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

Quantitative injectivity of the Fubini–Study map

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

We prove a quantitative version of the injectivity of the Fubini–Study map that is polynomial in the exponent of the ample line bundle, and correct the arguments in the author’s previous papers [yhhilb, yhextremal].

1. Introduction

Suppose that we have a polarised smooth projective variety (X,L)(X,L) over \mathbb{C} of complex dimension nn and a projective embedding ιk:X(H0(X,Lk))\iota_{k}:X\hookrightarrow\mathbb{P}(H^{0}(X,L^{k})^{\vee}), where LL is an ample line bundle over XX and kk is a large enough integers such that LkL^{k} is very ample. We may moreover assume, by replacing LL by a higher tensor power if necessary, that the automorphism group Aut0(X,L)\mathrm{Aut}_{0}(X,L) acts on (X,L)(X,L) via the ambient linear action of SL(H0(X,Lk))SL(H^{0}(X,L^{k})) fixing ιk(X)\iota_{k}(X). We fix a hermitian metric hh on LL which defines a Kähler metric ωh>0\omega_{h}>0 on XX, where ωhc1(L)\omega_{h}\in c_{1}(L). This in turn defines a positive definite hermitian form Hilb(hk)\mathrm{Hilb}(h^{k}) on H0(X,Lk)H^{0}(X,L^{k}) defined (see [donproj2]) as

Hilb(hk)(s1,s2):=NkVXhk(s1,s2)ωhnn!\mathrm{Hilb}(h^{k})(s_{1},s_{2}):=\frac{N_{k}}{V}\int_{X}h^{k}(s_{1},s_{2})\frac{\omega^{n}_{h}}{n!}

for s1,s2H0(X,Lk)s_{1},s_{2}\in H^{0}(X,L^{k}), where we wrote

Nk:=dimH0(X,Lk),V:=Xc1(L)n/n!.N_{k}:=\dim_{\mathbb{C}}H^{0}(X,L^{k}),\quad V:=\int_{X}c_{1}(L)^{n}/n!.

Choosing a Hilb(hk)\mathrm{Hilb}(h^{k})-orthonormal basis {si}i=1Nk\{s_{i}\}_{i=1}^{N_{k}}, we identify (H0(X,Lk))\mathbb{P}(H^{0}(X,L^{k})^{\vee}) with Nk1\mathbb{P}^{N_{k}-1}, and also identify the set k\mathcal{B}_{k} of all positive definite hermitian forms on H0(X,Lk)H^{0}(X,L^{k}) with positive definite Nk×NkN_{k}\times N_{k} hermitian matrices, noting that Hilb(hk)\mathrm{Hilb}(h^{k}) is the identity matrix with respect to this basis. The set k\mathcal{B}_{k} in turn can be identified with the symmetric space GL(Nk,)/U(Nk)GL(N_{k},\mathbb{C})/U(N_{k}). The Fubini–Study map FSk\mathrm{FS}_{k} associates to AkA\in\mathcal{B}_{k} a hermitian metric FSk(A)\mathrm{FS}_{k}(A), called the Fubini–Study metric, on LkL^{k}. Recall that FSk(A)\mathrm{FS}_{k}(A) is defined as a unique hermitian metric on LkL^{k} which satisfies

i=1NkFSk(A)(siA,siA)=1\sum_{i=1}^{N_{k}}\mathrm{FS}_{k}(A)(s^{A}_{i},s^{A}_{i})=1 (1)

over XX, for an AA-orthonormal basis {siA}i=1Nk\{s^{A}_{i}\}_{i=1}^{N_{k}} for H0(X,Lk)H^{0}(X,L^{k}). When we write Hk:=Hilb(hk)H_{k}:=\mathrm{Hilb}(h^{k}), the equation above immediately implies

FSk(A)=(i,j=1NkAij1FSk(Hk)(si,sj))1FSk(Hk),\mathrm{FS}_{k}(A)=\left(\sum_{i,j=1}^{N_{k}}A^{-1}_{ij}\mathrm{FS}_{k}(H_{k})(s_{i},s_{j})\right)^{-1}\mathrm{FS}_{k}(H_{k}),

by noting that {jAij1/2sj}i=1Nk\{\sum_{j}A^{-1/2}_{ij}s_{j}\}_{i=1}^{N_{k}} is an AA-orthonormal basis. For A,BkA,B\in\mathcal{B}_{k}, we define a smooth real function fk(A,B;Hk)f_{k}(A,B;H_{k}) on XX defined as

fk(A,B;Hk)\displaystyle f_{k}(A,B;H_{k}) :=FSk(Hk)FSk(A)FSk(Hk)FSk(B)\displaystyle:=\frac{\mathrm{FS}_{k}(H_{k})}{\mathrm{FS}_{k}(A)}-\frac{\mathrm{FS}_{k}(H_{k})}{\mathrm{FS}_{k}(B)}
=i,j=1Nk(Aij1Bij1)FSk(Hk)(si,sj).\displaystyle=\sum_{i,j=1}^{N_{k}}(A^{-1}_{ij}-B^{-1}_{ij})\mathrm{FS}_{k}(H_{k})(s_{i},s_{j}).

Given A,BkA,B\in\mathcal{B}_{k}, it is natural to conjecture that AA is close to BB in k\mathcal{B}_{k} if and only if fk(A,B;Hk)f_{k}(A,B;H_{k}) is close to zero in some appropriate norm. Indeed, Lempert [Lem21, Theorem 1.1] proved that the Fubini–Study map is injective, which is equivalent to saying that A=BA=B if and only if fk(A,B;Hk)0f_{k}(A,B;H_{k})\equiv 0, and also proved that its image is not closed [Lem21, Theorem 1.1, Example 6.2, Theorem 6.4]. In this paper, we consider the following quantitative version of injectivity.

Problem 1.1.

Can we bound an appropriately defined distance between AA and BB in k\mathcal{B}_{k} by Cklfk(A,B;Hk)Ck^{l}\|f_{k}(A,B;H_{k})\|, where C,l>0C,l>0 are some constants which depend only on hh?

The problem here is that the constant appearing in the estimate is at most polynomial in kk, as well as the right choice of the norm for fk(A,B;Hk)f_{k}(A,B;H_{k}). In [yhhilb, Lemma 3.1], the author claimed that this problem can be solved with the C0C^{0}-norm, but its proof does not hold since it is based on an erroneous claim that the Hilbert map is surjective [yhhilb, Theorem 1.1, Proposition 2.15]. Indeed, a counterexample was provided by Lempert in an email correspondence with the author (see [Finski22, Proposition 4.16] and also [Finski22s, Proposition 3.6]), and there is also a counterexample to the surjectivity of the Hilbert map by Sun [Sun22, §3, Example]. Both these counterexamples involve basis elements that are redundant in terms of the global generation of the line bundle (see the definition of a rather ample subspace in [Lem21, §6]). Sun [Sun22, Theorems 1.2 and 1.3] also described the closure of the image of the Hilbert map.

It is still plausible, however, that Problem 1.1 can be solved affirmatively once the right norm is chosen. The main result of this paper is the following.

Theorem 1.2.

There exist constants k0k_{0}\in\mathbb{N} and Ch>0C_{h}>0, depending only on hh, such that we have

A1B1HS(Hk)2Chknfk(A,B;Hk)W2,2(ωh)2\|A^{-1}-B^{-1}\|^{2}_{\mathrm{HS}(H_{k})}\leq C_{h}k^{n}\|f_{k}(A,B;H_{k})\|^{2}_{W^{2,2}(\omega_{h})}

for any A,BkA,B\in\mathcal{B}_{k} and for any kk0k\geq k_{0}, where HS(Hk)\|\cdot\|_{\mathrm{HS}(H_{k})} is the Hilbert–Schmidt norm with respect to Hk=Hilb(hk)H_{k}=\mathrm{Hilb}(h^{k}) and W2,2(ωh)\|\cdot\|_{W^{2,2}(\omega_{h})} is the Sobolev norm with respect to ωh\omega_{h}.

We stress here that A,BkA,B\in\mathcal{B}_{k} are regarded as matrices with respect to the HkH_{k}-orthonormal basis, when we consider the inverse matrices above.

The choice of the reference metric HkH_{k} is important in Theorem 1.2; see (6) for the formula with respect to a different reference. Indeed, the counterexamples of Lempert and Sun mentioned above deal with the case when the reference metric is close to being degenerate.

Acknowledgements. The author thanks Siarhei Finski, László Lempert, and Jingzhou Sun for very helpful discussions and immensely valuable examples. This research is partially supported by JSPS KAKENHI (Grant-in-Aid for Scientific Research (C)), Grant Number JP23K03120, and JSPS KAKENHI (Grant-in-Aid for Scientific Research (B)) Grant Number JP24K00524.

2. Proof of Theorem 1.2

In what follows, we write ωh,kkc1(L)\omega_{h,k}\in kc_{1}(L) for the Kähler metric on XX associated to FS(Hk)\mathrm{FS}(H_{k}), and ω~h,k\tilde{\omega}_{h,k} for the Fubini–Study metric on 𝒪Nk1(1)\mathcal{O}_{\mathbb{P}^{N_{k}-1}}(1) over Nk1=(H0(X,Lk))\mathbb{P}^{N_{k}-1}=\mathbb{P}(H^{0}(X,L^{k})^{\vee}) defined by HkH_{k}. Recall ωh,k=ιkω~h,k\omega_{h,k}=\iota^{*}_{k}\tilde{\omega}_{h,k}, and that we have

ωh,k=kωh12π¯logρk(ωh)\omega_{h,k}=k\omega_{h}-\frac{\sqrt{-1}}{2\pi}\partial\bar{\partial}\log\rho_{k}(\omega_{h})

by a result due to Rawnsley [rawnsley], where we wrote ρk(ωh)\rho_{k}(\omega_{h}) for the kk-th Bergman function with respect to ωh\omega_{h}. We also use the re-scaled version ρ¯k(ωh):=VNkρk(ωh)\bar{\rho}_{k}(\omega_{h}):=\frac{V}{N_{k}}\rho_{k}(\omega_{h}). We recall the asymptotic expansion of the Bergman function (see e.g. [mm])

ρk(ωh)=kn+O(kn1),orρ¯k(ωh)=1+O(1/k),\rho_{k}(\omega_{h})=k^{n}+O(k^{n-1}),\quad\text{or}\quad\bar{\rho}_{k}(\omega_{h})=1+O(1/k), (2)

where Nk=Vkn+O(kn1)N_{k}=Vk^{n}+O(k^{n-1}) by the asymptotic Riemann–Roch theorem. We also recall that the expansion above holds in CmC^{m}-norm for any mm\in\mathbb{N}. We thus have

ωh,k=kωh+O(1/k).\omega_{h,k}=k\omega_{h}+O(1/k).

We start with the following lemma concerning the second fundamental form of the projective embedding.

Lemma 2.1.

Suppose Hk=Hilb(hk)H_{k}=\mathrm{Hilb}(h^{k}), and let Ah,kA_{h,k} be the second fundamental form of TXTX in ιTNk1\iota^{*}T\mathbb{P}^{N_{k}-1} with respect to ω~h,k\tilde{\omega}_{h,k}. Then, there exist k0k_{0}\in\mathbb{N} and a constant λmin(h)=λmin>0\lambda_{\mathrm{min}}(h)=\lambda_{\mathrm{min}}>0 depending only on hh, such that the minimum eigenvalue of Ah,kAh,k-A_{h,k}^{*}\wedge A_{h,k} with respect to a gh,kg_{h,k}-orthonormal basis can be bounded below by λmin(h)\lambda_{\mathrm{min}}(h) for all kk0k\geq k_{0} and all xXx\in X. In particular, Ah,kA_{h,k} is nondegenerate everywhere on XX for all large enough kk.

Proof.

We recall that the curvature of the Fubini–Study metrics on the ambient projective space and XX can be related by

Ah,kAh,k=πT(F~h,k|TX)Fh,k-A_{h,k}^{*}\wedge A_{h,k}=\pi_{T}\circ(\tilde{F}_{h,k}|_{TX})-F_{h,k}

as given in [grihar, page 78] or [PS04, equation (5.27)], where we wrote F~h,k\tilde{F}_{h,k} for the curvature form of ω~h,k\tilde{\omega}_{h,k} on the ambient projective space and Fh,kF_{h,k} for that of ωh,k\omega_{h,k} on XX. The adjoint above is with respect to ω~h,k\tilde{\omega}_{h,k}. We recall that the curvature tensor of the Fubini–Study metric on Nk1\mathbb{P}^{N_{k}-1} can be written as

(F~h,k)ij¯lm¯=(g~h,k)ij¯(g~h,k)lm¯+(g~h,k)im¯(g~h,k)lj¯(\tilde{F}_{h,k})_{i\bar{j}l\bar{m}}=(\tilde{g}_{h,k})_{i\bar{j}}(\tilde{g}_{h,k})_{l\bar{m}}+(\tilde{g}_{h,k})_{i\bar{m}}(\tilde{g}_{h,k})_{l\bar{j}}

where g~h,k\tilde{g}_{h,k} is the metric tensor of ω~h,k\tilde{\omega}_{h,k} written with respect to its orthonormal basis.

By the Bergman kernel expansion (2), we find that the restriction of g~h,k\tilde{g}_{h,k} to TXTX satisfies the asymptotic expansion

gh,k=kgh+O(1/k),g_{h,k}=kg_{h}+O(1/k),

where ghg_{h} stands for the Kähler metric associated to hh. Thus we have

(πT(F~h,k|TX)Fh,k)ijlm¯=δijδlm¯+δljδim¯(Fh)ijlm¯+O(1/k){{(\pi_{T}\circ(\tilde{F}_{h,k}|_{TX})-F_{h,k})_{i}}^{j}}_{l\bar{m}}=\delta_{i}^{j}\delta_{l\bar{m}}+\delta_{l}^{j}\delta_{i\bar{m}}-{{(F_{h})_{i}}^{j}}_{l\bar{m}}+O(1/k)

with respect to any kghkg_{h}-orthonormal basis for TXTX (note that the curvature tensor FhF_{h} is of order 1 with respect to a kghkg_{h}-orthonormal basis, which is of order k1k^{-1} with respect to a ghg_{h}-orthonormal basis). Since this agrees with Ah,kAh,k-A_{h,k}^{*}\wedge A_{h,k}, the tensor above is positive semidefinite.

Suppose that Ah,kAh,k-A_{h,k}^{*}\wedge A_{h,k} is degenerate at xXx\in X. We then perturb hh locally around xx so that the perturbation hϵh_{\epsilon} satisfies

h(x)=hϵ(x),g(x)=gϵ(x)h(x)=h_{\epsilon}(x),\quad g(x)=g_{\epsilon}(x)

and the curvature tensor is perturbed by ϵδij(kgh)lm¯-\epsilon\delta_{i}^{j}(kg_{h})_{l\bar{m}} at xx, which is of order kϵk\epsilon locally at xx, and such that Hilb(hϵk)=Hilb(hk)+O(ϵ2)\mathrm{Hilb}(h_{\epsilon}^{k})=\mathrm{Hilb}(h^{k})+O(\epsilon^{2}). This is possible by considering the perturbation of the Kähler potential by ϵ|zi|2|zl|2\epsilon|z_{i}|^{2}|z_{l}|^{2} where (z1,,zn)(z_{1},\dots,z_{n}) are holomorphic normal coordinates around xx, multiplied by a cutoff function which decreases very rapidly away from xx so that the perturbation introduced in the integral inside Hilb\mathrm{Hilb} is small.

Applying the argument above to hϵh_{\epsilon} and writing with respect to a ghg_{h}-orthonormal basis, we find

Ah,k,ϵAh,k,ϵ=πT(F~h,k,ϵ|TX)Fh,k,ϵ\displaystyle-A_{h,k,\epsilon}^{*}\wedge A_{h,k,\epsilon}=\pi_{T}\circ(\tilde{F}_{h,k,\epsilon}|_{TX})-F_{h,k,\epsilon} =Ah,kAh,kϵδij(kgh)lm¯+O(ϵ2)+O(1/k)\displaystyle=-A_{h,k}^{*}\wedge A_{h,k}-\epsilon\delta_{i}^{j}(kg_{h})_{l\bar{m}}+O(\epsilon^{2})+O(1/k)
=Ah,kAh,kϵδijδlm¯+O(ϵ2)+O(1/k),\displaystyle=-A_{h,k}^{*}\wedge A_{h,k}-\epsilon\delta_{i}^{j}\delta_{l\bar{m}}+O(\epsilon^{2})+O(1/k),

where F~h,k,ϵ\tilde{F}_{h,k,\epsilon} is the curvature of the Fubini–Study metric on Nk1\mathbb{P}^{N_{k}-1} with respect to Hilb(hϵk)\mathrm{Hilb}(h^{k}_{\epsilon}), Fh,k,ϵF_{h,k,\epsilon} is the one with respect to its restriction to TXTX, and Ah,k,ϵA_{h,k,\epsilon} is the second fundamental form that is associated to it. This yields a contradiction for a sufficiently small ϵ>0\epsilon>0 and for all large enough kk, if we assume that Ah,kAh,k-A_{h,k}^{*}\wedge A_{h,k} is degenerate at xXx\in X.

The second claim is a consequence of the fact that, with respect to a kghkg_{h}-orthonormal basis, πT(F~h,k|TX)Fh,k\pi_{T}\circ(\tilde{F}_{h,k}|_{TX})-F_{h,k} converges to a positive definite form on XX by the previous argument and the Bergman kernel expansion gh,k=kgh+O(1/k)g_{h,k}=kg_{h}+O(1/k), together with the compactness of XX. ∎

Given A,BkA,B\in\mathcal{B}_{k}, we define a (not necessarily positive) hermitian matrix Λ\Lambda by

Λ:=A1B1,\Lambda:=A^{-1}-B^{-1},

and also define

f(Λ;Hk):=f(A,B;Hk).f(\Lambda;H_{k}):=f(A,B;H_{k}).

We write ξΛ\xi_{\Lambda} for the holomorphic vector field on Nk1\mathbb{P}^{N_{k}-1} generated by the linear action of Λ\Lambda. We recall that, when [Z1::ZNk][Z_{1}:\cdots:Z_{N_{k}}] is a set of homogeneous coordinates for Nk1\mathbb{P}^{N_{k}-1} consisting of HkH_{k}-orthonormal basis for H0(X,Lk)H^{0}(X,L^{k}), the Hamiltonian function for ξΛ\xi_{\Lambda} with respect to ω~h,k\tilde{\omega}_{h,k} is given by

f~(Λ;Hk):=i,j=1NkΛijZiZ¯jl=1Nk|Zl|2,\tilde{f}(\Lambda;H_{k}):=\sum_{i,j=1}^{N_{k}}\Lambda_{ij}\frac{Z_{i}\bar{Z}_{j}}{\sum_{l=1}^{N_{k}}|Z_{l}|^{2}},

which is a well-defined smooth function on Nk1\mathbb{P}^{N_{k}-1}. By the definition of FSk(Hk)\mathrm{FS}_{k}(H_{k}), we have

f(Λ;Hk)=ιkf~(Λ;Hk).f(\Lambda;H_{k})=\iota^{*}_{k}\tilde{f}(\Lambda;H_{k}).

We first consider the case when the trace of Λ\Lambda is zero, i.e. when Λ𝔰𝔩(Nk,)\Lambda\in\mathfrak{sl}(N_{k},\mathbb{C}). Following [PS04, §5], we consider a holomorphic vector bundle ιTNk1\iota^{*}T\mathbb{P}^{N_{k}-1} over XX, which we decompose as

ιTNk1=TX𝒩,\iota^{*}T\mathbb{P}^{N_{k}-1}=TX\oplus\mathcal{N}, (3)

where the direct sum is orthogonal with respect to the Fubini–Study metric on ιTNk1\iota^{*}T\mathbb{P}^{N_{k}-1} with respect to HkH_{k}. We write πT\pi_{T} and π𝒩\pi_{\mathcal{N}} for the projection to TXTX and 𝒩\mathcal{N}. By the defining equation

ι(ξΛ)ω~h,k=df~(Λ;Hk)\iota(\xi_{\Lambda})\tilde{\omega}_{h,k}=-d\tilde{f}(\Lambda;H_{k})

for the Hamiltonian function for ξΛ\xi_{\Lambda}, together with ωh,k=ιkω~h,k\omega_{h,k}=\iota^{*}_{k}\tilde{\omega}_{h,k} and (2), we find that πTξΛ\pi_{T}\xi_{\Lambda} is ωh,k\omega_{h,k}-metric dual to dfk(Λ;Hk)df_{k}(\Lambda;H_{k}) as pointed out in [Fine10, Lemma 20].

Recall that we have a faithful representation Aut0(X,L)SL(Nk,)\mathrm{Aut}_{0}(X,L)\hookrightarrow SL(N_{k},\mathbb{C}) by the linearisation of the action of Aut0(X,L)\mathrm{Aut}_{0}(X,L), which is unique up to an overall multiplicative constant (by replacing LL by a higher tensor power if necessary). We further decompose Λ=α+β\Lambda=\alpha+\beta according to 𝔰𝔩(Nk,)=𝔞𝔲𝔱(X,L)𝔞𝔲𝔱(X,L)\mathfrak{sl}(N_{k},\mathbb{C})=\mathfrak{aut}(X,L)\oplus\mathfrak{aut}(X,L)^{\perp}, where the orthogonality is defined with respect to the L2L^{2}-metric defined by ωh,k\omega_{h,k} which is the Fubini–Study metric on ιTNk1\iota^{*}T\mathbb{P}^{N_{k}-1} with respect to HkH_{k}. We then have

ξΛL2(ωh,k)2=ξα+βL2(ωh,k)2=ξαL2(ωh,k)2+ξβL2(ωh,k)2\|\xi_{\Lambda}\|^{2}_{L^{2}(\omega_{h,k})}=\|\xi_{\alpha+\beta}\|^{2}_{L^{2}(\omega_{h,k})}=\|\xi_{\alpha}\|^{2}_{L^{2}(\omega_{h,k})}+\|\xi_{\beta}\|^{2}_{L^{2}(\omega_{h,k})}

by definition. Note also that, according to the orthogonal decomposition (3), we have

ξβL2(ωh,k)2=πTξβL2(ωh,k)2+π𝒩ξβL2(ωh,k)2\|\xi_{\beta}\|^{2}_{L^{2}(\omega_{h,k})}=\|\pi_{T}\xi_{\beta}\|^{2}_{L^{2}(\omega_{h,k})}+\|\pi_{\mathcal{N}}\xi_{\beta}\|^{2}_{L^{2}(\omega_{h,k})}

because of the pointwise orthogonality, and also note ξα=πTξα\xi_{\alpha}=\pi_{T}\xi_{\alpha}.

Lemma 2.2.

Suppose ιk:X(H0(X,Lk))\iota_{k}:X\hookrightarrow\mathbb{P}(H^{0}(X,L^{k})^{\vee}), and that we consider the Fubini–Study metric FS(Hk)\mathrm{FS}(H_{k}) where Hk=Hilb(hk)H_{k}=\mathrm{Hilb}(h^{k}). Then there exists C1>0C_{1}>0 depending only on hh such that

π𝒩ξβL2(ωh,k)2C1¯(πTξβ)L2(ωh,k)2\|\pi_{\mathcal{N}}\xi_{\beta}\|^{2}_{L^{2}(\omega_{h,k})}\leq C_{1}\|\bar{\partial}(\pi_{T}\xi_{\beta})\|^{2}_{L^{2}(\omega_{h,k})}

holds for any β𝔞𝔲𝔱(X,L)\beta\in\mathfrak{aut}(X,L)^{\perp} and for all large enough kk.

Proof.

We prove

π𝒩ξβL2(ωh,k)2C1¯(π𝒩ξβ)L2(ωh,k)2=C1¯(πTξβ)L2(ωh,k)2.\|\pi_{\mathcal{N}}\xi_{\beta}\|^{2}_{L^{2}(\omega_{h,k})}\leq C_{1}\|\bar{\partial}(\pi_{\mathcal{N}}\xi_{\beta})\|^{2}_{L^{2}(\omega_{h,k})}=C_{1}\|\bar{\partial}(\pi_{T}\xi_{\beta})\|^{2}_{L^{2}(\omega_{h,k})}. (4)

The equality in (4) is straightforward, following [PS04, page 705], since

¯Nk1ξβ=0=¯Nk1(π𝒩ξβ)+¯Nk1(πTξβ),.\bar{\partial}_{\mathbb{P}^{N_{k}-1}}\xi_{\beta}=0=\bar{\partial}_{\mathbb{P}^{N_{k}-1}}(\pi_{\mathcal{N}}\xi_{\beta})+\bar{\partial}_{\mathbb{P}^{N_{k}-1}}(\pi_{T}\xi_{\beta}),.

which in turn implies

¯ξβ=0=¯(π𝒩ξβ)+¯(πTξβ),\bar{\partial}\xi_{\beta}=0=\bar{\partial}(\pi_{\mathcal{N}}\xi_{\beta})+\bar{\partial}(\pi_{T}\xi_{\beta}),

where ¯\bar{\partial} makes sense as a ¯\bar{\partial}-operator on the holomorphic vector bundle ιTNk1\iota^{*}T\mathbb{P}^{N_{k}-1}.

Thus it suffices to prove the inequality in (4), which is the reverse direction of [PS04, equation (5.19)], and we prove it by using the lower bound of the second fundamental form given in Lemma 2.1.

We reproduce parts of the argument in [PS04, §5] here for completeness. Fix a point xXx\in X and a local holomorphic frame {e1,,en,f1,,fm}\{e_{1},\dots,e_{n},f_{1},\dots,f_{m}\} of ιTNk1\iota^{*}T\mathbb{P}^{N_{k}-1} (with N1=n+mN-1=n+m) in a neighbourhood of xx such that it is an orthonormal basis for ιTxNk1\iota^{*}T_{x}\mathbb{P}^{N_{k}-1} with respect to g~h,k\tilde{g}_{h,k} and {e1,,en}\{e_{1},\dots,e_{n}\} form a local holomorphic frame of TXTX near xx. We then write

ξβ=i=1naiei+j=1mbjfj\xi_{\beta}=\sum_{i=1}^{n}a_{i}e_{i}+\sum_{j=1}^{m}b_{j}f_{j}

where a1,,an,b1,,bma_{1},\dots,a_{n},b_{1},\dots,b_{m} are (local) holomorphic functions. We then have

π𝒩ξβ=j=1mbj(fji=1nϕijei)\pi_{\mathcal{N}}\xi_{\beta}=\sum_{j=1}^{m}b_{j}\left(f_{j}-\sum_{i=1}^{n}\phi_{ij}e_{i}\right)

for some smooth functions ϕij\phi_{ij} vanishing at xx. Then we have

¯(π𝒩ξβ)=j=1mbj(i=1n(¯ϕij)ei.)\bar{\partial}(\pi_{\mathcal{N}}\xi_{\beta})=\sum_{j=1}^{m}b_{j}\left(-\sum_{i=1}^{n}(\bar{\partial}\phi_{ij})e_{i}.\right)

It is explained in [PS04, §5] that ¯ϕij\bar{\partial}\phi_{ij} is the second fundamental form of TXTX with respect to ω~k,h\tilde{\omega}_{k,h}. By Lemma 2.1, there exists a constant λmin(h)=λmin>0>0\lambda_{\mathrm{min}}(h)=\lambda_{\mathrm{min}}>0>0 depending only on hh, such that

i=1n|j=1N1nbj¯ϕij|2(x)λminj=1N1n|bj|2(x)\sum_{i=1}^{n}\left|\sum_{j=1}^{N-1-n}b_{j}\bar{\partial}\phi_{ij}\right|^{2}(x)\geq\lambda_{\mathrm{min}}\sum_{j=1}^{N-1-n}|b_{j}|^{2}(x)

holds for any xXx\in X. Integrating both sides over XX with respect to the volume form ωh,kn\omega^{n}_{h,k}, we get the required estimate (4).∎

Proof of Theorem 1.2.

We first recall that there exists C2>0C_{2}>0 depending only on hh (and independent of kk as long as it is large enough)

ΛHS(Hk)2C2kξΛL2(ωh,k)2=C2k(ξαL2(ωh,k)2+ξβL2(ωh,k)2)\|\Lambda\|^{2}_{\mathrm{HS}(H_{k})}\leq C_{2}k\|\xi_{\Lambda}\|^{2}_{L^{2}(\omega_{h,k})}=C_{2}k\left(\|\xi_{\alpha}\|^{2}_{L^{2}(\omega_{h,k})}+\|\xi_{\beta}\|^{2}_{L^{2}(\omega_{h,k})}\right)

by [PS04, (5.7)], where we note that the assumption (see [PS04, (5.1)]) for this estimate is satisfied for FSk(Hk)\mathrm{FS}_{k}(H_{k}) since the associated centre of mass is the identity matrix up to an error of order 1/k1/k (in the operator norm, after passing to a unitarily equivalent basis in which DkD_{k} and EkE_{k} are both diagonal) by the Bergman kernel expansion; note that Λ\Lambda is hermitian, and the Lie algebra 𝔰𝔲(N)\mathfrak{su}(N) is isomorphic to the vector space consisting of hermitian matrices.

Since ξα=πTξα\xi_{\alpha}=\pi_{T}\xi_{\alpha} is L2L^{2}-orthogonal to ξβ\xi_{\beta}, and pointwise orthogonal to π𝒩ξβ\pi_{\mathcal{N}}\xi_{\beta}, πTξα\pi_{T}\xi_{\alpha} is L2L^{2}-orthogonal to πTξβ\pi_{T}\xi_{\beta}. We thus find

πTξαL2(ωh,k)2+πTξβL2(ωh,k)2=πTξα+πTξβL2(ωh,k)2=πTξα+βL2(ωh,k)2\|\pi_{T}\xi_{\alpha}\|^{2}_{L^{2}(\omega_{h,k})}+\|\pi_{T}\xi_{\beta}\|^{2}_{L^{2}(\omega_{h,k})}=\|\pi_{T}\xi_{\alpha}+\pi_{T}\xi_{\beta}\|^{2}_{L^{2}(\omega_{h,k})}=\|\pi_{T}\xi_{\alpha+\beta}\|^{2}_{L^{2}(\omega_{h,k})}

by the linearity of πT\pi_{T}. Thus, combining these equalities we get

ξαL2(ωh,k)2+ξβL2(ωh,k)2\displaystyle\|\xi_{\alpha}\|^{2}_{L^{2}(\omega_{h,k})}+\|\xi_{\beta}\|^{2}_{L^{2}(\omega_{h,k})} =πTξαL2(ωh,k)2+πTξβL2(ωh,k)2+π𝒩ξβL2(ωh,k)2\displaystyle=\|\pi_{T}\xi_{\alpha}\|^{2}_{L^{2}(\omega_{h,k})}+\|\pi_{T}\xi_{\beta}\|^{2}_{L^{2}(\omega_{h,k})}+\|\pi_{\mathcal{N}}\xi_{\beta}\|^{2}_{L^{2}(\omega_{h,k})}
=πTξα+βL2(ωh,k)2+π𝒩ξβL2(ωh,k)2.\displaystyle=\|\pi_{T}\xi_{\alpha+\beta}\|^{2}_{L^{2}(\omega_{h,k})}+\|\pi_{\mathcal{N}}\xi_{\beta}\|^{2}_{L^{2}(\omega_{h,k})}.

By Lemma 2.2, we get

ΛHS(Hk)2C3k(πTξΛL2(ωh,k)2+¯(πTξΛ)L2(ωh,k)2),\|\Lambda\|^{2}_{\mathrm{HS}(H_{k})}\leq C_{3}k\left(\|\pi_{T}\xi_{\Lambda}\|^{2}_{L^{2}(\omega_{h,k})}+\|\bar{\partial}(\pi_{T}\xi_{\Lambda})\|^{2}_{L^{2}(\omega_{h,k})}\right),

for some C3>0C_{3}>0 which depends only on hh, where we used

¯(πTξβ)=¯(ξα+πTξβ)=¯(πTξΛ).\bar{\partial}(\pi_{T}\xi_{\beta})=\bar{\partial}(\xi_{\alpha}+\pi_{T}\xi_{\beta})=\bar{\partial}(\pi_{T}\xi_{\Lambda}).

Recalling that πTξΛ\pi_{T}\xi_{\Lambda} is ωh,k\omega_{h,k}-metric dual to dfk(Λ;Hk)df_{k}(\Lambda;H_{k}), we get

πTξΛL2(ωh,k)2=dfk(Λ;Hk)L2(ωh,k)22kn1dfk(Λ;Hk)L2(ωh)2\left\|\pi_{T}\xi_{\Lambda}\right\|^{2}_{L^{2}(\omega_{h,k})}=\left\|df_{k}(\Lambda;H_{k})\right\|^{2}_{L^{2}(\omega_{h,k})}\leq 2k^{n-1}\left\|df_{k}(\Lambda;H_{k})\right\|^{2}_{L^{2}(\omega_{h})}

for all large enough kk, by noting ωh,k=kωh+O(1/k)\omega_{h,k}=k\omega_{h}+O(1/k) which follows from the Bergman kernel expansion (2); we note that we used the natural dual metric of ωh,k\omega_{h,k} on TXT^{*}X, which contributes to the factor k1k^{-1} above. Similarly, we get

¯(πTξΛ)L2(ωh,k)22kn10,1dfk(Λ;Hk)L2(ωh)2,\left\|\bar{\partial}(\pi_{T}\xi_{\Lambda})\right\|^{2}_{L^{2}(\omega_{h,k})}\leq 2k^{n-1}\left\|\nabla^{0,1}df_{k}(\Lambda;H_{k})\right\|^{2}_{L^{2}(\omega_{h})},

where we wrote \nabla is the covariant derivative of ωh\omega_{h} and 0,1=¯\nabla^{0,1}=\bar{\partial} is its (0,1)(0,1)-part. Thus the estimates above give

ΛHS(Hk)22C3kn(fk(Λ;Hk)L2(ωh)2+fk(Λ;Hk)L2(ωh)2),\|\Lambda\|^{2}_{\mathrm{HS}(H_{k})}\leq 2C_{3}k^{n}\left(\|\nabla f_{k}(\Lambda;H_{k})\|^{2}_{L^{2}(\omega_{h})}+\|\nabla\nabla f_{k}(\Lambda;H_{k})\|^{2}_{L^{2}(\omega_{h})}\right),

when tr(Λ)=0\mathrm{tr}(\Lambda)=0.

In general, we write Λ=Λ0+cI\Lambda=\Lambda_{0}+cI where c:=tr(Λ)/Nkc:=\mathrm{tr}(\Lambda)/N_{k} and Λ0\Lambda_{0} is the trace-free part of Λ\Lambda. By diagonalising Λ0\Lambda_{0} and recalling the definition of the Hilbert–Schmidt norm, we find

ΛHS(Hk)2=Λ0HS(Hk)2+c2Nk.\|\Lambda\|^{2}_{\mathrm{HS}(H_{k})}=\|\Lambda_{0}\|^{2}_{\mathrm{HS}(H_{k})}+c^{2}N_{k}.

We likewise decompose

fk(Λ;Hk)=fk(Λ0;Hk)+ci=1Nk|si|FS(Hk)2=fk(Λ0;Hk)+cf_{k}(\Lambda;H_{k})=f_{k}(\Lambda_{0};H_{k})+c\sum_{i=1}^{N_{k}}|s_{i}|^{2}_{\mathrm{FS}(H_{k})}=f_{k}(\Lambda_{0};H_{k})+c

by recalling the definition (1) of FS(Hk)\mathrm{FS}(H_{k}). Writing ρ¯k(ωh)=VNkρk(ωh)\bar{\rho}_{k}(\omega_{h})=\frac{V}{N_{k}}\rho_{k}(\omega_{h}) and recalling FSk(Hk)=ρ¯k(ωh)1hk\mathrm{FS}_{k}(H_{k})=\bar{\rho}_{k}(\omega_{h})^{-1}h^{k} as in [rawnsley], the above equation gives

ρ¯k(ωh)fk(Λ;Hk)=i=1Nkλi|si|hk2+cρ¯k(ωh).\bar{\rho}_{k}(\omega_{h})f_{k}(\Lambda;H_{k})=\sum_{i=1}^{N_{k}}\lambda_{i}|s_{i}|^{2}_{h^{k}}+c\bar{\rho}_{k}(\omega_{h}).

Integrating this over XX, noting that {si}i=1Nk\{s_{i}\}_{i=1}^{N_{k}} is a Hilb(hk)\mathrm{Hilb}(h^{k})-orthonormal basis and that we have Xρ¯k(ωh)ωhn/n!=V\int_{X}\bar{\rho}_{k}(\omega_{h})\omega^{n}_{h}/n!=V, we get

cV=Xρ¯k(ωh)fk(Λ;Hk)ωhnn!cV=\int_{X}\bar{\rho}_{k}(\omega_{h})f_{k}(\Lambda;H_{k})\frac{\omega_{h}^{n}}{n!}

since tr(Λ0)=0\mathrm{tr}(\Lambda_{0})=0, and hence

|c|=1V|Xρ¯k(ωh)fk(Λ;Hk)ωhnn!|1Vρ¯k(ωh)L2(ωh)fk(Λ;Hk)L2(ωh)2fk(Λ;Hk)L2(ωh),|c|=\frac{1}{V}\left|\int_{X}\bar{\rho}_{k}(\omega_{h})f_{k}(\Lambda;H_{k})\frac{\omega_{h}^{n}}{n!}\right|\leq\frac{1}{V}\|\bar{\rho}_{k}(\omega_{h})\|_{L^{2}(\omega_{h})}\|f_{k}(\Lambda;H_{k})\|_{L^{2}(\omega_{h})}\leq 2\|f_{k}(\Lambda;H_{k})\|_{L^{2}(\omega_{h})},

by Cauchy–Schwarz and the Bergman kernel expansion (2).

Collecting all the estimates and setting C4:=2max{C3,2V}>0C_{4}:=2\max\{C_{3},2V\}>0, which depends only on hh and not on kk, we get

ΛHS(Hk)2\displaystyle\|\Lambda\|^{2}_{\mathrm{HS}(H_{k})} 2Nkfk(Λ;Hk)L2(ωh)2+2C3kn(fk(Λ;Hk)L2(ωh)2+fk(Λ;Hk)L2(ωh)2)\displaystyle\leq 2N_{k}\|f_{k}(\Lambda;H_{k})\|^{2}_{L^{2}(\omega_{h})}+2C_{3}k^{n}\left(\|\nabla f_{k}(\Lambda;H_{k})\|^{2}_{L^{2}(\omega_{h})}+\|\nabla\nabla f_{k}(\Lambda;H_{k})\|^{2}_{L^{2}(\omega_{h})}\right)
C4knfk(Λ;Hk)W2,2(ωh)2\displaystyle\leq C_{4}k^{n}\|f_{k}(\Lambda;H_{k})\|^{2}_{W^{2,2}(\omega_{h})}

for all large enough kk, since we have Nk=Vkn+O(kn1)N_{k}=Vk^{n}+O(k^{n-1}) by the asymptotic Riemann–Roch theorem. ∎

Remark 2.3.

A heuristic interpretation of Theorem 1.2 is as follows. The image of the Kodaira embedding ιk:X(H0(X,L))\iota_{k}:X\hookrightarrow\mathbb{P}(H^{0}(X,L)^{\vee}) is not contained in any proper linear subspace, and hence we expect that an appropriately defined norm of πTξΛ\pi_{T}\xi_{\Lambda} should be equivalent to ΛHS(Hk)\|\Lambda\|_{\mathrm{HS}(H_{k})}. The estimate depends on how “degenerate” ιk(X)\iota_{k}(X) is in the projective space, i.e. how close it is to being contained in a proper linear subspace. The argument above proves this intuition when we take the reference metric to be FSk(Hk)\mathrm{FS}_{k}(H_{k}), for which the centre of mass of the embedding is in fact close to the identity matrix by the Bergman kernel expansion (2). Since ιk\iota_{k} is an algebraic morphism, we can also expect that the constant depends at most polynomially in kk. This argument seems intuitively plausible, but no written proof seems to be available in the literature.

3. Correction to [yhextremal]

There are several places in [yhextremal] which requires corrections, since they depend on the results in [yhhilb] that are reproduced in [yhextremal, Lemmas 5 and 6]. [yhextremal, Lemma 5] is not directly used in the rest of the argument in [yhextremal], and only affects the paper through [yhextremal, Lemma 6]. The full quantitative statement of [yhextremal, Lemma 6] is used only in [yhextremal, page 3004 in §4.2, after (35)], and all the other places [yhextremal, pages 2982 and 3002] that need [yhextremal, Lemma 6] only need the injectivity of the Fubini–Study map which was also proved by Lempert [Lem21].

Thus it suffices to correct the argument in [yhextremal, page 3004], which deals with the equation

i=1Nk(di(1Fm,k4πkm+2))|si|h(m)k2=0,\sum_{i=1}^{N_{k}}\left(d_{i}-\left(1-\frac{F_{m,k}}{4\pi k^{m+2}}\right)\right)|s_{i}|^{2}_{h^{k}_{(m)}}=0, (5)

where d1,,dNkd_{1},\dots,d_{N_{k}}\in\mathbb{R}, mm\in\mathbb{N}, h(m)k=FSk(Hk,m)h^{k}_{(m)}=\mathrm{FS}_{k}(H_{k,m}), Hk,mkH_{k,m}\in\mathcal{B}_{k}, is a hermitian metric on LkL^{k} as constructed in [yhextremal, Proposition 1 and Proof of Corollary 1], and {si}i=1Nk\{s_{i}\}_{i=1}^{N_{k}} is an Hk,mH_{k,m}-orthonormal basis. We need to conclude that there exists C5>0C_{5}>0 such that

|di1|C5kc(n)m|d_{i}-1|\leq C_{5}k^{c(n)-m}

for all i=1,,Nki=1,\dots,N_{k}, and for all large enough kk, where c(n)c(n) is some constant which depends only on nn.

We apply the argument above. We first define two hermitian matrices A:=diag(d11,,dNk1)A:=\mathrm{diag}(d^{-1}_{1},\dots,d^{-1}_{N_{k}}) and B:=IB:=I with respect to the Hk,mH_{k,m}-orthonormal basis {si}i=1Nk\{s_{i}\}_{i=1}^{N_{k}}. Note that they take a different form when they are represented with respect to an HkH_{k}-orthonormal basis, where Hk=Hilb(hk)H_{k}=\mathrm{Hilb}(h^{k}) and ωh\omega_{h} is the extremal metric that is used as a reference metric in [yhextremal]. We find

fk(A,B;Hk,m)=i=1Nk(di1)|si|h(m)k2,f_{k}(A,B;H_{k,m})=\sum_{i=1}^{N_{k}}\left(d_{i}-1\right)|s_{i}|^{2}_{h^{k}_{(m)}},

and hence (5) can be re-written as

fk(A,B;Hk,m)=Fm,k4πkm+2.f_{k}(A,B;H_{k,m})=-\frac{F_{m,k}}{4\pi k^{m+2}}.

Taking Hk=Hilb(hk)H_{k}=\mathrm{Hilb}(h^{k}) as above, we observe

fk(A,B;Hk)=FSk(Hk)FSk(Hk,m)(FSk(Hk,m)FSk(A)FSk(Hk,m)FSk(B)),f_{k}(A,B;H_{k})=\frac{\mathrm{FS}_{k}(H_{k})}{\mathrm{FS}_{k}(H_{k,m})}\left(\frac{\mathrm{FS}_{k}(H_{k,m})}{\mathrm{FS}_{k}(A)}-\frac{\mathrm{FS}_{k}(H_{k,m})}{\mathrm{FS}_{k}(B)}\right),

which implies that

fk(A,B;Hk)=FSk(Hk)FSk(Hk,m)fk(A,B;Hk,m)=FSk(Hk)FSk(Hk,m)Fm,k4πkm+2.f_{k}(A,B;H_{k})=\frac{\mathrm{FS}_{k}(H_{k})}{\mathrm{FS}_{k}(H_{k,m})}f_{k}(A,B;H_{k,m})=-\frac{\mathrm{FS}_{k}(H_{k})}{\mathrm{FS}_{k}(H_{k,m})}\frac{F_{m,k}}{4\pi k^{m+2}}. (6)

We recall FSk(Hk)=ρ¯k(ωh)1hk\mathrm{FS}_{k}(H_{k})=\bar{\rho}_{k}(\omega_{h})^{-1}h^{k} [rawnsley], and

FSk(Hk,m)=ekϕ(m)hk\mathrm{FS}_{k}(H_{k,m})=e^{k\phi_{(m)}}h^{k}

in the terminology of [yhextremal, Proposition 1]; the only important point here is that kϕ(m)k\phi_{(m)} converges in CC^{\infty} as kk\to\infty, and hence its derivatives are all uniformly bounded for all large enough kk. Together with the expansion (2) of the Bergman function, we thus find that all derivatives of FSk(Hk)/FSk(Hk,m)\mathrm{FS}_{k}(H_{k})/\mathrm{FS}_{k}(H_{k,m}) are bounded uniformly for all large enough kk, and hence there exists C6>0C_{6}>0 such that

FSk(Hk)FSk(Hk,m)Fm,k4πkm+2W2,2(ωh)C6km2Fm,kW2,2(ωh)\left\|\frac{\mathrm{FS}_{k}(H_{k})}{\mathrm{FS}_{k}(H_{k,m})}\frac{F_{m,k}}{4\pi k^{m+2}}\right\|_{W^{2,2}(\omega_{h})}\leq C_{6}k^{-m-2}\|F_{m,k}\|_{W^{2,2}(\omega_{h})}

holds for all large enough kk. As pointed out in [yhextremal, page 2995, before the equation (20)], Fm,kW2,2(ωh)\|F_{m,k}\|_{W^{2,2}(\omega_{h})} is uniformly bounded for all large enough kk. Thus there exists a constant C7>0C_{7}>0 such that

fk(A,B;Hk)W2,2(ωh)C7km2.\|f_{k}(A,B;H_{k})\|_{W^{2,2}(\omega_{h})}\leq C_{7}k^{-m-2}. (7)

We now recall that A1=diag(d1,,dNk)A^{-1}=\mathrm{diag}(d_{1},\dots,d_{N_{k}}) and B1=IB^{-1}=I with respect to the Hk,mH_{k,m}-orthonormal basis {si}i=1Nk\{s_{i}\}_{i=1}^{N_{k}}. When we represent HkH_{k} as a matrix with respect to an Hk,mH_{k,m}-orthonormal basis, we find that each entry of HkH_{k} can be bounded uniformly for all large enough kk by the construction of Hk,mH_{k,m} as given in [yhextremal, Proof of Corollary 1] (see also [donnum, Appendix]), since it is defined as a Hilb(h~k)\mathrm{Hilb}(\tilde{h}^{k}) for some perturbation h~\tilde{h} of hh such that log(h~/h)=O(1/k)\log(\tilde{h}/h)=O(1/k). Thus there exists a constant C8>0C_{8}>0 such that

A1B1HS(Hm,k)C8NkA1B1HS(Hk),\|A^{-1}-B^{-1}\|_{\mathrm{HS}(H_{m,k})}\leq C_{8}N_{k}\|A^{-1}-B^{-1}\|_{\mathrm{HS}(H_{k})},

holds for all large enough kk. By recalling the asymptotic Riemann–Roch theorem Nk=Vkn+O(kn1)N_{k}=Vk^{n}+O(k^{n-1}), we find that there exists a constant C9>0C_{9}>0 such that

|di1|\displaystyle|d_{i}-1| A1B1HS(Hm,k)\displaystyle\leq\|A^{-1}-B^{-1}\|_{\mathrm{HS}(H_{m,k})}
C9knA1B1HS(Hk)\displaystyle\leq C_{9}k^{n}\|A^{-1}-B^{-1}\|_{\mathrm{HS}(H_{k})}
C9Chk2nfk(A,B;Hk)W2,2(ωh)\displaystyle\leq C_{9}C_{h}k^{2n}\|f_{k}(A,B;H_{k})\|_{W^{2,2}(\omega_{h})}
C7C9Chk2nm2\displaystyle\leq C_{7}C_{9}C_{h}k^{2n-m-2}

for all i=1,,Nki=1,\dots,N_{k} and for all large enough kk, by Theorem 1.2 and (7), as required.

References