0. (25)

for any H0H_{0}-hermitian A𝔤𝔩(H0(X,mKX))A\in\mathfrak{gl}(H^{0}(X,-mK_{X})).

We prove that the inequality (25) is strict unless A𝔞𝔲𝔱(X)rA\in\mathfrak{aut}(X)_{r}. Suppose for contradiction that there exists A𝔞𝔲𝔱(X)rA\not\in\mathfrak{aut}(X)_{r}, H0H_{0}-hermitian, such that

limt+ddt𝒟m(Ht)=0\lim_{t\to+\infty}\frac{d}{dt}\mathscr{D}_{m}(H_{t})=0 (26)

holds for the Bergman geodesic ray {Ht}t0\{H_{t}\}_{t\geq 0} as defined in (5). By applying the argument in [hk18, proof of Lemma 3.17], we can write AA as a sum

A=Aα+AβA=A_{\alpha}+A_{\beta} (27)

of two H0H_{0}-hermitian matrices such that, in the diagonalising basis for AA, we have

  1. (i)

    Aα=diag(α1,,αNm)A_{\alpha}=\mathrm{diag}(\alpha_{1},\dots,\alpha_{N_{m}}) with αi\alpha_{i}\in\mathbb{Q} for all ii and α1αNm\alpha_{1}\geq\cdots\geq\alpha_{N_{m}},

  2. (ii)

    Aβ=diag(β1,,βNm)A_{\beta}=\mathrm{diag}(\beta_{1},\dots,\beta_{N_{m}}) with βi\beta_{i}\in\mathbb{R} for all ii and β1βNm\beta_{1}\geq\cdots\geq\beta_{N_{m}},

  3. (iii)

    the flat limit (21) defined by AA, AαA_{\alpha}, and AβA_{\beta}, are all equal.

We define Ht,α:=exp(tAα)exp(tAα)H_{t,\alpha}:=\exp(-tA_{\alpha}^{*})\exp(-tA_{\alpha}) and Ht,β:=exp(tAβ)exp(tAβ)H_{t,\beta}:=\exp(-tA_{\beta}^{*})\exp(-tA_{\beta}). The properties above, together with Proposition 5.3, imply that we have

limt+ddt𝒟m(Ht)=limt+ddt𝒟m(Ht,α)+limt+ddt𝒟m(Ht,β).\lim_{t\to+\infty}\frac{d}{dt}\mathscr{D}_{m}(H_{t})=\lim_{t\to+\infty}\frac{d}{dt}\mathscr{D}_{m}(H_{t,\alpha})+\lim_{t\to+\infty}\frac{d}{dt}\mathscr{D}_{m}(H_{t,\beta}). (28)

Now (26), combined with (25) and (28), necessarily implies

limt+ddt𝒟m(Ht,α)=limt+ddt𝒟m(Ht,β)=0.\lim_{t\to+\infty}\frac{d}{dt}\mathscr{D}_{m}(H_{t,\alpha})=\lim_{t\to+\infty}\frac{d}{dt}\mathscr{D}_{m}(H_{t,\beta})=0.

We again take cc\in\mathbb{N} so that cAαcA_{\alpha} has integral eigenvalues, and write (𝒳cAα,cAα)(\mathcal{X}_{cA_{\alpha}},\mathcal{L}_{cA_{\alpha}}) for the test configuration associated to cAαcA_{\alpha}. We further note

limt+ddt𝒟m(Ht,α)=1c(Ding(𝒳cAα,cAα)+Chowm(𝒳cAα,cAα))=0,\lim_{t\to+\infty}\frac{d}{dt}\mathscr{D}_{m}(H_{t,\alpha})=\frac{1}{c}\left(\mathrm{Ding}(\mathcal{X}_{cA_{\alpha}},\mathcal{L}_{cA_{\alpha}})+\mathrm{Chow}_{m}(\mathcal{X}_{cA_{\alpha}},\mathcal{L}_{cA_{\alpha}})\right)=0,

which implies that (𝒳cAα,cAα)(\mathcal{X}_{cA_{\alpha}},\mathcal{L}_{cA_{\alpha}}) must be product by the FF-polystability. Thus 𝒳cAα\mathcal{X}_{cA_{\alpha}} is isomorphic to ι(X)×(H0(X,mKX))×\iota(X)\times\mathbb{C}\subset\mathbb{P}(H^{0}(X,-mK_{X})^{\vee})\times\mathbb{C} and its central fibre 𝒳0\mathcal{X}_{0}, which is the flat limit (21) defined by AαA_{\alpha}, must be isomorphic to ι(X)\iota(X). Noting that 𝒳cAαι(X)×\mathcal{X}_{cA_{\alpha}}\stackrel{{\scriptstyle\sim}}{{\to}}\iota(X)\times\mathbb{C} must be invariant under the action of etAαe^{-tA_{\alpha}}, we find Aα𝔞𝔲𝔱(X)A_{\alpha}\in\mathfrak{aut}(X) by the equivariant embedding theorem (9) and hence Aα𝔞𝔲𝔱(X)rA_{\alpha}\in\mathfrak{aut}(X)_{r} since AαA_{\alpha} is H0H_{0}-hermitian. Recall that Aβ𝔤𝔩(H0(X,mKX))A_{\beta}\in\mathfrak{gl}(H^{0}(X,-mK_{X})) was defined in such a way that its flat limit agrees with that of AαA_{\alpha}, which is 𝒳0=ι(X)\mathcal{X}_{0}=\iota(X). Noting that etAβe^{-tA_{\beta}} acts on 𝒳0=ι(X)\mathcal{X}_{0}=\iota(X) we find Aβ𝔞𝔲𝔱(X)rA_{\beta}\in\mathfrak{aut}(X)_{r}, again by (9), but this implies A=Aα+Aβ𝔞𝔲𝔱(X)rA=A_{\alpha}+A_{\beta}\in\mathfrak{aut}(X)_{r} and contradicts our original hypothesis A𝔞𝔲𝔱(X)rA\not\in\mathfrak{aut}(X)_{r}. Thus, we need to have

limt+ddt𝒟m(Ht)>0,\lim_{t\to+\infty}\frac{d}{dt}\mathscr{D}_{m}(H_{t})>0,

for all H0H_{0}-hermitian endomorphisms AA that are not contained in 𝔞𝔲𝔱(X)r\mathfrak{aut}(X)_{r}.

We now prove that 𝒟m\mathscr{D}_{m} is invariant under the Aut0(X)r\mathrm{Aut}_{0}(X)_{r}-action. Recalling Aut0(X)r=K0\mathrm{Aut}_{0}(X)_{r}=K_{0}^{\mathbb{C}} in the notation of §2.3, we first show that for any HmH\in\mathcal{B}_{m} and any A1Lie(K0)A\in\sqrt{-1}\mathrm{Lie}(K_{0}) we have ddt𝒟m(Ht)=0\frac{d}{dt}\mathscr{D}_{m}(H_{t})=0 for the family Ht:=etAHH_{t}:=e^{tA}\cdot H defined by the action (10). Noting that AA is diagonalisable, if AA has integral eigenvalues we find that the family {FS(Ht)}t0\{\mathrm{FS}(H_{t})\}_{t\geq 0} (which may not be associated to a geodesic emanating from HH with respect to the bi-invariant metric, unless AA is hermitian with respect to HH) admits (𝒳A,A)(\mathcal{X}_{A},\mathcal{L}_{A}) as its non-Archimedean limit (cf. Lemma 4.1) by Lemma 2.9, up to scaling if necessary, where (𝒳A,A)(\mathcal{X}_{A},\mathcal{L}_{A}) is the product test configuration generated by AA. We thus find, recalling Remark 2.8, that

limt+ddt𝒟m(Ht)=Ding(𝒳A,A)+Chowm(𝒳A,A)=0\lim_{t\to+\infty}\frac{d}{dt}\mathscr{D}_{m}(H_{t})=\mathrm{Ding}(\mathcal{X}_{A},\mathcal{L}_{A})+\mathrm{Chow}_{m}(\mathcal{X}_{A},\mathcal{L}_{A})=0

by the FF-polystability; Theorem 4.3 is stated only for the Bergman geodesic rays but the result above holds since {FS(Ht)}t0\{\mathrm{FS}(H_{t})\}_{t\geq 0} is a psh ray (which is in fact a geodesic in \mathcal{H} induced by the holomorphic vector field AA by Lemma 2.9) that admits (𝒳A,A)(\mathcal{X}_{A},\mathcal{L}_{A}) as its non-Archimedean limit (see also the proof of Lemma 5.5 presented later). We now observe the following: for any B𝔞𝔲𝔱(X)B\in\mathfrak{aut}(X) and the associated family Ht=etBHH_{t}=e^{tB}\cdot H, writing {Zi}i=1Nm\{Z^{\prime}_{i}\}_{i=1}^{N_{m}} for an HH-orthonormal basis and h~H\tilde{h}_{H} for the hermitian metric on the hyperplane bundle over Nm1\mathbb{P}^{N_{m}-1} with respect to HH, we have

ddt𝒟m(Ht)\displaystyle\frac{d}{dt}\mathscr{D}_{m}(H_{t}) =1mi,j=1Nm(Bij+Bij)(etBι(X)h~H(Zi,Zj)l=1NmZlh~H2h~H1/m1Nmδij)\displaystyle=\frac{1}{m}\sum_{i,j=1}^{N_{m}}(B_{ij}+B^{\dagger}_{ij})\left(\fint_{e^{-tB}\cdot\iota(X)}\frac{\tilde{h}_{H}(Z^{\prime}_{i},Z^{\prime}_{j})}{\sum_{l=1}^{N_{m}}\|Z^{\prime}_{l}\|_{\tilde{h}_{H}}^{2}}\tilde{h}_{H}^{1/m}-\frac{1}{N_{m}}\delta_{ij}\right)
=1mi,j=1Nm(Bij+Bij)(ι(X)h~H(Zi,Zj)l=1NmZlh~H2h~H1/m1Nmδij),\displaystyle=\frac{1}{m}\sum_{i,j=1}^{N_{m}}(B_{ij}+B^{\dagger}_{ij})\left(\fint_{\iota(X)}\frac{\tilde{h}_{H}(Z^{\prime}_{i},Z^{\prime}_{j})}{\sum_{l=1}^{N_{m}}\|Z^{\prime}_{l}\|_{\tilde{h}_{H}}^{2}}\tilde{h}_{H}^{1/m}-\frac{1}{N_{m}}\delta_{ij}\right), (29)

where all the matrices are represented with respect to {Zi}i=1Nm\{Z^{\prime}_{i}\}_{i=1}^{N_{m}} and BB^{\dagger} is the hermitian conjugate of BB with respect to HH, by recalling the proof of Lemma 2.5. We thus find that, for a general A1Lie(K0)A\in\sqrt{-1}\mathrm{Lie}(K_{0}) and the associated family Ht=etAHH_{t}=e^{tA}\cdot H, we have

ddt𝒟m(Ht)=lims+dds𝒟m(Hs)=0\frac{d}{dt}\mathscr{D}_{m}(H_{t})=\lim_{s\to+\infty}\frac{d}{ds}\mathscr{D}_{m}(H_{s})=0

for all t0t\geq 0, by approximating AA by an H0H_{0}-hermitian sequence {Ap}p=11Lie(K0)\{A_{p}\}_{p=1}^{\infty}\subset\sqrt{-1}\mathrm{Lie}(K_{0}) with rational eigenvalues and converging to AA, by using [hk18, Lemma 3.15] as before, noting that the associated flat limits are all equal to ι(X)\iota(X). Note moreover that, for any A1Lie(K0)A\in\sqrt{-1}\mathrm{Lie}(K_{0}) and any uK0u\in K_{0}, by replacing HH (resp. {Zi}i=1Nm\{Z^{\prime}_{i}\}_{i=1}^{N_{m}}) by uHu\cdot H (resp. {u1Zi}i=1Nm\{u^{-1}Z^{\prime}_{i}\}_{i=1}^{N_{m}}) in the above we have

ddt𝒟m(etAuH)=0,\frac{d}{dt}\mathscr{D}_{m}(e^{tA}u\cdot H)=0,

again by the FF-polystability, as {etAuH}t0\{e^{tA}u\cdot H\}_{t\geq 0} admits (𝒳A,A)(\mathcal{X}_{A},\mathcal{L}_{A}) (rotated by uu in Nm1\mathbb{P}^{N_{m}-1}) as its non-Archimedean limit if AA has integral eigenvalues. This proves the claimed Aut0(X)r\mathrm{Aut}_{0}(X)_{r}-invariance of 𝒟m\mathscr{D}_{m} by the global Cartan decomposition for Aut0(X)r=K0\mathrm{Aut}_{0}(X)_{r}=K_{0}^{\mathbb{C}}, since we have ddt𝒟m(etAH)=0\frac{d}{dt}\mathscr{D}_{m}(e^{tA^{\prime}}\cdot H)=0 for all ALie(K0)A^{\prime}\in\mathrm{Lie}(K_{0}) and all t0t\geq 0 by applying (29) to AA^{\prime} (otherwise it contradicts the compactness of K0K_{0}).

The conclusion of the argument so far is that 𝒟m\mathscr{D}_{m} is a smooth function on m\mathcal{B}_{m} that is convex along the Bergman geodesic rays emanating from H0H_{0} and invariant under the action of Aut0(X)r\mathrm{Aut}_{0}(X)_{r}, with a strictly positive asymptotic slope along a Bergman geodesic ray not contained in the Aut0(X)r\mathrm{Aut}_{0}(X)_{r}-orbit (or equivalently the Aut0(X)\mathrm{Aut}_{0}(X)-orbit, as long as we consider the Bergman geodesic rays emanating from H0H_{0} that are by definition generated by the H0H_{0}-hermitian endomorphisms). Since m\mathcal{B}_{m} is a complete Riemannian manifold with respect to the bi-invariant metric (whose geodesics are precisely the Bergman geodesics) on which Aut0(X)r\mathrm{Aut}_{0}(X)_{r} acts isometrically by Lemma 2.9, this implies that m\mathcal{B}_{m} admits a critical point which is unique modulo the action of Aut0(X)r\mathrm{Aut}_{0}(X)_{r}. In fact the critical point of 𝒟m\mathscr{D}_{m} is unique modulo the action of Aut0(X)\mathrm{Aut}_{0}(X); this can be seen by considering (29) for a general B𝔞𝔲𝔱(X)B\in\mathfrak{aut}(X) and the family Ht=etBHH_{t}=e^{tB}\cdot H emanating from the critical point HH of 𝒟m\mathscr{D}_{m}, for which the term inside the bracket is zero by Proposition 2.6.

For the proof of the reverse direction, suppose that (X,KX)(X,-K_{X}) admits an anticanonically balanced metric HH at level mm which is unique up to the Aut0(X)\mathrm{Aut}_{0}(X)-action. This in particular implies that 𝒟m\mathscr{D}_{m} must be bounded below over m\mathcal{B}_{m} by convexity (Theorem 4.2), which in turn implies that 𝒟m\mathscr{D}_{m} must be invariant under the group action Aut0(X)m\mathrm{Aut}_{0}(X)\curvearrowright\mathcal{B}_{m}, as otherwise there would exist an Aut0(X)\mathrm{Aut}_{0}(X)-orbit along which 𝒟m\mathscr{D}_{m} tends to -\infty, as we saw above in (29). Since 𝒟m\mathscr{D}_{m} is strictly convex along the Bergman geodesic rays not contained in the Aut0(X)\mathrm{Aut}_{0}(X)-orbit by Theorem 4.2, we find that for any H0mH^{\prime}_{0}\in\mathcal{B}_{m} and any Bergman geodesic ray {Ht}t0\{H^{\prime}_{t}\}_{t\geq 0} emanating from H0H^{\prime}_{0} we have

limt+ddt𝒟m(Ht)0,\lim_{t\to+\infty}\frac{d}{dt}\mathscr{D}_{m}(H^{\prime}_{t})\geq 0, (30)

with equality if and only if {Ht}t0\{H^{\prime}_{t}\}_{t\geq 0} is contained in the Aut0(X)\mathrm{Aut}_{0}(X)-orbit of H0H^{\prime}_{0}; this is a consequence of 𝒟m\mathscr{D}_{m} being proper modulo the Aut0(X)\mathrm{Aut}_{0}(X)-action over m\mathcal{B}_{m}, which in turn follows from the existence of the unique global minimum of 𝒟m\mathscr{D}_{m} over m\mathcal{B}_{m} up to Aut0(X)\mathrm{Aut}_{0}(X). Suppose that we have a very ample test configuration (𝒳A,A)(\mathcal{X}_{A},\mathcal{L}_{A}) of exponent mm, where AA is the generator of the one-parameter subgroup as in Proposition 3.3. We take a basis {si}i=1Nm\{s^{\prime}_{i}\}_{i=1}^{N_{m}} for H0(X,mKX)H^{0}(X,-mK_{X}) such that AA is diagonal (with integral eigenvalues), and define H0H^{\prime}_{0} to be the positive definite hermitian form such that {si}i=1Nm\{s^{\prime}_{i}\}_{i=1}^{N_{m}} is H0H^{\prime}_{0}-orthonormal. Arguing as in (24), we find (by Theorem 4.3)

limt+ddt𝒟m(Ht)=Ding(𝒳A,A)+Chowm(𝒳A,A),\lim_{t\to+\infty}\frac{d}{dt}\mathscr{D}_{m}(H^{\prime}_{t})=\mathrm{Ding}(\mathcal{X}_{A},\mathcal{L}_{A})+\mathrm{Chow}_{m}(\mathcal{X}_{A},\mathcal{L}_{A}),

which is nonnegative by (30), with equality if and only if A𝔞𝔲𝔱(X)A\in\mathfrak{aut}(X), i.e. (𝒳A,A)(\mathcal{X}_{A},\mathcal{L}_{A}) is product. This establishes the FF-polystability as required. ∎

We now note that the uniqueness of anticanonically balanced metrics modulo automorphism, if they exist, can also be proved directly from the second statement of Theorem 4.2 by Berndtsson.

Proposition 5.4.

If H,HmH,H^{\prime}\in\mathcal{B}_{m} are both critical points of 𝒟m\mathscr{D}_{m}, then there exists aAut0(X)a\in\mathrm{Aut}_{0}(X) such that H=aHH^{\prime}=a\cdot H.

We omit the proof, which follows from Theorem 4.2, Lemma 2.9, and [Lempert2021, Theorem 1.1]; the details are also written in [tak19, Proof of Proposition 3.3]. We are now ready to prove Corollary 1.2.

Proof of Corollary 1.2.

Since Aut0(X)\mathrm{Aut}_{0}(X) is assumed to be trivial, a test configuration is product if and only if it is trivial. Then δm>1\delta_{m}>1 easily follows from the FF-stability by Theorem 1.1 and Rubinstein–Tian–Zhang [rtz, Theorem 2.3 (ii)]. Conversely, if δm>1\delta_{m}>1 then there exists an anticanonically balanced metric at level mm by Rubinstein–Tian–Zhang [rtz, Theorem 2.3 (i)], which is unique by Proposition 5.4 since Aut0(X)\mathrm{Aut}_{0}(X) is trivial. Thus the claimed FF-stability follows from Theorem 1.1. ∎

We finally comment on the following fact, which is clearly well-known to the experts but we provide a proof as it will be referred to later.

Lemma 5.5.

If (X,KX)(X,-K_{X}) is FF-stable at level mm then Aut0(X)r\mathrm{Aut}_{0}(X)_{r} is trivial.

Proof.

Suppose 𝔞𝔲𝔱(X)r0\mathfrak{aut}(X)_{r}\neq 0. By taking a rational approximation as in the proof of Theorem 1.1, there exists a nonzero A1Lie(K0)A\in\sqrt{-1}\mathrm{Lie}(K_{0}) with integral eigenvalues (i.e. the generator of a one-parameter subgroup Aut0(X)rGL(H0(X,mKX))\mathbb{C}^{*}\to\mathrm{Aut}_{0}(X)_{r}\subset GL(H^{0}(X,-mK_{X})) in the category of algebraic groups), which gives rise to a product test configuration (𝒳,)(\mathcal{X},\mathcal{L}). We then observe Ding(𝒳,)=Ding(𝒳,)\mathrm{Ding}(\mathcal{X},\mathcal{L})=-\mathrm{Ding}(\mathcal{X}^{\prime},\mathcal{L}^{\prime}) and Chowm(𝒳,)=Chowm(𝒳,)\mathrm{Chow}_{m}(\mathcal{X},\mathcal{L})=-\mathrm{Chow}_{m}(\mathcal{X}^{\prime},\mathcal{L}^{\prime}) for the product test configuration (𝒳,)(\mathcal{X}^{\prime},\mathcal{L}^{\prime}) generated by A𝔞𝔲𝔱(X)r-A\in\mathfrak{aut}(X)_{r}; the first equality holds since in this case the Ding invariant, which is the asymptotic slope of 𝒟\mathscr{D} along a holomorphic vector field as we saw in the proof of Theorem 1.1, is equal to the classical Futaki invariant (see the original formula in [Futaki83], or [Berman16, (3.4)] and [dontoric, Proposition 2.2.2]), and the second equality follows since the Chow weight is the (genuine, finite dimensional) GIT weight of AA acting on the Chow point represented by ι(X)(H0(X,mKX))\iota(X)\subset\mathbb{P}(H^{0}(X,-mK_{X})^{\vee}) (see e.g. [rt07, Futaki12] for more details). Thus it is impossible for all the invariants above to be strictly positive. ∎

Note that the above proof is also related to the vanishing of the higher Futaki invariants [Futaki04], as discussed in e.g. [st19, Proposition 5.4].

5.2. Extension to Kähler–Ricci gg-solitons

We now prove a similar result for the Kähler–Ricci gg-soliton case. First note that we have, from Theorem 4.3 and (13),

limt+ddt𝒟mg(Ht)=𝒟mg,NA(𝒳A,A)\lim_{t\to+\infty}\frac{d}{dt}\mathscr{D}^{g}_{m}(H_{t})=\mathscr{D}^{g,\mathrm{NA}}_{m}(\mathcal{X}_{A},\mathcal{L}_{A})

for a Bergman geodesic ray {Ht}t0mT\{H_{t}\}_{t\geq 0}\subset\mathcal{B}^{T}_{m} emanating from H0H_{0}, generated by an H0H_{0}-hermitian matrix A𝔤𝔩(H0(X,mKX))A\in\mathfrak{gl}(H^{0}(X,-mK_{X})) with integral eigenvalues and commuting with the TT-action. Given this formula, it is straightforward to adapt the proof of Theorem 1.1 to the gg-soliton case.

Theorem 5.6.

A Fano manifold (X,KX)(X,-K_{X}) admits a TT-invariant anticanonically gg-balanced metric at level mm, which is unique up to Aut0(X,T)\mathrm{Aut}_{0}(X,T^{\mathbb{C}}), if and only if it is FF-polystable in the sense of gg-solitons at level mm.

Proof.

The proof is almost the same as that of Theorem 1.1, since the difference 𝒟mg(Ht)𝒟m(Ht)\mathscr{D}^{g}_{m}(H_{t})-\mathscr{D}_{m}(H_{t}) is linear in AA; we may even assume that mg(Ht)\mathscr{E}^{g}_{m}(H_{t}) is constantly equal to zero, by using the translation invariance of the functional 𝒟mg\mathscr{D}^{g}_{m} (cf. Remark 2.8). The only subtlety is that the domain of 𝒟mg\mathscr{D}^{g}_{m} is mT\mathcal{B}_{m}^{T} rather than m\mathcal{B}_{m}, but the argument works word by word since we only need to consider the matrices that commute with (i.e. preserve the weight decomposition of) TT^{\mathbb{C}}, and we can construct the rational approximation that also commutes with the TT^{\mathbb{C}}-action (recall also Lemma 3.12); this can be done by noting that AA can be diagonalised preserving the weight decomposition (12), since AA is assumed to commute with the TT^{\mathbb{C}}-action, and so we can repeat the previous argument which explicitly deals with the eigenvalues. We finally recall that mT\mathcal{B}^{T}_{m} is a Riemannian symmetric space with respect to the bi-invariant metric, as pointed out in §2.3. ∎

Remark 5.7.

The fact mentioned in Lemma 5.5 applies to the gg-solitons, by observing

𝒟mg,NA(𝒳,)=\displaystyle\mathscr{D}^{g,\mathrm{NA}}_{m}(\mathcal{X},\mathcal{L})= Ding(𝒳,)+Chowm(𝒳,)\displaystyle\mathrm{Ding}(\mathcal{X},\mathcal{L})+\mathrm{Chow}_{m}(\mathcal{X},\mathcal{L})
+tr(A+A)mh0(X,mKX)1mNmgm¯λPmg(λ/m)tr((A+A)|Rm,λ).\displaystyle+\frac{\mathrm{tr}(A+A^{*})}{mh^{0}(X,-mK_{X})}-\frac{1}{mN_{m}\overline{g_{m}}}\sum_{\lambda\in P_{m}}g(\lambda/m)\mathrm{tr}\left((A+A^{*})|_{R_{m,\lambda}}\right).

and arguing exactly the same as before, by noting that the last two terms clearly change the sign when AA is replaced by A-A. In particular, if (X,KX)(X,-K_{X}) is FF-stable in the sense of gg-solitons at level mm then Aut0(X,T)r\mathrm{Aut}_{0}(X,T^{\mathbb{C}})_{r} must be trivial; an important difference to Lemma 5.5 is that Aut0(X,T)r\mathrm{Aut}_{0}(X,T^{\mathbb{C}})_{r} can never be trivial if TT^{\mathbb{C}} is nontrivial. Thus (X,KX)(X,-K_{X}) cannot be FF-stable in the sense of gg-solitons if TT^{\mathbb{C}} is nontrivial.

Note that the above balanced metric is never unique if TT^{\mathbb{C}} is nontrivial, and unique only up to automorphisms that commute with TT^{\mathbb{C}}. This is a well-documented phenomenon and follows also from Remark 5.7; for there to be a balanced metric 𝒟mg\mathscr{D}^{g}_{m} must be bounded below by convexity, but this forces 𝒟mg\mathscr{D}^{g}_{m} to be invariant under Aut0(X,T)r\mathrm{Aut}_{0}(X,T^{\mathbb{C}})_{r} (the proof is exactly the same as the one for Theorem 1.1). Note furthermore that the analogue of Proposition 5.4 holds: if H,HmTH,H^{\prime}\in\mathcal{B}_{m}^{T} are both critical points of 𝒟mg\mathscr{D}^{g}_{m}, then there exists aAut0(X,T)a\in\mathrm{Aut}_{0}(X,T^{\mathbb{C}}) such that H=aHH^{\prime}=a\cdot H. Indeed, we connect H0:=HH_{0}:=H and H1:=HH_{1}:=H^{\prime} by a Bergman geodesic path {Ht}0t1\{H_{t}\}_{0\leq t\leq 1} in mT\mathcal{B}^{T}_{m}, and write Ht=eBtHeBtH_{t}=e^{B^{\dagger}t}He^{Bt} for some B𝔤𝔩(H0(X,mKX))B\in\mathfrak{gl}(H^{0}(X,-mK_{X})) that is hermitian with respect to HH and commutes with Lie(T)\mathrm{Lie}(T). We find B𝔞𝔲𝔱(X)B\in\mathfrak{aut}(X) exactly as in the proof of Proposition 5.4, and hence B𝔞𝔲𝔱(X,T)B\in\mathfrak{aut}(X,T^{\mathbb{C}}) as it commutes with Lie(T)\mathrm{Lie}(T).

With the above remark in mind we get the following corollary, by combining the above theorem with [rtz, Theorem 6.13]; see §A.2 for the definition of δmg\delta^{g}_{m}.

Corollary 5.8.

If (X,KX)(X,-K_{X}) is FF-polystable in the sense of gg-solitons then δmg1\delta^{g}_{m}\geq 1. If δmg>1\delta^{g}_{m}>1 then (X,KX)(X,-K_{X}) is FF-polystable in the sense of gg-solitons at level mm.

Note that Rubinstein–Tian–Zhang proved δmg=1\delta^{g}_{m}=1 for g1g\equiv 1 and X=nX=\mathbb{P}^{n} [rtz, Corollary 7.2].

5.3. Extension to coupled Kähler–Einstein metrics

We now prove the analogue of Theorem 1.1 for the coupled Kähler–Einstein case. We first prove the following result, which can be considered as an analogue of the argument that appears in [HWN19, §5], in which we do not necessarily assume 𝒴=𝒳1==𝒳k\mathcal{Y}=\mathcal{X}_{1}=\cdots=\mathcal{X}_{k} or 𝒴=i=1ki\mathcal{L}_{\mathcal{Y}}=\sum_{i=1}^{k}\mathcal{L}_{i} (see Remark 3.18).

Theorem 5.9.

Let {(H1,t,,Hk,t)}t0𝓑m\{(H_{1,t},\dots,H_{k,t})\}_{t\geq 0}\subset\bm{\mathcal{B}}_{m} be a kk-tuple of Bergman geodesic rays generated by (A1,,Ak)𝔤𝔩(H0(X,mL1)××𝔤𝔩(H0(X,mLk)(A_{1},\dots,A_{k})\in\mathfrak{gl}(H^{0}(X,mL_{1})\times\cdots\times\mathfrak{gl}(H^{0}(X,mL_{k}), which defines a kk-tuple of psh rays (FS1(H1,t),,FSk(Hk,t))(\mathrm{FS}_{1}(H_{1,t}),\dots,\mathrm{FS}_{k}(H_{k,t})) in 𝓗\bm{\mathcal{H}}. Suppose that AiA_{i} is hermitian, with integral eigenvalues, with respect to the reference hermitian form Hi,0H_{i,0} for each i=1,,ki=1,\dots,k, corresponding to a kk-tuple of very ample test configurations (𝒳A1,A1),,(𝒳Ak,Ak)(\mathcal{X}_{A_{1}},\mathcal{L}_{A_{1}}),\dots,(\mathcal{X}_{A_{k}},\mathcal{L}_{A_{k}}) of exponent mm. Then we have

limt+𝒟coupled(FS1(H1,t),,FSk(Hk,t))t=Ding((𝒳Ai,Ai)i=1k).\lim_{t\to+\infty}\frac{\mathscr{D}^{\mathrm{coupled}}(\mathrm{FS}_{1}(H_{1,t}),\dots,\mathrm{FS}_{k}(H_{k,t}))}{t}=\mathrm{Ding}\left((\mathcal{X}_{A_{i}},\mathcal{L}_{A_{i}})_{i=1}^{k}\right).
Proof.

First recall that we have Hi,t=eAiteAitH_{i,t}=e^{-A^{*}_{i}t}e^{-A_{i}t} as in (5), and we define

𝑯t:=H1,tHk,t.\bm{H}_{t}:=H_{1,t}\otimes\cdots\otimes H_{k,t}.

Writing {𝒔𝒋}𝒋\{\bm{s}_{\bm{j}}\}_{\bm{j}} for the reference basis for H0(X,mL1)H0(X,mLk)H^{0}(X,mL_{1})\otimes\cdots\otimes H^{0}(X,mL_{k}), as in (17), and defining

𝒈t:=eA1teAkt,\bm{g}_{t}:=e^{A_{1}t}\otimes\cdots\otimes e^{A_{k}t},

we find that {𝒈t𝒔𝒋}𝒋\{\bm{g}_{t}\cdot\bm{s}_{\bm{j}}\}_{\bm{j}} is an 𝑯t\bm{H}_{t}-orthonormal basis, where the action \cdot is the natural tensor product action on H0(X,mL1)H0(X,mLk)H^{0}(X,mL_{1})\otimes\cdots\otimes H^{0}(X,mL_{k}). After an appropriate scaling if necessary, we find by Lemma 4.1 that the Bergman geodesic ray {FS(𝑯t)}t0\{\mathrm{FS}(\bm{H}_{t})\}_{t\geq 0}\subset\mathcal{H} admits a non-Archimedean metric ϕNA\phi^{\mathrm{NA}} on KX-K_{X} as non-Archimedean limit, where ϕNA\phi^{\mathrm{NA}} is represented by the very ample test configuration defined as the Zariski closure of ιcoupled(X)(H0(X,mL1)H0(X,mLk))\iota_{\mathrm{coupled}}(X)\subset\mathbb{P}(H^{0}(X,mL_{1})^{\vee}\otimes\cdots\otimes H^{0}(X,mL_{k})^{\vee}) under the one-parameter subgroup 𝒈t=τA1τAk\bm{g}^{\vee}_{t}=\tau^{A_{1}}\otimes\cdots\otimes\tau^{A_{k}}, by recalling the identification τ=et\tau=e^{-t} (and the sign change by taking the dual) as in (22). The resulting test configuration is exactly the one that is generated by the \mathbb{C}^{*}-actions of (𝒳Ai,Ai)i=1k(\mathcal{X}_{A_{i}},\mathcal{L}_{A_{i}})_{i=1}^{k}, which is denoted by (𝒴,𝒴)(\mathcal{Y},\mathcal{L}_{\mathcal{Y}}).

Note moreover that Lemma 2.18 gives coupled(FS1(H1,t),,FSk(Hk,t))=(FS(𝑯t))\mathscr{L}^{\mathrm{coupled}}(\mathrm{FS}_{1}(H_{1,t}),\dots,\mathrm{FS}_{k}(H_{k,t}))=\mathscr{L}(\mathrm{FS}(\bm{H}_{t})). Thus we find, by Theorem 4.3, that

limt+coupled(FS1(H1,t),,FSk(Hk,t))t=limt+(FS(𝑯t))t=1+lct(𝒴ν,D𝒴ν,𝒴ν;𝒴0ν).\lim_{t\to+\infty}\frac{\mathscr{L}^{\mathrm{coupled}}(\mathrm{FS}_{1}(H_{1,t}),\dots,\mathrm{FS}_{k}(H_{k,t}))}{t}=\lim_{t\to+\infty}\frac{\mathscr{L}(\mathrm{FS}(\bm{H}_{t}))}{t}=-1+\mathrm{lct}(\mathcal{Y}^{\nu},D_{\mathcal{Y}^{\nu},\mathcal{L}_{\mathcal{Y}^{\nu}}};\mathcal{Y}^{\nu}_{0}).

Combined with the asymptotic slopes for (FSi(Hi,t))\mathscr{E}(\mathrm{FS}_{i}(H_{i,t})), again by Theorem 4.3, we get the result. ∎

The above theorem immediately implies the following.

Lemma 5.10.

Let {(H1,t,,Hk,t)}t0𝓑m\{(H_{1,t},\dots,H_{k,t})\}_{t\geq 0}\subset\bm{\mathcal{B}}_{m} be a kk-tuple of Bergman geodesic rays defined exactly as in Theorem 5.9. Then

limt+ddt𝒟mcoupled(H1,t,,Hk,t)=Ding((𝒳Ai,Ai)i=1k)+i=1kChowm(𝒳Ai,Ai).\lim_{t\to+\infty}\frac{d}{dt}\mathscr{D}^{\mathrm{coupled}}_{m}(H_{1,t},\dots,H_{k,t})=\mathrm{Ding}\left((\mathcal{X}_{A_{i}},\mathcal{L}_{A_{i}})_{i=1}^{k}\right)+\sum_{i=1}^{k}\mathrm{Chow}_{m}(\mathcal{X}_{A_{i}},\mathcal{L}_{A_{i}}).
Proof.

The claim is obvious by noting, as in [st19], that 𝒟mcoupled(H1,t,,Hk,t)\mathscr{D}^{\mathrm{coupled}}_{m}(H_{1,t},\dots,H_{k,t}) can be written as

𝒟coupled(FS1(H1,t),,FSk(Hk,t))+i=1k((FSi(Hi,t))i,m(Hi)),\mathscr{D}^{\mathrm{coupled}}(\mathrm{FS}_{1}(H_{1,t}),\dots,\mathrm{FS}_{k}(H_{k,t}))+\sum_{i=1}^{k}\left(\mathscr{E}(\mathrm{FS}_{i}(H_{i,t}))-\mathscr{E}_{i,m}(H_{i})\right),

by recalling Lemma 2.18, and applying Lemma 2.5, Definition 3.6, Theorems 4.3 and 5.9. ∎

We now prove the following coupled version of Theorem 1.1.

Theorem 5.11.

(X,KX;L1,,Lk)(X,-K_{X};L_{1},\dots,L_{k}) admits a coupled anticanonically balanced metric at level mm, which is unique up to Aut0(X)\mathrm{Aut}_{0}(X), if and only if

Ding((𝒳i,i)i=1k)+i=1kChowm(𝒳i,i)0\mathrm{Ding}\left((\mathcal{X}_{i},\mathcal{L}_{i})_{i=1}^{k}\right)+\sum_{i=1}^{k}\mathrm{Chow}_{m}(\mathcal{X}_{i},\mathcal{L}_{i})\geq 0

holds for any kk-tuple of very ample test configurations (𝒳1,1),,(𝒳k,k)(\mathcal{X}_{1},\mathcal{L}_{1}),\dots,(\mathcal{X}_{k},\mathcal{L}_{k}) for (X,L1),,(X,Lk)(X,L_{1}),\dots,(X,L_{k}) respectively, each of exponent mm, with equality if and only if the test configuration (𝒴,𝒴)(\mathcal{Y},\mathcal{L}_{\mathcal{Y}}) generated by their \mathbb{C}^{*}-actions is product.

Proof.

The proof in §5.1 applies almost word by word: for each (A1,,Ak)𝔤𝔩(H0(X,mL1)××𝔤𝔩(H0(X,mLk)(A_{1},\dots,A_{k})\in\mathfrak{gl}(H^{0}(X,mL_{1})\times\cdots\times\mathfrak{gl}(H^{0}(X,mL_{k}) we choose an Hi,0H_{i,0}-orthonormal basis for H0(X,mLi)H^{0}(X,mL_{i}) such that AiA_{i} is diagonal, for i=1,,ki=1,\dots,k, and argue as before. The only nontrivial point, however, is the following: when we choose a rational approximation {Ai,p}p=1\{A_{i,p}\}_{p=1}^{\infty} of AiA_{i} such that the associated flat limit defined by Ai,pA_{i,p} is the same as the one defined by AiA_{i} for all large enough pp, the flat limit 𝒴0(H0(X,mL1)H0(X,mLk))\mathcal{Y}_{0}\subset\mathbb{P}(H^{0}(X,mL_{1})^{\vee}\otimes\cdots\otimes H^{0}(X,mL_{k})^{\vee}) defined by eA1teAkte^{-A_{1}t}\otimes\cdots\otimes e^{-A_{k}t} is also the same as the one defined by eA1,pteAk,pte^{-A_{1,p}t}\otimes\cdots\otimes e^{-A_{k,p}t} for all large enough pp. This is indeed true, since the weight of (A1,,Ak)(A_{1},\dots,A_{k}) on H0(X,mL1)H0(X,mLk)H^{0}(X,mL_{1})^{\vee}\otimes\cdots\otimes H^{0}(X,mL_{k})^{\vee}, with Ai=diag(λi,1,,λi,Ni,m)A_{i}=\mathrm{diag}(\lambda_{i,1},\dots,\lambda_{i,N_{i,m}}), is given by the sum (and hence a \mathbb{Z}-linear combination) of these weights as i=1kλi,ji\sum_{i=1}^{k}\lambda_{i,j_{i}} (ji{1,,Ni,m}j_{i}\in\{1,\dots,N_{i,m}\} and i=1,,ki=1,\dots,k); thus the proof of [hk18, Lemmas 3.15 and 3.17], which was used in the proof of Theorem 1.1, applies word by word. Likewise, when we choose Ai,αA_{i,\alpha} and Ai,βA_{i,\beta} as (27) in the proof of Theorem 1.1, we can show that the flat limits defined by eA1teAkte^{-A_{1}t}\otimes\cdots\otimes e^{-A_{k}t}, eA1,αteAk,αte^{-A_{1,\alpha}t}\otimes\cdots\otimes e^{-A_{k,\alpha}t}, and eA1,βteAk,βte^{-A_{1,\beta}t}\otimes\cdots\otimes e^{-A_{k,\beta}t} are all equal. This ensures that the proof given in §5.1 applies word by word. ∎

We also get the following coupled version of Corollary 1.2; see §A.3 for the definition of δmcoupled\delta^{\textup{coupled}}_{m}.

Corollary 5.12.

Suppose that Aut0(X)\mathrm{Aut}_{0}(X) is trivial. Then δmcoupled>1\delta^{\textup{coupled}}_{m}>1 if and only if

Ding((𝒳i,i)i=1k)+i=1kChowm(𝒳i,i)>0\mathrm{Ding}\left((\mathcal{X}_{i},\mathcal{L}_{i})_{i=1}^{k}\right)+\sum_{i=1}^{k}\mathrm{Chow}_{m}(\mathcal{X}_{i},\mathcal{L}_{i})>0

holds for any kk-tuple of very ample test configurations (𝒳1,1),,(𝒳k,k)(\mathcal{X}_{1},\mathcal{L}_{1}),\dots,(\mathcal{X}_{k},\mathcal{L}_{k}) for (X,L1),,(X,Lk)(X,L_{1}),\dots,(X,L_{k}) respectively, each of exponent mm and none of which are trivial.

Proof.

Exactly the same as that of Corollary 1.2, by using [rtz, Theorem A.12] and [tak19, Proposition 3.3]; recall from §3.4 that (𝒳1,1),,(𝒳k,k)(\mathcal{X}_{1},\mathcal{L}_{1}),\dots,(\mathcal{X}_{k},\mathcal{L}_{k}) are all nontrivial if and only if the test configuration (𝒴,𝒴)(\mathcal{Y},\mathcal{L}_{\mathcal{Y}}) generated by their \mathbb{C}^{*}-actions is. ∎

Appendix A The δm\delta_{m}-invariant of Fujita–Odaka and its extensions

A.1. Original invariant δm\delta_{m}

We recall the invariants δm\delta_{m} and δ\delta, following [bj20, fo18]. We start by quickly recalling the log canonical thresholds, following the exposition in [Fujita18, §2.3].

Definition A.1.

Let YY be a normal variety and Δ\Delta be an \mathbb{R}-divisor on YY such that KY+ΔK_{Y}+\Delta is \mathbb{R}-Cartier, where Δ\Delta is not necessarily effective. The pair (Y,Δ)(Y,\Delta) is said to be sub log canonical if a(E,Y,Δ)1a(E,Y,\Delta)\geq-1 holds for any proper birational morphism σ:Y~Y\sigma:\tilde{Y}\to Y with Y~\tilde{Y} normal and any prime divisor EE on Y~\tilde{Y}, where a(E,Y,Δ):=ordE(KY~σ(KY+Δ))a(E,Y,\Delta):=\mathrm{ord}_{E}(K_{\tilde{Y}}-\sigma^{*}(K_{Y}+\Delta)). We say that (Y,Δ)(Y,\Delta) is log canonical if it is sub log canonical and Δ\Delta is effective.

For the pair (Y,Δ)(Y,\Delta) as above, and for a nonzero effective \mathbb{R}-Cartier divisor BB on YY, the log canonical threshold of BB with respect to (Y,Δ)(Y,\Delta) is defined as

lct(Y,Δ;B):=sup{c(Y,Δ+cB) is sub log canonical.};\mathrm{lct}(Y,\Delta;B):=\sup\{c\in\mathbb{R}\mid(Y,\Delta+cB)\text{ is sub log canonical}.\};

if (Y,Δ+cB)(Y,\Delta+cB) is not sub log canonical for any cc\in\mathbb{R}, we set lct(Y,Δ;B)=\mathrm{lct}(Y,\Delta;B)=-\infty.

If YY is further assumed to be \mathbb{Q}-Fano, i.e. KY-K_{Y} is an ample \mathbb{Q}-Cartier divisor and YY has at worst log terminal singularities, and DD is an effective \mathbb{Q}-Cartier divisor on YY, we further define the log canonical threshold of YY along DD as

lct(Y;D):=sup{c0(Y,cD) is log canonical.}.\mathrm{lct}(Y;D):=\sup\{c\in\mathbb{R}_{\geq 0}\mid(Y,cD)\text{ is log canonical}.\}.

While it is good to have the above general definitions, in this appendix we apply them exclusively to the case when the variety in question is a smooth Fano variety. As in the body of the text, XX stands for a Fano manifold of complex dimension nn.

Definition A.2.

Let XX be a Fano manifold and mm be a positive integer, and define Nm:=dimH0(X,mKX)N_{m}:=\dim_{\mathbb{C}}H^{0}(X,-mK_{X}). We say that a divisor DD on XX is of mm-basis type if there exists a basis {si}i=1Nm\{s_{i}\}_{i=1}^{N_{m}} for H0(X,mKX)H^{0}(X,-mK_{X}) such that

D=1mNmi=1Nmdiv(si).D=\frac{1}{mN_{m}}\sum_{i=1}^{N_{m}}\mathrm{div}(s_{i}).
Definition A.3.

Let XX be a Fano manifold and mm be a positive integer. The δm\delta_{m}-invariant of Fujita–Odaka is defined by

δm:=infDlct(X;D)\delta_{m}:=\inf_{D}\mathrm{lct}(X;D)

where the infimum is taken over all divisors on XX that are of mm-basis type. The δ\delta-invariant is defined by

δ:=lim supmδm.\delta:=\limsup_{m\to\infty}\delta_{m}.

Blum–Jonsson [bj20] provided a valuative formulation of the above invariants, which we now recall.

Definition A.4.

FF is said to be a prime divisor over XX if there exists a normal variety YY and a proper birational morphism σ:YX\sigma:Y\to X such that FF is a reduced irreducible Weil divisor on YY.

Writing

AX(F):=1+ordF(KY/X)A_{X}(F):=1+\mathrm{ord}_{F}(K_{Y/X})

for the log discrepancy of XX along a prime divisor FF over XX, where KY/X=KYσKXK_{Y/X}=K_{Y}-\sigma^{*}K_{X} is the relative canonical bundle, it is well-known (see e.g. [bj20, §1.8] and [bhj1, §1.5]) that we can write

lct(X;D)=infFAX(F)ordF(D),\mathrm{lct}(X;D)=\inf_{F}\frac{A_{X}(F)}{\mathrm{ord}_{F}(D)},

where the infimum is taken over all prime divisors over XX with ordF(D)>0\mathrm{ord}_{F}(D)>0.

For a prime divisor FF over XX, with FYF\subset Y and a birational morphism σ:YX\sigma:Y\to X, we write

Sm(F):=1mNmj=1dimH0(Y,σ(mKX)jF),S_{m}(F):=\frac{1}{mN_{m}}\sum_{j=1}^{\infty}\dim_{\mathbb{C}}H^{0}(Y,\sigma^{*}(-mK_{X})-jF),

where we observe that all but finitely many summands are zero. A formula in [fo18, proof of Lemma 2.2] shows

Sm(F)=supDordF(D)S_{m}(F)=\sup_{D}\mathrm{ord}_{F}(D)

where the supremum is taken over all mm-basis type divisors; in fact, [fo18, proof of Lemma 2.2] also shows that the supremum is attained by an mm-basis type divisor DFD_{F} which is compatible with the filtration of H0(X,mKX)H^{0}(X,-mK_{X}) defined by FF. Moreover [fo18, proof of Theorem 2.1] shows

limmSm(F)=1Vol(KX)0+Vol(KXxF)dx=:S(F),\lim_{m\to\infty}S_{m}(F)=\frac{1}{\mathrm{Vol}(-K_{X})}\int_{0}^{+\infty}\mathrm{Vol}(-K_{X}-xF)dx=:S(F),

where Vol(KXxF):=n!lim supmdimH0(Y,σ(mKX)xF)/mn\mathrm{Vol}(-K_{X}-xF):=n!\limsup_{m\to\infty}\dim_{\mathbb{C}}H^{0}(Y,\sigma^{*}(-mK_{X})-xF)/m^{n}. Blum–Jonsson [bj20, Proposition 4.3] proved that

δm=infFAX(F)Sm(F),\delta_{m}=\inf_{F}\frac{A_{X}(F)}{S_{m}(F)},

where the infimum is taken over all prime divisors over XX. They moreover proved [bj20, Theorem 4.4] that the limit limmδm\lim_{m\to\infty}\delta_{m} exists, and that

δ=limmδm=infFAX(F)S(F),\delta=\lim_{m\to\infty}\delta_{m}=\inf_{F}\frac{A_{X}(F)}{S(F)},

where the infimum is again taken over all prime divisors over XX.

A.2. Kähler–Ricci gg-solitons version δmg\delta_{m}^{g}

The invariants for the Kähler–Ricci gg-solitons were defined by Rubinstein–Tian–Zhang [rtz, Definition 6.4].

Definition A.5.

Suppose that an algebraic torus TT^{\mathbb{C}} acts on XX, which induces a weight decomposition

H0(X,mKX)=λPmRm,λH^{0}(X,-mK_{X})=\bigoplus_{\lambda\in P_{m}}R_{m,\lambda}

as in §2.3, where PmP_{m} is the character lattice and TT^{\mathbb{C}} acts by the character λ\lambda on Rm,λR_{m,\lambda}. We say that DD is an (m,g)(m,g)-basis divisor if there exists a basis {sα(λ)}α=1Nm,λ\{s_{\alpha}^{(\lambda)}\}_{\alpha=1}^{N_{m,\lambda}} for each Rm,λR_{m,\lambda} such that

D=1mNmgm¯λPmα=1Nm,λg(λ/m){sα(λ)=0}.D=\frac{1}{mN_{m}\overline{g_{m}}}\sum_{\lambda\in P_{m}}\sum_{\alpha=1}^{N_{m,\lambda}}g(\lambda/m)\{s_{\alpha}^{(\lambda)}=0\}.

The invariant δmg\delta_{m}^{g} is defined by

δmg:=infDlct(X;D)\delta_{m}^{g}:=\inf_{D}\mathrm{lct}(X;D)

where the infimum is over all (m,g)(m,g)-basis divisors. The invariant δg\delta^{g} is defined by

δg:=lim supmδmg.\delta^{g}:=\limsup_{m\to\infty}\delta^{g}_{m}.

Analogues of the results in §A.1 are also given in [rtz]. They define

Smg(F):=1mNmgm¯λPma=1g(λ/m)dimordFaRm,λS^{g}_{m}(F):=\frac{1}{mN_{m}\overline{g_{m}}}\sum_{\lambda\in P_{m}}\sum_{a=1}^{\infty}g(\lambda/m)\dim_{\mathbb{C}}\mathcal{F}^{\geq a}_{\mathrm{ord}_{F}}R_{m,\lambda}

for a TT^{\mathbb{C}}-invariant prime divisor over XX, where ordFaRm,λ\mathcal{F}^{\geq a}_{\mathrm{ord}_{F}}R_{m,\lambda} is the subspace of Rm,λR_{m,\lambda} consisting of sections that vanish along FF with order at least aa. We then have, by [rtz, Lemma 6.5],

δmg=infFAX(F)Smg(F)\delta^{g}_{m}=\inf_{F}\frac{A_{X}(F)}{S^{g}_{m}(F)}

where the infimum is over all TT^{\mathbb{C}}-invariant prime divisors over XX.

Setting

Volg(KXxF):=lim supmn!mnλPmg(λ/m)dimordFxRm,λ\mathrm{Vol}^{g}(-K_{X}-xF):=\limsup_{m\to\infty}\frac{n!}{m^{n}}\sum_{\lambda\in P_{m}}g(\lambda/m)\dim_{\mathbb{C}}\mathcal{F}^{\geq x}_{\mathrm{ord}_{F}}R_{m,\lambda}

and defining

Sg(F):=0+Volg(KXxF)𝑑x,S^{g}(F):=\int_{0}^{+\infty}\mathrm{Vol}^{g}(-K_{X}-xF)dx,

we have, by [rtz, Proposition 6.14],

δg=limmδmg=infFAX(F)Sg(F)\delta^{g}=\lim_{m\to\infty}\delta^{g}_{m}=\inf_{F}\frac{A_{X}(F)}{S^{g}(F)}

where the infimum is over all TT^{\mathbb{C}}-invariant prime divisors over XX.

K. Zhang [Zhang21, Remark 5.3] recently proved that δg>1\delta^{g}>1 implies the existence of Kähler–Ricci gg-solitons.

A.3. Coupled Kähler–Einstein version δmcoupled\delta^{\mathrm{coupled}}_{m}

The coupled version of the δm\delta_{m}-invariant was defined by Rubinstein–Tian–Zhang [rtz, Definition A.2] as follows.

Definition A.6.

The coupled δm\delta_{m}-invariant is defined by

δmcoupled:=infD1,,Dklct(X;i=1kDi),\delta^{\text{coupled}}_{m}:=\inf_{D_{1},\dots,D_{k}}\mathrm{lct}\left(X;\sum_{i=1}^{k}D_{i}\right),

where the infimum is over a kk-tuple of divisors (D1,,Dk)(D_{1},\dots,D_{k}) such that DiLiD_{i}\sim_{\mathbb{Q}}L_{i} is an mm-basis divisor for each i=1,,ki=1,\dots,k. The coupled δ\delta-invariant is defined by

δcoupled:=lim supmδmcoupled.\delta^{\text{coupled}}:=\limsup_{m\to\infty}\delta^{\text{coupled}}_{m}.

Rubinstein–Tian–Zhang [rtz] considers a more general definition in which DiD_{i}’s are allowed to have different exponents, i.e. DiLiD_{i}\sim_{\mathbb{Q}}L_{i} being an mim_{i}-basis divisor for possibly distinct integers m1,,mkm_{1},\dots,m_{k}, but it is not necessary for the results in this paper.

The valuative formula for δm\delta_{m} and δ\delta naturally generalises to the coupled case. For a prime divisor FF over XX, with a proper birational morphism σ:YX\sigma:Y\to X, we define

Sm(Li;F):=1mh0(X,mLi)j=1dimH0(Y,σ(mLi)jF).S_{m}(L_{i};F):=\frac{1}{mh^{0}(X,mL_{i})}\sum_{j=1}^{\infty}\dim_{\mathbb{C}}H^{0}(Y,\sigma^{*}(-mL_{i})-jF).

We have [rtz, Lemma A.3]

δmcoupled=infFAX(F)i=1kSm(Li;F)\delta^{\text{coupled}}_{m}=\inf_{F}\frac{A_{X}(F)}{\sum_{i=1}^{k}S_{m}(L_{i};F)}

where the infimum runs over all prime divisors over XX; the proof is exactly the same as above, by noting

lct(X,i=1kDi)=infFAX(F)i=1kordF(Di).\mathrm{lct}\left(X,\sum_{i=1}^{k}D_{i}\right)=\inf_{F}\frac{A_{X}(F)}{\sum_{i=1}^{k}\mathrm{ord}_{F}(D_{i})}.

Likewise, repeating the argument in [bj20, Proof of Theorem 4.4] immediately implies that the limit δcoupled:=limmδmcoupled\delta^{\text{coupled}}:=\lim_{m\to\infty}\delta^{\text{coupled}}_{m} exists, and equals

δcoupled=infFAX(F)i=1kS(Li;F)\delta^{\text{coupled}}=\inf_{F}\frac{A_{X}(F)}{\sum_{i=1}^{k}S(L_{i};F)}

where the infimum is over all prime divisors over XX and

S(Li;F):=1Vol(Li)0+Vol(LixF)𝑑x.S(L_{i};F):=\frac{1}{\mathrm{Vol}(L_{i})}\int_{0}^{+\infty}\mathrm{Vol}(L_{i}-xF)dx.

It was proved by K. Zhang [Zhang21, Remark 5.3] that δcoupled>1\delta^{\text{coupled}}>1 implies the existence of coupled Kähler–Einstein metrics.

References