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

A note on a question of Markman

Shengyuan Huang Shengyuan Huang, Mathematics Department, University of Wisconsin–Madison, 480 Lincoln Drive, Madison, WI 53706–1388, USA.
Email: shuang279@wisc.edu
Abstract

Abstract: Let \displaystyle\mathcal{F} be a vector bundle on a complex projective algebraic variety X\displaystyle X. If \displaystyle\mathcal{F} deforms along a first order deformation of X\displaystyle X, its Mukai vector remains of Hodge type along this deformation. We prove an analogous statement for all polyvector fields, not only for those in H1(X,TX)\displaystyle H^{1}(X,T_{X}) corresponding to deformations of the complex structure. This answers a question of Markman. We also explore a Lie theoretic analogue of the statement above.

Mathematics Subject Classification (2020). 14F06.

Key words. Hochschild homology and cohomology, HKR isomorphisms, deformation.

1 Introduction

1.1 Let X\displaystyle X be a smooth complex algebraic variety. Consider the first order deformation X~\displaystyle\tilde{X} of X\displaystyle X associated to a class α~H1(X,TX)\displaystyle\tilde{\alpha}\in H^{1}(X,T_{X}).

In general, a vector bundle \displaystyle\mathcal{F} may not deform to a bundle ~\displaystyle\tilde{\mathcal{F}} on X~\displaystyle\tilde{X}. The obstruction αExt2(,)\displaystyle\alpha_{\mathcal{F}}\in\mathrm{Ext}^{2}(\mathcal{F},\mathcal{F}) to the existence of a vector bundle ~\displaystyle\tilde{\mathcal{F}} on X~\displaystyle\tilde{X} such that ~|X\displaystyle\tilde{\mathcal{F}}|_{X}\cong\mathcal{F} was described in [B94, T09] as the contraction

α=α~atExt2(,).\displaystyle\alpha_{\mathcal{F}}=\tilde{\alpha}\lrcorner\,at_{\mathcal{F}}\in\mathrm{Ext}^{2}(\mathcal{F},\mathcal{F}).

Here atExt1(,ΩX)\displaystyle at_{\mathcal{F}}\in\mathrm{Ext}^{1}(\mathcal{F},\mathcal{F}\otimes\Omega_{X}) is the Atiyah class of \displaystyle\mathcal{F}.

Moreover, if \displaystyle\mathcal{F} does deform, then its Chern classes and hence its Mukai vector stay of Hodge type on the deformed space X~\displaystyle\tilde{X}. This implies that the class

α~v()HΩ(X)=defqp=Hp(X,qΩX)\displaystyle\tilde{\alpha}\lrcorner\,v(\mathcal{F})\in\mathrm{H\Omega}_{*}(X)\overset{\mathrm{def}}{=}\bigoplus_{q-p=*}H^{p}(X,\wedge^{q}\Omega_{X})

vanishes, where v()\displaystyle v(\mathcal{F}) is the Mukai vector of \displaystyle\mathcal{F}.

Thus, in the simple case where α~H1(X,TX)\displaystyle\tilde{\alpha}\in H^{1}(X,T_{X}) we conclude that if α~at\displaystyle\tilde{\alpha}\lrcorner\,at_{\mathcal{F}} is zero, then α~v()\displaystyle\tilde{\alpha}\lrcorner\,v(\mathcal{F}) is zero.

I

n an email correspondence, Eyal Markman asked if the above statement can be generalized to the case where α~\displaystyle\tilde{\alpha} is an arbitrary polyvector field in HT(X)=p+q=Hp(X,qTX)\displaystyle\mathrm{HT}^{*}(X)=\bigoplus_{p+q=*}H^{p}(X,\wedge^{q}T_{X}). According to Markman, this question is central to his study of the deformations of hyperkähler manifolds. We provide an answer to this question in this paper.

F

irst, there are two ways to generalize the class that appears in (LABEL:1.1) above. For a class α~HT(X)\displaystyle\tilde{\alpha}\in\mathrm{HT}^{*}(X) we can define classes in Ext(,)\displaystyle\mathrm{Ext}^{*}(\mathcal{F},\mathcal{F}) in the following two ways.

The first is defined by using the HKR isomorphism

IHKR:HT(X)HH(X),\displaystyle I^{HKR}:\mathrm{HT}^{*}(X)\rightarrow\mathrm{HH}^{*}(X),

where the latter is the Hochschild cohomology of X\displaystyle X. Since HH(X)\displaystyle\mathrm{HH}^{*}(X) can be interpreted as natural transformations of the identity functor at the dg level, this yields a natural map HH(X)Ext(,)\displaystyle\mathrm{HH}^{*}(X)\rightarrow\mathrm{Ext}^{*}(\mathcal{F},\mathcal{F}).

The second construction was defined by Toda [T09]. Consider the exponential Atiyah class

exp(at)=1+at++(at)kk!+,\displaystyle\exp(at_{\mathcal{F}})=1+at_{\mathcal{F}}+\cdots+\frac{(at_{\mathcal{F}})^{k}}{k!}+\cdots,

where (at)kExtk(,kΩX)\displaystyle(at_{\mathcal{F}})^{k}\in\mathrm{Ext}^{k}(\mathcal{F},\mathcal{F}\otimes\wedge^{k}\Omega_{X}). Let α~p,kHp(X,kTX)\displaystyle\tilde{\alpha}^{p,k}\in H^{p}(X,\wedge^{k}T_{X}) be the homogenous degree (p,k)\displaystyle(p,k) part of α~\displaystyle\tilde{\alpha}. We can contract α~p,k\displaystyle\tilde{\alpha}^{p,k} with (at)kk!\displaystyle\frac{(at_{\mathcal{F}})^{k}}{k!} to get an element in Extp+k(,)\displaystyle\mathrm{Ext}^{p+k}(\mathcal{F},\mathcal{F}). Taking the sum over all (p,k)\displaystyle(p,k), we get the desired class which will be denoted by αexp(at)~Ext(,)\displaystyle\tilde{\alpha\lrcorner\exp(at_{\mathcal{F}})}\in\mathrm{Ext}^{*}(\mathcal{F},\mathcal{F}). When α~\displaystyle\tilde{\alpha} is a class in H1(X,TX)\displaystyle H^{1}(X,T_{X}), we recover the previous contraction α~at\displaystyle\tilde{\alpha}\lrcorner\,at_{\mathcal{F}}.

Our first result is below.

Theorem A.

The two classes defined above are the same. In other words the diagram

HH(X)\displaystyle\textstyle{\mathrm{HH}^{*}(X)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Ext(,)\displaystyle\textstyle{\mathrm{Ext}^{*}(\mathcal{F},\mathcal{F})}HT(X)\displaystyle\textstyle{\mathrm{HT}^{*}(X)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}IHKR\displaystyle\scriptstyle{I^{HKR}}()exp(at)\displaystyle\scriptstyle{(-)\lrcorner\exp(at_{\mathcal{F}})}

is commutative.

There is an analogous result for Hopf algebras. See Theorem 2.7 in [CI17] and see [PW09] for more details.

T

he space HΩ(X)\displaystyle\mathrm{H\Omega}_{*}(X) is naturally a module over HT(X)\displaystyle\mathrm{HT}^{*}(X), mimicking the module structure of Hochschild homology over cohomology. For an object \displaystyle\mathcal{F} in the derived category of X\displaystyle X, its Mukai vector v()\displaystyle v(\mathcal{F}) lies in HΩ(X)\displaystyle\mathrm{H\Omega}_{*}(X). Thus we can act with the class α~\displaystyle\tilde{\alpha} to obtain α~v()HΩ(X)\displaystyle\tilde{\alpha}\lrcorner\,v(\mathcal{F})\in\mathrm{H\Omega}_{*}(X).

Theorem B. If α~exp(at)=0\displaystyle\tilde{\alpha}\lrcorner\exp(at_{\mathcal{F}})=0, then we have

D(α~)v()=0.\displaystyle D(\tilde{\alpha})\lrcorner\,v(\mathcal{F})=0.

Here D\displaystyle D is the Duflo operator,

D(α~)=Td(X)12α~,\displaystyle D(\tilde{\alpha})=Td(X)^{\frac{1}{2}}\lrcorner\,\tilde{\alpha},

where Td(X)\displaystyle Td(X) is the Todd class of X\displaystyle X.

Remark 1.1.

We are using the contraction symbol \displaystyle\lrcorner in three different ways in this paper.

  • A polyvector field α~HT(X)\displaystyle\tilde{\alpha}\in\mathrm{HT}^{*}(X) acts on a class vHΩ(X)\displaystyle v\in\mathrm{H\Omega}_{*}(X). This action is denoted by α~vHΩ(X)\displaystyle\tilde{\alpha}\lrcorner\,v\in\mathrm{H\Omega}_{*}(X).

  • A class vHΩ(X)\displaystyle v\in\mathrm{H\Omega}_{*}(X) acts on a polyvector field α~HT(X)\displaystyle\tilde{\alpha}\in\mathrm{HT}^{*}(X). This action yields an element vα~HT(X)\displaystyle v\lrcorner\,\tilde{\alpha}\in\mathrm{HT}^{*}(X). We only use the second contraction in the Duflo operator D(α~)=Td(X)12α~\displaystyle D(\tilde{\alpha})=Td(X)^{\frac{1}{2}}\lrcorner\,\tilde{\alpha} in this paper. Note that D\displaystyle D is an automorphism of HT(X)\displaystyle\mathrm{HT}^{*}(X). The inverse operator is D1(α~)=Td(X)12α~\displaystyle D^{-1}(\tilde{\alpha})=Td(X)^{-\frac{1}{2}}\lrcorner\,\tilde{\alpha}.

  • The third contraction map is βexp(at)Ext(,)\displaystyle\beta\lrcorner\exp(at_{\mathcal{F}})\in\mathrm{Ext}^{*}(\mathcal{F},\mathcal{F}) for βHT(X)\displaystyle\beta\in\mathrm{HT}^{*}(X). An element β\displaystyle\beta in Hp(X,kTX)\displaystyle H^{p}(X,\wedge^{k}T_{X}) can only contract with the term (at)kk!\displaystyle\frac{(at_{\mathcal{F}})^{k}}{k!} in the Taylor expansion of exp(at)\displaystyle\exp(at_{\mathcal{F}}). It is easy to distinguish this map from the previous two maps.

T

he inspiration for Theorem A comes from a similar statement in Lie theory. Let 𝔤\displaystyle\mathfrak{g} be a finite dimensional Lie algebra and V\displaystyle V be a finite dimensional representation. One can draw the diagram

(U𝔤)𝔤\displaystyle\textstyle{(U\mathfrak{\mathfrak{g}})^{\mathfrak{g}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Hom(V,V)\displaystyle\textstyle{\mathrm{Hom}(V,V)}(S𝔤)𝔤\displaystyle\textstyle{(S\mathfrak{g})^{\mathfrak{g}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}

which is similar to the one in Theorem A. We will provide more details and prove that the diagram above is commutative in section 2.

N

ote that our statement in Theorem B appears to be different from the original one, which did not have the Duflo operator D\displaystyle D. We will prove that the original statement follows easily from ours.

Plan of the paper.

Section 2 contains the proof of Theorem A and of its Lie theoretic analogue.

Section 3 is devoted to the proof of Theorem B. It is a consequence of Theorem A. At the end we prove that we can recover the result in (LABEL:1.1) from Theorem B.

Acknowledgments.

I would like to thank Andrei Căldăraru for discussing details with me during our weekly meeting. I have benefited from stimulating email correspondence with Dror Bar-Natan. I am also grateful to Eyal Markman for many valuable comments. The author is partially supported by the National Science Foundation under Grant No. DMS-1811925.

2 The proof of Theorem A

We prove Theorem A in this section. The diagram in Theorem A has a Lie theoretic background. We draw the corresponding diagram for Lie algebras and we explain the similarity between the Lie algebra diagram and the diagram in Theorem A. We provide a proof for the commutativity of the Lie algebra diagram and explain that the proof can be generalized to the diagram in Theorem A.

A similar diagram for Lie algebras.

Let 𝔤\displaystyle\mathfrak{g} be a finite dimensional Lie algebra over a field of characteristic zero and let V\displaystyle V be a finite dimensional representation of 𝔤\displaystyle\mathfrak{g}. There is a diagram

(U𝔤)𝔤\displaystyle\textstyle{(U\mathfrak{g})^{\mathfrak{g}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Hom(V,V)\displaystyle\textstyle{\mathrm{Hom}(V,V)}(S𝔤)𝔤.\displaystyle\textstyle{(S\mathfrak{g})^{\mathfrak{g}}.\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}PBW

The PBW map from the symmetric algebra S𝔤\displaystyle S\mathfrak{g} to the universal enveloping algebra U𝔤\displaystyle U\mathfrak{g} is defined on the degree n\displaystyle n-th component of S𝔤\displaystyle S\mathfrak{g} as follows

x1xn1n!σSnxσ(1)xσ(n).\displaystyle x_{1}\cdots x_{n}\rightarrow\frac{1}{n!}\sum_{\sigma\in S_{n}}x_{\sigma(1)}\cdots x_{\sigma(n)}.

Here Sn\displaystyle S_{n} is the symmetric group on a finite set of n symbols. The universal enveloping algebra U𝔤\displaystyle U\mathfrak{g} acts naturally on V\displaystyle V. This natural action defines the map (U𝔤)𝔤Hom(V,V)\displaystyle(\mathrm{U}\mathfrak{g})^{\mathfrak{g}}\rightarrow\mathrm{Hom}(V,V) on the top of the diagram above. The map (S𝔤)𝔤Hom(V,V)\displaystyle(\mathrm{S}\mathfrak{g})^{\mathfrak{g}}\rightarrow\mathrm{Hom}(V,V) is defined as follows. We can rewrite the representation map 𝔤VV\displaystyle\mathfrak{g}\otimes V\rightarrow V as a map Λ:VV𝔤\displaystyle\Lambda:V\rightarrow V\otimes\mathfrak{g}^{*}. Take the exponent

exp(Λ)=idV+Λ++Λkk!+\displaystyle\exp(\Lambda)=id_{V}+\Lambda+\cdots+\frac{\Lambda^{k}}{k!}+\cdots

of the map Λ\displaystyle\Lambda. Then we can contract exp(Λ)\displaystyle\exp(\Lambda) with S𝔤\displaystyle S\mathfrak{g}.

In algebraic geometry, Kapranov and Kontsevich [K99] observed that the shifted tangent bundle TX[1]\displaystyle T_{X}[-1] has a Lie algebra structure in the derived category of X\displaystyle X. Roberts and Willerton [RW10] proved that the category of representations of TX[1]\displaystyle T_{X}[-1] is the derived category of X\displaystyle X and the universal enveloping algebra of TX[1]\displaystyle T_{X}[-1] is the Hochschild cochain complex om(Δ𝒪X,Δ𝒪X)\displaystyle\mathcal{RH}om(\Delta_{*}\mathcal{O}_{X},\Delta_{*}\mathcal{O}_{X}), where Δ\displaystyle\Delta is the diagonal embedding Δ:XX×X\displaystyle\Delta:X\hookrightarrow X\times X. The functor ()𝔤\displaystyle(-)^{\mathfrak{g}} is the 0-th Lie algebra cohomology which is similar to H(X,)\displaystyle H^{*}(X,-). Setting 𝔤\displaystyle\mathfrak{g} to be equal to TX[1]\displaystyle T_{X}[-1] in the Lie algebra diagram, we get the diagram in Theorem A for a smooth complex variety X\displaystyle X. The Hochschild cohomology HH(X)\displaystyle\mathrm{HH}^{*}(X) plays the role of (U𝔤)𝔤\displaystyle(U\mathfrak{g})^{\mathfrak{g}}, HT(X)\displaystyle\mathrm{HT}^{*}(X) plays the role of (S𝔤)𝔤\displaystyle(S\mathfrak{g})^{\mathfrak{g}}, and the HKR map is precisely the PBW map.

{Proof}

[Proof of the commutativity for the Lie algebra diagram.] We can prove that the diagram in (2) is commutative even before taking 𝔤\displaystyle\mathfrak{g}-invariants, i.e., the diagram

U𝔤\displaystyle\textstyle{U\mathfrak{g}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Hom(V,V)\displaystyle\textstyle{\mathrm{Hom}(V,V)}S𝔤\displaystyle\textstyle{S\mathfrak{g}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}PBW

is commutative. The map PBW factors through the tensor algebra T𝔤\displaystyle T\mathfrak{g}

PBW:S𝔤\displaystyle\textstyle{\mbox{PBW}:S\mathfrak{g}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ψ\displaystyle\scriptstyle{~{}~{}~{}~{}~{}~{}\psi}T𝔤\displaystyle\textstyle{T\mathfrak{g}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}U𝔤,\displaystyle\textstyle{U\mathfrak{g},}

so we can replace U𝔤\displaystyle U\mathfrak{g} at the top left corner of the diagram by T𝔤\displaystyle T\mathfrak{g}. It is easy to check that the map S𝔤Hom(V,V)\displaystyle S\mathfrak{g}\rightarrow\mathrm{Hom}(V,V) is equal to the following map

S𝔤\displaystyle\textstyle{S\mathfrak{g}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ψ\displaystyle\scriptstyle{\psi}T𝔤\displaystyle\textstyle{T\mathfrak{g}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}φ\displaystyle\scriptstyle{\varphi~{}~{}~{}~{}~{}~{}~{}~{}}Hom(V,V),\displaystyle\textstyle{\mathrm{Hom}(V,V),}

where the map φ:T𝔤Hom(V,V)\displaystyle\varphi:T\mathfrak{g}\rightarrow\mathrm{Hom}(V,V) is defined as follows. Rewrite the representation map 𝔤VV\displaystyle\mathfrak{g}\otimes V\rightarrow V as a map Λ:VV𝔤\displaystyle\Lambda:V\rightarrow V\otimes\mathfrak{g}^{*}. Instead of taking the exponential of the map Λ\displaystyle\Lambda, we compose the map Λ\displaystyle\Lambda with itself k\displaystyle k times. We get a map Λk:VV(𝔤)k\displaystyle\Lambda^{\otimes k}:V\rightarrow V\otimes(\mathfrak{g}^{*})^{\otimes k} in this way. Contract Λk\displaystyle\Lambda^{\otimes k} with 𝔤k\displaystyle\mathfrak{g}^{\otimes k} and get a map 𝔤kHom(V,V)\displaystyle\mathfrak{g}^{\otimes k}\rightarrow\mathrm{Hom}(V,V). Adding the k\displaystyle k-th components for all k\displaystyle k\in\mathbb{N}, we obtain the desired map φ:T𝔤Hom(V,V)\displaystyle\varphi:T\mathfrak{g}\rightarrow\mathrm{Hom}(V,V).

Now we have two maps T𝔤Hom(V,V)\displaystyle T\mathfrak{g}\rightarrow\mathrm{Hom}(V,V). One of them is the map φ\displaystyle\varphi, and the other one is Θ:T𝔤U𝔤Hom(V,V)\displaystyle\Theta:T\mathfrak{g}\rightarrow U\mathfrak{g}\rightarrow\mathrm{Hom}(V,V). We want to show that they agree. This follows from Lemma 2 below by setting W1\displaystyle W_{1} to be V\displaystyle V and W2\displaystyle W_{2} to be 𝔤k\displaystyle\mathfrak{g}^{\otimes k}.

{Lemma}

Let W1\displaystyle W_{1} and W2\displaystyle W_{2} be finite dimensional vector spaces over a field k\displaystyle k and f\displaystyle f be a map W2W1W1\displaystyle W_{2}\otimes W_{1}\rightarrow W_{1}. Rewrite the map as g:W1W2W1\displaystyle g:W_{1}\rightarrow W_{2}^{*}\otimes W_{1} by the adjunction formula Hom(W2kW1,W1)=Hom(W1,W2kW1)\displaystyle\mathrm{Hom}(W_{2}\otimes_{k}W_{1},W_{1})=\mathrm{Hom}(W_{1},W_{2}^{*}\otimes_{k}W_{1}). Fix an element xW2\displaystyle x\in W_{2}. Then f(x)\displaystyle f(x\otimes-) is a map from W1\displaystyle W_{1} to W1\displaystyle W_{1}. This map is precisely g\displaystyle g followed by the contraction with x\displaystyle x.

Proof 2.1.

This is due to the adjunction property

Hom(W2kW1,W1)=Hom(W1,W2kW1).\displaystyle\mathrm{Hom}(W_{2}\otimes_{k}W_{1},W_{1})=\mathrm{Hom}(W_{1},W_{2}^{*}\otimes_{k}W_{1}).
{Proof}

[Proof of Theorem A.] The proof above reduces the commutativity of the Lie algebra diagram to a statement about tensor algebras. The statement about tensor algebras remains valid in the case of derived categories.

One can define a map Sym(TX[1])T(TX[1])\displaystyle\mathrm{Sym}(T_{X}[-1])\rightarrow T(T_{X}[-1]) given by the formula

x1xn1n!σSn(1)sgn(σ)xσ(1)xσ(n),\displaystyle x_{1}\wedge\cdots\wedge x_{n}\rightarrow\frac{1}{n!}\sum_{\sigma\in S_{n}}(-1)^{sgn(\sigma)}x_{\sigma(1)}\cdots x_{\sigma(n)},

where T(TX[1])\displaystyle T(T_{X}[-1]) is the tensor algebra on TX[1]\displaystyle T_{X}[-1].

The map above is a differential graded version of the map ψ\displaystyle\psi in (2). Let X(1)\displaystyle X^{(1)} be the first order neighborhood of X\displaystyle X in X×X\displaystyle X\times X. There are embeddings i:XX(1)\displaystyle i:X\hookrightarrow X^{(1)} and j:X(1)X×X\displaystyle j:X^{(1)}\hookrightarrow X\times X. Arinkin and Căldăraru [AC12] showed that T(TX[1])\displaystyle T(T_{X}[-1]) is isomorphic to (ii𝒪X)\displaystyle(i^{*}i_{*}\mathcal{O}_{X})^{\vee}, where ()\displaystyle(-)^{\vee} is the dual. The map

(ii𝒪X)(ijji𝒪X)=(ΔΔ𝒪X)=om(Δ𝒪X,Δ𝒪X)\displaystyle(i^{*}i_{*}\mathcal{O}_{X})^{\vee}\rightarrow(i^{*}j^{*}j_{*}i_{*}\mathcal{O}_{X})^{\vee}=(\Delta^{*}\Delta_{*}\mathcal{O}_{X})^{\vee}=\mathcal{RH}om(\Delta_{*}\mathcal{O}_{X},\Delta_{*}\mathcal{O}_{X})

is defined by the adjunction jj\displaystyle j^{*}\dashv j_{*}. The composite map

Sym(TX[1])T(TX[1])(ii𝒪X)(ijji𝒪X)\displaystyle\mathrm{Sym}(T_{X}[-1])\rightarrow T(T_{X}[-1])\cong(i^{*}i_{*}\mathcal{O}_{X})^{\vee}\rightarrow(i^{*}j^{*}j_{*}i_{*}\mathcal{O}_{X})^{\vee}
=(ΔΔ𝒪X)=om(Δ𝒪X,Δ𝒪X)\displaystyle=(\Delta^{*}\Delta_{*}\mathcal{O}_{X})^{\vee}=\mathcal{RH}om(\Delta_{*}\mathcal{O}_{X},\Delta_{*}\mathcal{O}_{X})

is the sheaf version HKR isomorphism as showed in [AC12]. Taking cohomology on both sides of the equality above, we get the HKR isomorphism

IHKR:HT(X)=p+q=Hp(X,qTX)HH(X).\displaystyle I^{HKR}:\mathrm{HT}^{*}(X)=\bigoplus_{p+q=*}H^{p}(X,\wedge^{q}T_{X})\rightarrow\mathrm{HH}^{*}(X).

Now it is clear that we have a commutative diagram

om(Δ𝒪X,Δ𝒪X)\displaystyle\textstyle{\mathcal{RH}om(\Delta_{*}\mathcal{O}_{X},\Delta_{*}\mathcal{O}_{X})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}om(,)\displaystyle\textstyle{\mathcal{RH}om(\mathcal{F},\mathcal{F})}T(TX[1])\displaystyle\textstyle{T(T_{X}[-1])\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Sym(TX[1]),\displaystyle\textstyle{\mathrm{Sym}(T_{X}[-1]),\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}

which is similar to the Lie algebra diagram in (2). Taking cohomology on the diagram above, we get the diagram

HH(X)\displaystyle\textstyle{\mathrm{HH}^{*}(X)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Ext(,)\displaystyle\textstyle{\mathrm{Ext}^{*}(\mathcal{F},\mathcal{F})}HT(X)=p+q=Hp(X,qTX)\displaystyle\textstyle{\mathrm{HT}^{*}(X)=\displaystyle{\bigoplus_{p+q=*}H^{p}(X,\wedge^{q}T_{X})}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}IHKR\displaystyle\scriptstyle{I^{HKR}}()exp(at)\displaystyle\scriptstyle{(-)\lrcorner\exp(at_{\mathcal{F}})}

that we start with in Theorem A.

3 The proof of Theorem B

We use Theorem A to prove Theorem B in this section.

D

enote Ihkr(α~)\displaystyle I^{hkr}(\tilde{\alpha}) by αHH(X)=ExtX×X(𝒪Δ,𝒪Δ)\displaystyle\alpha\in\mathrm{HH}^{*}(X)=\mathrm{Ext}_{X\times X}^{*}(\mathcal{O}_{\Delta},\mathcal{O}_{\Delta}), where 𝒪Δ=Δ𝒪X\displaystyle\mathcal{O}_{\Delta}=\Delta_{*}\mathcal{O}_{X}. Denote the image of α\displaystyle\alpha in Ext(,)\displaystyle\mathrm{Ext}^{*}(\mathcal{F},\mathcal{F}) by α\displaystyle\alpha_{\mathcal{F}}. For any vector bundle \displaystyle\mathcal{F} on X\displaystyle X, Căldăraru and Willerton [CW10] defined an abstract Chern character ch()\displaystyle\mathrm{ch}(\mathcal{F}) which lies in the degree zero part of the Hochschild homology HH(X)=ExtX×X(SΔ1,𝒪Δ)\displaystyle\mathrm{HH}_{*}(X)=\mathrm{Ext}_{X\times X}^{*}(S^{-1}_{\Delta},\mathcal{O}_{\Delta}), where SΔ1=Δ(ωX[dimX])\displaystyle S^{-1}_{\Delta}=\Delta_{*}(\omega_{X}^{\vee}[-\mathrm{dim}X]). There is an HKR isomorphism for Hochschild homology

IHKR:HH(X)HΩ(X)=qp=Hp(X,qΩX).\displaystyle I_{HKR}:\mathrm{HH}_{*}(X)\rightarrow\mathrm{H\Omega}_{*}(X)=\bigoplus_{q-p=*}H^{p}(X,\wedge^{q}\Omega_{X}).

The image of the abstract Chern character under the map IHKR\displaystyle I_{HKR} is the usual Chern character of \displaystyle\mathcal{F} [C05]. We need the lemma below.

{Lemma}

If α\displaystyle\alpha_{\mathcal{F}} is zero, then αch()\displaystyle\alpha\circ\mathrm{ch}(\mathcal{F}) is zero. Here \displaystyle\circ is the composition of morphisms in 𝐃b(X×X)\displaystyle\mathbf{D}^{b}(X\times X) and ch()\displaystyle\mathrm{ch}(\mathcal{F}) is the abstract Chern character.

Proof 3.1.

The proof is known in an email correspondence with Eyal Markman. Let β\displaystyle\beta be any class in ExtX×X(𝒪Δ,SΔ)\displaystyle\mathrm{Ext}_{X\times X}^{*}(\mathcal{O}_{\Delta},S_{\Delta}), where SΔ=Δ(ωX[dimX])\displaystyle S_{\Delta}=\Delta_{*}(\omega_{X}[\mathrm{dim}X]). Similar to the definition of the class α\displaystyle\alpha_{\mathcal{F}} associated to αExtX×X(𝒪Δ,𝒪Δ)\displaystyle\alpha\in\mathrm{Ext}_{X\times X}^{*}(\mathcal{O}_{\Delta},\mathcal{O}_{\Delta}), we get a class βExtX(,SX)\displaystyle\beta_{\mathcal{F}}\in\mathrm{Ext}_{X}^{*}(\mathcal{F},S_{X}\mathcal{F}), where SX()=ωX[dimX]\displaystyle S_{X}(-)=\omega_{X}[\mathrm{dim}X]\otimes-. It is shown in [C05] that the class ch()\displaystyle\mathrm{ch}(\mathcal{F}) is characterized by the identity

TrX×X(βch())=TrX(β).\displaystyle\mathrm{Tr}_{X\times X}(\beta\circ\mathrm{ch}(\mathcal{F}))=\mathrm{Tr}_{X}(\beta_{\mathcal{F}}).

Due to the equality above, we have

TrX×X(γαch())=TrX((γα))=TrX(γα)\displaystyle\mathrm{Tr}_{X\times X}(\gamma\circ\alpha\circ\mathrm{ch}(\mathcal{F}))=\mathrm{Tr}_{X}((\gamma\circ\alpha)_{\mathcal{F}})=\mathrm{Tr}_{X}(\gamma_{\mathcal{F}}\circ\alpha_{\mathcal{F}})

for any γExtX×X(𝒪Δ,SΔ)\displaystyle\gamma\in\mathrm{Ext}_{X\times X}^{*}(\mathcal{O}_{\Delta},S_{\Delta}). The right hand side is zero since we assume that α\displaystyle\alpha_{\mathcal{F}} is zero. We can conclude that αch()\displaystyle\alpha\circ\mathrm{ch}(\mathcal{F}) is zero because the equality TrX×X(γαch())=0\displaystyle\mathrm{Tr}_{X\times X}(\gamma\circ\alpha\circ\mathrm{ch}(\mathcal{F}))=0 holds for any γ\displaystyle\gamma and Tr()\displaystyle\mathrm{Tr}(-) is non-degenerate.

The two HKR isomorphisms IHKR\displaystyle I_{HKR} and IHKR\displaystyle I^{HKR} can be twisted by the Todd class. We denote the resulting twisted isomorphisms by IK\displaystyle I_{K} and IK\displaystyle I^{K}

IK:HH(X)HΩ(X)=qp=Hp(X,qΩX),\displaystyle I_{K}:\mathrm{HH}_{*}(X)\rightarrow\mathrm{H\Omega}_{*}(X)=\displaystyle{\bigoplus_{q-p=*}H^{p}(X,\wedge^{q}\Omega_{X})},
IK:HT(X)=p+q=Hp(X,qTX)HH(X).\displaystyle I^{K}:\mathrm{HT}^{*}(X)=\bigoplus_{p+q=*}H^{p}(X,\wedge^{q}T_{X})\rightarrow\mathrm{HH}^{*}(X).

They are given by the formula IK=(Td(X)12)IHKR\displaystyle I_{K}=(-\wedge Td(X)^{\frac{1}{2}})\circ I_{HKR} and IK=IHKRD1\displaystyle I^{K}=I^{HKR}\circ D^{-1}, where D1\displaystyle D^{-1} is the inverse of the Duflo operator.

The Mukai vector v()\displaystyle v(\mathcal{F}) of \displaystyle\mathcal{F} is IK(ch())\displaystyle I_{K}(\mathrm{ch}(\mathcal{F})) by definition. There are natural ring structures on HH(X)\displaystyle\mathrm{HH}^{*}(X) and HT(X)\displaystyle\mathrm{HT}^{*}(X): the product on HH(X)\displaystyle\mathrm{HH}^{*}(X) is the Yoneda product, and the product on HT(X)\displaystyle\mathrm{HT}^{*}(X) is the wedge product. Kontsevich [Kont03] claimed that the map IK\displaystyle I^{K} is a ring isomorphism. This statement was proved by Calaque and Van den Bergh [CV10]. The Hochschild homology is a module over the Hochschild cohomology and similarly HΩ(X)\displaystyle\mathrm{H\Omega}_{*}(X) is a module over HT(X)\displaystyle\mathrm{HT}^{*}(X). Calaque, Rossi, and Van den Bergh [CRV12] proved that the maps IK\displaystyle I_{K} and IK\displaystyle I^{K} respect the module structures. {Proof}[Proof of Theorem B.] The commutative diagram in Theorem A shows that

α~exp(at)=α,\displaystyle\tilde{\alpha}\lrcorner\exp(at_{\mathcal{F}})=\alpha_{\mathcal{F}},

which is zero under the assumption of Theorem B. We conclude that αch()\displaystyle\alpha\circ\mathrm{ch}(\mathcal{F}) is zero by Lemma 3. Since IK\displaystyle I_{K} and IK\displaystyle I^{K} respect the module structures, we have

0=IK(αch())=(IK)1(α)IK(ch())=(IK)1(α)v().\displaystyle 0=I_{K}(\alpha\circ\mathrm{ch}(\mathcal{F}))=(I^{K})^{-1}(\alpha)\lrcorner\,I_{K}(\mathrm{ch}(\mathcal{F}))=(I^{K})^{-1}(\alpha)\lrcorner\,v(\mathcal{F}).

The inverse map of IK\displaystyle I^{K} is the composite map

(IK)1:HH(X)\displaystyle\textstyle{(I^{K})^{-1}:\mathrm{HH}^{*}(X)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}(IHKR)1\displaystyle\scriptstyle{~{}~{}~{}~{}~{}~{}(I^{HKR})^{-1}}HT(X)\displaystyle\textstyle{\mathrm{HT}^{*}(X)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}D\displaystyle\scriptstyle{D}HT(X).\displaystyle\textstyle{\mathrm{HT}^{*}(X).}

As a consequence

0=IK(αch())=(IK)1(α)v()=D(α~)v().\displaystyle 0=I_{K}(\alpha\circ\mathrm{ch}(\mathcal{F}))=(I^{K})^{-1}(\alpha)\lrcorner\,v(\mathcal{F})=D(\tilde{\alpha})\lrcorner\,v(\mathcal{F}).

The special case when α~H1(X,TX)\displaystyle\tilde{\alpha}\in H^{1}(X,T_{X}).

The result in (LABEL:1.1) says that α~v()\displaystyle\tilde{\alpha}\lrcorner\,v(\mathcal{F}) is zero if α~exp(at)\displaystyle\tilde{\alpha}\lrcorner\exp(at_{\mathcal{F}}) is zero for any α~H1(X,TX)\displaystyle\tilde{\alpha}\in H^{1}(X,T_{X}). We end this paper by proving that Theorem B implies the result in (LABEL:1.1).

From now on let α~\displaystyle\tilde{\alpha} be an element in H1(X,TX)\displaystyle H^{1}(X,T_{X}). The only term in exp(at)=1+at+(at)22!+\displaystyle\exp(at_{\mathcal{F}})=1+at_{\mathcal{F}}+\frac{(at_{\mathcal{F}})^{2}}{2!}+\cdots that can contract with α~\displaystyle\tilde{\alpha} is at\displaystyle at_{\mathcal{F}}, so α~exp(at)=α~at\displaystyle\tilde{\alpha}\lrcorner\exp(at_{\mathcal{F}})=\tilde{\alpha}\lrcorner\,at_{\mathcal{F}} in this case.

Choose =OX\displaystyle\mathcal{F}=O_{X}. We have α~exp(atOX)=0\displaystyle\tilde{\alpha}\lrcorner\exp(at_{O_{X}})=0. Therefore

D(α~)v(OX)=(Td(X)12α~)Td(X)12=0\displaystyle D(\tilde{\alpha})\lrcorner\,v(O_{X})=(Td(X)^{\frac{1}{2}}\lrcorner\,\tilde{\alpha})\lrcorner\,Td(X)^{\frac{1}{2}}=0

according to Theorem B.

Expand the Todd class Td(X)\displaystyle Td(X) as 1+c12+c12+c212+\displaystyle 1+\frac{c_{1}}{2}+\frac{c_{1}^{2}+c_{2}}{12}+\cdots, and note that the only term of (Td(X)12α~)Td(X)12\displaystyle(Td(X)^{\frac{1}{2}}\lrcorner\,\tilde{\alpha})\lrcorner\,Td(X)^{\frac{1}{2}} in H2(X,OX)\displaystyle H^{2}(X,O_{X}) is α~c12\displaystyle\tilde{\alpha}\lrcorner\,\frac{c_{1}}{2}. Since (Td(X)12α~)Td(X)12=0\displaystyle(Td(X)^{\frac{1}{2}}\lrcorner\,\tilde{\alpha})\lrcorner\,Td(X)^{\frac{1}{2}}=0, we can conclude that α~c1\displaystyle\tilde{\alpha}\lrcorner\,c_{1} is zero for any α~H1(X,TX)\displaystyle\tilde{\alpha}\in H^{1}(X,T_{X}). The fact that α~c1=0\displaystyle\tilde{\alpha}\lrcorner\,c_{1}=0 for α~H1(X,TX)\displaystyle\tilde{\alpha}\in H^{1}(X,T_{X}) is also known due to Griffiths. Consider the first order deformation of X\displaystyle X corresponding to α~\displaystyle\tilde{\alpha}. The vanishing of α~c1\displaystyle\tilde{\alpha}\lrcorner\,c_{1} is equivalent to the class c1\displaystyle c_{1} remaining of type (p,p)\displaystyle(p,p).

The term α~c14\displaystyle\tilde{\alpha}\lrcorner\,\frac{c_{1}}{4} is exactly the difference between D(α~)\displaystyle D(\tilde{\alpha}) and α~\displaystyle\tilde{\alpha} because

D(α~)=Td(X)12α~=(1+c14+)α~=α~+c14α~+0.\displaystyle D(\tilde{\alpha})=Td(X)^{\frac{1}{2}}\lrcorner\,\tilde{\alpha}=(1+\frac{c_{1}}{4}+\cdots)\lrcorner\,\tilde{\alpha}=\tilde{\alpha}+\frac{c_{1}}{4}\lrcorner\,\tilde{\alpha}+0.

We conclude that α~v()\displaystyle\tilde{\alpha}\lrcorner\,v(\mathcal{F}) is zero if and only if D(α~)v()\displaystyle D(\tilde{\alpha})\lrcorner\,v(\mathcal{F}) is zero for α~H1(X,TX)\displaystyle\tilde{\alpha}\in H^{1}(X,T_{X}).

References

  • [AC12] D. Arinkin, and A. Căldăraru, When is the self-intersection a fibration?, Adv. Math. 𝟐𝟑𝟏\displaystyle\mathbf{231} (2012), no.2, 815-842.
  • [B94] I. Biswas, A remark on a deformation theory of Green and Lazarsfeld, J. Reine Angew Math. 𝟒𝟒𝟗\displaystyle\mathbf{449} (1994), 103-124.
  • [C05] A. Căldăraru, The Mukai pairing II, The Hochschild-Kostant-Rosenberg isomorphism, Adv. Math. 𝟏𝟗𝟒\displaystyle\mathbf{194} (2005), no.1, 34-66.
  • [CI17] J. Carlson and S. Iyengar, Hopf algebra structures and tensor products for group algebras, New York J. Math. 𝟐𝟑\displaystyle\mathbf{23} (2017), 351-364.
  • [CRV12] D. Calaque, C. A. Rossi, and M. Van den Bergh, Căldăraru’s conjecture and Tsygan’s formality, Annals of Mathematics 𝟏𝟕𝟔\displaystyle\mathbf{176} (2012), 865-923.
  • [CV10] D. Calaque and M. Van den Bergh, Hochschild cohomology and Atiyah classes, Adv. Math. 𝟐𝟐𝟒\displaystyle\mathbf{224} (2010), no. 5, 1839-1889.
  • [CW10] A. Căldăraru, and S. Willerton, The Mukai pairing, I a categorical approach, New York J. Math. 𝟏𝟔\displaystyle\mathbf{16} (2010), 61-98.
  • [K99] M. Kapranov, Rozansky-Witten weight invariants via Atiyah classes, Compositio Math. 𝟏𝟏𝟓\displaystyle\mathbf{115} (1999), no.1, 71-113.
  • [Kont03] M. Kontsevich, Deformation quantization of Poisson manifolds, I, Lett.Math.Phys. 𝟔𝟔\displaystyle\mathbf{66} (2003), 157-216.
  • [PW09] J. Pevtsova and S. Witherspoon, Varieties for modules of quantum elementary abelian groups, Algebr. Represent. Theory 𝟏𝟐\displaystyle\mathbf{12} (2009), no.6 567-595.
  • [RW10] J. Roberts, and S. Willerton, On the Rozansky-Witten weight systems, Algebr. Geom. Topol. 𝟏𝟎\displaystyle\mathbf{10} (2010), no.3, 1455-1519.
  • [T09] Y. Toda, Deformations and Fourier-Mukai transforms, J. Differential Geom. 𝟖𝟏\displaystyle\mathbf{81} (2009), no.1, 197-224.