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

Some examples of noncommutative projective Calabi-Yau schemes

Yuki Mizuno [email protected] Department of Mathematics, School of Science and Engineering, Waseda University, Ohkubo 3-4-1, Shinjuku, Tokyo 169-8555, Japan
Abstract.

In this article, we construct some examples of noncommutative projective Calabi-Yau schemes by using noncommutative Segre products and quantum weighted hypersurfaces. We also compare our constructions with commutative Calabi-Yau varieties and examples constructed in [11]. In particular, we show that some of our constructions are essentially new examples of noncommutative projective Calabi-Yau schemes.

Key words and phrases:
Noncommutative algebraic geometry, Calabi-Yau varieties
2020 Mathematics Subject Classification:
14A22, 14J32

1. Introduction

Calabi-Yau varieties are rich objects and play an important role in mathematics and physics. In noncommutative geometry, (skew) Calabi-Yau algebras are often treated as noncommutative analogues of Calabi-Yau varieties. Calabi-Yau algebras have a deep relationship with quiver algebras ([7], [29]). For example, many known Calabi-Yau algebras are constructed by using quiver algebras. They are also used to characterize Artin-Schelter regular algebras ([24], [23]). In particular, a connected graded algebra AA over a field kk is Artin-Schelter regular if and only if AA is skew Calabi-Yau.

On the other hand, a triangulated subcategory of the derived category of a cubic fourfold in 5\mathbb{P}^{5}, which is obtained by some semiorthogonal decompositions, has the 2-shift functor [2][2] as the Serre functor. Moreover, the structure of Hochschild (co)homology is the same as that of a projective K3 surface ([12]). However, some such categories are not obtained as the derived categories of coherent sheaves of projective K3 surfaces and called noncommutative K3 surfaces.

Artin and Zhang constructed a framework of noncommutative projective schemes in [1], which are defined from noncommutative graded algebras. In this framework, we can think of Artin-Schelter regular algebras as noncommutative analogues of projective spaces, which are called quantum projective spaces. Our objective is to produce examples of noncommutative projective Calabi-Yau schemes that are not obtained from commutative Calabi-Yau varieties. In the future, it would be an interesting question to compare the derived category of a noncommutative projective Calabi-Yau scheme created in the framework of Artin-Zhang’s noncommutative projective schemes with a noncommutative K3 surface obtained as a triangulated subcategory of the derived category of a cubic fourfold.

As the definition of noncommutative projective Calabi-Yau schemes, we adopt the definition introduced by Kanazawa ([11]). His definition is a direct generalization of the definition of commutative Calabi-Yau varieties to noncommutative projective schemes. He also constructed the first examples of noncommutative projective Calabi-Yau schemes that are not isomorphic to commutative Calabi-Yau varieties as hypersurfaces of quantum projective spaces. Recently, some examples constructed by Kanazawa play an important role in noncommutative Donaldson-Thomas theory ([13], [14]).

In this paper, we construct new examples of noncommutative projective Calabi-Yau schemes by using noncommutative Segre products and weighted hypersurfaces. There are many known examples of Calabi-Yau varieties in algebraic geometry. Some of them are complete intersections in Segre embeddings of products of projective spaces. Moreover, Reid gave a list of Calabi-Yau surfaces, which are hypersurfaces in weighted projective spaces ([10], [22]). Motivated by these two facts, we construct noncommutative analogues of the two types of examples of Calabi-Yau varieties (Theorem 3.3, Theorem 3.15) in Section 3.

In order to prove that a noncommutative projective scheme is Calabi-Yau, we use the methods of Kanazawa. However, they are not sufficient because the algebras we treat are more complicated than the ones he considered. In order to construct noncommutative projective Calabi-Yau schemes as noncommutative analogues of complete intersections in Segre products, we perform a more detailed analysis of noncommutative projective schemes defined by 2\mathbb{Z}^{2}-graded algebras, which were studied by Van Rompay ([31]). A different approach to noncommutative Segre products is also studied in [8]. In order to construct noncommutative projective Calabi-Yau schemes as noncommutative analogues of weighted hypersurfaces, we consider quotients of weighted quantum polynomial rings. In commutative algebraic geometry, the projective spectrum Proj(k[x0,,xn])\mathrm{Proj}(k[x_{0},\cdots,x_{n}]) of a weighted polynomial ring is not necessarily isomorphic to qgr(k[x0,,xn])\mathrm{qgr}(k[x_{0},\cdots,x_{n}]), where qgr(k[x0,,xn])\mathrm{qgr}(k[x_{0},\cdots,x_{n}]) is the quotient category associated to k[x0,,xn]k[x_{0},\cdots,x_{n}] constructed in [1]. However, qgr(k[x0,,xn])\mathrm{qgr}(k[x_{0},\cdots,x_{n}]) is thought of as a nonsingular model of Proj(k[x0,,xn])\mathrm{Proj}(k[x_{0},\cdots,x_{n}]) (see [25, Example 4.9]). We use this idea to construct new noncommutative projective Calabi-Yau schemes. In addition, it should be noted that local structures of noncommutative projective schemes of quotients of weighted quantum polynomial rings are somewhat complicated. An analysis of the local structures was performed by Smith ([25]). We show that the local structure obtained in [25] is described by the notion of quasi-Veronese algebras introduced by Mori ([17]).

In Section 4, we compare our constructions from weighted hypersurfaces in Section 3 with commutative Calabi-Yau varieties and the first examples constructed in [11], focusing on noncommutative projective Calabi-Yau schemes of dimensions 2. We show that some of our constructions in Section 3 are not isomorphic to any of the commutative Calabi-Yau varieties and the first examples constructed in [11] (Proposition 4.8). When we consider moduli spaces of point modules of noncommutative projective schemes obtained from weighted hypersurfaces in Section 3, there is a problem, which is in general weighted quantum polynomial rings are not generated in degree 11. So, the notion of point modules is not necessarily useful in this case. In this paper, we use theories of closed points studied in [18], [26] and [27], etc. A different approach to closed points of weighted quantum polynomial rings is studied in [28]. The notion of point modules defined in [28] corresponds to those of ordinary and thin points in [18]. To show that some of our constructions are not isomorphic to the examples obtained in [11], we use Morita theory of noncommutative schemes, which is established in [4] (see also [1, Section 6]). In the theory, we need to calculate the centers of noncommutative rings. By using these calculations, we can do a detailed analysis and some classifications of noncommutative projective Calabi-Yau surfaces.

2. Preliminaries

Notation and Terminology 2.1.

In this article, kk means an algebraically closed field of characteristic 0. We suppose \mathbb{N} contains 0. Let AA be a kk-algebra, MM be an AA-bimodule and ψ,ϕ\psi,\phi be algebra automorphisms of AA. Then, we denote the associated AA-bimodule by Mϕψ{}^{\psi}M^{\phi}, i.e. Mϕψ=M{}^{\psi}M^{\phi}=M as kk-modules and the new bimodule structure is given by amb:=ψ(a)mϕ(b)a*m*b:=\psi(a)m\phi(b) for all a,bAa,b\in A and all mMm\in M. Let 𝒞\mathcal{C} be a kk-linear abelian category. We denote the global dimension of 𝒞\mathcal{C} by gl.dim(𝒞)\mathrm{gl.dim}(\mathcal{C}). An \mathbb{N}-graded kk-algebra AA is connected if A0=kA_{0}=k.

For any \mathbb{N}-graded kk-algebra A=i=0AiA=\bigoplus_{i=0}^{\infty}A_{i}, we denote the category of graded right AA-modules (resp. finitely generated graded right AA-modules) by Gr(A)\mathrm{Gr}(A) (resp. gr(A)\mathrm{gr}(A)). Let MGr(A)M\in\mathrm{Gr}(A) and AA^{\circ} be the opposite algebra of AA. We define the Matlis dual MGr(A)M^{*}\in\mathrm{Gr}(A^{\circ}) by Mi:=Homk(Mi,k)M^{*}_{i}:=\mathrm{Hom}_{k}(M_{-i},k) and the shift M(n)Gr(A)M(n)\in\mathrm{Gr}(A) by M(n)i:=Mi+n(i,n)M(n)_{i}:=M_{i+n}\ (i,n\in\mathbb{Z}). For M,NGr(A)M,N\in\mathrm{Gr}(A), we write HomA(M,N):=nHomGr(A)(M,N(n))Gr(A){\mathrm{Hom}}_{A}(M,N):=\bigoplus_{n\in\mathbb{Z}}\mathrm{Hom}_{\mathrm{Gr(A)}}(M,N(n))\in\mathrm{Gr}(A). For MGr(A)M\in\mathrm{Gr}(A) and a homogeneous element mMm\in M, we denote the degree of mm by deg(m)\mathrm{deg}(m). We define the truncation Mn:=inMiGr(A)(n)M_{\geq n}:=\bigoplus_{i\geq n}M_{i}\in\mathrm{Gr}(A)\ (n\in\mathbb{Z}). An element mMm\in M is called torsion if mAn=0mA_{\geq n}=0 for n0n\gg 0. We say MM is a torsion module if any element of MM is torsion. We denote the subcategory of torsion modules in Gr(A)\mathrm{Gr}(A) (resp. gr(A)\mathrm{gr}(A)) by Tor(A)\mathrm{Tor}(A) (resp. tor(A)\mathrm{tor}(A)).

Definition 2.2 ([1, Section 2]).

Let AA be a right noetherian \mathbb{N}-graded kk-algebra. We define the quotient categories QGr(A):=Gr(A)/Tor(A)\mathrm{QGr(A)}:=\mathrm{Gr}(A)/\mathrm{Tor}(A) and qgr(A):=gr(A)/tor(A)\mathrm{qgr(A)}:=\mathrm{gr}(A)/\mathrm{tor}(A). We denote the projection functor by π\pi and its right adjoint functor by ω.\omega. The general (resp. noetherian) projective scheme of AA is defined as Proj(A):=(QGr(A),π(A))\mathrm{Proj}(A):=(\mathrm{QGr}{(A)},\pi(A)) (resp. proj(A):=(qgr(A),π(A))\mathrm{proj}(A):=(\mathrm{qgr}{(A)},\pi(A))).

Definition 2.3 ([1, Section 2], [26, Chapter 3]).

A quasi-scheme over kk is a pair (𝒞,𝒪)(\mathcal{C},\mathcal{O}) where 𝒞\mathcal{C} is a kk-linear abelian category and 𝒪\mathcal{O} is an object in 𝒞\mathcal{C}. A morphism from a quasi-scheme (𝒞,𝒪)(\mathcal{C},\mathcal{O}) to another quasi-scheme (𝒞,𝒪)(\mathcal{C}^{\prime},\mathcal{O}^{\prime}) is a pair (F,φ)(F,\varphi) consisting of a kk-linear right exact functor F:𝒞𝒞F:\mathcal{C}\rightarrow\mathcal{C}^{\prime} and an isomorphism φ:F(𝒪)𝒪\varphi:F(\mathcal{O})\overset{\simeq}{\rightarrow}\mathcal{O}^{\prime}. We call (F,φ)(F,\varphi) is an isomorphism if FF is an equivalence.

When AA is as in Definition 2.2, we think of proj(A)=(qgr(A),π(A))\mathrm{proj(A)}=(\mathrm{qgr}(A),\pi(A)) as a quasi-scheme. For any (commutative) noetherian scheme XX, (Coh(X),𝒪X)(\mathrm{Coh}(X),\mathcal{O}_{X}) is also a quasi-scheme. From this observation, we regard XX as a quasi-scheme.

Definition 2.4 ([30, Section 4], [34, Section4]).

Let A,BA,B be \mathbb{N}-graded kk-algebras and mAm_{A} be A1A_{\geq 1}. We define the torsion functor ΓmA:Gr(AkB)Gr(AkB)\Gamma_{m_{A}}:\mathrm{Gr}(A\otimes_{k}B^{\circ})\rightarrow\mathrm{Gr}(A\otimes_{k}B^{\circ}) by ΓmA(M):={mMmAn=0 for some n}\Gamma_{m_{A}}(M):=\{m\in M\mid mA_{\geq n}=0\text{ for some }n\in\mathbb{N}\}. We write HmAi:=RiΓmAH^{i}_{m_{A}}:=\mathrm{R}^{i}\Gamma_{m_{A}}.

Definition 2.5 ([30, Definition 6.1, 6.2], [34, Definition 3.3, 4.1] ).

Let AA be a right and left noetherian connected \mathbb{N}-graded kk-algebra and AeA^{e} be the enveloping algebra of AA. Let RR be an object of Db(Gr(Ae))\mathrm{D^{b}}(\mathrm{Gr}(A^{e})). Then, RR is called a dualizing complex of AA if (1) RRhas finite injective dimension over AA and AA^{\circ}, (2) The cohomologies of RR are finitely generated as both AA and AA^{\circ}-modules, (3) The natural morphisms ARHomA(R,R)A\rightarrow\mathrm{RHom}_{A}(R,R) and ARHomA(R,R)A\rightarrow\mathrm{RHom}_{A^{\circ}}(R,R) are isomorphisms in Db(Gr(Ae))\mathrm{D^{b}}(\mathrm{Gr}(A^{e})). Moreover, RR is called balanced if RΓmA(R)A\mathrm{R}\Gamma_{m_{A}}(R)\simeq A^{*} and RΓmA(R)A\mathrm{R}\Gamma_{m_{A^{\circ}}}(R)\simeq A^{*} in Db(Gr(Ae))\mathrm{D^{b}}(\mathrm{Gr}(A^{e})).

3. Calabi-Yau conditions

Definition 3.1 ([11, Section 2.2]).

Let AA be a connected right noetherian \mathbb{N}-graded kk-algebra. Then, proj(A)\mathrm{proj}(A) is a projective Calabi-Yau nn scheme if the global dimension of qgr(A)\mathrm{qgr}(A) is nn and the Serre functor of the derived category Db(qgr(A))\mathrm{D^{b}}(\mathrm{qgr}(A)) is the nn-shift functor [n][n].

Remark 3.2.

Actually, we do not need the condition that the global dimension of qgr(A)\mathrm{qgr}(A) is nn. If the Serre functor of the derived category Db(qgr(A))\mathrm{D^{b}}(\mathrm{qgr}(A)) is the nn-shift functor [n][n], then we can easily show that this condition holds. However, when we prove the existence of the Serre functor of Db(qgr(A))\mathrm{D^{b}}(\mathrm{qgr}(A)), we essentially need the condition that the global dimension of qgr(A)\mathrm{qgr}(A) is nn (cf. [5, Theorem A.4, Corollary A.5], Lemma 3.10).

3.1. 2\mathbb{Z}^{2}-graded algebras and Segre products

In commutative algebraic geometry, when XX is the Segre embedding of n×m\mathbb{P}^{n}\times\mathbb{P}^{m} into nm+n+m\mathbb{P}^{nm+n+m}, a smooth complete intersection YXY\subset X of bidegrees (n+1,0)(n+1,0) and (0,m+1)(0,m+1) provides a Calabi-Yau variety. We also have a little more complicated example that gives a Calabi-Yau variety. That is a smooth complete intersection of bidegrees (n,0)(n,0) (resp. (n+1,0)(n+1,0)) and (1,n+1)(1,n+1) in n×n\mathbb{P}^{n}\times\mathbb{P}^{n} (resp. n+1×n\mathbb{P}^{n+1}\times\mathbb{P}^{n}). We construct noncommutative analogues of these examples.

Let CC be an 2\mathbb{N}^{2}-graded kk-algebra. We denote the category of 2\mathbb{Z}^{2}-graded right CC-modules (resp. finitely generated 2\mathbb{Z}^{2}-graded right CC-modules) by BiGr(C)\mathrm{BiGr}(C) (resp. bigr(C)\mathrm{bigr}(C)). Let MBiGr(C)M\in\mathrm{BiGr}(C). We denote by CC^{\circ} (resp. CeC^{e}) the opposite (resp. enveloping) algebra of CC. We define the Matlis dual MBiGr(C)M^{*}\in\mathrm{BiGr}(C^{\circ}) by Mi,j:=Homk(Mi,j,k)M^{*}_{i,j}:=\mathrm{Hom}_{k}(M_{-i,-j},k) and the shift M(n,m)BiGr(C)M(n,m)\in\mathrm{BiGr}(C) by M(m,n)i,j:=Mi+m,j+n(m,n,i,j)M(m,n)_{i,j}:=M_{i+m,j+n}\ (m,n,i,j\in\mathbb{Z}). For M,NBiGr(C)M,N\in\mathrm{BiGr}(C), we write HomC(M,N):=m,nHomBiGr(C)(M,N(m,n)){\mathrm{Hom}}_{C}(M,N):=\bigoplus_{m,n\in\mathbb{Z}}\mathrm{Hom}_{\mathrm{BiGr(C)}}(M,N(m,n)). For a bihomogeneous element mMm\in M, we denote the bidegree of mm by bideg(m)\mathrm{bideg}(m).

Let MBiGr(C)M\in\mathrm{BiGr}(C). We define the truncation Mn,n:=in,jnMi,jBiGr(C)(n)M_{\geq n,\geq n}:=\bigoplus_{i\geq n,j\geq n}M_{i,j}\in\mathrm{BiGr}(C)\ (n\in\mathbb{Z}). We say mMm\in M is torsion if mCn,n=0mC_{\geq n,\geq n}=0 for n0n\gg 0. If all mMm\in M are torsion, then MM is called a torsion CC-module. We denote the category of 2\mathbb{Z}^{2}-graded torsion CC-modules by Tor(C)\mathrm{Tor}(C). We also define tor(C)\mathrm{tor}(C) to be the intersection of bigr(C)\mathrm{bigr}(C) and Tor(C)\mathrm{Tor}(C). When we assume that CC is right noetherian, we have the quotient categories QBiGr(C):=BiGr(C)/Tor(C)\mathrm{QBiGr}(C):=\mathrm{BiGr}(C)/\mathrm{Tor}(C) and qbigr(C):=bigr(C)/tor(C)\mathrm{qbigr}(C):=\mathrm{bigr}(C)/\mathrm{tor}(C) (cf. [31, Section 2]). We denote the projection functor by π\pi and its right adjoint functor by ω\omega. We can define the general (resp. noetherian) projective scheme Proj(C)\mathrm{Proj}(C) (resp. proj(C)\mathrm{proj}(C)) associated to CC and the notion of projective Calabi-Yau schemes as in the case of \mathbb{N}-graded algebras.

Let DD be an 2\mathbb{N}^{2}-graded algebra. We define mC++:=C1,1m_{C_{++}}:=C_{\geq 1,\geq 1} and the torsion functor ΓmC++:BiGr(CkD)BiGr(CkD)\Gamma_{m_{C_{++}}}:\mathrm{BiGr}(C\otimes_{k}D^{\circ})\rightarrow\mathrm{BiGr}(C\otimes_{k}D^{\circ}) by ΓmC++(M):={mMmCn,n=0 for some n}\Gamma_{m_{C_{++}}}(M):=\{m\in M\mid mC_{\geq n,\geq n}=0\text{ for some }n\in\mathbb{N}\}. We write mC:=i+j1Ci,jm_{C}:=\bigoplus_{i+j\geq 1}C_{i,j} and define another torsion functor ΓmC:BiGr(CkD)BiGr(CkD)\Gamma_{m_{C}}:\mathrm{BiGr}(C\otimes_{k}D^{\circ})\rightarrow\mathrm{BiGr}(C\otimes_{k}D^{\circ}) by ΓmC(M):={mMmCn=0 for some n}\Gamma_{m_{C}}(M):=\{m\in M\mid mC_{\geq n}=0\text{ for some }n\in\mathbb{N}\}, where Cn:=i+jnCi,jBiGr(C)C_{\geq n}:=\bigoplus_{i+j\geq n}C_{i,j}\in\mathrm{BiGr}(C). See [23, Section 3] for details of ΓmC\Gamma_{m_{C}}. We write HmC++i:=RiΓmC++H^{i}_{m_{C_{++}}}:=\mathrm{R}^{i}\Gamma_{m_{C_{++}}} and HmCi:=RiΓmCH^{i}_{m_{C}}:=\mathrm{R}^{i}\Gamma_{m_{C}}.

Theorem 3.3.

Let A:=kx0,,xn/(xjxiqjixixj)i,jA:=k\langle x_{0},\cdots,x_{n}\rangle/(x_{j}x_{i}-q_{ji}x_{i}x_{j})_{i,j}, B:=ky0,,ym/(yjyiqjiyiyj)i,jB:=k\langle y_{0},\cdots,y_{m}\rangle/(y_{j}y_{i}-q^{\prime}_{ji}y_{i}y_{j})_{i,j} and C:=AkBC:=A\otimes_{k}B, where qji,qjik×q_{ji},q^{\prime}_{ji}\in k^{\times} for all i,ji,j. We regard CC as an 2\mathbb{N}^{2}-graded algebra with bideg(xi)=(1,0)\mathrm{bideg}(x_{i})=(1,0) and bideg(yi)=(0,1)\mathrm{bideg}(y_{i})=(0,1) for all ii.

  1. (1)

    Let f:=i=0nxin+1f:=\sum_{i=0}^{n}x_{i}^{n+1} and g:=i=0myim+1g:=\sum_{i=0}^{m}y_{i}^{m+1}. We assume that (i) qii=qijqji=qijn+1=1q_{ii}=q_{ij}q_{ji}=q_{ij}^{n+1}=1for all i,ji,j, (ii) qii=qijqji=qijm+1=1q^{\prime}_{ii}=q^{\prime}_{ij}q^{\prime}_{ji}=q_{ij}^{\prime m+1}=1for all i,ji,j.

    Then, proj(C/(f,g))\mathrm{proj}(C/(f,g)) is a projective Calabi-Yau (n+m2)(n+m-2) scheme if and only if i=0nqij\prod_{i=0}^{n}q_{ij} and i=0mqij\prod_{i=0}^{m}q^{\prime}_{ij} are independent of jj, respectively.

  2. (2)

    Suppose that m=n+1m=n+1 (resp. m=nm=n) and qij=1q^{\prime}_{ij}=1 for all i,ji,j. Let f:=i=0nxin+1yif:=\sum_{i=0}^{n}x_{i}^{n+1}y_{i} and g:=i=0n+1yin+1g:=\sum_{i=0}^{n+1}y_{i}^{n+1} (resp. i=0nyin\sum_{i=0}^{n}y_{i}^{n}). We assume that qii=qijqji=qijn+1=1q_{ii}=q_{ij}q_{ji}=q_{ij}^{n+1}=1 for all i,ji,j.

    Then, proj(C/(f,g))\mathrm{proj}(C/(f,g)) is a projective Calabi-Yau (2n1)(resp. (2n2))(2n-1)(\text{resp. }(2n-2)) scheme if and only if i=0nqij\prod_{i=0}^{n}q_{ij} is independent of jj.

Notation 3.4.

For simplicity, we denote the bidegrees of f,gf,g in the theorem by (d0,d1),(e0,e1)(d_{0},d_{1}),(e_{0},e_{1}), respectively.

Remark 3.5.
  • f,gf,g are central elements in CC because of the choices of {qij},{qij}\{q_{ij}\},\{q^{\prime}_{ij}\}.

  • We have n+m2=d0+d1+e0+e14n+m-2=d_{0}+d_{1}+e_{0}+e_{1}-4 in (1). We have 2n1(resp. 2n2)=d0+d1+e0+e142n-1(\text{resp. }2n-2)=d_{0}+d_{1}+e_{0}+e_{1}-4 also in (2).

  • In (2) of the theorem, even if we do not assume qij=1q^{\prime}_{ij}=1, the condition for f,gf,g to be central in CC implies qij=1q_{ij}^{\prime}=1 for all i,ji,j after all.

To prove the theorem, we need to show some lemmas. Perhaps some experts may understand the following lemmas. However, to the best of the author’s knowledge, there are no references written on those lemmas, so the proofs are given below. In addition, the following proofs do not depend on whether (1) or (2) in the theorem is considered (except for Lemma 3.8).

Lemma 3.6.

Let :=π(RΓmC/(f,g)++(C/(f,g)))\mathcal{R}:=\pi(\mathrm{R}\Gamma_{m_{{C/(f,g)}_{++}}}(C/(f,g))^{*}) and :=π(RΓmC/(f,g)(C/(f,g)))\mathcal{R}^{\prime}:=\pi(\mathrm{R}\Gamma_{m_{C/(f,g)}}(C/(f,g))^{*}). Then, the functors 𝕃-\otimes^{\mathbb{L}}\mathcal{R} and 𝕃[1]-\otimes^{\mathbb{L}}\mathcal{R}^{\prime}[-1] between D(QBiGr(C/(f,g)))\mathrm{D(QBiGr}(C/(f,g))) and itself are naturally isomorphic.

Proof.

Let I1,I2I_{1},I_{2} be the ideals generated by {x0,,xn},{y0,ym}\{x_{0},\cdots,x_{n}\},\{y_{0}\cdots,y_{m}\}, respectively. Then, we have mC/(f,g)++=I1I2m_{{C/(f,g)}_{++}}=I_{1}\cap I_{2}, mC/(f,g)=I1+I2m_{C/(f,g)}=I_{1}+I_{2} and have the following long exact sequence in BiGr(C/(f,g)e)\mathrm{BiGr}(C/(f,g)^{e})

HmC/(f,g)i(C/(f,g))HI1i(C/(f,g))HI2i(C/(f,g))HmC/(f,g)++i(C/(f,g))\cdots\rightarrow H^{i}_{m_{C/(f,g)}}(C/(f,g))\rightarrow H^{i}_{I_{1}}(C/(f,g))\oplus H^{i}_{I_{2}}(C/(f,g))\rightarrow H^{i}_{m_{{C/(f,g)}_{++}}}(C/(f,g))\rightarrow\cdots

by using the Mayer-Vietoris sequence, where ΓIj(j=1,2)\Gamma_{I_{j}}(j=1,2) is defined not by using the degrees of IjI_{j} but by using powers of IjI_{j} (i.e., ΓIj(M):={mMmIjn=0 for some n}\Gamma_{I_{j}}(M):=\{m\in M\mid mI_{j}^{n}=0\text{ for some }n\}). Note that we can use the Mayer-Vietoris sequence in our case because I1,I2I_{1},I_{2} are generated by normal elements and this implies that I1,I2I_{1},I_{2} satisfy Artin-Rees property. We also have the exact triangle in D(BiGr(C/(f,g)e))\mathrm{D}(\mathrm{BiGr}(C/(f,g)^{e}))

RΓmC/(f,g)(C/(f,g))RΓI1(C/(f,g))RΓI2(C/(f,g))RΓmC/(f,g)++(C/(f,g)).\mathrm{R}\Gamma_{m_{C/(f,g)}}(C/(f,g))\rightarrow\mathrm{R}\Gamma_{I_{1}}(C/(f,g))\oplus\mathrm{R}\Gamma_{I_{2}}(C/(f,g))\rightarrow\mathrm{R}\Gamma_{m_{{C/(f,g)}_{++}}}(C/(f,g)).

Moreover, HI1i(C/(f,g))H^{i}_{I_{1}}(C/(f,g))^{*} and HI2i(C/(f,g))H^{i}_{I_{2}}(C/(f,g))^{*} are torsion modules for mC/(f,g)++m_{C/(f,g)_{++}} from Sub-Lemma 3.7. So, the cohomologies of RΓI1(C/(f,g))RΓI2(C/(f,g))\mathrm{R}\Gamma_{I_{1}}(C/(f,g))^{*}\oplus\mathrm{R}\Gamma_{I_{2}}(C/(f,g))^{*} are torsion. Combining this result with the above triangle, we get the claim.

Sub-Lemma 3.7.

Let I1,I2I_{1},I_{2} be as in the proof of Lemma 3.6. HI1i(C/(f,g))H^{i}_{I_{1}}(C/(f,g))^{*} and HI2i(C/(f,g))H^{i}_{I_{2}}(C/(f,g))^{*} are torsion modules for mC/(f,g)++m_{{C/(f,g)_{++}}} for any ii.

Proof.

We only show that HI1i(C/(f,g))H^{i}_{I_{1}}(C/(f,g))^{*} are torsion modules for mC/(f,g)++m_{C/(f,g)_{++}}. We can show that HI2i(C/(f,g))H^{i}_{I_{2}}(C/(f,g))^{*} are torsion in the same way.

First, we prove that HI1i(C)H^{i}_{I_{1}}(C)^{*} is torsion. We have ΓI1=ΓI1n+1\Gamma_{I_{1}}=\Gamma_{I_{1}^{n+1}}. Moreover, if J1J_{1} is the ideal generated by x0n+1,,xnn+1x_{0}^{n+1},\cdots,x_{n}^{n+1}, then we have ΓI1n+1=ΓJ1\Gamma_{I_{1}^{n+1}}=\Gamma_{J_{1}}. Note that x0n+1,,xnn+1x_{0}^{n+1},\cdots,x_{n}^{n+1} are central elements in CC from the choice of {qij}\{q_{ij}\}.

Let MGr(C)M\in\mathrm{Gr}(C) be injective. Then, we have a surjective localization map MM[xi(n+1)]M\rightarrow M[x_{i}^{-(n+1)}] for any ii and ΓJ1(M)\Gamma_{J_{1}}(M) is injective in Gr(C)\mathrm{Gr}(C) because J1J_{1} satisfies Artin-Rees property (cf. [6, Lemma A1.4]). When MM^{\prime} is injective in Gr(Ce)\mathrm{Gr}(C^{e}), then MM^{\prime} is injective in Gr(C)\mathrm{Gr}(C), where ResC:Gr(Ce)Gr(C)\mathrm{Res}_{C}:\mathrm{Gr}(C^{e})\rightarrow\mathrm{Gr}(C) is the restriction functor ([34, Lemma 2.1]). Thus, we can calculate ResC(HJ1i(C))\mathrm{Res}_{C}(H^{i}_{J_{1}}(C)) by using a Cˇ\check{\mathrm{C}}ech complex 𝒞(x0n+1,,xnn+1;C)\mathscr{C}(x_{0}^{n+1},\cdots,x_{n}^{n+1};C) (cf. [6, Theorem A1.3], [16, Chapter 2, 3]). Then, we have 𝒞(x0n+1,,xnn+1;C)=𝒞(x0n+1,,xnn+1;A)kB\mathscr{C}(x_{0}^{n+1},\cdots,x_{n}^{n+1};C)=\mathscr{C}(x_{0}^{n+1},\cdots,x_{n}^{n+1};A)\otimes_{k}B. This induces that ResC(HJ1i(C))HmAi(A)kB\mathrm{Res}_{C}({H^{i}_{J_{1}}(C)})\simeq H^{i}_{m_{A}}(A)\otimes_{k}B. Because HmAi(A)>0=0H^{i}_{m_{A}}(A)_{>0}=0 ([11, Proposition 2.4]), HI1i(C)H^{i}_{I_{1}}(C)^{*} is torsion.

Finally, we consider the exact sequences of CC-bimodules

0C(d0,d1)×fCC/(f)0,\displaystyle 0\rightarrow C(-d_{0},-d_{1})\overset{\times f}{\rightarrow}C\rightarrow C/(f)\rightarrow 0, (3.1.1)
0C/(f)(e0,e1)×gC/(f)C/(f,g)0.\displaystyle 0\rightarrow C/(f)(-e_{0},-e_{1})\overset{\times g}{\rightarrow}C/(f)\rightarrow C/(f,g)\rightarrow 0. (3.1.2)

Then, we take the long exact sequence for ΓI1\Gamma_{I_{1}} and we get the claim since HI1i(C)H^{i}_{I_{1}}(C)^{*} is torsion. ∎

Lemma 3.8.

gl.dim(qbigr(C/(f,g)))=d0+d1+e0+e14\mathrm{gl.dim}(\mathrm{qbigr}(C/(f,g)))=d_{0}+d_{1}+e_{0}+e_{1}-4.

Proof.

We show the proposition only in (1) of the theorem. In (2) of the theorem, the proposition can be shown in the same way (cf. Remark 3.9). We consider a bigraded (commutative) algebra D:=k[s0,,sn,t0,,tm]/(i=0nsi,i=0mti)D:=k[s_{0},\cdots,s_{n},t_{0},\cdots,t_{m}]/(\sum_{i=0}^{n}s_{i},\sum_{i=0}^{m}t_{i}) with si=xin+1,ti=yim+1s_{i}=x_{i}^{n+1},t_{i}=y_{i}^{m+1} and the projective spectrum biProj(D)\mathrm{biProj}(D) in the sense of [9, Section 1]. Then, C/(f,g)C/(f,g) is a finite DD-module. So, qbigr(C/(f,g))\mathrm{qbigr}(C/(f,g)) can be thought of as the category of modules over a sheaf 𝒜\mathcal{A} of 𝒪biProj(D)\mathcal{O}_{\mathrm{biProj}(D)}-algebras, where 𝒜\mathcal{A} is the sheaf on biProj(D)\mathrm{biProj}(D) which is locally defined by the algebra (k[x0,,xn,y0,,ym]/(f,g)xiyj)(0,0)(k[x_{0},\cdots,x_{n},y_{0},\cdots,y_{m}]/(f,g)_{x_{i}y_{j}})_{(0,0)} on each open affine scheme D+(sitj)Spec((Dsitj)(0,0))\mathrm{D}_{+}(s_{i}t_{j})\simeq\mathrm{Spec}((D_{s_{i}t_{j}})_{(0,0)}). Hence, it is enough to prove that the global dimension of (k[x0,,xn,y0,,ym]/(f,g)xiyj)(0,0)=d0+d1+e0+e14=n+m2(k[x_{0},\cdots,x_{n},y_{0},\cdots,y_{m}]/(f,g)_{x_{i}y_{j}})_{(0,0)}=d_{0}+d_{1}+e_{0}+e_{1}-4=n+m-2.

We can complete the rest of the proof in the same way as in [11, Section 2.3]. We give its sketch. For simplicity, we prove the claim when i=j=0i=j=0. We define a kk-algebra EE by E:=k[S1,,Sn,T1,,Tm]/(1+i=1nSi,1+i=0mTi)E:=k[S_{1},\cdots,S_{n},T_{1},\cdots,T_{m}]/(1+\sum_{i=1}^{n}S_{i},1+\sum_{i=0}^{m}T_{i}) with Si=si/s0,Ti=ti/t0S_{i}=s_{i}/s_{0},T_{i}=t_{i}/t_{0}. We also define an EE-algebra FF by F:=kX1,,Xn,Y1,,Ym/(XiXj(q0iqijqj0)XjXi,YiYj(q0iqijqj0)YjYi,1+i=1nXin+1,1+i=1mYim+1)F:=k\langle X_{1},\cdots,X_{n},Y_{1},\cdots,Y_{m}\rangle/(X_{i}X_{j}-(q_{0i}q_{ij}q_{j0})X_{j}X_{i},Y_{i}Y_{j}-(q^{\prime}_{0i}q^{\prime}_{ij}q^{\prime}_{j0})Y_{j}Y_{i},1+\sum_{i=1}^{n}X_{i}^{n+1},1+\sum_{i=1}^{m}Y_{i}^{m+1}) with Xi=xi/x0,Yi=yi/y0X_{i}=x_{i}/x_{0},Y_{i}=y_{i}/y_{0}. The module structure of FF is given by the identifications Si=Xin+1,Ti=Yim+1S_{i}=X_{i}^{n+1},T_{i}=Y_{i}^{m+1}. Let Fm~F_{\tilde{m}} be the localization of FF at a maximal ideal m~:=(S1a1,,Snan,T1b1,,Tmbm)\tilde{m}:=(S_{1}-a_{1},\cdots,S_{n}-a_{n},T_{1}-b_{1},\cdots,T_{m}-b_{m}) of EE with 1+i=1nai=1+i=1mbi=0(ai,bik)1+\sum_{i=1}^{n}a_{i}=1+\sum_{i=1}^{m}b_{i}=0\ (a_{i},b_{i}\in k). Then, it is enough to prove that the global dimension of Fm~F_{\tilde{m}} is n+m2n+m-2 ([11, Lemma 2.6, 2.7]).

If all ai,bia_{i},b_{i} are not 0, then F/m~FF/\tilde{m}F is a twisted group ring and hence semisimple. Moreover, S1a1,,Snan,T1b1,,TmbmS_{1}-a_{1},\cdots,S_{n}-a_{n},T_{1}-b_{1},\cdots,T_{m}-b_{m} is a regular sequence in Fm~F_{\tilde{m}}. This induces the claim ([15, Theorem 7.3.7]).

On the other hand, we assume that one of {a1,,an,b1,,bm}\{a_{1},\cdots,a_{n},b_{1},\cdots,b_{m}\} is 0. For example, we assume a1=0a_{1}=0. We consider F/(X1)F/(X_{1}). Then, we can show that the global dimension of (F/(X1))m~=n+m3(F/(X_{1}))_{\tilde{m}}=n+m-3 because pdF(S)=pdF/(X1)(S)+1\mathrm{pd}_{F}(S)=\mathrm{pd}_{F/(X_{1})}(S)+1 for any simple FF-module SS with Ann(S)=m~\mathrm{Ann}(S)=\tilde{m} ([15, Theorem 7.3.5]). If some other ai,bja_{i},b_{j} are 0, we repeat taking quotients and can reduce to considering the global dimension of the algebra k[X,Y]/(Xn+1+1,Ym+1+1)k[X,Y]/(X^{n+1}+1,Y^{m+1}+1), which are 0. ∎

Remark 3.9.

To prove Lemma 3.8 in (2) of the theorem, consider the projective spectrum X:=biProj(k[s0,,sn,t0,,tn+1]/X:=\mathrm{biProj}(k[s_{0},\cdots,s_{n},t_{0},\cdots,t_{n+1}]/ (i=0nsiti,i=0n+1tin+1))(\sum_{i=0}^{n}s_{i}t_{i},\sum_{i=0}^{n+1}t_{i}^{n+1})) (resp. biProj(k[s0,,sn,t0,,tn]/\mathrm{biProj}(k[s_{0},\cdots,s_{n},t_{0},\cdots,t_{n}]/ (i=0nsiti,i=0ntin))(\sum_{i=0}^{n}s_{i}t_{i},\sum_{i=0}^{n}t_{i}^{n}))) and the sheaf 𝒜\mathcal{A} of algebras on XX associated to C/(f,g)C/(f,g).

Proof of Theorem 3.3.

First, we calculate RΓmC/(f,g)(C/(f,g))\mathrm{R}\Gamma_{m_{C/(f,g)}}(C/(f,g))^{*}. From [11, Proposition 2.4] (or [23, Example 5.5]) and the proof of [23, Lemma 6.1], we have

RΓmC(C)\displaystyle\mathrm{R}\Gamma_{m_{C}}(C)^{*} RΓmA(A)RΓmB(B)\displaystyle\simeq\mathrm{R}\Gamma_{m_{A}}(A)^{*}\otimes\mathrm{R}\Gamma_{m_{B}}(B)^{*}
A1ϕ(d0e0)kB1ψ(d1e1)[d0+d1+e0+e1],\displaystyle\simeq{}^{\phi}A^{1}(-d_{0}-e_{0})\otimes_{k}{}^{\psi}B^{1}(-d_{1}-e_{1})[d_{0}+d_{1}+e_{0}+e_{1}],

where ϕ\phi (resp. ψ\psi) is the graded automorphism of AA (resp. BB) which maps xji=0nqjixjx_{j}\mapsto\prod_{i=0}^{n}q_{ji}x_{j} (resp. yji=0mqjiyjy_{j}\mapsto\prod_{i=0}^{m}q_{ji}^{\prime}y_{j}). Then, we consider the distinguished triangles

RΓmC(C(d0,d1))×fRΓmC(C)RΓmC/(f)(C/(f)),\displaystyle\mathrm{R}\Gamma_{m_{C}}(C(-d_{0},-d_{1}))\overset{\times f}{\longrightarrow}\mathrm{R}\Gamma_{m_{C}}(C)\longrightarrow\mathrm{R}\Gamma_{m_{C/(f)}}(C/(f)),
RΓmC/(f)((C/(f))(e0,e1))×gRΓmC/(f)(C/(f))RΓmC/(f,g)(C/(f,g))\displaystyle\mathrm{R}\Gamma_{m_{C/(f)}}((C/(f))(-e_{0},-e_{1}))\overset{\times g}{\longrightarrow}\mathrm{R}\Gamma_{m_{C/(f)}}(C/(f))\longrightarrow\mathrm{R}\Gamma_{m_{C}/(f,g)}(C/(f,g))

obtained from the exact sequences 3.1.1 and 3.1.2 of CC-bimodules. Hence, we have

RΓmC/(f,g)(C/(f,g))(AkB/(f,g))1ϕψ[d0+d1+e0+e12].\displaystyle\mathrm{R}\Gamma_{m_{C/(f,g)}}(C/(f,g))^{*}\simeq{}^{\phi\otimes\psi}(A\otimes_{k}B/(f,g))^{1}[d_{0}+d_{1}+e_{0}+e_{1}-2]. (3.1.3)

In addition, we have the Serre duality in Db(qbigr(C/(f,g)))\mathrm{D^{b}}(\mathrm{qbigr}(C/(f,g))) from Lemma 3.10. Thus, 𝕃π(RΓmC/(f,g)++(C/(f,g)))[1]-\otimes^{\mathbb{L}}\pi(\mathrm{R}\Gamma_{m_{C/(f,g)_{++}}}(C/(f,g))^{*})[-1] is the Serre functor of Db(qbigr(C/(f,g)))\mathrm{D^{b}}(\mathrm{qbigr}(C/(f,g))) because this functor induces an equivalence from Lemma 3.6 and the formula 3.1.3. Finally, the Serre functor 𝕃π(RΓmC/(f,g)++(C/(f,g)))[1]-\otimes^{\mathbb{L}}\pi(\mathrm{R}\Gamma_{m_{C/(f,g)_{++}}}(C/(f,g))^{*})[-1] induces the [d0+d1+e0+e14][d_{0}+d_{1}+e_{0}+e_{1}-4]-shift functor if and only if i=0nqij\prod_{i=0}^{n}q_{ij} and i=0mqij\prod_{i=0}^{m}q^{\prime}_{ij} are independent of jj (cf. [11, Remark 2.5]). This completes the proof.

The following lemma is well-known in the case of \mathbb{N}-graded algebras (for example, see [5], [35]).

Lemma 3.10 (Local Duality and Serre Duality for 2\mathbb{N}^{2}-graded algebras).

Let DD be a connected right noetherian 2\mathbb{N}^{2}-graded kk-algebra (connected means D0,0=kD_{0,0}=k). Let EE be a connected 2\mathbb{N}^{2}-graded kk-algebra. We assume that ΓmD++\Gamma_{m_{D_{++}}} has finite cohomological dimension.

  1. (1)

    Let Q:=ωπ:BiGr(D)BiGr(D)Q:=\omega\circ\pi:\mathrm{BiGr}(D)\rightarrow\mathrm{BiGr}(D). Let MD(BiGr(DkE))M\in\mathrm{D}(\mathrm{BiGr}(D\otimes_{k}E^{\circ})). Then,

    RΓmD++(M)\displaystyle\mathrm{R}\Gamma_{m_{D_{++}}}(M)^{*} RHomD(M,RΓmD++(D)),\displaystyle\simeq\mathrm{R}{\mathrm{Hom}}_{D}(M,\mathrm{R}\Gamma_{m_{D_{++}}}(D)^{*}), (a)
    RQ(M)\displaystyle\mathrm{R}Q(M)^{*} RHomD(M,RQ(D))\displaystyle\simeq\mathrm{R}{\mathrm{Hom}}_{D}(M,\mathrm{R}Q(D)^{*}) (b)

    in D(BiGr(DkE))\mathrm{D}(\mathrm{BiGr}(D\otimes_{k}E^{\circ})), where we denote the natural extension of QQ to a functor between BiGr(DkE)\mathrm{BiGr}(D\otimes_{k}E^{\circ}) and itself by the same notation.

  2. (2)

    We assume that qbigr(D)\mathrm{qbigr}(D) has finite global dimension. Let :=π(M)\mathcal{M}:=\pi(M), 𝒩:=π(N)\mathcal{N}:=\pi(N) (M,NDb(bigr(D)))(M,N\in\mathrm{D^{b}}(\mathrm{bigr}(D))). Suppose D:=π(RΓmD++(D))Db(qbigr(D)){\mathcal{R}}_{D}:=\pi(\mathrm{R}\Gamma_{m_{D_{++}}}(D)^{*})\in\mathrm{D^{b}}(\mathrm{qbigr}(D)). Then, 𝒩𝕃DDb(qbigr(D))\mathcal{N}\otimes^{\mathbb{L}}{\mathcal{R}}_{D}\in\mathrm{D^{b}}(\mathrm{qbigr}(D)) and

    HomDb(qbigr(D))(𝒩,)HomDb(qbigr(D))(,(𝒩𝕃D)[1]),\mathrm{Hom}_{\mathrm{D^{b}}(\mathrm{qbigr}(D))}(\mathcal{N},\mathcal{M})\simeq\mathrm{Hom}_{\mathrm{D^{b}}(\mathrm{qbigr}(D))}(\mathcal{M},(\mathcal{N}\otimes^{\mathbb{L}}{\mathcal{R}}_{D})[-1])^{\prime},

    which is functorial in \mathcal{M} and 𝒩\mathcal{N}. Here, ()(-)^{\prime} denotes the kk-dual.

Proof.

Since RiΓmD++()limnExti(D/Dn,n,)\mathrm{R}^{i}\Gamma_{m_{D_{++}}}(-)\simeq\lim_{n\to\infty}\mathrm{Ext}^{i}(D/D_{\geq n,\geq n},-) and DD is right noetherian, one can check that RiΓmD++()\mathrm{R}^{i}\Gamma_{m_{D_{++}}}(-) commutes with direct limits as in [33, Proposition 16.3.19]. In addition, if KK is a complex of graded free right DD-modules and LL is a complex of graded right DeD^{e}-modules, then ΓmD++(KDL)KDΓmD++(L)\Gamma_{m_{D_{++}}}(K\otimes_{D}L)\simeq K\otimes_{D}\Gamma_{m_{D_{++}}}(L) (cf. [21, Lemma 6.10]). So, we can apply the argument of [30, Theorem 5.1] (or [20, Theorem 2.1]) to prove (a) of (1).

In order to prove (b) of (1), note that we have the canonical exact sequence and the isomorphism (see also [3, Lemma 4.1.4, 4.1.5])

0ΓmD++(M)MQ(M)limnExt1(D/Dn,n,M)0,\displaystyle 0\rightarrow\Gamma_{m_{D_{++}}}(M)\rightarrow M\rightarrow Q(M)\rightarrow\lim_{n\to\infty}\mathrm{Ext}^{1}(D/D_{\geq n,\geq n},M)\rightarrow 0,
RiQ(M)Ri+1ΓmD++(M),(1i,MBiGr(D)).\displaystyle\mathrm{R}^{i}Q(M)\simeq\mathrm{R}^{i+1}\Gamma_{m_{D_{++}}}(M),\quad(1\leq i,M\in\mathrm{BiGr}(D)).

So, from the previous paragraph, QQ has finite cohomological dimension, RiQ\mathrm{R}^{i}Q commutes with direct limits. We also have Q(KDL)KDQ(L)Q(K\otimes_{D}L)\simeq K\otimes_{D}Q(L), where K,LK,L are as above (cf. [19, Lemma 3.28]). Hence, we can also apply the argument of [30, Theorem 5.1] (or [19, Theorem 3.29]) to prove (b) of (1).

We can prove (2) in the same way as in [5, Lemma A.1, Theorem A.4] by using (b) of (1). Note that we have a natural equivalence Db(qbigr(D))Dfb(QBiGr(D))\mathrm{D^{b}}(\mathrm{qbigr}(D))\simeq\mathrm{D^{b}_{f}}(\mathrm{QBiGr}(D)), where Dfb(QBiGr(D))\mathrm{D^{b}_{f}}(\mathrm{QBiGr}(D)) is the full subcategory of Db(QBiGr(D))\mathrm{D^{b}}(\mathrm{QBiGr}(D)) consisting of complexes with cohomology in qbigr(D)\mathrm{qbigr}(D) ([5, Lemma 2.2]).

As a corollary of Theorem 3.3, we construct examples of noncommutative projective Calabi-Yau schemes by using Segre products. Let A,B,fA,B,f and gg be as in Theorem 3.3.

Definition 3.11.
  1. (1)

    The Segre product ABA\circ B of AA and BB is the \mathbb{N}-graded kk-algebra with (AB)i=AikBi(A\circ B)_{i}=A_{i}\otimes_{k}B_{i}.

  2. (2)

    Let Mbigr(C)M\in\mathrm{bigr}(C) . We define a right graded ABA\circ B-module MΔM_{\Delta} as the graded ABA\circ B-module with (MΔ)i=Mi,i(M_{\Delta})_{i}=M_{i,i}.

Lemma 3.12 ([31, Theorem 2.4]).

We have the following natural isomorphism

qbigr(C)\textstyle{\mathrm{qbigr}(C)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}qgr(AB),\textstyle{\mathrm{qgr}(A\circ B),}π(M)\textstyle{\pi(M)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}π(MΔ).\textstyle{\pi(M_{\Delta}).}

In addition, the functor defined by ABC-\otimes_{A\circ B}C is the inverse of this equivalence.

Remark 3.13.

Let J:=(f,g)bigr(C)J:=(f,g)\in\mathrm{bigr}(C). We similarly obtain an equivalence

qbigr(C/J)qgr(AB/JΔ).\mathrm{qbigr}(C/J)\simeq\mathrm{qgr}(A\circ B/J_{\Delta}).

Combining Theorem 3.3 with Remark 3.13, we get the following.

Corollary 3.14.

Let J:=(f,g)bigr(C)J:=(f,g)\in\mathrm{bigr}(C). Then, proj(AB/JΔ)\mathrm{proj}(A\circ B/J_{\Delta}) is a projective Calabi-Yau scheme.

3.2. Weighted hypersurfaces

Reid produced the list of all commutative weighted Calabi-Yau hypersurfaces of dimensions 2 (for example, see [10], [22]). In this section, we construct noncommutative projective Calabi-Yau schemes from noncommutative weighted projective hypersurfaces. Let AA be a right noetherian \mathbb{N}-graded kk-algebra. Then, the rr-th Veronese algebra A(r)A^{(r)} is the \mathbb{N}-graded kk-algebra with Ai(r)=AriA^{(r)}_{i}=A_{ri}. We consider the (commutative) weighted polynomial ring A=k[x0,,xn]A=k[x_{0},\cdots,x_{n}] with deg(xi)=di\mathrm{deg}(x_{i})=d_{i}. Then, Coh(Proj(A))\mathrm{Coh}(\mathrm{Proj}(A)) is in general not equivalent to qgr(A)\mathrm{qgr}(A), but to qgr(A(n+1)lcm(d0,,dn))\mathrm{qgr}(A^{(n+1)\mathrm{lcm}(d_{0},\cdots,d_{n})}). However, we can think of qgr(A)\mathrm{qgr}(A) as a resolution of singularities of Coh(Proj(A))\mathrm{Coh}(\mathrm{Proj}(A)) (cf. [25, Example 4.9]). Moreover, we have qgr(A)Coh([(Spec(A)\{0})/𝔾m])\mathrm{qgr}(A)\simeq\mathrm{Coh}([(\mathrm{Spec}(A)\backslash\{0\})/\mathbb{G}_{m}]) and [(Spec(A)\{0})/𝔾m][(\mathrm{Spec}(A)\backslash\{0\})/\mathbb{G}_{m}] is a smooth Deligne-Mumford stack whose coarse moduli space is Proj(A)\mathrm{Proj}(A).

Theorem 3.15.

Let (d0,,dn)>0n+1(d_{0},\cdots,d_{n})\in\mathbb{Z}_{>0}^{n+1} and d:=i=0ndid:=\sum_{i=0}^{n}d_{i} such that dd is divisible by did_{i} for all ii. Let C:=kx0,,xn/(xjxiqjixixj)i,jC:=k\langle x_{0},\cdots,x_{n}\rangle/(x_{j}x_{i}-q_{ji}x_{i}x_{j})_{i,j}, where qjik×,deg(xi)=diq_{ji}\in k^{\times},\mathrm{deg}(x_{i})=d_{i} for all i,ji,j. Let f:=i=0nxihif:=\sum_{i=0}^{n}x_{i}^{h_{i}}, where hi:=d/dih_{i}:=d/d_{i}.

We assume that qii=qijqji=qijhi=qijhj=1q_{ii}=q_{ij}q_{ji}=q_{ij}^{h_{i}}=q_{ij}^{h_{j}}=1 for all i,ji,j. Then, proj(C/(f))\mathrm{proj}(C/(f)) is a projective Calabi-Yau (n1)(n-1) scheme if and only if there exists ckc\in k such that cdj=i=0nqijc^{d_{j}}=\prod_{i=0}^{n}q_{ij} for all jj.

Remark 3.16.
  • ff is a central element in CC from the choice of {qij}\{q_{ij}\}.

  • Theorem 3.15 is a generalization of [11, Theorem 1.1].

Lemma 3.17.

The balanced dualizing complex of C/(f)C/(f) is isomorphic to (C/(f))1ϕ[n]{{}^{\phi}}(C/(f))^{1}[n], where ϕ\phi is a graded automorphism of CC which maps xji=0nqjixjx_{j}\mapsto\prod_{i=0}^{n}q_{ji}x_{j}.

Proof.

Since CC is Artin-Schelter regular, CC is skew Calabi-Yau ([23, Lemma 1.2]). This induces that the balanced dualizing complex of CC is isomorphic to C1ϕ(d)[n+1]{}^{\phi}C^{1}(-d)[n+1], where ϕ\phi is the Nakayama automorphism of CC. From [23, Example 5.5], the automorphism ϕ\phi is the map which maps xji=0nqjixjx_{j}\mapsto\prod_{i=0}^{n}q_{ji}x_{j}.

By using this result, we can obtain the claim in the same way as in the proof of Theorem 3.3 after Remark 3.9. ∎

To calculate the global dimension of qgr(C/(f))\mathrm{qgr}(C/(f)), we recall the notion of quasi-Veronese algebras. In detail, see [17, Section 3].

Definition 3.18 ([17, Section 3]).

Let AA be an \mathbb{N}-graded kk-algebra. The ll-th quasi-Veronese algebra A[l]A^{[l]} of AA is a graded kk-algebra defined by

A[l]:=iAi[l]:=i(AliAli+1Ali+l1Ali1AliAli+l2Alil+1Alil+2Ali.).A^{[l]}:=\bigoplus_{i\in\mathbb{N}}A^{[l]}_{i}:=\bigoplus_{i\in\mathbb{N}}\begin{pmatrix}A_{li}&A_{li+1}&\cdots&A_{li+l-1}\\ A_{li-1}&A_{li}&\cdots&A_{li+l-2}\\ \vdots&\vdots&\ddots&\vdots\\ A_{li-l+1}&A_{li-l+2}&\cdots&A_{li}.\end{pmatrix}.
Remark 3.19.
  1. (1)

    We have Gr(A)Gr(A[l])\mathrm{Gr}(A)\simeq\mathrm{Gr}(A^{[l]}) ([17, Lemma 3.9]). The equivalence is obtained by the functor Q:Gr(A)Gr(A)Q:\mathrm{Gr}(A)\rightarrow\mathrm{Gr}(A), which is defined by Q(M):=i(j=0l1Mlij)Q(M):=\bigoplus_{i\in\mathbb{Z}}\left(\bigoplus_{j=0}^{l-1}M_{li-j}\right)

  2. (2)

    When AA is right noetherian, A[l]0i,jn1A(ji)(l)gr(A(l))A^{[l]}\simeq\bigoplus_{0\leq i,j\leq n-1}A(j-i)^{(l)}\in\mathrm{gr}(A^{(l)}), where A(l)A^{(l)} is the ll-th Veronese algebra of AA and the A(l)A^{(l)}-module structure of A[l]A^{[l]} is given by the natural inclusion A(l)A[l]A^{(l)}\subset A^{[l]} (cf. the proof of [18, Proposition 4.11]). Then, A[l]A^{[l]} is also right noetherian since A(l)A^{(l)} is right noetherian. In this case, QQ induces an equivalence between qgr(A)\mathrm{qgr}(A) and qgr(A[l])\mathrm{qgr}(A^{[l]}).

Lemma 3.20.

Let AA be an \mathbb{N}-graded kk-algebra which is generated by homogeneous elements y0,,yhy_{0},\cdots,y_{h} with deg(yi)>0\mathrm{deg}(y_{i})>0 as an A0A_{0}-algebra. Let lmax{deg(y0),,deg(yh)}l\geq\mathrm{max}\{\mathrm{deg}(y_{0}),\cdots,\mathrm{deg}(y_{h})\}. Then, A[l]A^{[l]} is generated in degree 0 and 11.

Proof.

For any ii\in\mathbb{N} and any a,b{0,1,,l1}a,b\in\{0,1,\cdots,l-1\}, it is enough to show that every homogeneous element mm of the form

m=(m0,0m0,βm0,l1mα,0mα,βmα,l1ml1,0ml1,βml1,l1)Ai[l],(mα,β(Ai[l])α,β:=Ali+βα,mα,β=0 when (α,β)(a,b)0α,βl1)m=\begin{pmatrix}m_{0,0}&\dots&m_{0,\beta}&\dots&m_{0,l-1}\\ \vdots&&\vdots&&\vdots\\ m_{\alpha,0}&\dots&m_{\alpha,\beta}&\dots&m_{\alpha,l-1}\\ \vdots&&\vdots&&\vdots\\ m_{l-1,0}&\dots&m_{l-1,\beta}&\dots&m_{l-1,l-1}\end{pmatrix}\in A^{[l]}_{i},\quad\left(\begin{gathered}m_{\alpha,\beta}\in\left(A^{[l]}_{i}\right)_{\alpha,\beta}:=A_{li+\beta-\alpha},\\ \ m_{\alpha,\beta}=0\text{ when }(\alpha,\beta)\neq(a,b)\\ 0\leq\alpha,\beta\leq l-1\end{gathered}\right)

is generated in degree 0 and 11. Moreover, we can assume that ma,b=j=0n1yij(ij{0,,h},n1)m_{a,b}=\prod_{j=0}^{n_{1}}y_{i_{j}}\ (i_{j}\in\{0,\cdots,h\},n_{1}\in\mathbb{N}).

If ma,bm_{a,b} is decomposed into j=0n1yij=j=0n2yijj=n2+1n1yij(n2)\prod_{j=0}^{n_{1}}y_{i_{j}}=\prod_{j=0}^{n_{2}}y_{i_{j}}\prod_{j=n_{2}+1}^{n_{1}}y_{i_{j}}\ (n_{2}\in\mathbb{N}) such that ladeg(j=1n2yij)2la1l-a\leq\mathrm{deg}(\prod_{j=1}^{n_{2}}y_{i_{j}})\leq 2l-a-1, then we have j=0n2yij(A1[l])a,c=Al+ca\prod_{j=0}^{n_{2}}y_{i_{j}}\in(A^{[l]}_{1})_{a,c}=A_{l+c-a} and j=n2+1n1yij(Ai1[l])c,b=Al(i1)+bc\prod_{j=n_{2}+1}^{n_{1}}y_{i_{j}}\in(A^{[l]}_{i-1})_{c,b}=A_{l(i-1)+b-c} (0cl1)(0\leq{}^{\exists}c\leq l-1). In this case, we can show the claim by using induction on the degree of mm. So, it is sufficient to show that we have such a decomposition for all mm. Indeed, we can find at least one such decomposition from (2la1)(la)+1=l(2l-a-1)-(l-a)+1=l and the choice of ll. In detail, we have ladeg(yi0)2la1l-a\leq\mathrm{deg}(y_{i_{0}})\leq 2l-a-1 or there exists n3n_{3}\in\mathbb{N} such that deg(yi0yi1yin3)<la\mathrm{deg}(y_{i_{0}}y_{i_{1}}\cdots y_{i_{n_{3}}})<l-a and ladeg(yi0yi1yin3yin3+1)2la1l-a\leq\mathrm{deg}(y_{i_{0}}y_{i_{1}}\cdots y_{i_{n_{3}}}y_{i_{n_{3}}+1})\leq 2l-a-1 since 0<deg(yi)l0<\mathrm{deg}(y_{i})\leq l. ∎

Lemma 3.21.

gl.dim(qgr(C/(f)))=n1\mathrm{gl.dim}(\mathrm{qgr}(C/(f)))=n-1.

Proof.

We use the idea of the proof in Lemma 3.8. We consider an \mathbb{N}-graded kk-algebra B:=k[s0,,sn]/(i=0nsi)B:=k[s_{0},\cdots,s_{n}]/(\sum_{i=0}^{n}s_{i}) with si=xihis_{i}=x_{i}^{h_{i}}. Then, A[d]A^{[d]} is right noetherian and qgr(C/(f))qgr((C/(f))[d])\mathrm{qgr}(C/(f))\simeq\mathrm{qgr}((C/(f))^{[d]}) from Remark 3.19. So, it is enough to prove that gl.dim(qgr((C/(f))[d]))=n1\mathrm{gl.dim}(\mathrm{qgr}((C/(f))^{[d]}))=n-1. Because (C/(f))(d)(C/(f))^{(d)} is finite over BB, (C/(f))[d](C/(f))^{[d]} is also finite over BB. In addition, (C/(f))[d](C/(f))^{[d]} is generated in degrees 0 and 11 from Lemma 3.20. So, qgr((C/(f))[d])\mathrm{qgr}((C/(f))^{[d]}) is equivalent to the category of coherent modules over a sheaf 𝒜\mathcal{A} of 𝒪Proj(B)\mathcal{O}_{\mathrm{Proj}(B)}-algebra, where 𝒜\mathcal{A} is the sheaf on the projective spectrum Proj(B)\mathrm{Proj}(B) which is locally defined by a tiled matrix algebra

Ni=(Ei,0Ei,1Ei,d1Ei,1Ei,0Ei,d2Ei,d+1Ei,d+2Ei,0)N_{i}=\begin{pmatrix}E_{i,0}&E_{i,1}&\cdots&E_{i,d-1}\\ E_{i,-1}&E_{i,0}&\cdots&E_{i,d-2}\\ \vdots&\vdots&\cdots&\vdots\\ E_{i,-d+1}&E_{i,-d+2}&\cdots&E_{i,0}\\ \end{pmatrix}

on each D+(si)D_{+}(s_{i}). Here, Ei:=(C/(f))[xi1]E_{i}:=(C/(f))[x_{i}^{-1}] and Ei,jE_{i,j} is the degree jj part of EiE_{i}. As in the proof of Lemma 3.8, it is enough to show that the global dimension of NiN_{i} is n1n-1 for all ii.

On the other hand, R1:=EiEi(1)Ei(d2)Ei(d1)R_{1}:=E_{i}\oplus E_{i}(1)\oplus\cdots\oplus E_{i}(d-2)\oplus E_{i}(d-1) and R2:=EiEi(1)Ei(di2)Ei(di1)R_{2}:=E_{i}\oplus E_{i}(1)\oplus\cdots\oplus E_{i}(d_{i}-2)\oplus E_{i}(d_{i}-1) are progenerators in Gr(Ei)\mathrm{Gr}(E_{i}). So, the category of right Endgr(R1)\mathrm{End}_{\mathrm{gr}}(R_{1})-modules and the category of right Endgr(R2)\mathrm{End}_{\mathrm{gr}}(R_{2})-modules are equivalent because they are equivalent to the category of graded right EiE_{i}-modules (cf. [26, Lemma 4.8], [25, Remarks after Proposition 4.5]). We also have Endgr(R1)Ni\mathrm{End}_{\mathrm{gr}}(R_{1})\simeq N_{i} and

Endgr(R2)Mi:=(Ei,0Ei,1Ei,di1Ei,1Ei,0Ei,di2Ei,di+1Ei,di+2Ei,0).\mathrm{End}_{\mathrm{gr}}(R_{2})\simeq M_{i}:=\begin{pmatrix}E_{i,0}&E_{i,1}&\cdots&E_{i,d_{i}-1}\\ E_{i,-1}&E_{i,0}&\cdots&E_{i,d_{i}-2}\\ \vdots&\vdots&\cdots&\vdots\\ E_{i,-d_{i}+1}&E_{i,-d_{i}+2}&\cdots&E_{i,0}\\ \end{pmatrix}.

So, it is sufficient to prove the global dimension of MiM_{i} is n1n-1 for each ii.

For simplicity, we assume i=0i=0. When i0i\neq 0, we can show the claim in the same way. Let D=k[S1,,Sn]/(1+j=0nSj)D=k[S_{1},\cdots,S_{n}]/(1+\sum_{j=0}^{n}S_{j}) with Sj=sj/s0S_{j}=s_{j}/s_{0}. We show that the global dimension of the DD-algebra M0M_{0} is n1n-1. The module structure of M0M_{0} is given by the identification Sj=(xjhj/x0h0)Id0M0S_{j}=(x_{j}^{h_{j}}/x_{0}^{h_{0}})I_{d_{0}}\in M_{0}, where Id0I_{d_{0}} is the (d0×d0)(d_{0}\times d_{0})-identity matrix. Let m~=(S1a1,Snan)(ajk)\tilde{m}=(S_{1}-a_{1},\cdots S_{n}-a_{n})\ (a_{j}\in k) be a maximal ideal of DD with 1+j=1naj=01+\sum_{j=1}^{n}a_{j}=0. It is sufficient to show that gl.dim((M0)m~)=n1\mathrm{gl.dim}((M_{0})_{\tilde{m}})=n-1, where (M0)m~(M_{0})_{\tilde{m}} is the localization of M0M_{0} at m~\tilde{m} (cf. the second paragraph of the proof of Lemma 3.8). We divide the proof of this claim into two cases.

Case (a) : all aja_{j} are not 0. Because S1a1,,SnanS_{1}-a_{1},\cdots,S_{n}-a_{n} is a regular sequence in (M0)m~(M_{0})_{\tilde{m}}, we show that the global dimension of (M0)m~/m~(M0)m~M0/m~M0(M_{0})_{\tilde{m}}/\tilde{m}(M_{0})_{\tilde{m}}\simeq M_{0}/\tilde{m}M_{0} is 0 (cf. the third paragraph of the proof of Lemma 3.8).

First, the category of M0/m~M0M_{0}/\tilde{m}M_{0}-modules is equivalent to the category of graded E0:=E0/(x1h1/x0h0a1,,xnhn/x0h0an)E0E_{0}^{\prime}:=E_{0}/(x_{1}^{h_{1}}/x_{0}^{h_{0}}-a_{1},\cdots,x_{n}^{h_{n}}/x_{0}^{h_{0}}-a_{n})E_{0}-modules. This is a Morita equivalence obtained from the isomorphism Endgr(E0)M0/m~M0\mathrm{End}_{\mathrm{gr}}(E_{0}^{\prime})\simeq M_{0}/\tilde{m}M_{0} (cf. the three previous paragraph).

Next, we see that E0E_{0}^{\prime} is strongly graded. Since E0(C[x01])/(1+(x1h1/x0h0)++(xnhn/x0h0))E_{0}\simeq(C[x_{0}^{-1}])/(1+(x_{1}^{h_{1}}/x_{0}^{h_{0}})+\cdots+(x_{n}^{h_{n}}/x_{0}^{h_{0}})), we have E0(C[x01])/(x1h1/x0h0a1,,xnhn/x0h0an)E_{0}^{\prime}\simeq(C[x_{0}^{-1}])/(x_{1}^{h_{1}}/x_{0}^{h_{0}}-a_{1},\cdots,x_{n}^{h_{n}}/x_{0}^{h_{0}}-a_{n}). For any ll\in\mathbb{Z}, if x~:=x0l0x1l1xnln(E0)l(l0,l1,ln)\tilde{x}:=x_{0}^{l_{0}}x_{1}^{l_{1}}\cdots x_{n}^{l_{n}}\in(E_{0}^{\prime})_{l}\ (l_{0}\in\mathbb{Z},l_{1},\cdots l_{n}\in\mathbb{N}), then there exist k1,,knk_{1},\cdots,k_{n}\in\mathbb{N} such that x~:=x0(ki)h0l0x1k1h1l1xnknhnln(E0)l\tilde{x}^{\prime}:=x_{0}^{(-\sum k_{i})h_{0}-l_{0}}x_{1}^{k_{1}h_{1}-l_{1}}\cdots x_{n}^{k_{n}h_{n}-l_{n}}\in(E_{0}^{\prime})_{-l}. Because x~x~k\tilde{x}\tilde{x}^{\prime}\in k^{*}, we get 1(E0)l(E0)l1\in(E_{0}^{\prime})_{l}(E_{0}^{\prime})_{-l} and E0E_{0}^{\prime} is strongly graded.

Since E0E_{0}^{\prime} is strongly graded, we have Gr(E0)Mod((E0)0)\mathrm{Gr}(E_{0}^{\prime})\simeq\mathrm{Mod}((E_{0}^{\prime})_{0}). Then, (E0)0(E_{0}^{\prime})_{0} is a twisted group algebra, where a kk-basis of (E0)0(E_{0}^{\prime})_{0} is {x0e0x1e1x2e2xnen(E0)0j=0nejdj=0 and 0ej<hj(j=1,2,,n)}\{x_{0}^{e_{0}}x_{1}^{e_{1}}x_{2}^{e_{2}}\cdots x_{n}^{e_{n}}\in(E_{0}^{\prime})_{0}\mid\sum_{j=0}^{n}e_{j}d_{j}=0\text{ and }0\leq e_{j}<h_{j}\ ({}^{\forall}j=1,2,\cdots,n)\}. In particular, (E0)0(E_{0}^{\prime})_{0} is semisimple. Hence, the graded global dimension of E0E_{0}^{\prime} is 0 and gl.dim(M0/m~M0)=0\mathrm{gl.dim}(M_{0}/\tilde{m}M_{0})=0.

Case (b) : some of aja_{j} are 0. For example, we assume a1=0a_{1}=0. Then, (x1h1/x0h0)Id0(x_{1}^{h_{1}}/x_{0}^{h_{0}})I_{d_{0}} is an annihilator of any simple M0M_{0}-module NN. On the other hand, we have a unique integer r1r_{1} such that 0deg(x1/x0r1)d010\leq\mathrm{deg}(x_{1}/x_{0}^{r_{1}})\leq d_{0}-1. If deg(x1/x0r1)=0\mathrm{deg}(x_{1}/x_{0}^{r_{1}})=0, then J=x1/x0r1Id0J=x_{1}/x_{0}^{r_{1}}I_{d_{0}} annihilates NN. Otherwise, the matrix

J=(Ox1/x0r1x1/x0r1\hdashlinex1/x0r1+1x1/x0r1+1O)M0J=\left(\begin{array}[]{c:c}{\Huge O}&{\begin{array}[]{ccc}x_{1}/x_{0}^{r_{1}}&&\\ &\ddots&\\ &&x_{1}/x_{0}^{r_{1}}\end{array}}\\ \hdashline{\begin{array}[]{ccc}x_{1}/x_{0}^{r_{1}+1}&&\\ &\ddots&\\ &&x_{1}/x_{0}^{r_{1}+1}\end{array}}&{\Huge O}\end{array}\right)\in M_{0}

annihilates NN because nJ{}^{\exists}n_{J}\in\mathbb{N} such that JnJ=(x1h1/x0h0)Id0J^{n_{J}}=(x_{1}^{h_{1}}/x_{0}^{h_{0}})I_{d_{0}} (the reduction of NiN_{i} to MiM_{i} is used here). Thus, it is enough to prove that the global dimension of (M0/JM0)m~=n2(M_{0}/JM_{0})_{\tilde{m}}=n-2 (cf. the fourth paragraph of the proof of Lemma 3.8). Note that we have

M0/JM0(F0,0F0,1F0,d01F0,1F0,0F0,d02F0,d0+1F0,d0+2F0,0),\displaystyle M_{0}/JM_{0}\simeq\begin{pmatrix}F_{0,0}&F_{0,1}&\cdots&F_{0,d_{0}-1}\\ F_{0,-1}&F_{0,0}&\cdots&F_{0,d_{0}-2}\\ \vdots&\vdots&\cdots&\vdots\\ F_{0,-d_{0}+1}&F_{0,-d_{0}+2}&\cdots&F_{0,0}\end{pmatrix}, (3.2.1)

where F0=E0/x1E0kx0,x2,,xn/(xjxiqjixixj,x0h0+x2h2++xnhn)i,j[x01]F_{0}=E_{0}/x_{1}E_{0}\simeq k\langle x_{0},x_{2},\cdots,x_{n}\rangle/(x_{j}x_{i}-q_{ji}x_{i}x_{j},x_{0}^{h_{0}}+x_{2}^{h_{2}}+\cdots+x_{n}^{h_{n}})_{i,j}[x_{0}^{-1}] and F0,jF_{0,j} is the degree jj part of F0F_{0}.

If any of a2,,ana_{2},\cdots,a_{n} is not 0, we can reduce to the case (a) from 3.2.1. If some of a2,,ana_{2},\cdots,a_{n} are 0, repeat the above process until we can reduce to the case (a). ∎

Proof of Theorem 3.15.

gl.dim(qgr(C/(f)))\mathrm{gl.dim}(\mathrm{qgr}(C/(f))) is finite. So, the balanced dualizing complex (C/(f))1ϕ[n]{}^{\phi}(C/(f))^{1}[n] of C/(f)C/(f) induces the Serre functor of qgr(C/(f))\mathrm{qgr}(C/(f)) from [5, Theorem A.4]. We complete the proof as in the proof of Theorem 3.3. ∎

4. Comparison and closed points

In this section, we calculate closed points of noncommutative projective Calabi-Yau schemes of dimension 22 obtained in Section 3.2 and compare our examples with commutative Calabi-Yau varieties and the first examples constructed in [11]. In particular, we show that a noncommutative projective Calabi-Yau scheme in Section 3.2 gives essentially a new example of noncommutative projective Calabi-Yau schemes.

Example 4.1.

Any weight (d0,d1,d2,d3)(d_{0},d_{1},d_{2},d_{3}) of noncommutative projective Calabi-Yau 2 schemes in Theorem 3.15 such that gcd(d0,d1,d2,d3)=1\mathrm{gcd}(d_{0},d_{1},d_{2},d_{3})=1 is one of the following (obtained by using a computer):

(d0,d1,d2,\displaystyle(d_{0},d_{1},d_{2}, d3)=(1,1,1,1),(1,1,1,3),(1,1,2,2),(1,1,2,4),(1,1,4,6),(1,2,2,5),\displaystyle d_{3})=(1,1,1,1),(1,1,1,3),(1,1,2,2),(1,1,2,4),(1,1,4,6),(1,2,2,5),
(1,2,3,6),(1,2,6,9),(1,3,4,4),(1,6,14,21),(2,3,3,4),(2,3,10,15).\displaystyle(1,2,3,6),(1,2,6,9),(1,3,4,4),(1,6,14,21),(2,3,3,4),(2,3,10,15).

From now, we focus on the closed points of noncommutatative projective Calabi-Yau 2 schemes in Theorem 3.15 whose weights are of type (1,1,a,b)(1,1,a,b). We recall the notion of closed points of noncommutative projective schemes.

For simplicity, we often call an \mathbb{N}-graded kk-algebra of the form kz0,,zm/(zjzipjizizj)i,j(pjik×,m)k\langle z_{0},\cdots,z_{m}\rangle/(z_{j}z_{i}-p_{ji}z_{i}z_{j})_{i,j}\ (p_{ji}\in k^{\times},m\in\mathbb{N}) with deg(zi)>0\mathrm{deg}(z_{i})>0 a weighted quantum polynomial ring. (pji)(p_{ji}) is called the quantum parameter.

Definition 4.2 ([18, Section 3.1]).

Let AA be a finitely generated right noetherian connected \mathbb{N}-graded kk-algebra. A closed point of proj(A)\mathrm{proj}(A) is an object of qgr(A)\mathrm{qgr}(A) represented by a 11-critical module of AA. In particular, if AA is a quotient of a weighted quantum polynomial ring, then every point is one of the following:

  1. (1)

    An ordinary point, which is represented by a finitely generated 11-critical module of multiplicity 11.

  2. (2)

    A fat point, which is represented by a finitely generated 11-critical module of multiplicity >1>1.

  3. (3)

    A thin point, which is represented by a finitely generated 11-critical module of multiplicity <1<1.

For the definitions of 11-critical modules and multiplicities, see [18, Definition 3.1, 3.10]. Note that if AA is generated in degree 11, the notion of ordinary points and that of point modules are the same, and there is no thin point. We denote by |proj(A)||\mathrm{proj}(A)| the set of closed points of proj(A)\mathrm{proj(A)}.

Let C:=kx0,x1,x2,x3/(xjxiqjixixj)i,jC:=k\langle x_{0},x_{1},x_{2},x_{3}\rangle/(x_{j}x_{i}-q_{ji}x_{i}x_{j})_{i,j} whose weight is of type (d0,d1,d2,d3)=(1,1,a,b)(0<ab)(d_{0},d_{1},d_{2},d_{3})=(1,1,a,b)\ (0<a\leq b). We assume that qijqji=qii=1q_{ij}q_{ji}=q_{ii}=1 for all i,ji,j. Since d0=1d_{0}=1, C[x01]C[x_{0}^{-1}] is strongly graded. So, from [18, Theorem 4.20], we have

|proj(C)|=|spec(C[x01]0)||proj(C/(x0))|,|\mathrm{proj}(C)|=|\mathrm{spec}(C[x_{0}^{-1}]_{0})|\bigsqcup|\mathrm{proj}(C/(x_{0}))|,

where we denote by |spec(C[x01]0)||\mathrm{spec}(C[x_{0}^{-1}]_{0})| the set of simple modules of C[x01]0C[x_{0}^{-1}]_{0}. In this equality, the 11 (resp. n>1n>1)-dimensional simple modules of spec(C[x01]0)\mathrm{spec}(C[x_{0}^{-1}]_{0}) correspond to ordinary (resp. fat) points in proj(C)\mathrm{proj}(C). Similarly, we have

|proj(C)|\displaystyle|\mathrm{proj}(C)| =|spec(C[x01]0)||spec(C/(x0)[x11]0)||proj(C/(x0,x1))|.\displaystyle=|\mathrm{spec}(C[x_{0}^{-1}]_{0})|\bigsqcup|\mathrm{spec}(C/(x_{0})[x_{1}^{-1}]_{0})|\bigsqcup|\mathrm{proj}(C/(x_{0},x_{1}))|.

It easy to see that C[x01]0C[x_{0}^{-1}]_{0} is isomorphic to kX1,X2,X3/(XjXiqjiXiXj)i,jk\langle X_{1},X_{2},X_{3}\rangle/(X_{j}X_{i}-q^{\prime}_{ji}X_{i}X_{j})_{i,j}, where qji:=q0jdiqjiqi0dj(i,j0)q^{\prime}_{ji}:=q_{0j}^{d_{i}}q_{ji}q_{i0}^{d_{j}}\ (i,j\neq 0). C/(x0)[x11]0C/(x_{0})[x_{1}^{-1}]_{0} is also isomorphic to kY2,Y3/(Y3Y2p32Y2Y3)k\langle Y_{2},Y_{3}\rangle/(Y_{3}Y_{2}-p_{32}Y_{2}Y_{3}), where p32:=q13d2q32q21d3p_{32}:=q_{13}^{d_{2}}q_{32}q_{21}^{d_{3}}.

Let C1:=kx0,x1,x2,x3/(xjxiqjixixj)i,jC_{1}:=k\langle x_{0}^{\prime},x_{1}^{\prime},x_{2}^{\prime},x_{3}^{\prime}\rangle/(x_{j}^{\prime}x_{i}^{\prime}-q^{\prime}_{ji}x_{i}^{\prime}x_{j}^{\prime})_{i,j}, where deg(xi)=1\mathrm{deg}(x_{i}^{\prime})=1, q0i=qj0=1q^{\prime}_{0i}=q^{\prime}_{j0}=1 for all i,ji,j. Let C2:=ky1,y2,y3/(yjyipjiyiyj)i,jC_{2}:=k\langle y_{1},y_{2},y_{3}\rangle/(y_{j}y_{i}-p_{ji}y_{i}y_{j})_{i,j}, where deg(yi)=1\mathrm{deg}(y_{i}^{\prime})=1, p1i=pj1=1p_{1i}=p_{j1}=1 for all i,ji,j. Then, we can consider the point scheme of proj(C1)\mathrm{proj}(C_{1}) (resp. proj(C2)\mathrm{proj}(C_{2})), which is isomorphic to the set of ordinary points |proj(C1)|ord|\mathrm{proj}(C_{1})|_{\text{ord}} (resp. |proj(C2)|ord|\mathrm{proj}(C_{2})|_{\text{ord}}) as sets. Thus, we regard |proj(C1)|ord|\mathrm{proj}(C_{1})|_{\text{ord}} (resp. |proj(C2)|ord|\mathrm{proj}(C_{2})|_{\text{ord}}) as the point scheme of proj(C1)\mathrm{proj}(C_{1}) (resp. proj(C2)\mathrm{proj}(C_{2})).

Let |spec(C[x01]0)|1|\mathrm{spec}(C[x_{0}^{-1}]_{0})|_{1} (resp. |spec(C/(x0)[x11]0)|1|\mathrm{spec}(C/(x_{0})[x_{1}^{-1}]_{0})|_{1}) be the set of 11-dimensional simple modules of C[x01]0C[x_{0}^{-1}]_{0} (resp. C/(x0)[x11]0C/(x_{0})[x_{1}^{-1}]_{0}). Because C1[x01]0C[x01]0C_{1}[{x^{\prime}_{0}}^{-1}]_{0}\simeq C[x_{0}^{-1}]_{0} and C2[y11]0C/(x0)[x11]0C_{2}[{y_{1}}^{-1}]_{0}\simeq C/(x_{0})[x_{1}^{-1}]_{0}, we can think of |spec(C[x01]0)|1|\mathrm{spec}(C[x_{0}^{-1}]_{0})|_{1} (resp. |spec(C/(x0)[x11]0)|1|\mathrm{spec}(C/(x_{0})[x_{1}^{-1}]_{0})|_{1}) as a locally closed subscheme of |proj(C1)|ord|\mathrm{proj}(C_{1})|_{\text{ord}} (resp. |proj(C2)|ord|\mathrm{proj}(C_{2})|_{\text{ord}}) from [18, Theorem 4.20].

Lemma 4.3.
  1. (1)

    If qji1q^{\prime}_{ji}\neq 1 for all i,j0i,j\neq 0, |spec(C[x01]0)|1|\mathrm{spec}(C[x_{0}^{-1}]_{0})|_{1} is a union of three affine lines.

  2. (2)

    If p321p_{32}\neq 1, |spec(C/(x0)[x11]0)|1|\mathrm{spec}(C/(x_{0})[x_{1}^{-1}]_{0})|_{1} is a union of two affine lines. Otherwise, |spec(C/(x0)[x11]0)|1𝔸2|\mathrm{spec}(C/(x_{0})[x_{1}^{-1}]_{0})|_{1}\simeq\mathbb{A}^{2}.

Proof.

(2) is well-known (for example, see [26, Section 4.3]). Regarding (1), under the assumption of the lemma, proj(C1)\mathrm{proj}(C_{1}) belongs to case (3) or case (4) in [32, Corollary 5.1]. This shows that |spec(C1[x01]0)|1|\mathrm{spec}(C_{1}[{x^{\prime}_{0}}^{-1}]_{0})|_{1} is isomorphic to ijZ(Xi,Xj)𝔸3=Spec(k[X1,X2,X3])\bigcup_{i\neq j}Z(X^{\prime}_{i},X^{\prime}_{j})\subset\mathbb{A}^{3}=\mathrm{Spec}(k[X^{\prime}_{1},X^{\prime}_{2},X^{\prime}_{3}]) (cf. [32, Proposition 4.2] or [2, Theorem 1]). ∎

Remark 4.4.

We consider the weights (1,1,a,b)(1,1,a,b) and the quantum parameters which give noncommutative projective Calabi-Yau 2 schemes in Theorem 3.15. Then, we can check that if p321p_{32}\neq 1, then qji1q^{\prime}_{ji}\neq 1 for all i,j0i,j\neq 0 by using a computer. Moreover, if p32=1p_{32}=1, then qji=1q^{\prime}_{ji}=1 for all i,j0i,j\neq 0. In this case, |spec(C[x01]0)|1𝔸3|\mathrm{spec}(C[x_{0}^{-1}]_{0})|_{1}\simeq\mathbb{A}^{3}.

We consider C/(x0,x1)=kx2,x3/(x3x2q32x2x3)C/(x_{0},x_{1})=k\langle x_{2},x_{3}\rangle/(x_{3}x_{2}-q_{32}x_{2}x_{3}). Then, it is known that a weighted quantum polynomial ring of 22 variables is a twisted algebra of a commutative weighted polynomial ring k[x,y]k[x,y] with deg(x)=a>0,deg(y)=b>0\mathrm{deg}(x)=a>0,\mathrm{deg}(y)=b>0 (for example, see [28, Example 4.1] or [36, Example 3.6]). So, it is enough to consider the closed points of proj(k[x,y])\mathrm{proj}(k[x,y]). We want to study the closed points of proj(k[x,y])\mathrm{proj}(k[x,y]) in the case of (a,b)=(2,2),(2,4) or (4,6)(a,b)=(2,2),(2,4)\text{ or }(4,6). Note that when (a,b)=(1,1)(a,b)=(1,1) or (1,3)(1,3), they are classified in [18, Theorem 3.16]. We treat a more general setting below.

Lemma 4.5.

Let R=k[x,y]R=k[x,y] be a commutative weighted polynomial ring with deg(x)=a>0,deg(y)=b>0\mathrm{deg}(x)=a>0,\mathrm{deg}(y)=b>0. Let g:=gcd(a,b),a:=a/gg:=\mathrm{gcd}(a,b),a^{\prime}:=a/g and b:=b/gb^{\prime}:=b/g. Then, every closed point of proj(R)\mathrm{proj}(R) is one of the following:

  1. (1)

    πR/(x)(i),i=0,,b1\pi R/(x)(-i),\;i=0,\cdots,b-1.

  2. (2)

    πR/(y)(j),j=0,,a1\pi R/(y)(-j),\;j=0,\cdots,a-1.

  3. (3)

    πR/(βxbαya)(k),\pi R/(\beta x^{b^{\prime}}-\alpha y^{a^{\prime}})(-k), where (α,β)1\{(0,1),(1,0)}(\alpha,\beta)\in\mathbb{P}^{1}\backslash\{(0,1),(1,0)\} and k=0,,g1k=0,\cdots,g-1.

Moreover, all of them are not isomorphic in proj(R)\mathrm{proj}(R).

Proof.

The proof is almost the same as the proof of [18, Lemma 3.15, Theorem 3.16]. We give the sketch of the proof.

Firstly, every closed point of proj(R)\mathrm{proj}(R) is represented by a cyclic critical Cohen-Macaulay module of depth 11. Then, Mgr(R)M\in\mathrm{gr}(R) satisfies these conditions and is generated in degree 0 if and only if MM is isomorphic to one of R/(x),R/(y) or R/(βxbαya)(α,βk×)R/(x),R/(y)\text{ or }R/(\beta x^{b^{\prime}}-\alpha y^{a^{\prime}})\;(\alpha,\beta\in k^{\times}). Since being cyclic critical Cohen-Macaulay of depth 11 is invariant under shifting, any closed point is represented by some shifts of one of the above modules (that is, R/(x)(l),R/(y)(l),R/(βxbαya)(l),lR/(x)(-l),R/(y)(-l),R/(\beta x^{b^{\prime}}-\alpha y^{a^{\prime}})(-l),\ l\in\mathbb{Z}).

Finally, we classify the isomorphic classes of these modules in proj(R)\mathrm{proj}(R). We have no isomorphisms between the three types of closed points by considering their Hilbert polynomials and multiplicities. Then, we have πR/(βxbαya)πR/(βxbαya)(gl),(l,(α,β)1\{(1,0),(0,1)})\pi R/(\beta x^{b^{\prime}}-\alpha y^{a^{\prime}})\simeq\pi R/(\beta x^{b^{\prime}}-\alpha y^{a^{\prime}})(-gl),\ ({}^{\forall}l\in\mathbb{Z},{}^{\forall}(\alpha,\beta)\in\mathbb{P}^{1}\backslash\{(1,0),(0,1)\}). We also have πR/(βxbαya)πR/(βxbαya)\pi R/(\beta x^{b^{\prime}}-\alpha y^{a^{\prime}})\simeq\pi R/(\beta^{\prime}x^{b^{\prime}}-\alpha^{\prime}y^{a^{\prime}}) if and only if (α,β)=(α,β)(\alpha,\beta)=(\alpha^{\prime},\beta^{\prime}) in 1\{(1,0),(0,1)}\mathbb{P}^{1}\backslash\{(1,0),(0,1)\}. In addition, we can show that πR/(x)πR/(x)(i)\pi R/(x)\simeq\pi R/(x)(-i) (resp. πR/(y)πR/(y)(j)\pi R/(y)\simeq\pi R/(y)(-j)) if and only if i0(modb)i\equiv 0\;(\mathrm{mod}\;b) (resp. j0(moda)j\equiv 0\;(\mathrm{mod}\;a)). From these discussions, we get the claim. ∎

We can study ordinary and thin points of noncommutative projective Calabi-Yau 2 schemes in Theorem 3.15 by using the above investigations. We give examples of noncommutative projective Calabi-Yau schemes whose moduli of ordinary closed points are different from those in [11, Proposition 3.4] and commutative Calabi-Yau varieties.

Example 4.6.

We consider the weight (1,1,2,2)(1,1,2,2) and the quantum parameter

𝐪=(qij)=(111ω211ω211ω11ω111),ω:=1+i32.\mathbf{q}=(q_{ij})={\small\begin{pmatrix}1&1&1&\omega^{2}\\ 1&1&\omega^{2}&1\\ 1&\omega&1&1\\ \omega&1&1&1\\ \end{pmatrix}},\quad\omega:=\frac{-1+i\sqrt{3}}{2}.

Then, we have

𝐪=(qij)=(1ω2ωω1ω2ω2ω1),q13d2q32q21d3=ω2.\displaystyle\mathbf{q}^{\prime}=(q_{ij}^{\prime})={\small\begin{pmatrix}1&\omega^{2}&\omega\\ \omega&1&\omega^{2}\\ \omega^{2}&\omega&1\end{pmatrix}},\quad q_{13}^{d_{2}}q_{32}q_{21}^{d_{3}}=\omega^{2}.

From Lemma 4.3 and Lemma 4.5, the set of ordinary and thin points

|proj(C/(f))|ord & thin\displaystyle|\mathrm{proj}(C/(f))|_{\text{ord \& thin}} =|spec(C/(f)[x01]0)|1|spec(C/(f,x0)[x11]0)|1\displaystyle=|\mathrm{spec}(C/(f)[x_{0}^{-1}]_{0})|_{1}\bigsqcup|\mathrm{spec}(C/(f,x_{0})[x_{1}^{-1}]_{0})|_{1}
|proj(C/(f,x0,x1))|\displaystyle\bigsqcup|\mathrm{proj}(C/(f,x_{0},x_{1}))|

is 2424 points. To be more precise, we have |spec(C/(f)[x01]0)|1=ijZ(Xi,Xj,1+X16+X23+X33)𝔸3|\mathrm{spec}(C/(f)[x_{0}^{-1}]_{0})|_{1}=\bigsqcup_{i\neq j}Z(X_{i},X_{j},1+X_{1}^{6}+X_{2}^{3}+X_{3}^{3})\subset\mathbb{A}^{3}, |spec(C/(f,x0)[x11]0)|1=i=1,2Z(Yi,1+Y23+Y33)|\mathrm{spec}(C/(f,x_{0})[x_{1}^{-1}]_{0})|_{1}=\bigsqcup_{i=1,2}Z(Y_{i},1+Y_{2}^{3}+Y_{3}^{3}) and |proj(C/(f,x0,x1))|={3pts}{3pts}|\mathrm{proj}(C/(f,x_{0},x_{1}))|=\{3\text{pts}\}\sqcup\{3\text{pts}\}.

This calculation shows that for a fixed weight, if the number of the set of ordinary and thin points of proj(C/(f))\mathrm{proj}(C/(f)) is finite, then the number is independent of the quantum parameters.

From the method in Example 4.6, Remark 4.4 and a direct computation, we have the following.

Proposition 4.7.

For a weight (1,1,a,b)(1,1,a,b) in Example 4.1 and a quantum parameter 𝐪\mathbf{q} which gives a noncommutative projective Calabi-Yau scheme, if the set of ordinary and thin points of proj(C/(f))\mathrm{proj}(C/(f)) is finite, then the number of the set is always 2424.

The following proposition shows that some of noncommutative projective Calabi-Yau 2 schemes in Theorem 3.15 are essentially new examples.

Proposition 4.8.

There exists a noncommutative projective Calabi-Yau 2 scheme which is obtained in Theorem 3.15 and not isomorphic to either commutative Calabi-Yau surfaces or noncommutative projective Calabi-Yau 2 schemes obtained in [11].

Proof.

We divide the proof into four steps.

Step 1. We choose the weight (1,1,a,b)(1,1,a,b) and the quantum parameter 𝐪\mathbf{q} as in Example 4.6. Then, the number of ordinary and thin points of proj(C/(f))\mathrm{proj}(C/(f)) is finite. So, proj(C/f)\mathrm{proj}(C/f) is not isomorphic to any commutative Calabi-Yau surfaces.

Step 2. We prove that proj(C/(f))\mathrm{proj}(C/(f)) is not isomorphic to any noncommutative projective Calabi-Yau 2 schemes in [11]. To prove this, we use the theory established in [4]. First, note that we can think of qgr(C/(f))\mathrm{qgr}(C/(f)) as the category of coherent modules of a sheaf 𝒜\mathcal{A} of algebras on the projective spectrum Proj(k[s0,s1,s2,s3]/(s0+s1+s2+s3))\mathrm{Proj}(k[s_{0},s_{1},s_{2},s_{3}]/(s_{0}+s_{1}+s_{2}+s_{3})) (cf. the proof of Lemma 3.21). We define a sheaf 𝒵𝒜\mathcal{Z}_{\mathcal{A}} to be the sheaf whose sections are

Γ(U,𝒵𝒜)={sΓ(U,𝒜)s|VZ(Γ(V,𝒜)),VU:open}\Gamma(U,\mathcal{Z}_{\mathcal{A}})=\{s\in\Gamma(U,\mathcal{A})\mid s|_{V}\in Z(\Gamma(V,\mathcal{A})),{}^{\forall}V\subset U:\text{open}\}

for all open subsets UU (cf. [4, Proposition 2.11]). In particular, if UU is affine, Γ(U,𝒵𝒜)=Z(Γ(U,𝒜))\Gamma(U,\mathcal{Z}_{\mathcal{A}})=Z(\Gamma(U,\mathcal{A})). Then, we show that Spec(Z(Γ(D+(si),𝒜)))\mathrm{Spec}(Z(\Gamma(D_{+}(s_{i}),\mathcal{A}))) has 44 singular points when i=0,1i=0,1 and a 11-dimensional singular locus when i=2,3i=2,3. In the following, we verify this claim for i=0,2i=0,2. Similarly, the claim is proved for i=1,3i=1,3. In the following, we write ZiZ_{i} as Z(Γ(D+(si),𝒜))Z(\Gamma(D_{+}(s_{i}),\mathcal{A})) for any ii. We also use the notations in the proof of Lemma 3.21.

When i=0i=0, any mZ0m\in Z_{0} is of the form m=(μ1e00μ2e)N0,(eE0,0,μ1,μ2k×)m=\left(\begin{smallmatrix}\mu_{1}e&0\\ 0&\mu_{2}e\end{smallmatrix}\right)\in N_{0},\ (e\in E_{0,0},\mu_{1},\mu_{2}\in k^{\times}) from the definition of 𝒜\mathcal{A}. We have E0,0kX1,X2,X3(XjXiqjiXiXj,1+X16+X23+X33)i,jE_{0,0}\simeq k\langle X_{1},X_{2},X_{3}\rangle(X_{j}X_{i}-q^{\prime}_{ji}X_{i}X_{j},1+X_{1}^{6}+X_{2}^{3}+X_{3}^{3})_{i,j}, which is obtained from the identifications X1=x1x01,X2=x2x02X_{1}=x_{1}x_{0}^{-1},X_{2}=x_{2}x_{0}^{-2} and X3=x3x02X_{3}=x_{3}x_{0}^{-2}. Here, the qjiq^{\prime}_{ji} are as in Example 4.6. So, Z(E0,0)k[Y,Z,W,U]/(1+Y2+Z+W,YZWλ1U3)(λ1k×)Z(E_{0,0})\simeq k[Y,Z,W,U]/(1+Y^{2}+Z+W,YZW-\lambda_{1}U^{3})\ (\lambda_{1}\in k^{\times}), which is obtained from the identifications Y=(x1x01)3,Z=(x2x02)3,W=(x3x02)3Y=(x_{1}x_{0}^{-1})^{3},Z=(x_{2}x_{0}^{-2})^{3},W=(x_{3}x_{0}^{-2})^{3} and U=(x1x01)(x2x02)(x3x02)U=(x_{1}x_{0}^{-1})(x_{2}x_{0}^{-2})(x_{3}x_{0}^{-2}). On the other hand, we define the inclusion ϕ:Z(E0,0)N0\phi:Z(E_{0,0})\rightarrow N_{0} in which Y,Z,WY,Z,W are mapped naturally and UU to (U00ωU)\left(\begin{smallmatrix}U&0\\ 0&\omega U\end{smallmatrix}\right). It is easy to see that ϕ(Z(E0,0))Z0\phi(Z(E_{0,0}))\subset Z_{0}. Because the choice of μ1\mu_{1} determines μ2\mu_{2} in the above form of mm, the map ϕ\phi induces Z0Z(E0,0)Z_{0}\simeq Z(E_{0,0}). Thus, one can show that Spec(Z0)\mathrm{Spec}(Z_{0}) has 44 singular points by using the Jacobi criterion.

When i=2i=2, any mZ2m\in Z_{2} is of the form m=(μ1e00μ2e)N2,(eE2,0,μ1,μ2k×)m=\left(\begin{smallmatrix}\mu_{1}e&0\\ 0&\mu_{2}e\end{smallmatrix}\right)\in N_{2},\ (e\in E_{2,0},\mu_{1},\mu_{2}\in k^{\times}) from the definition of 𝒜\mathcal{A}. We also have E2,0kX0,X1,X2,X3/(XjXiqji′′XiXj,1+X06+X16+X33,X0X1λ2X22)i,j(λ2k×)E_{2,0}\simeq k\langle X_{0},X_{1},X_{2},X_{3}\rangle/(X_{j}X_{i}-q^{\prime\prime}_{ji}X_{i}X_{j},1+X_{0}^{6}+X_{1}^{6}+X_{3}^{3},X_{0}X_{1}-\lambda_{2}X_{2}^{2})_{i,j}\ (\lambda_{2}\in k^{\times}), which is obtained from the identifications X0=x02x21,X1=x12x21,X2=x0x1x21X_{0}=x_{0}^{2}x_{2}^{-1},X_{1}=x_{1}^{2}x_{2}^{-1},X_{2}=x_{0}x_{1}x_{2}^{-1} and X3=x3x21X_{3}=x_{3}x_{2}^{-1}. Here, the qij′′q^{\prime\prime}_{ij} are defined by the matrix

(qij′′)=(1ωω2ωω21ωω2ωω211ω2ω11).(q^{\prime\prime}_{ij})=\left(\begin{smallmatrix}1&\omega&\omega^{2}&\omega\\ \omega^{2}&1&\omega&\omega^{2}\\ \omega&\omega^{2}&1&1\\ \omega^{2}&\omega&1&1\end{smallmatrix}\right).

So, Z(E2,0)k[X,Y,W,U,V]/(X+Y+1+W,XYλ3U2,XYWλ4V2)(λ3,λ4k×)Z(E_{2,0})\simeq k[X,Y,W,U,V]/(X+Y+1+W,XY-\lambda_{3}U^{2},XYW-\lambda_{4}V^{2})\ (\lambda_{3},\lambda_{4}\in k^{\times}), which is obtained from the identifications X=(x02x21)3,Y=(x12x21)3,W=(x3x21)3,U=(x0x1x21)3X=(x_{0}^{2}x_{2}^{-1})^{3},Y=(x_{1}^{2}x_{2}^{-1})^{3},W=(x_{3}x_{2}^{-1})^{3},U=(x_{0}x_{1}x_{2}^{-1})^{3} and V=(x0x1x21)(x3x21)V=(x_{0}x_{1}x_{2}^{-1})(x_{3}x_{2}^{-1}). On the other hand, we define the inclusion ϕ:Z(E2,0)N2\phi:Z(E_{2,0})\rightarrow N_{2} in which X,Y,W,UX,Y,W,U are mapped naturally and VV to (V00ωV)\left(\begin{smallmatrix}V&0\\ 0&\omega V\end{smallmatrix}\right). It is easy to see that ϕ(Z(E2,0))Z2\phi(Z(E_{2,0}))\subset Z_{2}. Because the choice of μ1\mu_{1} determines μ2\mu_{2} in the above form of mm, the map ϕ\phi induces Z2Z(E2,0)Z_{2}\simeq Z(E_{2,0}). Thus, one can show that Spec(Z2)\mathrm{Spec}(Z_{2}) has a 11-dimensional singular locus by using the Jacobi criterion.

Step 3. We consider the weight (1,1,1,1)(1,1,1,1) and take a quantum parameter which gives a noncommutative projective Calabi-Yau 2 scheme proj(C/(f))\mathrm{proj}(C^{\prime}/(f^{\prime})) whose point scheme is finite. qgr(C/(f))\mathrm{qgr}(C^{\prime}/(f^{\prime})) is thought of as the category of coherent modules of a sheaf \mathcal{B} of algebras on the projective spectrum Proj(k[t0,t1,t2,t3]/(t0+t1+t2+t3))\mathrm{Proj}(k[t_{0},t_{1},t_{2},t_{3}]/(t_{0}+t_{1}+t_{2}+t_{3})).

The number of the choices of quantum parameters (qij)(q_{ij}) which satisfy the conditions of Theorem 3.15 and give a noncommutative projective Calabi-Yau scheme whose moduli space of point modules is finite is 2020 except permutating variables (we get the list below by using a computer and hand calculations):

1.(1111111111111111),2.(111111ii1i1i1ii1),3.(1111111111111111),4.(111111ii1i1i1ii1),\displaystyle 1.\left(\begin{smallmatrix}1&1&1&1\\ 1&1&-1&-1\\ 1&-1&1&-1\\ 1&-1&-1&1\end{smallmatrix}\right),2.\left(\begin{smallmatrix}1&1&1&1\\ 1&1&-i&i\\ 1&i&1&-i\\ 1&-i&i&1\end{smallmatrix}\right),3.\left(\begin{smallmatrix}1&1&1&-1\\ 1&1&-1&1\\ 1&-1&1&1\\ -1&1&1&1\end{smallmatrix}\right),4.\left(\begin{smallmatrix}1&1&1&-1\\ 1&1&-i&-i\\ 1&i&1&i\\ -1&i&-i&1\end{smallmatrix}\right),
5.(111i111i111iiii1),6.(111i11i11i11i111),7.(111i111i111iiii1),8.(111i11i11i11i111),\displaystyle 5.\left(\begin{smallmatrix}1&1&1&i\\ 1&1&-1&-i\\ 1&-1&1&-i\\ -i&i&i&1\end{smallmatrix}\right),6.\left(\begin{smallmatrix}1&1&1&i\\ 1&1&-i&-1\\ 1&i&1&1\\ -i&-1&1&1\end{smallmatrix}\right),7.\left(\begin{smallmatrix}1&1&1&-i\\ 1&1&-1&i\\ 1&-1&1&i\\ i&-i&-i&1\end{smallmatrix}\right),8.\left(\begin{smallmatrix}1&1&1&-i\\ 1&1&-i&1\\ 1&i&1&-1\\ i&1&-1&1\end{smallmatrix}\right),
9.(111111ii1i1i1ii1),10.(111i11i11i11i111),11.(111i11i11i11i111),12.(11ii11iiii11ii11),\displaystyle 9.\left(\begin{smallmatrix}1&1&-1&-1\\ 1&1&-i&i\\ -1&i&1&i\\ -1&-i&-i&1\end{smallmatrix}\right),10.\left(\begin{smallmatrix}1&1&-1&i\\ 1&1&i&-1\\ -1&-i&1&-1\\ -i&-1&-1&1\end{smallmatrix}\right),11.\left(\begin{smallmatrix}1&1&-1&-i\\ 1&1&-i&-1\\ -1&i&1&-1\\ i&-1&-1&1\end{smallmatrix}\right),12.\left(\begin{smallmatrix}1&1&i&i\\ 1&1&-i&-i\\ -i&i&1&-1\\ -i&i&-1&1\end{smallmatrix}\right),
13.(11ii11iiii11ii11),14.(1111111111111111),15.(111111ii1i1i1ii1),16.(111i111i111iiii1),\displaystyle 13.\left(\begin{smallmatrix}1&1&i&-i\\ 1&1&-i&i\\ -i&i&1&1\\ i&-i&1&1\end{smallmatrix}\right),14.\left(\begin{smallmatrix}1&-1&-1&-1\\ -1&1&-1&-1\\ -1&-1&1&-1\\ -1&-1&-1&1\end{smallmatrix}\right),15.\left(\begin{smallmatrix}1&-1&-1&-1\\ -1&1&-i&i\\ -1&i&1&-i\\ -1&-i&i&1\end{smallmatrix}\right),16.\left(\begin{smallmatrix}1&-1&-1&i\\ -1&1&-1&i\\ -1&-1&1&i\\ -i&-i&-i&1\end{smallmatrix}\right),
17.(111i111i111iiii1),18.(11ii11iiii11ii11),19.(1iiii1iiii1iiii1),20.(1iiii1iiii1iiii1).\displaystyle 17.\left(\begin{smallmatrix}1&-1&-1&-i\\ -1&1&-1&-i\\ -1&-1&1&-i\\ i&i&i&1\end{smallmatrix}\right),18.\left(\begin{smallmatrix}1&-1&i&i\\ -1&1&i&i\\ -i&-i&1&-1\\ -i&-i&-1&1\end{smallmatrix}\right),19.\left(\begin{smallmatrix}1&i&i&i\\ -i&1&-i&i\\ -i&i&1&-i\\ -i&-i&i&1\end{smallmatrix}\right),20.\left(\begin{smallmatrix}1&i&i&-i\\ -i&1&-i&-i\\ -i&i&1&i\\ i&i&-i&1\end{smallmatrix}\right).

When we choose one (qij)(q_{ij}) of the above 20 quantum parameters, then for any ll, Γ(D+(tl),)kY1,Y2,Y3/(YiYjqijYjYi,Y14+Y24+Y34+1)1i,j3\Gamma(D_{+}(t_{l}),\mathcal{B})\simeq k\langle Y_{1},Y_{2},Y_{3}\rangle/(Y_{i}Y_{j}-q_{ij}^{\prime}Y_{j}Y_{i},Y_{1}^{4}+Y_{2}^{4}+Y_{3}^{4}+1)_{1\leq i,j\leq 3}, where (qij)(q^{\prime}_{ij}) is represented by one of the following matrices (we can verify this with direct calculations) :

(a).(111111111),(b).(1iii1iii1).(a).\left(\begin{smallmatrix}1&-1&-1\\ -1&1&-1\\ -1&-1&1\end{smallmatrix}\right),\qquad(b).\left(\begin{smallmatrix}1&-i&i\\ i&1&-i\\ -i&i&1\end{smallmatrix}\right).

We write Zl:=Z(Γ(D+(tl),))Z_{l}^{\prime}:=Z(\Gamma(D_{+}(t_{l}),\mathcal{B})). When (qij)(q^{\prime}_{ij}) is type (a), Spec(Zl)\mathrm{Spec}(Z^{\prime}_{l}) has 6 singular points because ZlZ_{l}^{\prime} is generated by Y12,Y22,Y32Y_{1}^{2},Y_{2}^{2},Y_{3}^{2} and Y1Y2Y3Y_{1}Y_{2}Y_{3} as a kk-algebra. When (qij)(q^{\prime}_{ij}) is type (b), Spec(Zl)\mathrm{Spec}(Z_{l}^{\prime}) has 3 singular points because ZlZ_{l}^{\prime} is generated by Y14,Y24,Y34Y_{1}^{4},Y_{2}^{4},Y_{3}^{4} and Y1Y2Y3Y_{1}Y_{2}Y_{3} as a kk-algebra. Moreover, for any (qij)(q_{ij}) in the above table, if \mathcal{B} is type (a) (resp. (b)) on D+(tl)D_{+}(t_{l}) for some ll, it is also type (a) (resp. (b)) on D+(tl)D_{+}(t_{l}) for any other ll.

Step 4. If qgr(C/(f))\mathrm{qgr}(C/(f)) is equivalent to qgr(C/(f))\mathrm{qgr}(C^{\prime}/(f^{\prime})) then, we must have an isomorphism of schemes between Spec(𝒵𝒜)\mathrm{Spec}(\mathcal{Z}_{\mathcal{A}}) and Spec(𝒵)\mathrm{Spec}(\mathcal{Z}_{\mathcal{B}}) by [4, Theorem 4.4] (cf. [1, Section 6]). Since Spec(𝒵𝒜)\mathrm{Spec}(\mathcal{Z}_{\mathcal{A}}) has infinite singular points, but, Spec(𝒵)\mathrm{Spec}(\mathcal{Z}_{\mathcal{B}}) has finite singular points, such a situation does not happen. Hence, we complete the proof. ∎

Acknowledgements

The author would like to express his gratitude to his supervisor Professor Hajime Kaji for his encouragement. He is also grateful to Professor Atsushi Kanazawa for telling him the articles [13], [14]. He would like to thank Professor Izuru Mori, Professor Balázs Szendroi, Professor Ryo Ohkawa, Professor Shinnosuke Okawa and Professor Kenta Ueyama for helpful comments. In addition, he thanks Niklas Lemcke for proofreading his English. This work is supported by Grant-in-Aid for JSPS Fellows (Grant Number 22KJ2923).

References

  • [1] M. Artin and J. J. Zhang “Noncommutative Projective Schemes” In Adv. Math. 109.2, 1994, pp. 228–287
  • [2] Pieter Belmans, Kevin De Laet and Lieven Le Bruyn “The point variety of quantum polynomial rings” In Journal of Algebra 463 Elsevier, 2016, pp. 10–22
  • [3] A. Bondal and M. Bergh “Generators and representability of functors in commutative and noncommutative geometry” In Mosc. Math. J. 3.1, 2003, pp. 1–36, 258
  • [4] I. Burban and Y. Drozd “Morita theory for non–commutative Noetherian schemes” In Adv. Math. 399 Elsevier, 2022, pp. 1–42
  • [5] K. De Naeghel and M. Bergh “Ideal classes of three-dimensional Sklyanin algebras” In J. Algebra 276.2 Elsevier, 2004, pp. 515–551
  • [6] D. Eisenbud “The geometry of syzygies: A second course in commutative algebra and algebraic geometry” Springer, 2005
  • [7] V. Ginzburg “Calabi-Yau algebras” In arXiv:0612139, 2006
  • [8] J.-W. He and K. Ueyama “Twisted Segre products” In J. Algebra 611 Elsevier, 2022, pp. 528–560
  • [9] E. Hyry “The diagonal subring and the Cohen-Macaulay property of a multigraded ring” In Trans. Amer. Math. Soc. 351.6, 1999, pp. 2213–2232
  • [10] A. R. Iano-Fletcher “Working with weighted complete intersections” In Explicit Birational Geometry of 3-folds 281, London Mathematical Society Lecture Note Series Cambridge University Press, 2000, pp. 101–174
  • [11] A. Kanazawa “Non-commutative projective Calabi–Yau schemes” In J. Pure Appl. Algebra 219.7, 2015, pp. 2771–2780
  • [12] A. Kuznetsov “Calabi–Yau and fractional Calabi–Yau categories” In J. Reine Angew. Math. 2019.753 De Gruyter, 2019, pp. 239–267
  • [13] Y.-H. Liu “Donaldson–Thomas theory of quantum Fermat quintic threefolds I” In arXiv.1911.07949, 2019
  • [14] Y.-H. Liu “Donaldson–Thomas theory of quantum Fermat quintic threefolds II” In arXiv:2004.10346, 2020
  • [15] J. C. McConnell, J. C. Robson and L. W. Small “Noncommutative Noetherian rings” American Mathematical Soc., 2001
  • [16] R. Mckemey “Relative local cohomology”, Thesis, University of Manchester, 2012
  • [17] I. Mori “B-construction and C-construction” In Comm. Algebra 41.6 Taylor & Francis, 2013, pp. 2071–2091
  • [18] I. Mori “Regular modules over 2-dimensional quantum Beilinson algebras of Type S” In Math. Z. 279.3 Springer, 2015, pp. 1143–1174
  • [19] I. Mori and A. Nyman “A categorical characterization of quantum projective \mathbb{Z}-spaces” In arXiv:2307.15253, 2023
  • [20] I Mori and A Nyman “Corrigendum to “Local duality for connected \mathbb{Z}-algebras” [J. Pure Appl. Algebra 225 (9)(2019) 106676]” In J. Pure Appl. Algebra 228.3 Elsevier, 2024, pp. 107493
  • [21] I. Mori and A. Nyman “Local duality for connected \mathbb{Z}-algebras” In J. Pure Appl. Algebra 225.9, 2021, pp. Paper No. 106676, 22
  • [22] M. Reid “Canonical 3-folds”, Journees de geometrie algebrique, Angers/France 1979, 273-310, 1980
  • [23] M. Reyes, D. Rogalski and J. J. Zhang “Skew Calabi–Yau algebras and homological identities” In Adv. Math. 264, 2014, pp. 308–354
  • [24] M. L. Reyes and D. Rogalski “Graded twisted Calabi–Yau algebras are generalized Artin–Schelter regular” In Nagoya Math. J. 245 Cambridge University Press, 2022, pp. 100–153
  • [25] S. P. Smith “Maps between non-commutative spaces” In Trans. Amer. Math. Soc. 356.7, 2004, pp. 2927–2944
  • [26] S. P. Smith “Noncommutaive algebraic geometry”, Lecture Notes, University of Washington, 2000
  • [27] S. P. Smith “Subspaces of non-commutative spaces” In Trans. Amer. Math. Soc. 354.6, 2002, pp. 2131–2171
  • [28] D. R. Stephenson “Quantum planes of weight (1,1,n)(1,1,n) In J. Algebra 225.1, 2000, pp. 70–92
  • [29] M. Van den Bergh “Calabi-Yau algebras and superpotentials” In Selecta Math. (N.S.) 21.2, 2015, pp. 555–603
  • [30] M. Van den Bergh “Existence Theorems for Dualizing Complexes over Non-commutative Graded and Filtered Rings” In J. Algebra 195.2, 1997, pp. 662–679
  • [31] K. Van Rompay “Segre Product of Artin–Schelter Regular Algebras of Dimension 2 and Embeddings in Quantum 3\mathbb{P}^{3}’s” In J. Algebra 180.2, 1996, pp. 483–512
  • [32] Jorge Vitoria “Equivalences for noncommutative projective spaces” In arXiv:1001.4400, 2010
  • [33] A. Yekutieli “Derived Categories” Cambridge University Press, 2019
  • [34] A. Yekutieli “Dualizing complexes over noncommutative graded algebras” In J. Algebra 153.1, 1992, pp. 41–84
  • [35] A. Yekutieli and J. J. Zhang “Serre duality for noncommutative projective schemes” In Proc. Amer. Math. Soc. 125.3, 1997, pp. 697–707
  • [36] J. J. Zhang “Twisted graded algebras and equivalences of graded categories” In Proc. Lond. Math. Soc. 3.2 Oxford Academic, 1996, pp. 281–311