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

A short proof of the Hanlon-Hicks-Lazarev Theorem

Michael K. Brown  and  Daniel Erman
Abstract.

We give a short, new proof of a recent result of Hanlon-Hicks-Lazarev about toric varieties. As in their work, this leads to a proof of a conjecture of Berkesch-Erman-Smith on virtual resolutions and to a resolution of the diagonal in the simplicial case.

The second author was supported by NSF grant DMS-2200469.

1. Main result

We give a short, new proof of a recent result of Hanlon-Hicks-Lazarev about toric varieties and their multigraded Cox rings. Throughout, we let XX be a simplicial, projective toric variety over an algebraically closed field kk with Cl(X)\operatorname{Cl}(X)-graded Cox ring SS. Our main result (Theorem 1.2) was first proven in [HHL], but our proof is independent from their methods. Our approach is more algebraic and simpler, while their approach is more explicit and connects to a wider range of topics, including symplectic geometry and homological mirror symmetry. See also the work of Favero-Huang [FH], which was completed simultaneously with [HHL] and whose main results coincide with some of Hanlon-Hicks-Lazarev’s.

Our interest in these topics begins with a program to extend results on syzygies to multigraded or toric settings. The basic perspective, introduced by Berkesch-Erman-Smith in [BES], is that many classical results about minimal free resolutions will have strong analogues in the toric setting, as long as one replaces minimal free resolutions with the more flexible notion of a virtual resolution.

Definition 1.1.

Let MM be a finitely generated Cl(X)\operatorname{Cl}(X)-graded SS-module. A virtual resolution of MM is a free complex FF_{\bullet} of SS-modules such that there is a quasi-isomorphism F~M~\widetilde{F_{\bullet}}\xrightarrow{\simeq}\widetilde{M} of complexes of 𝒪X\mathcal{O}_{X}-modules.111If XX is smooth, then F~\widetilde{F_{\bullet}} consists of sums of line bundles and is sometimes called a line bundle resolution. See Remark 3.4 regarding the simplicial case.

The following is a consequence of Hanlon-Hicks-Lazarev’s main result [HHL, Theorem A].

Theorem 1.2.

Let YY be a normal toric variety and YXY\hookrightarrow X a closed immersion that is a toric morphism [CLS, Definition 3.3.3]. Denote by II the defining ideal of YXY\hookrightarrow X (Definition 2.1). The SS-module S/IS/I admits a virtual resolution of length codim(YX)\operatorname{codim}(Y\subseteq X).

And here is our short proof of Theorem 1.2. The proof relies on some elementary facts about toric varieties that we recall in Lemma 2.2 below.

Proof of Theorem 1.2.

The Cox ring SS of XX is positively Cl(X)\operatorname{Cl}(X)-graded [BE, Definition A.1, Example A.2], and so we may consider Cl(X)\operatorname{Cl}(X)-graded minimal free resolutions of Cl(X)\operatorname{Cl}(X)-graded SS-modules. Let RR be the normalization of S/IS/I and FF_{\bullet} the minimal free resolution of RR as an SS-module. Since YY is normal, R~=𝒪Y\widetilde{R}={\mathcal{O}}_{Y} as a sheaf on XX, and so FF_{\bullet} is a virtual resolution of S/IYS/I_{Y}. By Lemma 2.2(1) and [CLS, Theorem 1.1.17 and Proposition 1.3.8], the ring RR is a product of affine semigroup rings of the same dimension. Hochster’s Theorem therefore implies that each component of RR is a Cohen-Macaulay ring [hochster, Theorem 1]. It follows that RR is also a Cohen-Macaulay S/IS/I-module: indeed, we have dim(R)=dim(S/I)\dim(R)=\dim(S/I), and since RR is a finitely generated S/IS/I-module [eisenbudbook, Theorem 4.14], any system of parameters on S/IS/I is a system of parameters on each component of RR and hence a regular sequence. The length of FF_{\bullet} is the projective dimension of RR, which, by the Auslander-Buchsbaum formula [eisenbudbook, Theorem 19.9], is equal to 0ptS(S)0ptS(R)=dim(S)dim(S/I)0pt_{S}(S)-0pt_{S}(R)=\dim(S)-\dim(S/I) (while the version of the Auslander-Buchsbaum formula we cite pertains to local rings, the desired result for the polynomial ring SS follows by [BH, Proposition 1.5.15]). Lemma 2.2(2) therefore implies that the length of FF is equal to codim(YX)\operatorname{codim}(Y\subseteq X). ∎

We now describe applications of Theorem 1.2 and their history. For a fuller discussion, see [HHL, §1]. We start with a special case, first proven by Hanlon-Hicks-Lazarev:

Theorem 1.3 ([HHL] Corollary B).

The coordinate ring of the diagonal embedding XX×XX\subseteq X\times X admits a virtual resolution of length dimX\dim X.

Special cases of Theorem 1.3 were studied in [BE, brown-sayrafi, canonaco], and  [BPS, anderson] study closely related questions. It was known that this result would immediately yield proofs of two conjectures that also had received independent interest. The first conjecture is due to Berkesch-Erman-Smith [BES, Question 1.3] and was proven by Hanlon-Hicks-Lazarev:

Theorem 1.4 ([HHL] Corollary C).

Any module MM as in Definition 1.1 has a virtual resolution of length dimX\leq\dim X.

Hilbert’s Syzygy Theorem gives a bound of dimS=dimX+rankCl(X)\dim S=\dim X+\operatorname{rank}\operatorname{Cl}(X); Theorem 1.4 implies that the added flexibility of virtual resolutions allows for significantly shorter resolutions, especially when rankCl(X)\operatorname{rank}\operatorname{Cl}(X) is large. See [BES, HNV, berkesch-klein-loper-yang] and elsewhere for many examples of this phenomenon. Prior to [HHL], Theorem 1.4 had been proven in several special cases: when rankPic(X)=1\operatorname{rank}\operatorname{Pic}(X)=1 it essentially follows from Hilbert’s Syzygy Theorem; for products of projective spaces it was shown in [BES, Theorem 1.2] (see also [EES, Corollary 1.14]); Yang proved it for any monomial ideal in the Cox ring of a smooth toric variety [yang]; and Brown-Sayrafi proved it for smooth projective toric varieties of Picard rank 2 [brown-sayrafi].

The second conjecture, due to Orlov, is the special case of [orlov, Conjecture 10] for toric varieties. This was first proven by Favero-Huang in [FH, Theorem 1.2], and independently and essentially simultaneously in  [HHL, Corollary E].

Theorem 1.5.

The Rouquier dimension of Db(X)D^{b}(X) equals dimX\dim X.

Special cases of Theorem 1.5 had been established in [BC, BF, BDM, BFK] before Favero-Huang and Hanlon-Hicks-Lazarev proved it in general. The full version of Orlov’s Conjecture states that Theorem 1.5 extends to any smooth quasi-projective variety; see [BC, §1.2] for a list of known cases of this conjecture.

Theorem 1.2 easily implies Theorems 1.3,  1.4 and 1.5. To prove Theorem 1.3, observe that the diagonal XX×XX\subseteq X\times X satisfies the conditions of Theorem 1.2. To prove Theorem 1.4, one can simply follow the method of [BES, Proof of Theorem 1.2]. For Theorem 1.5, one can use standard techniques on derived categories; see, e.g., the proof of  [HHL, Corollary E].

Our proof of Theorem 1.2 is quite simple, perhaps embarrassingly so given the prior partial results on these questions cited above. It is not yet clear how to compare our resolutions to those obtained in [HHL], but we believe that the two constructions agree in the case of Theorem 1.3. Their work gives a creative perspective on building these resolutions, drawing motivation from the symplectic side of the mirror symmetry functor and involving a wide array of ideas.222In a different direction, we refer to Borisov’s work [borisov] for an alternative proof of Hochster’s Theorem [hochster, Theorem 1]—the main ingredient of our proof of Theorem 1.2—and an explanation of how the techniques used there relate to mirror symmetry. The resolutions they obtain are quite explicit; indeed, their resolution of the diagonal yields a canonical generating set for the derived category of any normal toric variety, proving a claim of Bondal [HHL, Corollary D]. However, some algebraic aspects of their constructions are harder to determine. For instance, if FF_{\bullet} is the free complex of SS-modules corresponding to one of their resolutions, their work implies that the modules Hi(F)H_{i}(F_{\bullet}) correspond to the zero sheaf on XX for all i>0i>0, but it is not clear whether Hi(F)H_{i}(F_{\bullet}) equals the zero module on the nose, i.e. it is not clear if FF_{\bullet} is acyclic as a complex of SS-modules. The SS-module that arises as H0(F)H_{0}(F_{\bullet}) is also unclear. By comparison, the complexes that arise in our construction are always acyclic, and they resolve normalizations of coordinate rings. However, we are not able to give as explicit of a description of the terms. It would be very interesting to better compare these complexes, and to compare them with those in [BE, brown-sayrafi]. Favero-Huang’s approach [FH] can almost certainly yield all of the above results as well, and it would be interesting to compare to those resolutions too.

Remark 1.6.

As our resolutions from Theorem 1.2 rely only on standard algebraic constructions, they can be directly computed in Macaulay2 [M2]. The constructions in [HHL] are explicit, but due to their novelty, computing them in practice requires more effort. Of course, if one could show that the two constructions coincide, this would shed more light on both.

2. Some elementary facts about toric varieties

Definition 2.1.

Let XX, YY, and SS be as in Theorem 1.2, BSB\subseteq S the irrelevant ideal of XX, and ZZ the closure in Spec(S)\operatorname{Spec}(S) of the inverse image of YY under the canonical surjection π:Spec(S)V(B)X\pi\colon\operatorname{Spec}(S)\setminus V(B)\to X. The defining ideal of YY in XX is the radical ideal ISI\subseteq S corresponding to the closed subset ZSpec(S)Z\subseteq\operatorname{Spec}(S).

Lemma 2.2.

Let ZZ and II be as in Definition 2.1.

  1. (1)

    The irreducible components of ZZ are affine toric varieties of the same dimension. Furthermore, if the divisor class group Cl(X)\operatorname{Cl}(X) is torsion-free, then ZZ is irreducible.

  2. (2)

    We have dim(S)dim(S/I)=codim(YX)\dim(S)-\dim(S/I)=\operatorname{codim}(Y\subseteq X).

Proof.

Since YXY\hookrightarrow X is a toric morphism, it induces an embedding TYTXT_{Y}\hookrightarrow T_{X} on tori and hence a surjection p:MXMYp\colon M_{X}\twoheadrightarrow M_{Y} of lattices. Taking the pushout of the surjection pp and the canonical map MXdimSM_{X}\to\mathbb{Z}^{\dim{S}} yields the morphism

(2.3) 0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}MX\textstyle{M_{X}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}p\scriptstyle{p}dimS\textstyle{\mathbb{Z}^{\dim S}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}q\scriptstyle{q}Cl(X)\textstyle{\operatorname{Cl}(X)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\cong}0\textstyle{0}0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}MY\textstyle{M_{Y}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}M\textstyle{M^{\prime}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Cl(X)\textstyle{\operatorname{Cl}(X)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0\textstyle{0}

of exact sequences. The abelian group MM^{\prime} is isomorphic to rA\mathbb{Z}^{r}\oplus A, where rr is defined to be dim(S)dim(X)+dim(Y)\dim(S)-\dim(X)+\dim(Y), and AA is some finite abelian group. We observe that II coincides with the radical of Jker(Sk[dimS]𝑞k[M])J\coloneqq\ker(S\hookrightarrow k[\mathbb{Z}^{\dim{S}}]\xrightarrow{q}k[M^{\prime}]); note that k[M]k[M^{\prime}] need not be reduced when char(k)0\operatorname{char}(k)\neq 0, since MM^{\prime} may have torsion, and so JJ need not be radical. Let us verify that I=rad(J)I=\operatorname{rad}(J): since pp is surjective, the Snake Lemma implies that qq is surjective, and so JJ is the defining ideal of the closure of Spec(k[M])\operatorname{Spec}(k[M^{\prime}]) in Spec(S)\operatorname{Spec}(S). Diagram (2.3) induces the following morphism of short exact sequences of algebraic groups:

0\textstyle{0}TX\textstyle{\ignorespaces\ignorespaces\ignorespaces\ignorespaces T_{X}}Spec(k[dimS])\textstyle{\ignorespaces\ignorespaces\ignorespaces\ignorespaces\operatorname{Spec}(k[\mathbb{Z}^{\dim S}])}α\scriptstyle{\alpha}ker(α)\textstyle{\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ker(\alpha)}0\textstyle{\ignorespaces\ignorespaces\ignorespaces\ignorespaces 0}0\textstyle{0}TY\textstyle{\ignorespaces\ignorespaces\ignorespaces\ignorespaces T_{Y}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Spec(k[M])\textstyle{\ignorespaces\ignorespaces\ignorespaces\ignorespaces\operatorname{Spec}(k[M^{\prime}])\ignorespaces\ignorespaces\ignorespaces\ignorespaces}β\scriptstyle{\beta}ker(β)\textstyle{\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ker(\beta)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\cong}0.\textstyle{\ignorespaces\ignorespaces\ignorespaces\ignorespaces 0.}

It follows that α1(TY)\alpha^{-1}(T_{Y}) is equal to the image of Spec(k[M])\operatorname{Spec}(k[M^{\prime}]) in Spec(k[dimS])\operatorname{Spec}(k[\mathbb{Z}^{\dim S}]). Since ZZ is equal to the closure of α1(TY)\alpha^{-1}(T_{Y}) in Spec(S)\operatorname{Spec}(S), we conclude that I=rad(J)I=\operatorname{rad}(J).

Writing R=k[r]R=k[\mathbb{Z}^{r}] and A=i=1t/(ni)A=\bigoplus_{i=1}^{t}\mathbb{Z}/(n_{i}), we have

k[M]R[z1,,zt]/(z1n11,,ztnt1).k[M^{\prime}]\cong R[z_{1},\dots,z_{t}]/(z_{1}^{n_{1}}-1,\dots,z_{t}^{n_{t}}-1).

The quotient of k[M]k[M^{\prime}] by its nilradical is therefore a product of copies of RR, and so II is a finite intersection of prime ideals arising as kernels of ring homomorphisms SRS\to R. It therefore follows from [CLS, Proposition 1.1.8] that the irreducible components of ZZ are affine toric varieties of dimension rr. If Cl(X)\operatorname{Cl}(X) is torsion-free, then the bottom row of Diagram (2.3) splits, and so A=0A=0, which means II is prime. This proves (1). As for (2): we have shown that dim(Z)=r\dim(Z)=r, which is precisely dim(S)codim(YX)\dim(S)-\operatorname{codim}(Y\subseteq X). ∎

3. Examples

Example 3.1.

Let X=nX=\mathbb{P}^{n} and T=k[x0,,xn,y0,,yn]T=k[x_{0},\dots,x_{n},y_{0},\dots,y_{n}], the Cox ring of X×XX\times X. Let IΔTI_{\Delta}\subseteq T be the defining ideal (Definition 2.1) of the diagonal XX×XX\subseteq X\times X, i.e. the ideal corresponding to the closure of the set of points in Spec(T)\operatorname{Spec}(T) of the form (x0,,xn,tx0,,txn)(x_{0},\dots,x_{n},tx_{0},\dots,tx_{n}), where tkt\in k^{*}. One easily checks that IΔI_{\Delta} is the kernel of the map Sk[x0,,xn,y0,,yn,t]S\to k[x_{0},\dots,x_{n},y_{0},\dots,y_{n},t] given by xixix_{i}\mapsto x_{i} and yitxiy_{i}\mapsto tx_{i}, and so T/IΔT/I_{\Delta} is isomorphic to the normal semigroup ring k[x0,,xn,tx0,,txn]k[x_{0},\dots,x_{n},tx_{0},\dots,tx_{n}]. The ideal IΔI_{\Delta} is generated by the 2×22\times 2 minors of the matrix (x0x1xny0y1yn).\begin{pmatrix}x_{0}&x_{1}&\cdots&x_{n}\\ y_{0}&y_{1}&\cdots&y_{n}\end{pmatrix}. More specifically: these minors vanish on Δ\Delta, and since this is a generic matrix, the ideal of 2×22\times 2 minors is prime of codimension nn. As T/IΔT/I_{\Delta} is already normal, the virtual resolution of T/IΔT/I_{\Delta} arising from Theorem 1.2 is just the minimal free resolution of T/IΔT/I_{\Delta}, which is given by the Eagon-Northcott complex on this matrix.

Example 3.2.

Let XX be the weighted projective space (1,1,2)\mathbb{P}(1,1,2) and TT the Cox ring k[x0,x1,x2,y0,y1,y2]k[x_{0},x_{1},x_{2},y_{0},y_{1},y_{2}] of X×XX\times X. By a calculation similar to Example 3.1, the ring T/IΔT/I_{\Delta} is isomorphic to the semigroup ring k[x0,x1,x2,tx0,tx1,t2x2],k[x_{0},x_{1},x_{2},tx_{0},tx_{1},t^{2}x_{2}], which is not normal because tx2tx_{2} lies in the fraction field and satisfies the integral equation (tx2)2x2(t2x2)=0(tx_{2})^{2}-x_{2}\cdot(t^{2}x_{2})=0. Let RR be the normalization of T/IΔT/I_{\Delta}. A presentation matrix for RR as a TT-module is given as follows, where the rows correspond to the generators 11 and tx2tx_{2}:

1( x1y0x0y1x2y0x2y1x0y2x1y2) tx20x0x1y0y1.\bordermatrix{&&&&&\cr 1&x_{1}y_{0}-x_{0}y_{1}&x_{2}y_{0}&x_{2}y_{1}&x_{0}y_{2}&x_{1}y_{2}\cr tx_{2}&0&-x_{0}&-x_{1}&-y_{0}&-y_{1}}.

The free resolution of RR as a TT-module is given by:

(3.3) TT(1,1)[x1y0x0y1x2y0x2y1x0y2x1y20x0x1y0y1]T(1,1)T(2,1)2T(1,2)2[x20y2x1y10x0y000x1y10x0y0]T(3,1)T(2,2)T(1,3)0.\begin{matrix}T\\ \oplus\\ T(-1,-1)\end{matrix}\xleftarrow{\left[\begin{smallmatrix}x_{1}y_{0}-x_{0}y_{1}&x_{2}y_{0}&x_{2}y_{1}&x_{0}y_{2}&x_{1}y_{2}\\ 0&-x_{0}&-x_{1}&-y_{0}&-y_{1}\end{smallmatrix}\right]}\begin{matrix}T(-1,-1)\\ \oplus\\ T(-2,-1)^{2}\\ \oplus\\ T(-1,-2)^{2}\end{matrix}\xleftarrow{\left[\begin{smallmatrix}-x_{2}&0&-y_{2}\\ x_{1}&-y_{1}&0\\ -x_{0}&y_{0}&0\\ 0&-x_{1}&-y_{1}\\ 0&x_{0}&y_{0}\end{smallmatrix}\right]}\begin{matrix}T(-3,-1)\\ \oplus\\ T(-2,-2)\\ \oplus\\ T(-1,-3)\end{matrix}\leftarrow 0.

Additionally: we have the short exact sequence 0T/IΔRQ0,0\to T/I_{\Delta}\to R\to Q\to 0, and Q=tx2k[x2,y2]Q=tx_{2}\cdot k[x_{2},y_{2}]. One can directly compute that the sheaf Q~\widetilde{Q} corresponding to QQ is the zero sheaf on X×XX\times X. In fact, since QQ is annihilated by x0,x1,y0x_{0},x_{1},y_{0} and y1y_{1}, we can reduce to checking that Q~\widetilde{Q} is also zero on the affine patch D(x2y2)D(x_{2}y_{2}). The global sections of Q~\widetilde{Q} on this patch are Q[x21,y21](0,0)=0Q[x_{2}^{-1},y_{2}^{-1}]_{(0,0)}=0, and thus Q~=0\widetilde{Q}=0 as desired.

Remark 3.4.

Since 𝒪(1)\mathcal{O}(-1) and 𝒪(3)\mathcal{O}(-3) are not vector bundles on (1,1,2)\mathbb{P}(1,1,2), the resolution (3.3) does not induce a locally free resolution of the diagonal. Indeed, virtual resolutions are not guaranteed to induce locally free resolutions of 𝒪X\mathcal{O}_{X}-modules unless XX is smooth. Alternatively, as in [HHL], one could consider the corresponding toric stack.

Remark 3.5.

In many of the prior known cases of Theorem 1.4, a slightly stronger result was proven. Namely, it was shown that for any such MM, there exists another module MM^{\prime} satisfying M~=M~\widetilde{M}=\widetilde{M^{\prime}} and pdim(M)dimX\operatorname{pdim}(M^{\prime})\leq\dim X; see [EES, bruce-heller-sayrafi, yang]. It would be interesting to determine if this was true in general.

Acknowledgments

We are very grateful to Andrew Hanlon, Jeff Hicks, and Oleg Lazarev for patiently talking to us about their work and for several inspiring conversations. We only found this approach because of our efforts to understand their beautiful results. We also thank Christine Berkesch, Lauren Cranton Heller, Mahrud Sayrafi, and Jay Yang for helpful comments and discussions. Finally, we thank the anonymous referee for many helpful suggestions.

References


Department of Mathematics and Statistics, Auburn University, Auburn, AL

E-mail address: [email protected]

Department of Mathematics, University of Hawai‘i at Mānoa, Honolulu, HI

E-mail address: [email protected]