1,209
Views
2
CrossRef citations to date
0
Altmetric
Original Articles

Gorenstein Formats, Canonical and Calabi–Yau Threefolds

, &

Abstract

Gorenstein formats present the equations of regular canonical, Calabi–Yau and Fano varieties embedded by subcanonical divisors. We present a new algorithm for the enumeration of these formats based on orbifold Riemann–Roch and knapsack packing-type algorithms. We apply this to extend the known lists of threefolds of general type beyond the well-known classes of complete intersections and also to find classes of Calabi–Yau threefolds with canonical singularities.

2010 Mathematics Subject Classification:

1. Introduction

General smooth K3 surfaces of genus 5 embed as complete intersections of three quadrics S2,2,2P5 in codimension 3. Altınok [CitationAltınok 05] discovered 69 others families of K3 surfaces that also embed as projectively Gorenstein varieties in codimension 3 in weighted projective spaces, SP(a0,,a5) for various weights 1a0a5. These are non-complete intersections, each defined by five equations that arise as the Pfaffians of skew 5 × 5 matrices. Corti and Reid [CitationCorti and Reid 02] and Grojnowski develop a general theoretical framework of weighted Grassmannians encompassing these cases: the equations arise as regular pullbacks from various weighted Grassmannians wGr(2,5)wP9, each of which describes a kind of systematic structure, or “format,” for the equations of a variety (see Definition 3.1; also Stevens [CitationStevens 03, Section 12]). This paper applies knapsack packing-type algorithms to enumerate new varieties embedded in various formats.

While this paper focuses on constructing threefolds, the methods apply without change to construct polarized d-dimensional orbifolds X, A with canonical class KX = kA, for any integer k, that have zero-dimensional orbifold locus; such a polarizing divisor A is termed subcanonical. The orbifold restriction is imposed only because we do not know the contribution to orbifold Riemann–Roch of higher dimensional orbifold strata; but see Zhou [CitationZhou 11] and Selig [CitationSelig 15] for progress. Computer code that can make such searches systematically, written for the computational algebra system [CitationBosma et al. 97], is available for download in [CitationBrown and Kasprzyk].

1.1. The equations of canonical threefolds

This paper focuses on threefolds, that is, complex three-dimensional projective varieties with Q-factorial canonical singularities. A canonical threefold is one that has ample canonical class. For example, a nonsingular sextic hypersurface X6P4 is a canonical threefold, with canonical ring R(X,KX) (see Section 2) isomorphic to its homogeneous coordinate ring. The canonical ring is rarely generated in degree one: the double cover of P3 branched in a nonsingular surface of degree ten is a hypersurface X10P(1,1,1,1,5) whose canonical ring is again its homogeneous coordinate ring, in this case generated in degrees 1,1,1,1,5. Iano-Fletcher [CitationIano-Fletcher 00, Table 3] lists 23 families of such weighted canonical hypersurfaces, the most exotic being X46P(4,5,6,7,23).

Table 3. Codimension three.

Iano-Fletcher [CitationIano-Fletcher 00, Section 16.7] also lists 59 families of canonical threefolds Xd1,d2P(a0,,a5) in codimension 2. His method is to work systematically through all possible ai, up to ai100, and d1, d2 satisfying d1+d2=1+ai. Since the results all have relatively small ai (the biggest, X12,28P(3,4,5,6,7,14), has ai=39) he conjectures [CitationIano-Fletcher 00, Section 18.19] that the lists are the complete classification; this is proved by Chen–Chen–Chen [CitationChen et al. 11, Theorem 7.4], classifying all (general) canonical threefold complete intersections.

After formidable calculation, Corti–Reid [CitationCorti and Reid 02] discovered a canonical threefold defined similarly by five equations in wGr(2,5) format. Our first result extends this to 18 cases, treating the Corti–Reid framework as a format for the equations of a variety.

Theorem 1.1.

There are 18 deformation families of canonical threefolds whose general member embeds pluricanonically as a codimension three subvariety XP(a0,,a6) with equations in weighted Grassmannian Gr(2,5) format for which ai70. These 18 families are described in .

This result extends the classification of Iano-Fletcher and Chen–Chen–Chen to the first case of non-complete intersections (that is, the case of lowest codimension in the pluricanonical embedding).

These 18 are striking, but the main point is that one can go much further with these constructions using different formats: we consider both the intersection of a wGr(2,5) format by a residual hypersurface (which mimics the equation format of the six equations of the canonical model of a non-trigonal curve of genus 6 with no g52), and the equations of OGr(5,10) in codimension 5 (which mimic the 10 equations of canonical models of curves of genus 7 with no g41 [CitationMukai 95, Main Theorem]).

Theorem 1.2.

  1. There are 57 families of canonical threefolds whose general member embeds pluricanonically as a codimension four subvariety XP(a0,,a6) with 6 equations in format Gr(2,5)H, that is, weighted Grassmannian Gr(2,5) format with a residual intersection hypersurface, for which ai45.

  2. There are 21 families of canonical threefold whose general member embeds pluricanonically as a codimension five subvariety XP(a0,,a8) with equations in weighted orthogonal Grassmannian OGr(5,10) format for which ai147. These 21 families are described in .

It is at least possible that the 18 families of Theorem 1.1 realize all canonical threefolds in codimension 3, without the restriction ai70, other than complete intersections and their degenerations. While the search space is infinite, there are only finitely many solutions, and there is some indication that all solutions arise early in the search; see Section 1.2 for discussion. Similarly, it is possible that the 21 families of Theorem 1.2(b) give the complete list of canonical threefolds in codimension 5 OGr(5,10) format. In contrast, the 57 families of Theorem 1.2(a) is certainly not the complete list of varieties in that format: we expect many other families of canonical threefold in codimension 4 with six equations, and we give an example of one in Section 1.2. As a weaker statement, it follows from Theorem 3.11 that the lists of possibilities in the two theorems is complete for each pair a0, a1 that appears.

To make a comparison with known results, we apply our methods to surfaces of general type (see Section 2.1, ); the resulting surfaces are not new, but many surfaces with small pg and K2 that are central to the classification appear readily. Further results appear in (discussed in Section 1.2), in Section 5.2, and on the online Graded Ring Database [CitationBrown and Kasprzyk], with computer code that can be used to generate many other cases.

Such classification results are particular applications of the main part of this paper, which is devoted to describing our computational approach (Section 4). These techniques apply automatically to any prescribed format, and work in any dimension. Crucially, our method of searching is both systematic and exhaustive.

Qureshi and Szendrői [CitationQureshi 15CitationQureshi and Szendrői 12] develop other formats based on other classical groups (these are included in our computer package [CitationBrown and Kasprzyk]). They too apply them to finding varieties by an approach based on the singularity baskets. One difference is that these baskets are part of the output of our method, rather than the input; this is a key advantage when baskets get large or complicated, as they can do (see ).

Table 1. The number of cases of Fano, Calabi–Yau, and canonical 3-dimensional orbifolds in various formats. All were computed allowing isolated canonical quotient singularities. The column klast gives the largest adjunction number for which a result was found; kmax gives the largest degree searched; #raw gives the number of candidates found by the computer; #results gives the number of candidates after removing obvious failures. (The 317 Calabi–Yau hypersurfaces are taken from [CitationKreuzer and Skarke 00] for completeness, since the method we use here is not effective in that case.)

Table 2. Examples of surfaces S of general type, polarized by A=1kKS, in various formats: #results gives the number of numerical types that arise early in the search, and the right-most column lists these as pairs of invariants, pg and KS2, that are realized by surfaces. The general member of each family with k = 1 is smooth; Z/2 canonical quotient points (A1 singularities where A is not Cartier) often appear when k = 2.

Table 4. Codimension five.

1.2. Results for threefolds: understanding

The method we describe also constructs varieties other than canonical varieties. summarizes results for other threefolds to illustrate the flexibility and limits of our approach. It lists the number of “candidates” for varieties. A candidate is essentially a set of ambient weights a0,,an and baskets of quotient singularities compatible with a Hilbert series; a candidate may or may not be realized by a variety in the chosen format (see Definition 3.4).

is generated by a systematic computer search in order of increasing adjunction number k=ai, the adjunction number of the ambient space. The search continues until the calculations become unwieldy. The table indicates this stopping point: kmax is the largest adjunction number up to which the search is complete. It also records the largest adjunction number, denoted klast(kmax), for which a candidate was found. records the number of candidates found, denoted #raw. In a few case, it is easy to see that there cannot be a quasismooth realization of a candidate. For example, any threefold (1–1) X6,30P(1,2,3,4,10,15),(1–1) has a nonterminal singularity at XP(10,15); the degree six equation cannot give a tangent term there. The final column #results records the number of results after removing such cases that obviously fail.

When kmax is much larger than klast, it is conceivable that we have found all the results. For example, in the cases of canonical threefolds in codimensions 3 and 5, the gap kmaxklast where no new results appear compares with the similar gap in Iano–Fletcher’s calculations for complete intersections [CitationIano-Fletcher 00]. It is only in this sense that we may imagine that those two lists may be complete.

When the two numbers kmax and klast are close, almost certainly we are only part of the way through the complete list. For example, a general codimension 4 variety XP(4,5,6,6,7,7,8,9) defined by an equation of degree 18 and the maximal Pfaffians of a 5 × 5 antisymmetric matrix with degrees (45676788910) is a quasismooth canonical threefold with adjunction number k = 53, which exceeds kmax in this case, and so does not appear in .

1.3. The method of computation

Our proof of the theorems above is based on the orbifold Riemann–Roch formula of Buckley, Reid, and Zhou [CitationBuckley et al. 13], which we state in our context as Theorem 3.8. We show that the terminal singularities arising on canonical threefolds make strictly positive contributions to this formula (Theorem 3.11), which bounds the number of possible baskets of singularities for given invariants.

The crucial novelty of our approach is that we do not search through the space of weights ai or possible baskets of singularities, but instead solve for the ai and the singularities; in particular, there are no assumptions about the number of singularities. The primary objects we enumerate are Gorenstein formats, essentially the graded Betti data of a free resolution, as in Definition 3.1. Section 4.1 explains how this leads to a knapsack-style problem for the other numerical data. Solving this presents small numbers of numerical candidates for varieties that we then consider case by case.

2. Graded rings of varieties

We explain the more general setup. If A is an ample divisor on a projective variety X (A is not assumed to be effective), one may consider the graded ring R(X,A)=m0H0(X,mA) of the polarized variety (X, A). Since A is ample, XProjR(X,A), and if R(X, A) is generated in degrees a0,,an with homogeneous relations f1,,fs, then X(f1==fs=0)P(a0,,an).

Denoting the weighted degree of each weighted homogeneous polynomial fi by di=deg(fi), we slightly abuse notation and abbreviate the data by X=Xd1,,dsP(a0,,an).

We refer to the codimension of X as its codimension ndim(X) in this embedding (which depends on A). When X is a complete intersection, dim(X)=ns and this unambiguously describes a general such X.

We consider cases for which KX = kA for some kZ. Goto and Watanabe [CitationGoto and Watanabe 78] characterize such graded rings.

Theorem 2.1

([CitationGoto and Watanabe 1978, 5.1.9–11]). Let X be a projective variety and A an ample divisor. Set R=R(X,A), the corresponding graded ring, so that X=ProjR. If R is Cohen–Macaulay then

  1. Hi(X,OX)=0 for 0<i<dimX;

  2. R is Gorenstein if and only if KX = kA for some integer k.

By the minimal model program [CitationBirkar et al. 10, CitationMori 88], each birational equivalence class of varieties includes a variety X that either has KX nef (that is, KXC0 for every complete curve CX) or admits a morphism f:XY with KX relatively ample (that is, KXC>0 for every complete curve CX contracted by f). The three possibilities KX ample, KX = 0 and KX ample are the three extreme cases, and these are particularly important from the point of view of birational classification.

The first of these three classes is vast: any variety V of general type is birational to its unique canonical model X=ProjR(V,KV) which has KX ample; the finite generation of R(V,KV) in this case is the celebrated result of [CitationBirkar et al. 10]. Thus birational classification is equivalent to listing canonical models. Although the number of generators of the canonical ring R(V,KV) is not bounded, we may hope to classify those cases with few generators, up to some bound; Theorems 1.1 and 1.2 take this approach.

The second class KX = 0 of, loosely speaking, Calabi–Yau varieties have been studied in examples defined by explicit equations since at least Hirzebruch [CitationHirzebruch 87]; we discuss this case in Section 5.2.1. The third class, Fano varieties, is known to be bounded (under additional conditions on singularities), and so an attempt at explicit classification describing varieties by small sets of equations may ultimately provide the whole classification—see, for example, [Altınok Citation02], and the foreword to [CitationCorti and Reid 00].

2.1. The equations of regular surfaces of general type

Canonical surfaces S=Proj(S,KS) with KS ample have been studied intensively for decades, very often using explicit descriptions. We assume in addition that S is regular, that is, q=h1(S,OS)=0. Following Persson [CitationPersson 87, Section 2], the set of all such surfaces is often understood as a “geography” by plotting pg=h0(S,KS) against KS2 (or equivalently χ(OS)=1+pg against KS2, or the Euler characteristic c2(X) against c1(S)2).

In , we follow the program described in §4.1 in dimension 2 for a few steps as a comparison with the threefold case, which is our main interest here. For surfaces, this is merely a crude first step, and these cases are well known to experts, especially among the canonical models (k = 1): for example, pg = 2, KS2=1 is realized by S10P(12,2,5), the famous case for which 4KS is not birational; pg = 1, KS2=1 is realized by S6,6P(1,22,32) (see [CitationCatanese 79]); and so on.

A single equation format does not usually describe all surfaces that realize given numerical invariants. For example, pg = 3, KS2=4 is realized by a complete intersection S4,4P(13,22), but there are also such surfaces in Gr(2,5) format in P(13,22,3), which are codimension 1 in moduli where |KS| picks up a base point, and others in codimension 4; see [CitationReid 89, Theorems 2.1, 3.1], [CitationDicks 88]. The celebrated case pg = 4, KS2=7 is yet more complex (see [CitationBauer 01, 5]) with several different formats across different components of moduli, while K2=8 is far from complete (see [CitationBauer and Roberto 09, CitationCatanese et al. 14]). The case pg = 6, K2=11 is in Ashikaga–Konno [CitationAshikaga and Konno 90] (see Example 3.6) while K2=13 is in Neves [CitationNeves 03].

also includes the first few cases with k = 2. These are surfaces polarized by A=12KS (not assumed to be effective or Cartier). These surfaces also have canonical models, and it varies from case to case whether the model with k = 1 or 2 is the simpler. For example, the case pg = 3, KS2=2 appears for both k = 1 and 2. A general such surface is a double cover of P2 branched over an octic, S8P(13,4) (k = 1). When the octic degenerates to the transverse union of a cubic and a quintic, S8 gains 15 ordinary double points above the intersections. Such surfaces admit a half-canonical model as T6,10P(23,3,5) (k = 2), where the 15 nodes are the 15 Z/2 quotient singularities; the map TS is simply the Veronese, which in this case, and rather untypically, happens to lie in smaller codimension.

3. Formats and candidate varieties

3.1. Regular pullbacks from key varieties

A format describes a presentation of the equations of a variety, for example, by saying that the equations are minors of some matrix. Informal notions of format for polynomial equations appear regularly, sometimes describing a component of a Hilbert scheme or capturing some other feature of the geometry, and there are more formal prescriptions such as [CitationStevens 03, Section 12]. We define format to suit our applications, loosely following Dicks and Reid [CitationReid 89, Theorem 3.3], [CitationReid 11, Section 1.5]:

Definition 3.1.

A Gorenstein format F of codimension c is a triple (V˜,χ,F) consisting of:

  1. A Gorenstein (in particular, Cohen–Macaulay) affine variety V˜Cn of codimension c, which we refer to as the key variety of the format;

  2. A diagonal C* action on V˜ with strictly positive weights χ, which we refer to as the key weights of the format;

  3. A graded minimal free resolution F of OV˜ as a graded OCn-module.

The C* actions on Cn that are compatible with its toric structure are parametrized by the character lattice NCn=Zn, and the positive actions are those lying strictly in the positive quadrant QNCn. A subset ΛNCn of these actions leave V˜ invariant, and condition (ii) asserts that ΛQ is not empty. We need a little more: that the given free resolution F is equivariant for the action. In many cases we consider, the key variety has monomial syzygies, so the homogeneity of the equations of V˜ is enough, and ΛQ is some (infinite) polyhedron in Q. We then iterate over the formats by enumerating the points of ΛQ.

Condition (iii) determines the Hilbert numerator Pnum(t) of the format: Pnum(t)=1tdi+tej+(1)ctk, where di are the degrees of the equations, ej the degrees of the first syzygies, and so on, and k is the adjunction number of F. This polynomial has Gorenstein symmetry: tkPnum(1/t)=(1)cPnum(t). It determines the Hilbert series, as in Proposition 3.3.

One could imagine other definitions of format, both weaker and stronger, but this one is well adapted to our applications.

Let F=(V˜,χ,F) be a Gorenstein format of codimension c. We construct Gorenstein varieties XPd+c(W) of codimension c and dimension d in weighted projective space, with weights W, as regular pullbacks, which we recall from [CitationReid 11, Section 1.5]:

Proposition 3.2

(Reid [CitationReid 11]). Let (V˜Cn,χ,F) be a Gorenstein format of codimension c. Let R be a polynomial ring and φ:SpecRCn a morphism. The following are equivalent:

  1. φ1(V˜)SpecR has codimension c;

  2. The pullback of F by φ is a free resolution of R-modules;

  3. xiφ*(xi) for i=1,,n form a regular sequence on SpecR×Cn, where x1,,xn are the coordinates of Cn.

If these conditions hold then φ1(V˜)SpecR is called a regular pullback of V˜, and is a Gorenstein affine variety. Furthermore, if R is graded by weights W and φ is graded of degree zero with respect to W and χ, then the pullback of F by φ is a graded minimal free resolution of R-modules with the same Hilbert numerator as F.

Fix any integer d > 0, the dimension of the varieties X that we seek. Let F=(V˜,χ,F) be a Gorenstein format of codimension c and fix a graded polynomial ring R with d+c+1 variables and strictly positive weights W. If φ:SpecRCn is graded of degree zero and φ1(V˜)SpecR is a regular pullback containing the origin OSpecR, then we define the projectivised regular pullback to be X=φ1(V˜)//W C*=(φ1(V˜)O)/C*P(W).

The next proposition follows immediately: the Hilbert series of X is determined by the graded Betti numbers of a free resolution, and since φ satisfies the conditions of Proposition 3.2 and has degree zero, the graded Betti numbers are exactly those of F with grading χ.

Proposition 3.3.

Let F=(V˜Cn,χ,F) be a Gorenstein format of codimension c, R a polynomial ring graded by strictly positive weights W with a morphism φ:SpecRCn graded of degree zero. Then every projectivised regular pullback XP(W) has Hilbert series PX(t)=Pnum(t)/aW(1ta) where Pnum(t) is the Hilbert numerator of the format F.

If, in addition, X is an irreducible variety that is well-formed as a subvariety of P(W) then the canonical sheaf of X is ωX=OX(kV˜α), where α is the sum of the weights W and kV˜=degPnum(t) is the adjunction number of F.

Recall that XP(W) is well formed if the intersection of X with any non-trivial orbifold locus of P(W) has codimension at least two in X; see [CitationIano-Fletcher 00, Definition 6.9].

Definition 3.4.

A candidate variety is a format F=(V˜,χ,F) of codimension c together with a morphism φ:SpecRCn of degree zero from a graded polynomial ring R that satisfies the equivalent conditions of Proposition 3.2. A candidate variety is well-formed if the projectivised regular pullback XP(W) is well-formed as a subvariety.

We think of a candidate variety as representing general members of a family of varieties in a common weighted projective space whose equations and syzygies are modeled on a common free resolution F. The condition only asks for a single map, although in the practical situations we encounter below any sufficiently general map will work. The space of maps SpecRCn of degree zero that give regular pullbacks may have more than one component, but we do not consider this question at all.

Example 3.5.

Following Corti and Reid [CitationCorti and Reid 02], let V˜=CGr(2,5)C10 be the affine cone over the Grassmannian Gr(2,5) in its Plücker embedding. The equations of V˜ are the maximal Pfaffians of a generic skew 5 × 5 matrix M=(x1x2x3x4x5x6x7x8x9x10) (we write only the strict upper-triangular part of such matrices). These equations are homogeneous with respect to a five-parameter system of weights Z5=ΛZ10, which one can determine by enforcing homogeneity of these Pfaffians.

We can use V˜ as a key variety to find K3 surfaces. Let χ=(3,4,4,5,5,5,6,6,7,7)Λ, which we understand better in matrix form as χ=(3445556677).

This has Hilbert numerator Pnum=1t92t10t11t12+t14+t15+2t16+t17t26.

Taking a suitable map of P(a0,,a5) with a0++a5=26 may describe a family of K3 surfaces, since at least the canonical class is right and h1(X,OX)=0 by Theorem 2.1. In this case, maps from either P(1,3,4,5,6,7) or P(2,3,4,5,5,7) work, and these are two families in Altınok’s list [CitationAltınok 05] of 69 codimension three K3 surfaces in Gr(2,5) format.

The weighted projective space P(1,3,4,5,6,7) also admits a map to a different Gr(2,5) format with grading χ=(1,3,4,5,4,5,6,7,8,9)Λ, with Pnum=1t8t9t10t12t13+t13+t14+t16+t17+t18t26, which realizes another family of K3 surfaces from [CitationAltınok 05].

These examples are not complete intersections in a weighted Grassmannian (V˜//χC*)H1H4, for quasilinear hypersurfaces Hi, since there are no variables of weights one or two in χ. To interpret these regular pullbacks as intersection, one can take a cone on the weighted Grassmannian, introducing additional variables of weights one and two, as in [CitationCorti and Reid, CitationQureshi and Szendrői 12]. More general complete intersections inside weighted homogeneous spaces are also common. The way we define “format,” taking hypersurface slices of one format describe a new format, a tensor-like combination of the existing format and a complete intersection; see Section 5.1.

Example 3.6.

There is no reason why format variables should be weighted positively. The role of the key variety is as a target for regular pullbacks, and these are defined on the affine cone, so there is no risk of taking Proj of a ring with nonpositive weights.

For example, consider the same key variety CGr(2,5)C10 as above, but with key weights χ=(0111111222).

A regular pullback to a nonsingular curve in P4 defines a curve of genus five in its canonical embedding. If φ*(x1)=0, then the curve is trigonal and lies on the scroll given by the minors of the upper 2 × 3 block of the matrix. Deforming φ*(x1)=λ away from zero moves the regular pullback off the trigonal locus to give a non-special canonical curve, a (2, 2, 2) complete intersection in P4. This example can be extended to P5, where the special pullback is the trigonal K3 surface extending this canonical curve.

In this format, the pullback by φ of the 5 × 5 matrix is the matrix of first syzygies among the equations, so this matrix must not have non-zero constant entries, otherwise, as in the example, the free resolution is not minimal and we fall into a different format. Such entries only happen when the key weight is zero, and in that case we only remain in the format if the corresponding pullback is the zero polynomial, giving a special element of the family.

As another example, the weights χ=(1111111333) admit a regular pullback to a canonical surface in P5, with pg = 6, K2=11, where necessarily φ*(x1)=0; as a sanity check, with these invariants Riemann–Roch gives PX(t)=13t2+2t32t4+3t5t7(1t)6.

For a general regular pullback, this is just a degree (3, 4) complete intersection in P1×P2 in the mild disguise of its Segre embedding, so is well known, but there are other cases that cannot be expressed in such straightforward terms. See Ashikaga–Konno [CitationAshikaga and Konno 90] for a complete analysis of this case; the description here appears as [CitationAshikaga and Konno 90] Theorem 1.5(4), with a=b=c=1, and the evident pencil of curves of genus 3 in the description here is typical.

It is easy to see that one cannot allow two key weights 0 that are pulled back to the zero polynomial. Below we note that even a single one cannot work for the kind of threefold we seek. For example, attempting to make a quasismooth Calabi–Yau threefold with key weights χ=(0222222444) and a regular pullback to P(1,1,1,2,2,2,3), we find no problem when φ*(x1)=0 except that X is then a complete intersection rather than in this Grassmannian format, but when φ*(x1)=0 the regular pullback is not quasismooth at the index three point.

We seek threefolds, and in this format negative key weights do not arise:

Proposition 3.7.

Let X be a variety in CGr(2,5) format with ambient weights χ. If X is of dimension 3 and quasismooth, then χ consists of strictly positive integers.

Proof.

If not, then without loss of generality φ*(x1)=0 and any point of X in the locus (φ*(x2)==φ*(x7)=0)X is a non-quasismooth point (the Jacobian has at most 2 non-zero rows here). This locus is necessarily non-empty if dimX3.

This Proposition does not rule out isolated singular points. For example, there could be a canonical threefold with non-quasismooth terminal singularities (these have embedding dimension one, by Mori [CitationMori 85] and Reid [CitationReid 83], which can achieved locally) but we do not construct one.

3.2. The Hilbert series of a canonical threefold

Let P=1r(r1,a,ra) be a terminal quotient singularity with r > 1 and 1a<r coprime integers. (The first weight is r – 1 since we consider varieties polarized by their canonical class.) Following [CitationBuckley et al. 13], we define A=1tr1t=1+t+t2++tr1 and B=bL1tb1t, and let C=C(t) be the Gorenstein symmetric polynomial with integral coefficients such that BC1(modA) whose exponents lie in the integer range {c/2+1,,c/2+r1} (we abbreviate this to ‘C is supported on [α,β]’ for appropriate integers α, β). In our case X is a threefold with terminal singularities polarized by KX, hence c = 5.

Theorem 3.8

([CitationBuckley et al. 13, Theorem 1.3]). Let X be a canonical threefold with singularity basket B. For a terminal quotient singularity Q=1r(r1,a,ra), define Porb(Q)=B(t)(1t)3(1tr), where B=B(t) is a polynomial supported on [3,r+1] which satisfies B×b[r1,a,ra]1tb1t1mod1tr1t.

Then the Hilbert series of X polarized by KX is PX=Pini+QBPorb(Q), where Pini=1+at+bt2+bt3+at4+t5(1t)4 for integers a:=P14 and b:=P24P1+6.

The relationship between a, b and plurigenera P1, P2 is determined by the expansion P=1+P1t+P2t2+=1+(a+4)t+(b+4a+10)t2+, since each series Porb(t) has no quadratic terms or lower.

Example 3.9.

Suppose that p=12(1,1,1). We have A=1+t and B = 1, so the inverse of B is 1 modulo A. The numerator of Porb(p) is supported in the range [3,3]. Observe that t31(mod(A)), so Porb(p)=t3(1t)3(1t2).

Expanded formally as a power series, Porb(p)=t33t47t510t6.

Example 3.10.

Suppose now that p=18(3,5,7). Observing that B=(1+t++t6)(1+t+t2)(1+t+t2+t3+t4)t7(t3t4t5t6t7)(1+t+t2+t3+t4)t2(1+t+t2+t3+t4)2, where the equivalence is taken modulo A=1+t++t7, it is clear that t3(1+t5+t10)(t5+t10+t15)Bt5(1+t++t14)(t5+t6++t19)t5·t15·t5·t151.

So we have an inverse for B. To shift this inverse into the desired range of exponents (and hence find C), we use the fact that t81(mod(A)): t3(1+t5+t2)(t5+t2+t7)t3(t5+t2+t7+t2+t7+t4+t7+t4+t)t3(32tt23t3t42t53t6).

Thus Porb(p)=3t32t4t53t6t72t83t9(1t)3(1t8).

Until the final step all the polynomials appearing had non-negative coefficients. Since the last subtraction was required only to eliminate the out-of-range t7 monomial, and since this monomial had the largest coefficient, we see that every coefficient of the numerator of Porb(p) is strictly negative. This is the case in general for canonically polarized terminal quotient singularities.

Theorem 3.11.

Let X be a canonically-polarized threefold, and pX be a terminal quotient singularity 1r(1,a,a) for coprime integers r > 1 and 1a<r. Define mZ by the conditions 0<mr/2 and am1(mod(r)). Then C(t)=c3t3++cr+1tr+1, where ci+3={iamif 0<iam,miaif m<ia2m1,motherwise.

Here 0<iar satisfies iaim(mod(r)). More concisely, ci+3=min{m,|mia|}.

Notice that it might be necessary to switch the roles of a and – a in order for such an m to exist – this is implicit in the statement of the theorem. For example, when considering Example 3.10, we are forced to take a = 5.

Theorem 3.11 computes Porb for singularities of the form Q=1r(1,a,a). Multiplying by the natural denominator, the leading terms are (1t)3(1tr)Porb(Q)=mt3min{m,r2m}t4, where m=1/a(mod(r)), as in the theorem.

Corollary 3.12.

Let Porb(p)=a0+a1t+a2t2+Z[[t]] for some terminal quotient singularity pX. Then a0=a1=a2=0 and ai<0 for all i3. In particular there exists a bound on the number of singularities of X in terms of pg and P2.

Proof of Theorem 3.11.

With notation as above, observe that B=(1+t++tr2)(1++ta1)(1++tra1)tr1(1++ta1)(tra++tr1)(mod(A))=t2ra1(1+t++ta1)2.

With m as defined in the theorem, t(1+ta+t2a++t(m1)a)(1+t+t2++ta1)=t+t2++tma, which is congruent to –1 modulo A. Hence Cta+1·t2(1+ta++t(m1)a)2(mod(A))=t3(1+ta+t2a++t(m1)a)(ta+t2a++tma).

We consider the product of factors C1=(1+ta+t2a++t(m1)a)(ta+t2a++tma).

Recall that the numerator C of Porb(p) is supported in [3,r+1]; we compute this by finding the integral polynomial equivalent to C1 modulo A supported in [0,r2].

The terms of C1 arise as a product tja with 0jm1 from the first factor and tka with 1km from the second. Hence, the coefficient of tia in the resulting expansion is given by {i,if 0<im;2mi,if m<i2m1.

Since a is coprime to r, the resulting monomials are equivalent modulo 1tr (and hence also modulo A) to distinct powers of t in the range t,,tr1 (recall that by definition 2m1r1). We obtain the equivalent polynomial C1c1t++cr1tr1(mod(A)), where ci={ia,if 0<iam;2mia,if m<ia2m1;0,otherwise.

Subtracting mA from this (to shift the degree down by one) gives the desired result. □

4. Enumeration of Hilbert series and varieties

We aim to construct d-dimensional varieties XP(W), for weights W, in a given format and with canonical class ωX=OX(k) for given k. Moreover we insist that the singularities appearing on X are those of some chosen family. This could be a meaningful complete family—terminal threefold singularities, say—or an arbitrary collection amenable to computation—isolated fourfold terminal quotient singularities, for example. We consider families for which we are able to compute their Porb.

4.1. The general process to find orbifolds

Fix a key variety V˜Cn of codimension c, and fix integers d,kZ with d2 and a class of singularities Q for which Porb(Q) is computable. We aim to construct d-dimensional varieties X in weighted projective space that have KX=OX(k), singularities in the chosen class, and key variety V˜. This pseudo-algorithm is similar in spirit to that of Corti and Reid [CitationCorti and Reid 02] and Qureshi and Szendrői [CitationQureshi and Szendrői 12], but differs in that here we determine the target Hilbert series first and then try to match a basket, rather than choosing a basket and computing the Hilbert series.

  1. Choose a grading χ on V˜. This determines a format F=(V˜,χ,F).

  2. List all possible ambient weights W for which there is a map φ:Cd+c+1Cn that is equivariant of degree zero with respect to the diagonal C* action with weights W in the domain and χ in the codomain; that is, φ is defined by a vector of n polynomials homogeneous with respect to W of weights exactly χ (and not a multiple of χ).

  3. Setting X˜=φ1(V˜), write out the Hilbert series PX(t) of X=X˜//W C*P(W), and determine the initial term Pini(t).

  4. Set R(t)=PX(t)Pini(t). Compute all ways of realizing R(t)=QBPorb(Q) for finite sets B of singularities of the chosen family. If there are no solutions, then a variety cannot be realized admitting only the given class of singularities.

  5. Accept or reject candidate Hilbert series according to whether or not there exists an orbifold in the given format that realizes it.

Apart from the final step (v), this process can be automated on any computer algebra system—it uses only standard tools such as rational functions and power series. Steps (i) and (v) rely on knowledge of the chosen format. The other steps are essentially independent of the format, and we discuss these first.

4.1.1. Step ii: Enumerating the ambient weights

The maximum key weight χmax is part of the format. For orbifolds (or canonical threefold with terminal singularities) no variable can be omitted from the equations, so the largest degree occurring in any ambient weight sequence W cannot exceed χmax. Together with the condition that aWa=kkV˜, this implies that there are only finitely many weight sequences W, and they can easily be computed with standard techniques. (One can immediately reject sequences that will lead to non-well-formed varieties, for example when W has a nontrivial common divisor.)

4.1.2. Step iii: Recovering the Hilbert series PX and Pini

For each choice of χ and of W, we suppose that suitable regular pullback φ exists, and write PX(t) using the formula of Proposition 3.3. As power series expansions, the Porb summands have terms that start in degree d+k+1+1, so that Pini agrees with PX in all degrees up to its center of Gorenstein symmetry. So to compute the numerator of Pini we need only determine whether any equations have low degrees and compensate appropriately in the corresponding coefficients of PX. For canonical threefold, the coefficients of t and t2 are enough.

4.1.3. Step iv: Polytopes and knapsack kernels

Next, we match the possible Porb contributions arising from the candidate singularities σ1,,σm to the Hilbert series, and so build the possible baskets. This is a “knapsack”-style search: summing non-negative multiples of a known collection of vectors to obtain a given solution. The first few terms of each possible Porb contribution, together with the target sequence PXPini, are used to construct a polyhedron in the positive orthant whose integer points (a1,,am)Z0m give solutions to aiPorb(σi)=PXPini. It is an important point that the resulting polyhedron may be infinite: it decomposes into a sum of a compact polytope Q and a (possibly empty) tail cone C. The points in Q correspond to the possible baskets for X, whilst the Hilbert basis of C describes the possible “kernels”; that is, collections of singularities whose net Porb contribution is zero, so can be added to any basket.

4.1.4. Remarks

The process described above does not even in principle give rigorous classification results—the key varieties we use have infinitely many diagonal C* actions. It is worth being clear about where the process is finite and determined, where it is infinite but under control, and where it contains essentially infinite searches.

  1. The ambient weights W are solutions to a “knapsack”-type problem—find a fixed number of strictly positive integers with a given sum. Such problems of course have a finite solution, with well-documented algorithms, if one wants to implement them.

    Our approach has a striking virtue: it is easier to solve for ambient weights W if one imposes additional conditions on the weights than if one does not. For example, to find cases of canonical threefold with empty bi-canonical linear system we can solve for W among integers 3. Such conditions dramatically simplify the problem; compare Section 1.2.

  2. As explained in Section 4.1.3, the list of possible baskets that solve the purely numerical problem of completing Pini to the Hilbert series PX can be infinite. But even then, it is represented by the points of a finitely determined polyhedron, and these points can be enumerated in a systematic order, from “small” baskets to “large” baskets. Any given candidate variety has known ambient weights and equation degrees, and so only finitely many of these baskets could possibly occur.

    The kind of elementary calculation one faces is this: if the ambient stratum that has an index three stabilizer is Γ=P(3,6), and if one of the equations has degree 12, then, unless the format forces this equation to vanish along Γ, there cannot be more than two orbifold points of index three, since this equation restricted to Γ is quadratic.

  3. Although many geometrically important searches will have a finite solution (compare [CitationJohnson and Kollár 01, Theorem 4.1] for quasismooth hypersurfaces), the search routine outlined above does not have a stopping condition, and we cannot know if or when all solutions have been found. This is in the same spirit as Iano-Fletcher’s original enumeration for Fano threefold in codimension two (retrospectively complete by [CitationChen et al. 11]), but differs from Reid’s computation of the 95 Fano hypersurfaces and Johnson–Kollár’s calculation of Fano complete intersections. For many of our searches, we simply continue searching until no new results appear; see the columns klast and kmax of .

  4. The process as stated works in any generality for any key variety. We describe the Gr(2,5) format in detail in Section 4.2, and sketch some other formats in Section 5.1.

  5. We have not used the condition that φ exists except to bound the weights appearing in W, nor have we enforced the condition that φ1(V˜) is Cohen–Macaulay. Both of these are postponed to the final step.

4.2. Canonical threefold in Gr(2,5) format

We make formats with the codimension three key variety V˜=CGr(2,5) of Example 3.5 and its usual Pfaffian free resolution.

4.2.1. Steps i–iv

Iterating over the possible gradings χ is one pass through an infinite loop. By Corti and Reid [CitationCorti and Reid 02], χ is determined by a vector (w1,,w5) with either all wiZ or all wi12+Z: for Plücker coordinates xij with 1i<j5, set degxij=wi+wj, and then χ=(χij). To enumerate all possible w, we may assume w1w5. By Proposition 3.7, when d3 all key variables have positive degrees, so w1+w2>0, and in particular w2>0. The adjunction number of the key variety is kV˜=2wi. A naive search routine now computes all w satisfying these conditions for a given kV˜ (which is finite), and the full search is carried out in increasing adjunction number kV˜=1,2,; this is the only point where the search is not finite.

The weights of the five equations, dj=(wi)w6j, are determined by the format and satisfy d1d5. For Step ii, we choose weights a0a6 of a potential ambient space P(a0,a1,,a6). To find canonical varieties, we choose ai=k1.

If XP(a0,a1,,a6) is a variety in this format, then its Hilbert series is PX(t)=Pnum/Π, where Π:=(1tai) and Pnum:=1td1td5+tkd5++tkd1tk with k=2wi.

It is easy to see that for canonical threefolds there will be no equations of degree two, and so the first two coefficients of the power series expansion PX=1+P1t+P2t2+ are P1 = c1 and P2=c2+12c1(c1+1), where cs is the number of ai equal to s.

4.2.2. Step v: Complete intersections in cones

In practice, it is often convenient to treat candidate varieties as complete intersections inside projective cones, even though the regular pullbacks we use can be more general. If possible we apply Bertini’s theorem. However, when there are many different weights bigger than one, the base loci appearing in successive ample systems tend to be large.

Example 4.1.

Number 4 in : XP(15,22). Let V1P(15,210) be the projective cone over V˜ with vertex P4, which is also the locus of non-quasismooth points. Then XV1 is the complete intersection of eight quadrics. The system of quadrics has empty base locus, and between them they miss the vertex, so X is quasismooth by Bertini’s theorem.

Numbers 1 and 2 in work in the same way: the complete intersection in the end has empty base locus because there are no coprime weights to be eliminated.

Example 4.2.

Number 6 in : XP(14,22,3). Let V1P(14,23,34,4) be the projective cone over V˜ with vertex P1. Consider V2V1, a general complete intersection of three cubics. Between them, these cubics miss V1P(34), since that is codimension one in P(34), and they miss the vertex too. But each cubic does have base locus V1P(23,4), which is codimension one in P(23,4), and is in fact a surface together with residual point. So at this stage, we know that V1P(14,23,3,4) is quasismooth away from that locus. (Eliminating the variables does not cause confusion, since the locus of concern is exactly where they all vanish, and so it does not move away from P(23,4) when we eliminate—that is obvious in this case, since that is the only stratum with any index two stabilizer, but we need to know this in other situations later too.)

Now let V3V2 be the locus of a general quartic. The linear system of quartics has base locus V2P(34), but that is empty. So V3P(14,23,3) is quasismooth away from a curve ΓP(23). Finally, XV3 is the locus of a general quadric. The system of quadrics has empty base locus on V3, so the only question remains about the point(s) where the quadric vanishes on Γ. But it is easy to write equations for a specific X that meets P(23) in a single point that is manifestly quasismooth, and so the general X is quasismooth as claimed.

Numbers 3, 5, and 7–11 in work in the same way: each new hypersurface cuts the existing base locus down, but there is new base locus to consider too.

Example 4.3.

Number 12 in : XP(12,22,32,4). Let V1P(12,22,33,44,5) be the projective cone over V˜ with vertex P1. The final variety X will simply be a 3,4,4,4,5 complete intersection in V1, but Bertini’s theorem is not so easy to apply since most low-degree linear systems have rather large base locus. Nevertheless, with care it can still be made to work.

First consider V2V1, a general complete intersection of three quartics. Between them, these quartics miss V1P(44), since that is codimension one there, and they miss the vertex too. But each quartic does have base locus V1P(33,5), which is a copy of P(32,5) and a residual index three point. (So far similar to the previous example.)

Now let V3V2 be the locus of a general quintic. It meets the previous base locus in V2P(33)—a line and a disjoint point—and it also has base locus of its own, namely (V2P(22,4))(V2P(33,4)).

We leave the first of these for now, but note that the second is a collection of finitely many points, none of which are at the index four point. At this stage, we have V3P(12,22,33,4), with the three groups of loci of concern.

Finally, XV3 is the locus of a general cubic. It misses all isolated base points, other than those lying in P(22,4), and cuts the index three line in a single point; calculation on an example shows this point to be 13(1,2,2) in general.

It remains to consider the locus V3P(22,4), since this is in the base locus of the linear system of cubics. Calculation on an example shows that this is finitely many 12(1,1,1) points, and a standard weighted Hilbert–Burch calculation confirms that there are four such points (necessarily, from the original orbifold Riemann–Roch calculation, if you prefer).

One could continue, but the calculations become rather fiddly, with many distinct base loci to keep track of. We settle, at this stage, for computing sufficiently general examples over the rational numbers and using computer algebra to check that their Jacobian ideals define the empty set. For example, number 18 in , XP(3,42,52,6,7), can be realized by the Pfaffians of the skew 5 × 5 matrix (ytvwvwxt+y2+z2xu+yzx3t2+u2).

4.2.3. Plurigenus invariants

We recall the plurigenus formula:

Theorem 4.4

([CitationReid 80, Theorem 5.5], [CitationFletcher 87, Theorem 2.5(4)]). Let X be a canonical threefold with singularity basket B and χ=χ(OX). Then h0(X,mKX)=(12m)χ+m(m1)(2m1)12K3+pBcm(P) where, for P=1r(1,a,a) and ab1(mod(r)), we have cm(P)=i=1m1ib¯(rib¯)2r.

Iano-Fletcher [CitationFletcher 87] gives four different expressions for the terms in the plurigenus formula. In fact, this formula holds exactly as stated for any projective threefold with canonical singularities. The plurigenus formula goes together with the Barlow–Kawamata formula [CitationKawamata 86] for KX·c2(X): π*KX·c2(Y)=Qr21r24χ(OX),for any resolution π:YX.

Corollary 4.5

(Basic numerology). Set Pm=h0(X,mKX) for mZ. It follows from Kawamata–Viehweg vanishing that Pm=χ(X,mKX), for m2, and from Theorem 2.1 that h1(X,KX)=h2(X,OX)=0 and h2(X,KX)=h1(X,OX)=0, so that P1=χ(X,KX)+1, or equivalently that χ(X,OX)=1P1.

We use the plurigenus formula to calculate KX3 and KX·c2(X) in and .

5. Other formats and varieties

5.1. Other formats

We can consider any affine Gorenstein variety that admits some C* actions to be a Gorenstein format, following Reid [CitationReid 11, 1.5], so there are very many. We describe those that appear in . The point V˜=V(x1==xn=0)Cn is a key variety, and regular pullbacks from formats based on this are complete intersections. Qureshi and Szendrői [CitationQureshi and Szendrői 11, CitationQureshi and Szendrői 12] use quasihomogeneous varieties for Lie groups as formats, extending those of Corti and Reid [CitationCorti and Reid 02]. Other formats that often arise in practice for varieties in codimension four are included in [CitationBrown et al. 12, Section 9] and [CitationBrown et al. 18]; the rolling factors format is described by Stevens [CitationStevens 01], and is used by Bauer et al. [CitationBauer et al. 06] to construct surfaces of general type.

We can take products of formats to make new ones. Given two formats V˜=V(f1,,fs)Cn with key weights χ=(χ1,,χn) and Hilbert numerator N(t), and U˜=V(g1,,gr)Cm with key weights ψ=(ψ1,,ψm) and Hilbert numerator M(t), we can make a format W˜=V(f1,,fs,g1,,gr)Cn+m with key weights (χ1,,χn,ψ1,ψm) and Hilbert numerator N(t)×M(t). (We omit the free resolution information here, since we do not need it for the calculations in .)

For example, the product of Gr(2,5) and a codimension one complete intersection describes (non-quasilinear) hypersurfaces inside weighted Grassmannian pullbacks, which have six equations and 10 first syzygies; in we denote this format by Gr(2,5)H. Non-special canonical curves of genus six are in this format.

5.1.1. Orthogonal Grassmannian in codimension five

We recall the weighted orthogonal Grassmannians of Corti and Reid [CitationCorti and Reid 02], and we list canonical threefold in this format in .

Let w=(w1,,w5) as above (wi all congruent modulo Z and have denominator one or two) and positive uZ. These parameters will determine certain weights. There are 16 indeterminates: x, x1,,x5, and xij for 1i<j5. The 10 equations are xxi=Pfi(M) and M(x1,,x5)t=(0,,0)t, where M is the antisymmetric 5 × 5 matrix with upper-triangular entries xij, and the signed maximal Pfaffians Pf1(M),,Pf5(M) of M are Pfi(M)=(1)i(xjkxlmxjlxkm+xjmxkl), where {i,j,k,l,m}={1,,5} and j<k<l<m.

These equations are homogeneous with respect to the weights wtx=u,wtxi=u+|w|wi,wtxij=wi+wj+u, so the 10 equations, respectively, have weights 2u+|w|wi and 2u+|w|+wi, for i=1,,5.

We may assume that u=wtx is smallest weight in the format and that w is ordered; these are normalizing conditions to prevent duplication of the same format (up to automorphism) for different choices of u and w. We enforce that wi+wj>0 for all i, j; in particular, only w1 may be negative.

The 10 equations define V˜=COGr(5,10), the affine cone over the orthogonal Grassmannian; the weights determine a C* action on V˜. We do not need to know more of the free resolution of the coordinate ring—in the given order, the Jacobian matrix is the matrix of first syzygies—except to note the canonical degree k which is kV˜=4|w|+8u.

The first example in appears as [CitationCorti and Reid 02, Example 5.1]. Arguing with Bertini’s theorem shows that the first five entries of the table really do exist as claimed. The argument becomes more involved, and we have not verified the remaining cases—although they do intersect the orbifold loci correctly—so they should be treated only as plausible candidates.

5.1.2. Comparison with known lists: the famous 95 and all that

We recalculated the known classifications of Fano threefolds that arise in the formats we compute. The classical Fano threefold of can be found in [CitationIskovskikh and Prokhorov 99]. The famous 95 hypersurfaces of [CitationReid 80], the 85 codimension two complete intersections of Iano-Fletcher [CitationIano-Fletcher 00], and Altınok’s 69 codimension three Gr(2,5) cases all appeared early in their respective searches. (If run for K3 surfaces, the trigonal K3 surface of Example 3.6 also appears.) We find the classical X2,2,2P6 in codimension 3, and [CitationChen et al. 11] prove that there are no more Fano complete intersections. Although we do not list them in the table, we also checked Suzuki’s index two Fano threefold: 26 in codimension two and two in codimension three in [CitationBrown and K. Suzuki 07] ( and ).

In higher codimensions, there will be many different formats, and any single format is likely to realize only a few of the possible varieties. In codimension 4, [CitationBrown and Kasprzyk] lists 145 Hilbert series of Fano threefolds, whereas the 6 × 10 codimension 4 format of Section 5.1 realizes only a single family. The remaining 144 do exist, usually as two or more families: see [CitationBrown et al. 12, CitationPapadakis 08]. In codimension 5, again the format we demonstrate realizes a single family, while [Brown and Kasprzyk] lists 164 possible Hilbert series.

Canonical threefolds that arise as complete intersections appear in [CitationIano-Fletcher 00], and those lists are proved complete in [CitationChen et al. 11]; in particular, there are no examples in codimension 6 or higher. The codimension two and three complete intersections we find include some interesting near misses. Seven of the raw results are elliptic fibrations over rational surfaces, so not of general type, and we removed these by hand (see the columns #raw and #results in ). Each one has a hyperquotient singularity of type 14(1,1,2,3;2) that is not terminal—but it takes more than numerical data to see that.

5.1.3. Hypersurfaces

Complete intersections in codimension one illustrate the limitations of this approach. Although we find the famous 95 easily, there are, also famously [CitationKreuzer and Skarke 00], 7555 quasismooth Calabi–Yau hypersurfaces, of which 317 have isolated quotient singularities. In theory, the algorithm will eventually find all of these 317 cases, but in practice our code finds only the first 194 of them before becoming unreasonably slow; we include this case in for completeness, but did not calculate it using this method.

There are other specialized algorithms that handle hypersurfaces more effectively. To find all 7555 independently of [CitationKreuzer and Skarke 00], one can use the well-known “quasismooth hypersurface” algorithm of [CitationJohnson and Kollár 01, CitationReid 80] that we implement in [CitationBrown and Kasprzyk 16]. That algorithm does not require the singularities to be isolated, but analyses all singular loci.

5.2. Other classes of variety

5.2.1. Calabi–Yau threefolds

A Calabi–Yau threefold is a threefold with KX = 0 and h1(X,OX)=h2(X,OX)=0 and canonical singularities. The Calabi–Yau map of Candelas, Lynker, and Schimmrigk [CitationCandelas et al. 90] which lists weighted hypersurfaces has been enormously influential, and, together with its famous extension to toric hypersurfaces by Kreuzer and Skarke [CitationKreuzer and Skarke 00], it continues to motivate the subject. Qureshi and Szendrői [CitationQureshi 15CitationQureshi and Szendrői 12] develop several formats in this context other than the few we describe in Section 3.1, and they find other new projective models of Calabi–Yau threefolds.

We restrict to orbifolds having only isolated orbifold points of the form 1r(a,b,c) with a+b+c0(mod(r)); these are the isolated three-dimensional cyclic quotient singularities that admit crepant resolutions, so each of our examples has a resolution by a Calabi–Yau manifold. Although we apply the same method, in contrast to canonical threefolds, the Riemann–Roch contributions of singularities need not be linearly independent; for example, the pair 13(1,1,1) and 13(2,2,2) make opposite contributions. This rarely causes confusion in the low-codimensional models we describe, but it does mean our purely numerical arguments can at the first sight have infinitely many possible baskets of singularities to report.

Another contrast with canonical threefolds is that lists of Calabi–Yau threefolds tend to be large. We certainly do not find all possible Calabi–Yau threefolds in the formats we consider. The rows k = 0 in have klastkmax small, so that examples were still appearing as the calculations became unreasonably slow; no doubt there will be more cases for higher values of k in most formats. Nevertheless, there has been a great deal of work to describe Calabi–Yau threefolds, and our examples extend some known lists already in the literature, such as the nonsingular examples of Tonoli [CitationTonoli 04] and [CitationBertin 09].

Some candidates cannot be realized by an orbifold; these are removed from the raw lists by hand, just as (Equation1–1) above. In most cases, their failure to be quasismooth occurs on the orbifold loci, so is easy to see. However, there are a few that are quasismooth at the orbifold locus but singular at some other point. For example, XP(1,1,2,5,8,13,19) defined with syzygy degrees (126881214131519) must contain the coordinate plane D=P(5,13,19): the first two rows and columns of this matrix necessarily lie in the ideal ID for reasons of degree. Any general such threefold X is still a Calabi–Yau threefold, but is not Q-factorial, and has single node lying on D. In the terminology of [CitationBrown et al. 12], DX is in Jerry12 format, and following the methods there it can be unprojected to give a quasismooth Calabi–Yau threefold YP(1,1,2,5,8,13,19,37), embedded in codimension 4, with a single 137(5,13,19) orbifold point: the birational map XY is the small D-ample resolution of the node followed by the contraction of D to the orbifold point. Unlike cases in [CitationBrown and Georgiadis 17, CitationBrown et al. 12], X cannot be deformed to quasismooth in its Pfaffian format: DX always appears as Jerry12, and Y is only realized as one deformation family. (As mentioned in [CitationBrown et al. 12], Jerry tends to have higher degree than Tom, so having Jerry with just one node makes it hard for Tom.)

5.2.2. Higher index threefold of general type: the case χ = 1

The same methods apply to varieties polarized by a Weil divisor A which satisfies KX = kA for some k > 1. Regular canonical threefold with χ>0, or equivalently h0(X,KX)=0, are fairly rare, but we can search for them directly by using weights W that do not include 1 (or 2,3,).

For example, setting k = 2, so that KX=2A, we find X18,35P(5,6,7,9,11,13) with {P1=P2=0,P3=1B={13(1,1,2),111(5,6,9),113(6,7,11)}KX3=8/429.

An example with KX=3A is given by X60P(4,5,7,11,30) with {P1=P2=0 and S|3KX| is not irreducibleB={12(1,1,1),2×15(1,2,4),17(2,4,5),111(4,7,8)}KX3=27/770, and similarly with KX=4A by X42P(5,6,7,9,11), which manages P2=0 despite having three variables in degree < 8.

Acknowledgments

This work was supported in part by EPSRC grant EP/E000258/1. AK is supported by EPSRC Fellowship EP/N022513/1. GB was supported by the Royal Society International short visit grant V00872850.

References

  • [Altınok 05] S. Altınok. “Constructing New K3 Surfaces.” Turkish J. Math., 29:2 (2005), 175–192.
  • [Altınok 02] S. Altınok, G. Brown, and M. Reid. “Fano 3-folds, K3 Surfaces and Graded Rings.” In Topology and Geometry: Commemorating SISTAG, v. 314 of Contemp. Math., pp. 25–53. Providence, RI: Amer. Math. Soc., 2002.
  • [Ashikaga and Konno 90] T. Ashikaga and K. Konno. “ Algebraic Surfaces of General Type With c12=3pg−7.” Tohoku Math. J. (2), 42:4 (1990), 517–536.
  • [Bauer 01] Ingrid C. Bauer. “Surfaces With K2=7 and pg = 4.” Mem. Amer. Math. Soc., 152 721: (2001), viii+79.
  • [Bauer et al. 06] I. C. Bauer, F. Catanese, and R. Pignatelli. “The Moduli Space of Surfaces With K2=6 and pg=4.” Math. Ann., 336 2: (2006), 421–438
  • [Bauer and Roberto 09] I. C. Bauer and P. Roberto. “Surfaces With K2=8,pg=4 and Canonical Involution.” Osaka J. Math., 46:3 2009, 799–820.
  • [Bertin 09] M.-A. Bertin. “ Examples of Calabi-Yau 3-Folds of P7 With ρ = 1.” Canad. J. Math., 5:61 (2009), 1050–1072.
  • [Birkar et al. 10] C. Birkar, P. Cascini, C. D. Hacon, and J. McKernan. “Existence of Minimal Models for Varieties of Log General Type.” J. Amer. Math. Soc., 23:2 (2010), 405–468.
  • [Bosma et al. 97] W. Bosma, J. Cannon, and C. Playoust. “The Magma Algebra System. I. The User Language.” J. Symbolic Comput. 24:3/4 (1997), 235–265. Computational algebra and number theory (London, 1993).
  • [Brown and Georgiadis 17] G. Brown and K. Georgiadis. “ Polarized Calabi–Yau 3-folds in Codimension 4.” Math. Nach., 290:5–6 (2017), 710–725.
  • [Brown and Kasprzyk 16] G. Brown and A. Kasprzyk. “Four-Dimensional Projective Orbifold Hypersurfaces.” Exp. Math., 25:2 (2016), 176–193.
  • [Brown and Kasprzyk] G. Brown and A. M. Kasprzyk. “The Graded Ring Database. Available at http://www.grdb.co.uk/.
  • [Brown et al. 18] G. Brown, A. M. Kasprzyk, and M. I. Qureshi. “ Fano 3-folds in P2×P2 format, Tom and Jerry.” Eur. J. Math., 4:1 (2018), 51–72.
  • [Brown et al. 12] G. Brown, M. Kerber, and M. Reid. “Fano 3-folds in Codimension 4, Tom and Jerry. Part I.” Compos. Math., 148:4 (2012), 1171–1194.
  • [Brown and K. Suzuki 07] G. Brown and K. Suzuki. “Fano 3-Folds with Divisible Anticanonical Class.” Manuscripta Math., 123:1 (2007), 37–51.
  • [Buckley et al. 13] A. Buckley, M. Reid, and S. Zhou. “Ice cream and orbifold Riemann-Roch.” Izv. Ross. Akad. Nauk Ser. Mat., 77:3 (2013), 29–54, .
  • [Candelas et al. 90] P. Candelas, M. Lynker, and R. Schimmrigk. Calabi-Yau Manifolds in Weighted P\sb4. Nuclear Phys. B, 341:2 (1990), 383–402.
  • [Catanese 79] F. Catanese. “Surfaces With K2=pg=1 and Their Period Mapping.” In Algebraic geometry (Proc. Summer Meeting, Univ. Copenhagen, Copenhagen, 1978), vol. 732, edited by K. Lonsted, Lecture Notes in Math., pp. 1–29. Berlin: Springer, 1979.
  • [Catanese et al. 14] F. Catanese, W. Liu, and R. Pignatelli. “The Moduli Space of Even Surfaces of General Type With K2=8, pg = 4 and q = 0.” J. Math. Pures Appl. (9), 101:6 (2014), 925–948.
  • [Chen et al. 11] J.-J. Chen, J. A. Chen, and M. Chen. “On Quasismooth Weighted Complete Intersections.” J. Algebraic Geom. 20: 2(2011), 239–262.
  • [Corti and Reid 00] A. Corti and M. Reid, editors. Explicit Birational Geometry of 3-folds, vol. 281, London Mathematical Society Lecture Note Series. Cambridge: Cambridge University Press, 2000.
  • [Corti and Reid 02] A. Corti and M. Reid. Weighted Grassmannians. In Algebraic Geometry, pp. 141–163. Berlin: de Gruyter, 2002.
  • [Dicks 88] D. Dicks. Surface with pg = 3 and K2=4 and extension-deformation theory. PhD thesis, University of Warwick, 1988.
  • [Fletcher 87] A. R. Fletcher. “Contributions to Riemann-Roch on Projective 3-folds With Only Canonical Singularities and Applications.” In Algebraic Geometry, Bowdoin, 1985 (Brunswick, Maine, 1985), vol. 46 of Proc. Sympos. Pure Math., edited by S. Bloch and C.H. Clemens, pp. 221–231. Providence, RI: Amer. Math. Soc., 1987.
  • [Goto and Watanabe 78] S. Goto and K. Watanabe. “ On Graded Rings. I.” J. Math. Soc. Japan, 30:2 (1978), 179–213.
  • [Hirzebruch 87] F. Hirzebruch. Some examples of threefolds with trivial canonical bundle. In Gesammelte Abhandlungen/Collected papers. II. 1963–1987, Springer Collected Works in Mathematics, pp. 757–770. Berlin: Springer-Verlag, 1987.
  • [Iano-Fletcher 00] A. R. Iano-Fletcher. “Working With Weighted Complete Intersections.” In Explicit Birational Geometry of 3-folds, vol. 281 London Math. Soc. Lecture Note Ser., pp. 101–173. Cambridge: Cambridge Univ. Press, 2000.
  • [Iskovskikh and Prokhorov 99] V. A. Iskovskikh and Yu. G. Prokhorov. “Fano Varieties.” In Algebraic Geometry, V, vol. 47, Encyclopaedia Math. Sci., pp. 1–247. Berlin: Springer, 1999.
  • [Johnson and Kollár 01] J. M. Johnson and J. Kollár. “Fano Hypersurfaces in Weighted Projective 4-spaces.” Experiment. Math., 10:1 (2001), 151–158.
  • [Kawamata 86] Y. Kawamata. “On the Plurigenera of Minimal Algebraic 3-folds With K≡ 0.” Math. Ann., 275:4 (1986), 539–546.
  • [Kreuzer and Skarke 00] M. Kreuzer and H. Skarke. “ Complete Classification of Reflexive Polyhedra in Four Dimensions.” Adv. Theor. Math. Phys., 4:6 (2000), 1209–1230.
  • [Mori 85] S. Mori. “ On 3-dimensional Terminal Singularities.” Nagoya Math. J., 98:43–66, 1985.
  • [Mori 88] S. Mori. “ Flip Theorem and the Existence of Minimal Models for 3-folds.” J. Amer. Math. Soc., 1:1 (1988), 117–253.
  • [Mukai 95] S. Mukai. “Curves and Symmetric Spaces.” I. Amer. J. Math., 117:6 (1995), 1627–1644.
  • [Neves 03] J. Neves. Halfcanonical Rings on Algebraic Curves and Applications to Surfaces of General Type. PhD thesis, University of Warwick, 2003.
  • [Papadakis 08] S. A. Papadakis . “The Equations of Type II1 Unprojection.” J. Pure Appl. Algebra, 212:10 (2008), 2194–2208.
  • [Persson 87] U. Persson. “An Introduction to the Geography of Surfaces of General Type.” In Algebraic Geometry, Bowdoin, 1985 (Brunswick, Maine, 1985), vol. 46, Proc. Sympos. Pure Math., pp. 195–218. Providence, RI: Amer. Math. Soc., 1987.
  • [Qureshi 15] M. I. Qureshi. “Constructing Projective Varieties in Weighted Flag Varieties II.” Math. Proc. Cambridge Philos. Soc., 158:2 (2015), 193–209.
  • [Qureshi 17] M. I. Qureshi. “ Computing Isolated Orbifolds in Weighted Flag Varieties.” J. Symbolic Comput., 79:part 2 (2017), 457–474.
  • [Qureshi and Szendrői 11] M. I. Qureshi and B. Szendrői. Constructing Projective Varieties in Weighted Flag Varieties. Bull. Lond. Math. Soc., 43:4 (2011), 786–798.
  • [Qureshi and Szendrői 12] M. I. Qureshi and B. Szendrői. “ Calabi-Yau threefolds in Weighted Flag Varieties.” Adv. High Energy Phys., Art. ID 547317, 14, 2012.
  • [Reid 80] M. Reid. “Canonical 3-folds.” In Journées de Géometrie Algébrique d’Angers, Juillet 1979/Algebraic Geometry, Angers, 1979, edited by A. Beauville, pp. 273–310. Alphen aan den Rijn: Sijthoff & Noordhoff, 1980.
  • [Reid 83] M. Reid. “Minimal Models of Canonical 3-folds.” In Algebraic Varieties and Analytic Varieties (Tokyo, 1981), vol. 1, Adv. Stud. Pure Math., pp. 131–180. Amsterdam: North-Holland, 1983.
  • [Reid 89] M. Reid. “Surfaces With pg = 3, K2=4 According to E. Horikawa and D. Dicks.” In Proceedings of Algebraic Geometry Mini-Symposium (Tokyo Univ., Dec 1989), pp. 1–22, December 1989.
  • [Reid 11] M. Reid. Fun in codimension 4. Preprint available online via the author’s webpage, 2011.
  • [Selig 15] M. Selig. On the Hilbert Series of Polarised Orbifolds. PhD thesis, University of Warwick, 2015.
  • [Stevens 01] J. Stevens. “Rolling Factors Deformations and Extensions of Canonical Curves.” Doc. Math. 6 (2001), 185–226 (electronic).
  • [Stevens 03] J. Stevens. Deformations of Singularities, vol. 1811, Lecture Notes in Mathematics. Berlin: Springer-Verlag, 2003.
  • [Tonoli 04] F. Tonoli. “ Construction of Calabi-Yau 3-folds in P6.” J. Algebraic Geom., 13:2 (2004), 209–232.
  • [Zhou 11] S. Zhou. Orbifold Riemann–Roch and Hilbert Series. PhD thesis, University of Warwick, 2011.