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

On the geography of 33-folds via asymptotic behavior of invariants

Yerko Torres-Nova [email protected] Facultad de Matemáticas, Pontificia Universidad Católica de Chile, Campus San Joaquín, Avenida Vicuña Mackenna 4860, Santiago, Chile.
Abstract.

Roughly speaking, the problem of geography asks for the existence of varieties of general type after we fix some invariants. In dimension 11, where we fix the genus, the geography question is trivial, but already in dimension 22 it becomes a hard problem in general. In higher dimensions, this problem is essentially wide open. In this paper, we focus on geography in dimension 33. We generalize the techniques that compare the geography of surfaces with the geography of arrangements of curves via asymptotic constructions. This involves resolutions of singularities and a certain asymptotic behavior of the associated Dedekind sums and continued fractions. We discuss the general situation with emphasis on dimension 33, analyzing the singularities and various resolutions that show up, and proving results about the asymptotic behavior of the invariants we fix.

Keywords: Complex algebraic geometry, Geography of threefolds, Chern numbers.
MSC2020: 14A99, 14C17, 14J30, 14J40.

1. Introduction

We work with normal projective varieties XX over the complex numbers {\mathbb{C}}. As usual, when dimX=1\dim X=1, 22, or d3d\geq 3 we say that XX is a curve, a surface, or a dd-fold respectively. The main purpose of this paper is to study the geography problem for 33-folds of general type. Following the philosophy of Hirzebruch and Sommese [Hir83][Som84] studying the slopes of Chern numbers c12/c2c_{1}^{2}/c_{2} for minimal surfaces of general type, in [Hun89] the author proposed the study of slopes

[c13,c1c2,c3]2[c_{1}^{3},c_{1}c_{2},c_{3}]\in{\mathbb{P}}^{2}_{{\mathbb{Q}}}

for non-singular minimal 33-folds of general type. This, allows us to work by charts, in this case, in the charts c130c_{1}^{3}\neq 0 or c1c20c_{1}c_{2}\neq 0.

Since minimal models may admit singularities in higher dimensions, we want to extend our problem, but we do not directly have the notion of Chern numbers. If XX is a non-singular 33-fold, then we have Chern numbers

c13(X)=KX3,c1c2(X)=24χ(𝒪X),c3(X)=e(X).c_{1}^{3}(X)=-K_{X}^{3},\quad c_{1}c_{2}(X)=24\chi({\mathcal{O}}_{X}),\quad c_{3}(X)=e(X).

From the birational geometry of 33-folds, the triple (KX3,χ(𝒪X),e(X))×2(K_{X}^{3},\chi({\mathcal{O}}_{X}),e(X))\in{\mathbb{Q}}\times{\mathbb{Z}}^{2} is invariant between minimal models of the same birational class [Mat10, 12.1.2] [Kol91, Th.3.2.2]. Then we are extending naturally to the singular case by taking slopes [KX3,χ(𝒪X),e(X)]2[-K_{X}^{3},\chi({\mathcal{O}}_{X}),e(X)]\in{\mathbb{P}}^{2}_{{\mathbb{Q}}}. There are several inequalities for minimal 33-folds, most of them for Gorenstein minimal 33-folds, since we have well-behaved dualities on the cohomology groups. For example, in this case χ(ωX)=χ(𝒪X)\chi(\omega_{X})=-\chi({\mathcal{O}}_{X}). The most important for us are the Miyaoka-Yau inequality KX372χ(𝒪X)-K_{X}^{3}\geq 72\chi({\mathcal{O}}_{X}) ([Miy87] [GKPT19]), and the classical general type inequalities KX3,χ(𝒪X)<0K_{X}^{3},\chi({\mathcal{O}}_{X})<0 where the last one is just for the Gorestein case. The search for Noether’s type inequalities is an active research area, see [CH06] [Hu18] [CCJ20].

Previous results in this research line are the following. In [Hun89], the author studied slopes of non-singular 33-folds with ample canonical class in the chart c1c20c_{1}c_{2}\neq 0. In [CL01] is proved that the Chern slopes [c13,c1c2,c3][c_{1}^{3},c_{1}c_{2},c_{3}] of non-singular 33-folds with ample canonical bundle define a bounded region by finding inequalities for the coordinate c3/c1c2c_{3}/c_{1}c_{2}. For a more general result, let N(d)N(d) be the number of partitions of dd. In [DS22], it is proved that Chern slopes [c1d,,cd]N(d)1[c_{1}^{d},\ldots,c_{d}]\in{\mathbb{P}}^{N(d)-1}_{{\mathbb{Q}}} of non-singular dd-folds with ample canonical bundle define a bounded region in the chart c1d0c_{1}^{d}\neq 0. See [TN23, p.14-18] to see an updated map in the chart c1c20c_{1}c_{2}\neq 0 for non-singular 33-folds with ample canonical class. The problem is that we do not know if those regions are optimal as we have for surfaces, i.e., the slopes c12/c2c_{1}^{2}/c_{2} are dense in the interval [1/5,3][1/5,3].

1.1. Results of this work

When working with slopes, it turns out that we can apply the tool of nn-th root coverings to investigate their behavior. The main reason is that, as nn grows to infinity, the slopes erase the effects of the degree nn of the cover. This allows us to control the slopes from the base of the cover, and this is the main technique to get density results for surfaces

Consider the following data. Let ZZ be a non-singular dd-fold, and let D1,,DrD_{1},\ldots,D_{r} be distinct prime divisors in ZZ. Assume that Dred:=D1++DrD_{red}:=D_{1}+\ldots+D_{r} is a simple normal crossing divisor (SNC). Let n>1n>1 be a positive integer, and let 0<νi<n0<\nu_{i}<n be a collection of rr integers coprime to nn. Assume that there exists a line bundle {\mathcal{L}} on ZZ such that

n𝒪Z(i=1rνiDi).{\mathcal{L}}^{\otimes n}\simeq{\mathcal{O}}_{Z}\left(\sum_{i=1}^{r}\nu_{i}D_{i}\right). (1)

Then, there exists an nn-th root cover hn:YnZh_{n}\colon Y_{n}\to Z branched along DredD_{red}, where YnY_{n} is a dd-fold (Section 2.2). These are the nn-th root covers developed by Esnault and Viehweg (Cf. [EV92]). For curves, the prime divisors D1,,DrD_{1},\ldots,D_{r} are distinct points on ZZ. Since points are isolated, we have YnY_{n} as a non-singular curve. By the Riemann-Hurwitz formula, we have

c1(Yn)n=nc1(Z)(n1)rn=c1(Z)r+rn=c¯1(Z,D)+rn.\frac{c_{1}(Y_{n})}{n}=\frac{nc_{1}(Z)-(n-1)r}{n}=c_{1}(Z)-r+\frac{r}{n}=\bar{c}_{1}(Z,D)+\frac{r}{n}.

Here c¯1(Z,D)\bar{c}_{1}(Z,D) is the first logarithmic Chern number of the pair (Z,Dred)(Z,D_{red}) (see Section 2.1). Hence, if we fix the points DredD_{red} and we consider partitions ν1++νr=n\nu_{1}+\ldots+\nu_{r}=n with n0n\gg 0 a prime number, then we asymptotically have c1(Yn)nc¯1(Z,D)c_{1}(Y_{n})\approx n\bar{c}_{1}(Z,D).

Question 1.1.

Does this asymptotic phenomenon happen in higher dimensions?

In Section 3, we prove that this phenomenon occurs in any dimension for logarithmic morphisms of degree nn with branch locus as a disjoint collection of non-singular distinct prime divisors D1,,DrD_{1},\ldots,D_{r}. As an application, for nn-root cover we have YnY_{n} as a non-singular projective variety, thus in Corollary 3.2 we get.

Theorem A.

Assume we have nn-th root covers h:YnZh\colon Y_{n}\to Z branched at D=jνjDjD=\sum_{j}\nu_{j}D_{j}. If DredD_{red} is non-singular, then for each partition i1++im=di_{1}+\ldots+i_{m}=d, the Chern numbers satisfy,

ci1cim(Yn)nc¯i1c¯im(Z,D),\frac{c_{i_{1}}\ldots c_{i_{m}}(Y_{n})}{n}\to\bar{c}_{i_{1}}\ldots\bar{c}_{i_{m}}(Z,D),

as nn\to\infty for prime numbers.

However, this theorem is restrictive for us in terms of geography. It is not easy to get the necessary hypothesis to construct a minimal dd-fold of general type. On the other hand, we do not drop the possibility of having applications in other contexts. Also, this is a cornerstone of our research and opens the following discussion.

For dimZ2\dim Z\geq 2, if the branch divisor Dred=D1++DrD_{red}=D_{1}+\dots+D_{r} has singularities, then YnY_{n} have rational singularities [Vie77]. In order to have well-behaved invariants, we can choose a (partial) resolution of singularities. For dimZ=2\dim Z=2, the asymptoticity of Chern numbers was proved in [Urz09] for random nn-th root covers. Let us explain briefly what random means (see Section 2.5 for more details). First, in dimension two, each singularity of YnY_{n} is a cyclic surface singularity of type 1n(q,1)\frac{1}{n}(q,1) for some 0<q<n0<q<n. Thus, we use the Hirzebruch-Jung algorithm (see Section 2.4.1) to resolve these singularities in a minimal way. For us, there are two important quantities, (1) the length of the resolution, i.e., the number of steps of the algorithm, and (2) the Dedekind sums,

d(q,1,n)=i=1n1((iqn))((in)),d(q,1,n)=\sum_{i=1}^{n-1}\left(\left(\frac{iq}{n}\right)\right)\left(\left(\frac{i}{n}\right)\right),

where (()):((\cdot)):{\mathbb{R}}\to{\mathbb{R}} is the saw-tooth function (see Section 2.4.2). We get a resolution of singularities XnYnX_{n}\to Y_{n}. However, the Chern numbers c12c_{1}^{2} and c2c_{2} depend on the lengths and the Dedekind sums coming from all cyclic singularities resolved. To guarantee asymptoticity, we have to consider asymptotic arrangements. Indeed, for each prime number n17n\geq 17, there exists a set On{1,,n}O_{n}\subset\{1,\ldots,n\} (Section 2.5) such that for each qOnq\in O_{n}, the lengths and Dedekind sums are bounded by cnc\sqrt{n} for a constant c>0c>0. Thus we say that DredD_{red} is an asymptotic arrangement if satisfies :

  • For prime numbers n0n\gg 0, there exists multiplicities 0<νj<n0<\nu_{j}<n such that the singularity over DjDkD_{j}\cap D_{k} of YnY_{n} is of type 1n(qjk,1)\frac{1}{n}(q_{jk},1) with qjkOnq_{jk}\in O_{n}.

  • For each nn there are line bundles Pic(Z){\mathcal{L}}\in\mbox{Pic}(Z) such that

    n𝒪Z(j=1rνjDj).{\mathcal{L}}^{\otimes n}\simeq{\mathcal{O}}_{Z}\left(\sum_{j=1}^{r}\nu_{j}D_{j}\right).

The main set-up is when there exists HPic(Z)H\in\mbox{Pic}(Z) such that DjHD_{j}\simeq H for each component of DredD_{red}. Observe that the condition (1) is satisfied as

D=ν1D1++νrDr(ν1++νr)HnH,D=\nu_{1}D_{1}+\ldots+\nu_{r}D_{r}\sim(\nu_{1}+\ldots+\nu_{r})H\sim nH,

thus we can construct nn-th root covers. In [Urz09] was proved that for a random partition ν1++νr=n\nu_{1}+\ldots+\nu_{r}=n, the probability of DD being an asymptotic arrangement tends to 11 as nn grows. In this way, for n0n\gg 0 we take random asymptotic partitions, and we get a family of random surfaces XnX_{n} with

c12(Xn)nc¯12(Z,D),c2(Xn)nc¯2(Z,D).c_{1}^{2}(X_{n})\approx n\bar{c}_{1}^{2}(Z,D),\quad c_{2}(X_{n})\approx n\bar{c}_{2}(Z,D).

Now we can discard the SNC property for the branch divisor for those one having only normal crossings. We consider a log\log resolution γ:ZZ\gamma\colon Z^{\prime}\to Z such that the reduced divisor defined by γDred\gamma^{*}D_{red} is SNC. If such a divisor turns out to be an asymptotic arrangement, then we have the asymptotic result for Chern numbers as above. For this see [TN23, p. 57]), and see [Urz16] to connect this result with minimal models. A direct application is a relation between Chern slopes of simply connected surfaces of general type and Chern slopes of arrangements of lines [EFU22]. In this way, in higher dimensions, we have several issues with achieving an analog asymptotic result. For example, the singularities of YnY_{n} are not cyclic, and the choice of a right (partial) resolution of YnY_{n} with good behavior as nn grows is a challenging problem.

Question 1.2.

Is there an analog of the asymptotic results in dimension two for dimension three?

For instance, if this question has a positive answer, then we would be able to study the geography of 33-folds using arrangements of planes in 3{\mathbb{P}}^{3}. The first result in this direction is in Section 4.1, where we find that the Chern number c1c2=24χc_{1}c_{2}=24\chi is asymptotic and independent of the chosen resolution.

Theorem B.

Let ZZ be any non-singular projective 33-fold, and let {D1,,Dr}\{D_{1},\ldots,D_{r}\} be an asymptotic arrangement. For prime numbers n0n\gg 0 there are projective non-singular 33-folds covers XnZX_{n}\to Z of degree nn such that

c1c2(Xn)nc¯1c¯2(Z,D),\frac{c_{1}c_{2}(X_{n})}{n}\to\bar{c}_{1}\bar{c}_{2}(Z,D),

as nn\to\infty

However, the canonical volume and the topological characteristic depend on the chosen resolution. In this way, the first issue is that the singularities of YnY_{n} are of order (multiplicity) n2n^{2}, too big. This means that to connect with a non-singular model, at a bad choice of resolution we could have big exceptional data with respect to nn, and so we would lose the asymptotic property. In Section 4.2, we introduce a prototype of first step, i.e., by toric methods we construct a local cyclic resolution In this way, we get singularities of multiplicity lower than nn, and of cyclic quotient type. We are interested in cyclic quotient singularities since they are log\log-terminal and have a well-known algorithm to resolve them: the Fujiki-Oka continuous fraction (Cf. [Ash19]). In Section 4.3 we globalize this local cyclic resolution, and we get a cyclic resolution XnYnX_{n}\to Y_{n} having the desired asymptotic property. The following is a summary of Theorem 4.2 (using Corollary 2.16), Theorem 4.18, and Theorem 4.24.

Theorem C.

Let ZZ be any non-singular 33-fold, and let {D1,,Dr}\{D_{1},\ldots,D_{r}\} be an asymptotic arrangement. For prime numbers n0n\gg 0 there are 33-folds XnZX_{n}\to Z with at most cyclic quotient singularities of order lower than nn such that

KXn3n,24χ(𝒪Xn)n,e(Xn)nc¯13(Z,D),c¯1c¯2(Z,D),c¯3(Z,D),\frac{-K_{X_{n}}^{3}}{n},\frac{24\chi({\mathcal{O}}_{X_{n}})}{n},\frac{e(X_{n})}{n}\approx\bar{c}_{1}^{3}(Z,D),\bar{c}_{1}\bar{c}_{2}(Z,D),\bar{c}_{3}(Z,D),

where c¯i(Z,D)\bar{c}_{i}(Z,D) are the Chern classes of the arrangement.

It is important to note that the above is an embedded {\mathbb{Q}}-resolution in the language of [ABMMOG12], introduced as an efficient resolution without useless data. In Section 5.1, as a by-product of the computations to get C, we construct minimal non-singular 33-folds of general type using as a base 33-folds Z4Z\hookrightarrow{\mathbb{P}}^{4} with 33 hyperplane sections. This allows us to prove.

Corollary 1.3.

For d>5d>5 and prime numbers n0n\gg 0 there are minimal non-singular 33-folds XnX_{n} of general type over ZZ with slopes

c13(Xn)c1c2(Xn)(d2)31(d2)(d1)2,c3(Xn)c1c2(Xn)(d5)(d2+2d+6)(d2)(d1)2,\frac{c_{1}^{3}(X_{n})}{c_{1}c_{2}(X_{n})}\approx\frac{(d-2)^{3}-1}{(d-2)(d-1)^{2}},\quad\frac{c_{3}(X_{n})}{c_{1}c_{2}(X_{n})}\approx\frac{(d-5)(d^{2}+2d+6)}{(d-2)(d-1)^{2}},

as nn\to\infty. In particular, as the degree of ZZ grows, the slopes have limit point (1,1)(1,1).

In Section 5.2, we see our resolution in terms of pairs (Xn,D~red)(Z,Dred)(X_{n},\tilde{D}_{red})\to(Z,D_{red}). As a corollary, we prove that for asymptotic arrangements of hyperplane sections on a minimal 33-fold of general type, the resolution preserves the bigness of the loglog-canonical divisor KXn+D~redK_{X_{n}}+\tilde{D}_{red}.

Corollary 1.4.

Let ZdZ\hookrightarrow{\mathbb{P}}^{d} be a minimal non-singular projective 33-fold of general type, and let {H1,,Hr}\{H_{1},\ldots,H_{r}\} be a collection of hyperplane sections in general position. Then, for prime numbers n0n\gg 0 there are finite morphisms of degree nn (Xn,D~red)(Z,Dred)(X_{n},\tilde{D}_{red})\to(Z,D_{red}) such that:

  1. (1)

    XnX_{n} is of log\log-general type, i.e., KXn+D~redK_{X_{n}}+\tilde{D}_{red} is big and nef,

  2. (2)

    KXn3>0K_{X_{n}}^{3}>0,

  3. (3)

    XnX_{n} has cyclic singularities (log\log-terminal) of order lower than nn, and

  4. (4)

    the slopes (K3/24χ,e/24χ)(-K^{3}/24\chi,e/24\chi) of XnX_{n} are arbitrarily near to (2,1/3)(2,1/3).

We point out that our cyclic resolution preserves the positivity of the canonical volume, i.e., we start with KZ3>0K_{Z}^{3}>0 and we get KXn3>0K_{X_{n}}^{3}>0. So in future work, if we are able to control the MMP of the chosen resolution, then we will have the asymptotic results with KXnK_{X_{n}} nef, so for n0n\gg 0, the varieties XnX_{n} will be minimal of general type. For us, the goal is that the asymptotic behavior of the slopes of XnX_{n} coincides with the slopes of its minimal models. For this, we need a good terminalization of the cyclic singularities obtained. In Section 6 we discuss what means the word good.

Acknowledgments: I am grateful to my Ph.D. thesis advisor Giancarlo Urzúa for his time, guidance, and support throughout this work. The results in this paper are part of my Ph.D. thesis at the Pontificia Universidad Católica de Chile. I would also like to thank Jungkai Alfred Chen for his hospitality during my stay at the National Center for Theoretical Sciences in Taiwan. Special thanks to Pedro Montero and Maximiliano Leyton, for many comments and suggestions to improve this work. I was funded by the Agencia Nacional de Investigación y Desarrollo (ANID) through the Beca Doctorado Nacional 2019.

2. Preliminaries

2.1. Logarithmic properties

For a non-singular projective variety ZZ of dimension dd, let A(Z)=e0Ae(Z)A(Z)=\bigoplus_{e\geq 0}A^{e}(Z) be its Chow ring, where Ae(Z)A^{e}(Z) are the Chow groups of ee-cycles. We denote its Chern classes by ce(Z)Ae(Z)c_{e}(Z)\in A^{e}(Z), i.e. the Chern classes of its tangent bundle 𝒯Z{\mathcal{T}}_{Z} [Har77, Appendix A]. We have that c1(Z)=KZc_{1}(Z)=-K_{Z}, where KZK_{Z} is the canonical class. We define the Chern numbers of ZZ as the degree of the top-intersection of its Chern classes

ci1(Z)cim(Z)Ad(Z),i1++im=d.c_{i_{1}}(Z)\ldots c_{i_{m}}(Z)\in A^{d}(Z),\quad i_{1}+\ldots+i_{m}=d.

In the following, if the context is understood, we abuse notation using the symbol ci1cim(Z)c_{i_{1}}\ldots c_{i_{m}}(Z) for Chern numbers or more simply the notation ci1cimc_{i_{1}}\ldots c_{i_{m}}.

We have c1d=(1)dKZdc_{1}^{d}=(-1)^{d}K_{Z}^{d}, and since we work over {\mathbb{C}}, it is well-known that cd=e(Z)c_{d}=e(Z), the topological Euler characteristic. These numbers are codified into the Todd class of 𝒯Z{\mathcal{T}}_{Z} by the following formal sum

td(𝒯Z)=1+c12+c12+c212+c1c224c144c12c23c22c1c3+c4720+.\mbox{td}({\mathcal{T}}_{Z})=1+\frac{c_{1}}{2}+\frac{c_{1}^{2}+c_{2}}{12}+\frac{c_{1}c_{2}}{24}-\frac{c_{1}^{4}-4c_{1}^{2}c_{2}-3c_{2}^{2}-c_{1}c_{3}+c_{4}}{720}+\ldots.

As a consequence of the Hirzebruch-Riemann-Roch Theorem, we have the Noether’s identities, i.e., the analytic Euler characteristic are equal to the dd-th summand of td(𝒯Z)\mbox{td}({\mathcal{T}}_{Z}). For example,

χ(𝒪Z)=c1c224,when d=3.\chi({\mathcal{O}}_{Z})=\dfrac{c_{1}c_{2}}{24},\quad\text{when }d=3.
Definition 2.1.

A simple normal crossing (SNC) divisor D=j=1rDjD=\sum_{j=1}^{r}D_{j} is a reduced effective divisor with distinct non-singular components DjD_{j} satisfying the following condition: for each pDp\in D there are local coordinates x1,,xdx_{1},\ldots,x_{d} on ZZ such that the equation defining DD on pp is x1xe=0x_{1}\ldots x_{e}=0, with ede\leq d.

From [Iit77] we introduce the following sheaf on ZZ.

Definition 2.2.

For a SNC divisor DD, the sheaf of log\log-differentials along DD, denoted by ΩZ1(logD)\Omega^{1}_{Z}(\log D), is the 𝒪Z{\mathcal{O}}_{Z}-submodule of ΩZ1𝒪Z(D)\Omega^{1}_{Z}\otimes{\mathcal{O}}_{Z}(D) described as follows. Let pZp\in Z be a point.

  • (i)

    If pDp\not\in D, then (ΩZ1(logD))p=ΩZ,p1(\Omega^{1}_{Z}(\log D))_{p}=\Omega^{1}_{Z,p}.

  • (ii)

    If pDp\in D, we choose local coordinates x1,,xdx_{1},\ldots,x_{d} on ZZ with x1xe=0x_{1}\ldots x_{e}=0 defining DD on pp. Then, (ΩZ1(logD))p(\Omega^{1}_{Z}(\log D))_{p} is generated as 𝒪Z,p{\mathcal{O}}_{Z,p}-module by

    dx1x1,,dxexe,dxe+1,,dxd.\frac{dx_{1}}{x_{1}},\ldots,\frac{dx_{e}}{x_{e}},dx_{e+1},\ldots,dx_{d}.

If D=j=1rνjDjD=\sum_{j=1}^{r}\nu_{j}D_{j} is a divisor on ZZ, whose associated reduced divisor Dred=jDjD_{red}=\sum_{j}D_{j} is a SNC divisor, then for simplicity we set

ΩZ1(logD):=ΩZ1(logDred).\Omega_{Z}^{1}(\log D):=\Omega_{Z}^{1}(\log D_{red}).

In the rest of this section, we assume D=j=1rνjDjD=\sum_{j=1}^{r}\nu_{j}D_{j} as a divisor with DredD_{red} a SNC divisor.

Definition 2.3.

The log\log-Chern classes of a pair (Z,D)(Z,D) are defined as

c¯i(Z,D)=ci(ΩZ(logD)).\bar{c}_{i}(Z,D)=c_{i}(\Omega_{Z}(\log D)^{\vee}).

The log\log-Chern numbers of a pair (Z,D)(Z,D) are defined as the degree of top-dimensional intersections

c¯i1c¯im:=c¯i1(Z,D)c¯im(Z,D),i1++im=d.\bar{c}_{i_{1}}\ldots\bar{c}_{i_{m}}:={\bar{c}}_{i_{1}}(Z,D)\ldots{\bar{c}}_{i_{m}}(Z,D),\quad i_{1}+\ldots+i_{m}=d.

We set ΩZe(logD):=eΩZ1(logD)\Omega^{e}_{Z}(\log D):=\bigwedge^{e}\Omega_{Z}^{1}(\log D) for any 1ed1\leq e\leq d. In this way, ΩZd(logD)=𝒪Z(KZ+Dred)\Omega^{d}_{Z}(\log D)={\mathcal{O}}_{Z}(K_{Z}+D_{red}), i.e., c¯1(Z,D)=c1(Z)Dred{\bar{c}}_{1}(Z,D)=c_{1}(Z)-D_{red}, and it is known that,

c¯d=e(Z)e(Dred),{\bar{c}}_{d}=e(Z)-e(D_{red}),

see [Iit78, Prop. 2].

Lemma 2.4.

We have a natural exact sequence

0ΩZ1ΩZ1(logD)i=1r𝒪Di0,0\to\Omega^{1}_{Z}\to\Omega^{1}_{Z}(\log D)\to\bigoplus_{i=1}^{r}{\mathcal{O}}_{D_{i}}\to 0,

which is known as the residual exact sequence.

Proof.

See [EV92, Proposition 2.3]. ∎

From the residual exact sequence, we can compute the Chern polynomial through the identity

ct(ΩZ1(logD))=ct(ΩZ1)j=1r(e=0dDjete).c_{t}(\Omega_{Z}^{1}(\log D))=c_{t}(\Omega_{Z}^{1})\prod_{j=1}^{r}\left(\sum_{e=0}^{d}D_{j}^{e}t^{e}\right). (2)

Let 0ed0\leq e\leq d, and let i1++im=ei_{1}+\ldots+i_{m}=e be any partition of positive integers. By convention, for the case e=0e=0 we assume the existence of a unique partition, i.e., i1=0i_{1}=0. We introduce the following notation,

D[i1,,im]:=j1<<jmDj1i1Djmim,D[0]=1.D^{[i_{1},...,i_{m}]}:=\sum_{j_{1}<\ldots<j_{m}}D_{j_{1}}^{i_{1}}\ldots D_{j_{m}}^{i_{m}},\quad D^{[0]}=1.

Examples of this notation are,

D[e]=j=1rDje,D[1,1]=j<kDjDk,D^{[e]}=\sum_{j=1}^{r}D_{j}^{e},\quad D^{[1,1]}=\sum_{j<k}D_{j}D_{k},

and in general,

j=1r(e=0dDe)=e=0d(i1++im=eD[i1,,im]).\prod_{j=1}^{r}\left(\sum_{e=0}^{d}D^{e}\right)=\sum_{e=0}^{d}\left(\sum_{i_{1}+\ldots+i_{m}=e}D^{[i_{1},\ldots,i_{m}]}\right). (3)
Corollary 2.5.

We have the identity,

c¯d(Z,D)=cd(Z)+e=1d(1)ecde(Z)(i1++im=eD[i1,,im]).\bar{c}_{d}(Z,D)=c_{d}(Z)+\sum_{e=1}^{d}(-1)^{e}c_{d-e}(Z)\left(\sum_{i_{1}+\ldots+i_{m}=e}D^{[i_{1},\ldots,i_{m}]}\right).
Proof.

Using identity (2), and the previous identity (3) we can compute c¯d\bar{c}_{d} as the degree dd element of the expression

(e=0dce(Z))(e=0d(i1++im=eD[i1,,im])).\left(\sum_{e=0}^{d}c_{e}(Z)\right)\left(\sum_{e=0}^{d}\left(\sum_{i_{1}+\ldots+i_{m}=e}D^{[i_{1},\ldots,i_{m}]}\right)\right).

Corollary 2.6.

Assume DD is non-singular, i.e., its non-singular components are pairwise disjoint. Then for each 1ed1\leq e\leq d we have c¯e(Z,D)=ce(Z)+Re(D)\bar{c}_{e}(Z,D)=c_{e}(Z)+R_{e}(D) where

Re(D)=k+l=eke(1)lD[l]ck(Z),R_{e}(D)=\sum_{\begin{subarray}{c}k+l=e\\ k\neq e\end{subarray}}(-1)^{l}D^{[l]}c_{k}(Z),

for each e=1,,de=1,\ldots,d.

Proof.

Since DD is non-singular, then DiDj=0D_{i}D_{j}=0 for all iji\neq j. Thus, from the identity (2) , we get

e=0d(1)ec¯e(Z,D)te=(e=0d(1)ece(Z)te)(e=0dD[e]te).\sum_{e=0}^{d}(-1)^{e}\bar{c}_{e}(Z,D)t^{e}=\left(\sum_{e=0}^{d}(-1)^{e}c_{e}(Z)t^{e}\right)\left(\sum_{e=0}^{d}D^{[e]}t^{e}\right).

Using the Cauchy product formula for polynomials we have

(1)ec¯e(Z,D)=k+l=e(1)kck(Z)D[l],(-1)^{e}\bar{c}_{e}(Z,D)=\sum_{k+l=e}(-1)^{k}c_{k}(Z)D^{[l]},

and from this, the formula follows. ∎

Corollary 2.7.

Consider a non-singular 33-fold ZZ, and D=jνjDjD=\sum_{j}\nu_{j}D_{j} on ZZ. We have

c¯2(Z,D)=c2(Z)Dred(c1Dred)D[1,1].{\bar{c}}_{2}(Z,D)=c_{2}(Z)-D_{red}(c_{1}-D_{red})-D^{[1,1]}.

Thus, the logarithmic Chern numbers for 33-folds are,

c¯13=c133c12Dred+3c1([D[2]+2D[1,1])+D[3]+3(D[1,2]+D[2,1])+6D[1,1,1]\bar{c}_{1}^{3}=c_{1}^{3}-3c_{1}^{2}D_{red}+3c_{1}([D^{[2]}+2D^{[1,1]})+D^{[3]}+3(D^{[1,2]}+D^{[2,1]})+6D^{[1,1,1]}
c¯1c¯2=c1c2Dred(c12+c2)+c1(2D[2]+3D[1,1])Dred(D[2]+D[1,1]).\bar{c}_{1}\bar{c}_{2}=c_{1}c_{2}-D_{red}(c_{1}^{2}+c_{2})+c_{1}(2D^{[2]}+3D^{[1,1]})-D_{red}(D^{[2]}+D^{[1,1]}).
c¯3=c3c2Dred+c1(D[2]+D[1,1])(D[3]+D[1,2]+D[2,1]+D[1,1,1]).\bar{c}_{3}=c_{3}-c_{2}D_{red}+c_{1}(D^{[2]}+D^{[1,1]})-\left(D^{[3]}+D^{[1,2]}+D^{[2,1]}+D^{[1,1,1]}\right).
Proof.

Using identity (2) for d=3d=3, and looking for the degree 22 terms we obtain c¯2(Z,D)\bar{c}_{2}(Z,D). The other formulas are direct computations using the above lemmas. ∎

Example 2.8.

For ZZ a non-singular curve, set D=νjP1++νrPrD=\nu_{j}P_{1}+\ldots+\nu_{r}P_{r} where PiZP_{i}\in Z are points. So for the unique log\log-Chern number we have

c¯1=c1r=(2g(X)2+r).\bar{c}_{1}=c_{1}-r=-(2g(X)-2+r).

Let ZZ be a non-singular surface, and D=j=1rνjDjD=\sum_{j=1}^{r}\nu_{j}D_{j} with DjD_{j} non-singular curves. Then

c¯12=c122c1Dred+Dred2,c¯2=c2+t2+2j=1r(g(Dj)2),\bar{c}_{1}^{2}=c_{1}^{2}-2c_{1}D_{red}+D_{red}^{2},\bar{c}_{2}=c_{2}+t_{2}+2\sum_{j=1}^{r}(g(D_{j})-2),

Where t2t_{2} is the number of nodes of DD. See [Urz09, Prop. 3.1]. As in the case of nodes for surfaces, let us denote the number of triple points of DD by t3t_{3}. Then we can rewrite

c¯3=c3c2Dred+c1(D[2]+D[1,1])(D[3]+D[1,2]+D[2,1]+t3).\bar{c}_{3}=c_{3}-c_{2}D_{red}+c_{1}(D^{[2]}+D^{[1,1]})-\left(D^{[3]}+D^{[1,2]}+D^{[2,1]}+t_{3}\right).
Example 2.9.

Let ZmZ\hookrightarrow{\mathbb{P}}^{m} be a non-singular projective 33-fold. Let H1,,HrHH_{1},\ldots,H_{r}\sim H hyperplane sections defining an SNC arrangement. We have

c¯13=c13(Z)r3deg(Z)3rc12H+3c1(Z)H2\bar{c}_{1}^{3}=c_{1}^{3}(Z)-r^{3}\deg(Z)-3rc_{1}^{2}H+3c_{1}(Z)H^{2}
c¯1c¯2=c1c2(Z)rH(c12+c2)(Z)+(2r+3(r2))c1(Z)H2deg(Z)r(r+(r2)),\bar{c}_{1}\bar{c}_{2}=c_{1}c_{2}(Z)-rH(c_{1}^{2}+c_{2})(Z)+\left(2r+3\binom{r}{2}\right)c_{1}(Z)H^{2}-\deg(Z)r\left(r+\binom{r}{2}\right),
c¯3=c3(Z)rHc2(Z)+(r+(r2))c1(Z)H2deg(Z)(r+2(r2)+(r3)).\bar{c}_{3}=c_{3}(Z)-rHc_{2}(Z)+\left(r+\binom{r}{2}\right)c_{1}(Z)H^{2}-\deg(Z)\left(r+2\binom{r}{2}+\binom{r}{3}\right).

Thus as rr grows, we have

limrc¯13c¯1c¯2=limrr3r(r2)=2,andlimrc¯3c¯1c¯2=limr(r3)r(r2)=13.\lim_{r\to\infty}\frac{\bar{c}_{1}^{3}}{\bar{c}_{1}\bar{c}_{2}}=\lim_{r\to\infty}\frac{r^{3}}{r\binom{r}{2}}=2,\ \ \text{and}\ \ \lim_{r\to\infty}\frac{\bar{c}_{3}}{\bar{c}_{1}\bar{c}_{2}}=\lim_{r\to\infty}\frac{\binom{r}{3}}{r\binom{r}{2}}=\frac{1}{3}.

Logarithmic Chern classes are well-behaved under logarithmic morphisms.

Definition 2.10.

Let ZZ be a non-singular projective variety and DD an effective divisor with DredD_{red} as SNC divisor. A surjective morphism h:YZh\colon Y\to Z between non-singular projective varieties is called a log\log-morphism, if Dred=(hD)redD^{\prime}_{red}=(h^{*}D)_{red} is a SNC divisor.

Lemma 2.11.

For any log\log-morphism h:(Y,D)(Z,D)h\colon(Y,D^{\prime})\to(Z,D), we have an injection

hΩZ(logD)ΩY(logD).h^{*}\Omega_{Z}(\log D)\hookrightarrow\Omega_{Y}(\log D^{\prime}).

Moreover, if hh is finite and ramified at DD, then we have isomorphism outside the singularities of DD.

Proof.

See [Vie82, Lemma 1.6]. ∎

2.2. nn-th root covers

In this section, we follow [EV92, Sec. 3]. Consider the following building data (Z,D,n,)(Z,D,n,{\mathcal{L}}) where

  1. (1)

    ZZ is a non-singular projective variety of dimension dd,

  2. (2)

    D=j=1rνjDjD=\sum_{j=1}^{r}\nu_{j}D_{j} is an effective divisor on ZZ, with Dred=j=1rDjD_{red}=\sum_{j=1}^{r}D_{j} a SNC divisor,

  3. (3)

    n2n\geq 2 is a prime number, and

  4. (4)

    {\mathcal{L}} a line bundle on ZZ such that 𝒪Z(D)n{\mathcal{O}}_{Z}(D)\simeq{\mathcal{L}}^{\otimes n}.

With this building data, we construct

f:Yn=SpecZi=0n1iZ,f^{\prime}\colon Y^{\prime}_{n}=\mbox{Spec}_{Z}\bigoplus_{i=0}^{n-1}{\mathcal{L}}^{-i}\to Z,

and the normalization f:YnYnZf\colon Y_{n}\to Y^{\prime}_{n}\to Z will be called the nn-th root covering associated to the building data (Z,D,n,)(Z,D,n,{\mathcal{L}}). Since the morphisms are finite of degree prime, both YnY^{\prime}_{n} and YnY_{n} are projective and irreducible. The morphism ff is branched at DD, and YnY_{n} has its singularities over the intersection of components of DD. Thus, YnY_{n} is non-singular if DjDk=D_{j}\cap D_{k}=\emptyset for all j,kj,k. On the other hand, if Dj1DjeD_{j_{1}}\cap\ldots\cap D_{j_{e}}\neq\emptyset, then the intersection is defined by a local equation

zj1νj1zjeνje=0,z_{j_{1}}^{\nu_{j_{1}}}\ldots z^{\nu_{j_{e}}}_{j_{e}}=0,

where z1,,zdz_{1},\ldots,z_{d} are local parameters for ZZ on pp. Thus the singularity of YnY_{n} over Dj1DjeD_{j_{1}}\cap\ldots\cap D_{j_{e}} is locally analytically isomorphic to the normalization of

Spec([z1,,zd,t]tnzj1νj1zjeνje).\mbox{Spec}\left(\frac{{\mathbb{C}}[z_{1},\ldots,z_{d},t]}{t^{n}-z_{j_{1}}^{\nu_{j_{1}}}\ldots z^{\nu_{j_{e}}}_{j_{e}}}\right).
Definition 2.12.

A partial resolution of singularities of YnY_{n} is a projective, surjective, birational morphism g:XYng:X\to Y_{n} with XX a projective normal variety having at most rational singularities. This last means that for any resolution of singularities g:XXg^{\prime}\colon X^{\prime}\to X we have Rig𝒪X=0R^{i}g^{\prime}_{*}{\mathcal{O}}_{X^{\prime}}=0 for i>0i>0. As usual, we omit the word partial if XX is non-singular.

Since the degree nn of ff is a prime number, we have fDj=nDjf^{*}D_{j}=nD^{\prime}_{j}, where Dj=(fDj)redD_{j}^{\prime}=(f^{*}D_{j})_{red}. Thus, for any partial resolution h:XYnZh\colon X\to Y_{n}\to Z we must have a ramification formula

hDj=nDj+Δj,h^{*}D_{j}=nD^{\prime}_{j}+\Delta_{j},

where Δj\Delta_{j} is a divisor supported in the exceptional divisors of hh over DjD_{j}.

Theorem 2.13.

For the nn-th root covering f:YnZf\colon Y_{n}\to Z we have

  1. (1)

    The morphism ff is flat.

  2. (2)

    The variety YnY_{n} has rational singularities.

  3. (3)

    We have the following decomposition on eigenspaces

    f𝒪Yn=i=0n1𝒪Z(L(i)),L(i)=iL+jiνjnDj.f_{*}{\mathcal{O}}_{Y_{n}}=\bigoplus_{i=0}^{n-1}{\mathcal{O}}_{Z}(-L^{(i)}),\quad L^{(i)}=-iL+\sum_{j}\left\lfloor\frac{i\nu_{j}}{n}\right\rfloor D_{j}.

    Indeed, Yn=Specf𝒪YnY_{n}=\mbox{Spec}f_{*}{\mathcal{O}}_{Y_{n}}. Therefore f:YnZf:Y_{n}\to Z is an affine morphism.

  4. (4)

    For any partial resolution of singularities g:XYng\colon X\to Y_{n}, the composition h=fgh=f\circ g satisfies the following {\mathbb{Q}}-numerical equivalence

    KXh(KZ+n1nDred)+Δ,K_{X}\sim_{{\mathbb{Q}}}h^{*}\left(K_{Z}+\frac{n-1}{n}D_{red}\right)+\Delta,

    where Δ\Delta is a divisor supported on the exceptional divisor of gg.

Proof.

For (1), (2), and (3) see [EV92, Ch. 3]. For (4), we use the fact that codim(Sing(Yn))2\operatorname{codim}(\mbox{Sing}(Y_{n}))\geq 2. In this way, there are non-singular open sets UXVZU\hookrightarrow X\to V\hookrightarrow Z avoiding the singularities of DD. From Lemma 2.11, we get hΩUd(logD)=ΩVd(logD)h^{*}\Omega^{d}_{U}(\log D)=\Omega^{d}_{V}(\log D^{\prime}) where DD^{\prime} is the strict transform of DD. In terms of divisors,

h(KV+Dred)=KU+DredKU+h(Dred)n,h^{*}(K_{V}+D_{red})=K_{U}+D^{\prime}_{red}\sim_{{\mathbb{Q}}}K_{U}+\frac{h^{*}(D_{red})}{n},

from where the result holds locally. After extending globally the exceptional term Δ\Delta appears. ∎

The following corollary will be useful in Section 5. First recall that on a variety XX a curve CC is KXK_{X}-negative, KXK_{X}-positive or KXK_{X}-trivial if its intersection with the canonical divisor is negative, positive, or zero.

Corollary 2.14.

The {\mathbb{Q}}-divisor h(n1ni=1rDi)+Δh^{*}\left(\frac{n-1}{n}\sum_{i=1}^{r}D_{i}\right)+\Delta is an effective {\mathbb{Z}}-divisor. Thus, if KZK_{Z} is nef, then the KXnK_{X_{n}}-negative curves of XnX_{n} are contained in the support of hDh^{*}D.

Proof.

The first assertion is a direct consequence of 4.5. Now assume that KZK_{Z} is nef, and let CC be a curve in XX. If CC is not contained in the support of hDh^{*}D, then

KXnC=(h(KZ+n1ni=1rDi)+Δ).C>0,K_{X_{n}}\,C=\left(h^{*}\left(K_{Z}+\frac{n-1}{n}\sum_{i=1}^{r}D_{i}\right)+\Delta\right).C>0,

since hKZh^{*}K_{Z} is nef (projection formula), and effectiveness of h(n1ni=1rDi)+Δh^{*}\left(\frac{n-1}{n}\sum_{i=1}^{r}D_{i}\right)+\Delta. Thus, if CC is negative must lie in the support of hDh^{*}D.

Lemma 2.15.

Let YY be a normal variety and g:XYg:X\to Y a proper, surjective, birational morphism. Assume that XX has rational singularities. Then g𝒪X=𝒪Yg_{*}{\mathcal{O}}_{X}={\mathcal{O}}_{Y} and Rig𝒪X=0R^{i}g_{*}{\mathcal{O}}_{X}=0 for all i>0i>0 if and only if YY has rational singularities.

Proof.

See [Vie77, Lemma 1].∎

Corollary 2.16.

For any partial resolution of singularities g:XYng:X\to Y_{n}, we have χ(𝒪X)=i=0n1χ(𝒪Z(L(i)))\chi({\mathcal{O}}_{X})=\sum_{i=0}^{n-1}\chi({\mathcal{O}}_{Z}(-L^{(i)})), i.e., the analytic Euler characteristic of XX is independent of the chosen partial resolution.

Proof.

Let g:XXg^{\prime}\colon X^{\prime}\to X be a resolution of singularities. Since YnY_{n} has rational singularities, by Lemma 2.15, we must have h𝒪X=g𝒪X=𝒪Yh_{*}{\mathcal{O}}_{X^{\prime}}=g_{*}{\mathcal{O}}_{X}={\mathcal{O}}_{Y}, and Rih𝒪X=Rig𝒪X=0R^{i}h_{*}{\mathcal{O}}_{X^{\prime}}=R^{i}g_{*}{\mathcal{O}}_{X}=0 for all i>0i>0. Thus, χ(𝒪X)=χ(𝒪X)=χ(𝒪Yn)\chi({\mathcal{O}}_{X^{\prime}})=\chi({\mathcal{O}}_{X})=\chi({\mathcal{O}}_{Y_{n}}). Then, we assume that gg is just a resolution of singularities, so

Hi(X,𝒪X)Hi(Yn,g𝒪X)Hi(Yn,𝒪Yn),i0.H^{i}(X,{\mathcal{O}}_{X})\cong H^{i}(Y_{n},g_{*}{\mathcal{O}}_{X})\cong H^{i}(Y_{n},{\mathcal{O}}_{Y_{n}}),\quad i\geq 0.

Since ff is an affine morphism, also we have Hi(Yn,𝒪Yn)Hi(Z,f𝒪Yn)H^{i}(Y_{n},{\mathcal{O}}_{Y_{n}})\cong H^{i}(Z,f_{*}{\mathcal{O}}_{Y_{n}}) for all i0i\geq 0. Thus, we have χ(𝒪X)=i=0n1χ(𝒪Z(L(i)))\chi({\mathcal{O}}_{X})=\sum_{i=0}^{n-1}\chi({\mathcal{O}}_{Z}(-L^{(i)})). ∎

Finally, we give a state about the connectedness of a partial resolution.

Proposition 2.17.

Any partial resolution g:XYng:X\to Y_{n} is irreducible.

Proof.

Since XX is normal we reduce the proof to show that XX is connected [Sta18, Tag. 0347]. From Corollary 2.16 we know that

h0(𝒪X)=1+i=1n1h0(𝒪Z(L(i))).h^{0}({\mathcal{O}}_{X})=1+\sum_{i=1}^{n-1}h^{0}({\mathcal{O}}_{Z}(-L^{(i)})).

If YY is not connected, then h0(𝒪X)>1h^{0}({\mathcal{O}}_{X})>1, so there exists a i1i\geq 1 such that h0(𝒪Z(L(i)))1h^{0}({\mathcal{O}}_{Z}(-L^{(i)}))\geq 1. We have,

nL(i)j=1r{iνj}nDj.-nL^{(i)}\sim\sum_{j=1}^{r}\{i\nu_{j}\}_{n}D_{j}.

So, we choose curves Γj\Gamma_{j} on ZZ such that DjΓj>0D_{j}\,\Gamma_{j}>0. Thus, we get a system of equations Aν0modnA\nu\equiv 0\mod n where ν=[ν1,,νr]T\nu=[\nu_{1},...,\nu_{r}]^{T} and A=(DjΓk)jkA=(D_{j}\Gamma_{k})_{jk}. Since nn is prime, ν0modn\nu\equiv 0\mod n, and since 0<νj<n0<\nu_{j}<n, we get a contradiction. ∎

2.3. Toric picture

In this section, for toric varieties we mainly follow the notation of [CLS11].

Let n>0n>0 be a prime number and 0ν1,,νd<n0\leq\nu_{1},\ldots,\nu_{d}<n integers. Choose a νk0\nu_{k}\neq 0, and let 0q1,,qd<n0\leq q_{1},\ldots,q_{d}<n be integers such that νj+qjνk0\nu_{j}+q_{j}\nu_{k}\equiv 0 modulo nn. In particular, qk=n1q_{k}=n-1. As usual set N=dN={\mathbb{Z}}^{d} and NdN_{{\mathbb{R}}}\cong{\mathbb{R}}^{d} with canonical basis e1,,ede_{1},\ldots,e_{d}, and M=NdM=N^{\vee}\cong{\mathbb{Z}}^{d} with M=dM_{{\mathbb{R}}}={\mathbb{R}}^{d}. Consider the semigroup

S=e1,,ek1,jkqjej+nek,ek+1,,ed,jkνj+qjνknej+νkek.S=\left\langle e_{1},\ldots,e_{k-1},\sum_{j\neq k}q_{j}e_{j}+ne_{k},e_{k+1},\ldots,e_{d},\sum_{j\neq k}\frac{\nu_{j}+q_{j}\nu_{k}}{n}e_{j}+\nu_{k}e_{k}\right\rangle_{{\mathbb{N}}}.

Since

jkνj+qjνknej+νkek=jkνjnej+νkn(jkqjej+nek),\sum_{j\neq k}\frac{\nu_{j}+q_{j}\nu_{k}}{n}e_{j}+\nu_{k}e_{k}=\sum_{j\neq k}\frac{\nu_{j}}{n}e_{j}+\frac{\nu_{k}}{n}\left(\sum_{j\neq k}q_{j}e_{j}+ne_{k}\right),

we have that the saturation [CLS11, p. 27] of SS is Ssat=σdS^{sat}=\sigma^{\vee}\cap{\mathbb{Z}}^{d} where

σ=C(e1,,ek1,jkqjej+nek,ek+1,,ed1,ed)M\sigma^{\vee}=C\left(e_{1},\ldots,e_{k-1},\sum_{j\neq k}q_{j}e_{j}+ne_{k},e_{k+1},\ldots,e_{d-1},e_{d}\right)\subset M_{{\mathbb{R}}}

is the simplicial dd-cone defined by those elements. It is the dual cone of

σ=C(ne1q1ek,,nek1qk1ek,ek,nek+1qk+1ek,,nedqdek)N.\sigma=C\left(ne_{1}-q_{1}e_{k},\ldots,ne_{k-1}-q_{k-1}e_{k},e_{k},ne_{k+1}-q_{k+1}e_{k},\ldots,ne_{d}-q_{d}e_{k}\right)\subset N_{{\mathbb{R}}}.

Observe that mult(σ)=nd1\mbox{mult}(\sigma)=n^{d-1} and mult(σ)=n\mbox{mult}(\sigma^{\vee})=n. Let PσP_{\sigma} the fundamental parallelepiped of σ\sigma, i.e., the points of σ\sigma with coordinates in [0,1)[0,1) respect its generators. Direct computations show that every element of vPσdv\in P_{\sigma}\cap{\mathbb{Z}}^{d} can written as

v=ijvi(neiqiej)+{ijviqi}nejn,0vi<n.v=\frac{\sum_{i\neq j}v_{i}(ne_{i}-q_{i}e_{j})+\left\{\sum_{i\neq j}v_{i}q_{i}\right\}_{n}e_{j}}{n},\quad 0\leq v_{i}<n. (4)
Proposition 2.18.

The toric variety associated with the semigroup SS is

Spec([S])=Spec([x1,,xd,t](tnx1ν1xdνd)).\mbox{Spec}({\mathbb{C}}[S])=\mbox{Spec}\left(\frac{{\mathbb{C}}[x_{1},\ldots,x_{d},t]}{(t^{n}-x_{1}^{\nu_{1}}\ldots x_{d}^{\nu_{d}})}\right).

Moreover, its normalization correspond with Spec([σd])\mbox{Spec}({\mathbb{C}}[\sigma^{\vee}\cap{\mathbb{Z}}^{d}]).

Proof.

For simplicity, we will prove the result in the case k=dk=d. Since Norm(Spec([S]))=Spec([Ssat])Norm(\mbox{Spec}({\mathbb{C}}[S]))=\mbox{Spec}({\mathbb{C}}[S^{sat}]) we will prove the first. Take the surjective morphism of semigroups ϕ:d+1S\phi\colon{\mathbb{N}}^{d+1}\mapsto S such that

ϕ(ej)=ej,ϕ(ed)=i=1d1qjej+ned,ϕ(ed+1)=j=1d1νj+qjνdnej+νded.\phi(e_{j})=e_{j},\quad\phi(e_{d})=\sum_{i=1}^{d-1}q_{j}e_{j}+ne_{d},\quad\phi(e_{d+1})=\sum_{j=1}^{d-1}\frac{\nu_{j}+q_{j}\nu_{d}}{n}e_{j}+\nu_{d}e_{d}.

It induces a surjective morphism of coordinate rings f:[x1,,xd,t][S]f\colon{\mathbb{C}}[x_{1},\ldots,x_{d},t]\to{\mathbb{C}}[S], and by [CLS11] in Proposition 1.1.9 it is known that

ker(f)=(x1a1xdadtad+1=x1b1xdbdtbd+1:ϕ(a)=ϕ(b),a,bd+1).\mbox{ker}(f)=(x_{1}^{a_{1}}\ldots x_{d}^{a_{d}}t^{a_{d+1}}=x_{1}^{b_{1}}\ldots x_{d}^{b_{d}}t^{b_{d+1}}:\phi(a)=\phi(b),a,b\in{\mathbb{N}}^{d+1}).

If we set xj=ajbjx_{j}=a_{j}-b_{j}, the condition ϕ(a)=ϕ(b)\phi(a)=\phi(b) gives equations

{x1+q1xd+ν1+q1νdnxd+1=0xd1+qd1xd+νd1+qd1νdnxd+1=0nxd+νdxd+1=0\left\{\begin{array}[]{ccc}x_{1}+q_{1}x_{d}+\dfrac{\nu_{1}+q_{1}\nu_{d}}{n}x_{d+1}&=&0\\ \ldots&\ldots&\ldots\\ x_{d-1}+q_{d-1}x_{d}+\dfrac{\nu_{d-1}+q_{d-1}\nu_{d}}{n}x_{d+1}&=&0\\ nx_{d}+\nu_{d}x_{d+1}&=&0\end{array}\right.

We can assume xd+1=ncx_{d+1}=nc with c>0c>0, then xd=νdcx_{d}=-\nu_{d}c, and the equations reduces to

{x1=ν1cxd1=νd1c{bj=aj+νjc,1jdad+1=bd+1+nc\left\{\begin{array}[]{ccc}x_{1}&=&-\nu_{1}c\\ \ldots&\ldots&\ldots\\ x_{d-1}&=&-\nu_{d-1}c\end{array}\right.\Rightarrow\left\{\begin{array}[]{cccc}b_{j}&=&a_{j}+\nu_{j}c,&1\leq j\leq d\\ a_{d+1}&=&b_{d+1}+nc&\end{array}\right.

So ker(f)\mbox{ker}(f) is generated by elements of the form

x1a1xdadtbd+1((x1ν1xdνd)c(tn)c),x_{1}^{a_{1}}\ldots x_{d}^{a_{d}}t^{b_{d+1}}((x_{1}^{\nu_{1}}\ldots x_{d}^{\nu_{d}})^{c}-(t^{n})^{c}),

and the result follows. ∎

Corollary 2.19.

The normalization of the affine varieties tn=x1ν1xdνdt^{n}=x_{1}^{\nu_{1}}\ldots x_{d}^{\nu_{d}} and tn=xjjkxinqjt^{n}=x_{j}\prod_{j\neq k}x_{i}^{n-q_{j}} are isomorphic.

Proof.

Observe that the cones defining both varieties are the same, equal to σ\sigma. ∎

Remark 2.20.

We point out the following. Assume that j=ej=e and νe+1==νd=0\nu_{e+1}=\ldots=\nu_{d}=0. Thus qe+1==qd=0q_{e+1}=\ldots=q_{d}=0, and we have a toric description of the normalization of the varieties tn=x1ν1xeνet^{n}=x_{1}^{\nu_{1}}...x_{e}^{\nu_{e}} embedded in 𝔸d{\mathbb{A}}^{d} for any 1ed1\leq e\leq d.

Remark 2.21.

It is known that the cone σ\sigma^{\vee} defines a toric variety isomorphic to the dd-dimensional quotient cyclic singularity of type

(nq1,,nqk1,1,nqk+1,,nqd)n.\frac{(n-q_{1},\ldots,n-q_{k-1},1,n-q_{k+1},\ldots,n-q_{d})}{n}.

We have Spec(σ3)d/ϕ\mbox{Spec}(\sigma\cap{\mathbb{Z}}^{3})\cong{\mathbb{C}}^{d}/\langle\phi\rangle, where ϕ:dd\phi\colon{\mathbb{C}}^{d}\to{\mathbb{C}}^{d} is defined by

ϕ:(z1,,zd)(ζnq1z1,,ζzj,,ζnqdzd),\phi\colon(z_{1},\ldots,z_{d})\mapsto(\zeta^{n-q_{1}}z_{1},\ldots,\zeta z_{j},\ldots,\zeta^{n-q_{d}}z_{d}),

with ζd=1\zeta^{d}=1 (Cf. [Ash15]). In this way, quotient cyclic singularities are geometrically dual to the singularities of nn-th root covers. We will denote a cyclic singularity of this type by Cq1,,qk^,,qdC_{q_{1},...,\hat{q_{k}},...,q_{d}}. In dimension 22, it occurs the accident that singularities of nn-th root covers are also cyclic quotient singularities.

2.4. Hirzebruch-Jung algorithm and Dedekind sums

2.4.1. Planar cones and Hirzebruch-Jung algorithm

Set NdN\cong{\mathbb{Z}}^{d} and M=NM=N^{\vee}. A planar cone τ\tau in NN_{{\mathbb{R}}} is a cone of dimension 22, i.e., it is generated by two rays defined by primitive generators v0,vs+1Nv_{0},v_{s+1}\in N (ss will have sense soon). Assume that mult(τ)=n\mbox{mult}(\tau)=n. It is known that if n>1n>1, then there exists some vτNv\in\tau\cap N such that v0,vv_{0},v generate τN\tau\cap N, i.e., |det(v0,v)|=1|\det(v_{0},v)|=1. If v=c1v0+c2vs+1v=c_{1}v_{0}+c_{2}v_{s+1}, with ci0c_{i}\in{\mathbb{Q}}_{\geq 0}, then

det(v0,v)=nc2=1c2=1/n.\det(v_{0},v)=nc_{2}=1\Leftrightarrow c_{2}=1/n.

On the other hand, since det(v,vs+1)\det(v,v_{s+1})\in{\mathbb{N}}, we have c11nc_{1}\in\frac{1}{n}{\mathbb{N}}. If we set q=nc1q=nc_{1}, we say that τ\tau is of type (n,q)(n,q) in direction v0v_{0} to vs+1v_{s+1}, or type (n,q)(n,q^{\prime}) in the opposite direction, where qq^{\prime} is the inverse modulo nn of qq.

Assume that n,qn,q are coprime, then consider the Hirzebruch-Jung algorithm of division for n/qn/q, i.e., a pair of sequences

m0=n>m1=q>>ms=1>ms+1=0,m_{0}=n>m_{1}=q>\ldots>m_{s}=1>m_{s+1}=0,
n0=0<n1=1<<ns=q<ns+1=n,n_{0}=0<n_{1}=1<\ldots<n_{s}=q^{\prime}<n_{s+1}=n,

which are related by

mα+1=kαmαmα1m_{\alpha+1}=k_{\alpha}m_{\alpha}-m_{\alpha-1}
nα+1=kαnαnα1,n_{\alpha+1}=k_{\alpha}n_{\alpha}-n_{\alpha-1},

where k1,,ksk_{1},\ldots,k_{s} are integers satisfying kα2k_{\alpha}\geq 2. Usually we denote n/q=[k1,,ks]n/q=[k_{1},\ldots,k_{s}]. These sequences define the Hirzebruch-Jung continuous fraction as

nq=k11k211ks.\frac{n}{q}=k_{1}-\cfrac{1}{k_{2}-\cfrac{1}{\ddots-\frac{1}{k_{s}}}}.
Remark 2.22.

Observe that sequence nαn_{\alpha} is the sequence mαm_{\alpha} for n/qn/q^{\prime}, i.e., if the pair (mα,nα)(m_{\alpha}^{\prime},n_{\alpha}^{\prime}) is the resolution of n/qn/q^{\prime} then (mα,nα)=(ns+1α,ms+1α)(m_{\alpha}^{\prime},n_{\alpha}^{\prime})=(n_{s+1-\alpha},m_{s+1-\alpha}).

Lemma 2.23.

For each α\alpha, we have the following relations

  1. (1)

    mαnα+1mα+1nα=n,m_{\alpha}n_{\alpha+1}-m_{\alpha+1}n_{\alpha}=n,

  2. (2)

    gcd(mα,mα+1)=1,\mbox{gcd}(m_{\alpha},m_{\alpha+1})=1,

  3. (3)

    gcd(mα,nα)=1.\mbox{gcd}(m_{\alpha},n_{\alpha})=1.

The non-singular resolution of the planar cone τ\tau is a refinement by adding the rays defined recursively by

vα=mαvα1+vs+1mα1=mαv0+nαvs+1n,1αs.v_{\alpha}=\frac{m_{\alpha}v_{\alpha-1}+v_{s+1}}{m_{\alpha-1}}=\frac{m_{\alpha}v_{0}+n_{\alpha}v_{s+1}}{n},\quad 1\leq\alpha\leq s.

See Figure 1. Each cone C(vα,vα+1)C(v_{\alpha},v_{\alpha+1}) is non-singular, since

det(vα,vα+1)=1n(mαnα+1mα+1nα)=1.\det(v_{\alpha},v_{\alpha+1})=\frac{1}{n}(m_{\alpha}n_{\alpha+1}-m_{\alpha+1}n_{\alpha})=1.

From Remark 2.22 observe that we have a dual non-singular resolution given by the sequence

vα=mαvα1+v0mα1=mαvs+1+nαv0n,1αs.v_{\alpha}^{\prime}=\frac{m^{\prime}_{\alpha}v_{\alpha-1}^{\prime}+v_{0}}{m^{\prime}_{\alpha-1}}=\frac{m^{\prime}_{\alpha}v_{s+1}+n^{\prime}_{\alpha}v_{0}}{n},\quad 1\leq\alpha\leq s.

from where we have vα=vs+1αv_{\alpha}=v^{\prime}_{s+1-\alpha}.

Refer to caption
Figure 1. Resolved planar cone

Let us change the {\mathbb{Z}}-base of NN such that e1=ve_{1}=v and e2=v0e_{2}=v_{0}. So, we have vs+1=ne1qe2v_{s+1}=ne_{1}-qe_{2}. We have,

τ=C(e1,pe1+ne2)i=3d3wi=C(w1,w2)d2,\tau^{\vee}=C(e_{1},pe_{1}+ne_{2})\oplus\bigoplus_{i=3}^{d-3}{\mathbb{R}}w_{i}=C(w_{1},w_{2})\oplus{\mathbb{R}}^{d-2},

where the wiMw_{i}\in M are such that v0,wi=vs+1,wi=0\langle v_{0},w_{i}\rangle=\langle v_{s+1},w_{i}\rangle=0 and ,w0\langle\cdot,w\rangle\geq 0 on τ\tau for any wC(w1,w2)w\in C(w_{1},w_{2}). Thus, in terms of toric varieties, Xτ=Cq×(×)d2X_{\tau}=C_{q}\times({\mathbb{C}}^{\times})^{d-2}, where Cq=Spec([C(w1,w2)M])C_{q}=\mbox{Spec}({\mathbb{C}}[C(w_{1},w_{2})\cap M]) is the cyclic quotient surface singularity of type 1n(q,1)\frac{1}{n}(q,1) (Remark 2.21). Thus, the constructed resolution is a blow-up h:Bl𝔞(Xτ)Xτh:\mbox{Bl}_{{\mathfrak{a}}}(X_{\tau})\to X_{\tau}, where 𝔞=𝔪[x3±1,,xd±1]{\mathfrak{a}}={\mathfrak{m}}\otimes{\mathbb{C}}[x_{3}^{\pm 1},...,x_{d}^{\pm 1}] and

𝔪=wC(w1,w2)(M0)χm.{\mathfrak{m}}=\bigoplus_{w\in C(w_{1},w_{2})\cap(M\setminus 0)}\chi^{m}.

For details see [CLS11, 11.3.6]. In particular, Bl𝔞(Xτ)=Bl𝔪(Xτ)×(×)d2\mbox{Bl}_{{\mathfrak{a}}}(X_{\tau})=\mbox{Bl}_{{\mathfrak{m}}}(X_{\tau})\times({\mathbb{C}}^{\times})^{d-2}. We can give a explicit description of 𝔪{\mathfrak{m}} noting that the projection Cq𝔸2C_{q}\to{\mathbb{A}}^{2} is given by the surjection [χw1,χw2][Cq]{\mathbb{C}}[\chi^{w_{1}},\chi^{w_{2}}]\to{\mathbb{C}}[C_{q}].

2.4.2. Dedekind sums

Consider the sawtooth function ((x)):((x))\colon{\mathbb{R}}\to{\mathbb{R}} is defined as

((x))={xx1/2x0x.((x))=\left\{\begin{array}[]{cc}x-\lfloor x\rfloor-1/2&x\in{\mathbb{R}}\setminus{\mathbb{Z}}\\ 0&x\in{\mathbb{Z}}\end{array}\right..

Observe that is an odd periodic function of period 11. For n3n\geq 3 prime and a1,,ada_{1},\ldots,a_{d}\in{\mathbb{Z}} define the Dedekind sum of dimension dd by

d(a1,,ad,n)=i=1n1((ia1n))((iadn)).d(a_{1},\ldots,a_{d},n)=\sum_{i=1}^{n-1}\left(\left(\frac{ia_{1}}{n}\right)\right)\cdots\left(\left(\frac{ia_{d}}{n}\right)\right).

By periodicity, we can reduce aia_{i}\in{\mathbb{Z}} to 0ai<n0\leq a_{i}<n, then

d(a1,,ad,n)=d({a1}n,,{ad}n,n).d(a_{1},\ldots,a_{d},n)=d(\{a_{1}\}_{n},\ldots,\{a_{d}\}_{n},n).

where {ai}n\{a_{i}\}_{n} is the residue modulo nn of aia_{i}. Since ((x))((x)) is an odd function, we always have

d(a1,,ad,n)=(1)dd(a1,,ad,n),x,(x,n)=1,d(-a_{1},\ldots,-a_{d},n)=(-1)^{d}d(a_{1},\ldots,a_{d},n),\quad\forall x\in{\mathbb{Z}},\quad(x,n)=1,

i.e., d(a1,,ad,n)=0d(a_{1},...,a_{d},n)=0 for any odd dimension dd. Therefore, the non-trivial Dedekind sums are those of even dimension. We can rewrite this sum as

d(a1,,ad,n)=1ndi=1n1({ia1}nn2)({iad}nn2).d(a_{1},\ldots,a_{d},n)=\frac{1}{n^{d}}\sum_{i=1}^{n-1}\left(\{ia_{1}\}_{n}-\frac{n}{2}\right)\cdots\left(\{ia_{d}\}_{n}-\frac{n}{2}\right).
Lemma 2.24.

We have the following relations,

i=1n1{ia}n{ib}n=n2d(a,b,n)+n2(n1)4,\sum_{i=1}^{n-1}\{ia\}_{n}\{ib\}_{n}=n^{2}d(a,b,n)+\frac{n^{2}(n-1)}{4},
i=1n1{ia}n2{ib}n=i=1n1{ia}n{ib}n2=n3d(a,b,n)+n2(n1)(2n1)12\sum_{i=1}^{n-1}\{ia\}_{n}^{2}\{ib\}_{n}=\sum_{i=1}^{n-1}\{ia\}_{n}\{ib\}_{n}^{2}=n^{3}d(a,b,n)+\frac{n^{2}(n-1)(2n-1)}{12}
i=1n1{ia}n{ib}n{ic}n=n32(d(a,b,n)+d(a,c,n)+d(b,c,n))+n3(n1)8,\sum_{i=1}^{n-1}\{ia\}_{n}\{ib\}_{n}\{ic\}_{n}=\frac{n^{3}}{2}\left(d(a,b,n)+d(a,c,n)+d(b,c,n)\right)+\frac{n^{3}(n-1)}{8},
Proof.

We have an identity

i=1n1{ia}nk=i=1n1ik,\sum_{i=1}^{n-1}\{ia\}_{n}^{k}=\sum_{i=1}^{n-1}i^{k},

for each k0k\geq 0 integer. Then, we use repeatedly this identity in the following expressions. The first formula follows from,

n2d(a,b,n)=i=1n1({ia}nn2)({ib}nn2),n^{2}d(a,b,n)=\sum_{i=1}^{n-1}\left(\{ia\}_{n}-\frac{n}{2}\right)\left(\{ib\}_{n}-\frac{n}{2}\right),

and using that 0=i=1n1({ia}nn2)({ib}nn2)({ic}nn2),0=\sum_{i=1}^{n-1}\left(\{ia\}_{n}-\frac{n}{2}\right)\left(\{ib\}_{n}-\frac{n}{2}\right)\left(\{ic\}_{n}-\frac{n}{2}\right), we get the other two.

A well-known result due to Barkan [Bar77] (see Holzapfel [Hol88]) is the following formula relating the length ss of the continued fraction and Dedekind sums,

Theorem 2.25.

Let nn be a prime number, and qq be an integer such that 0<q<n0<q<n. Let n/q=[k1,,ks]n/q=[k_{1},\ldots,k_{s}]. Then

d(1,q,n)+s=α=1s(kα2)+q+qn.d(1,q,n)+s=\sum_{\alpha=1}^{s}(k_{\alpha}-2)+\frac{q+q^{\prime}}{n}.

Remark 2.26.

In [Zag73] was studied the trigonometric version of the Dedekind sum treated here.

2.5. Asymptotic resolution in dimension 2

Let ZZ be a non-singular projective surface, and let DD be an effective divisor with SNC reduced divisor. Assume the necessary hypothesis to construct the normal nn-th root cover YnZY_{n}\to Z along DD (Section 2.2). We have to choose a resolution of singularities h:XnYnZh\colon X_{n}\to Y_{n}\to Z, and the Chern numbers c12,c2c_{1}^{2},c_{2} of XnX_{n} will depend on this resolution. The singularities of YnY_{n} over each point of DjDkD_{j}\cap D_{k} are analytically isomorphic to the normalization of

Spec([x,y,t]tnxnqjky),\mbox{Spec}\left(\frac{{\mathbb{C}}[x,y,t]}{t^{n}-x^{n-q_{jk}}y}\right),

where νj+qjkνk0\nu_{j}+q_{jk}\nu_{k}\equiv 0 modulo nn. This singularity is a cyclic quotient singularity of type 1n(qjk,1)\frac{1}{n}(q_{jk},1), and the singular point can be resolved by some weighted blow-ups. The exceptional data will be a chain of non-singular rational curves {E1,,Es}\{E_{1},\ldots,E_{s}\} with EjEj+1=1E_{j}E_{j+1}=1 and Ej2=kjE_{j}^{2}=-k_{j}, where the kj2k_{j}\geq 2 are the integers that define the negative regular continued fraction

nqjk=k11k211ks,\frac{n}{q_{jk}}=k_{1}-\cfrac{1}{k_{2}-\cfrac{1}{\ddots-\frac{1}{k_{s}}}},

usually called Hirzebruch-Jung continued fraction. See Section 2.4.1 for explicit computations. The number ss is called the length of the resolution and we denoted it by (qjk,n)\ell(q_{jk},n). In this way, we resolve all singularities of YnY_{n} obtaining a morphism g:XnYng\colon X_{n}\to Y_{n}, with composition h:XnZh\colon X_{n}\to Z.

In dimension 22, for the chosen resolution XnX_{n} Dedekind sums and lengths appear in the following formulas [Urz09],

χ(𝒪Xn)=nχ(𝒪Z)p2112nD[2]p14e(D)+j<kd(1,qjk,n)DjDk\chi({\mathcal{O}}_{X_{n}})=n\chi({\mathcal{O}}_{Z})-\frac{p^{2}-1}{12n}D^{[2]}-\frac{p-1}{4}e(D)+\sum_{j<k}d(1,q_{jk},n)D_{j}D_{k}
c2(Xn)=nc2(Z)(n1)e(D)+j<k(qjk,n)DjDk.c_{2}(X_{n})=nc_{2}(Z)-(n-1)e(D)+\sum_{j<k}\ell(q_{jk},n)D_{j}D_{k}.

Then, we can recover a formula for c12c_{1}^{2} by Noether’s identity. In [Gir03] and [Gir06], Girstmair proved that the lengths and the values of Dedekind sums have a particular asymptotical behavior.

Theorem 2.27 (Girstmair).

For n17n\geq 17 there exists a set On{0,,n}O_{n}\subset\{0,\ldots,n\} such that for any qOnq\in O_{n} we have |d(1,q,n)|3n+5,(q,n)3n+2|d(1,q,n)|\leq 3\sqrt{n}+5,\ell(q,n)\leq 3\sqrt{n}+2. Moreover |{0,,n}On|nlog(4n)|\{0,\ldots,n\}\setminus O_{n}|\leq\sqrt{n}\log(4n).

Remark 2.28.

Observe that from the Barkan-Holzapfel relation (Theorem 2.25), and combining it with the results of Girstmair, we have an asymptotic behavior for the coefficients of the Hirzebruch-Jung continued fraction in the following sense: For a prime number n0n\gg 0, and integers qOnq\in O_{n} with n/q=[k1,,ks]n/q=[k_{1},\ldots,k_{s}], we have

α=1s(kα2)6n+7.\sum_{\alpha=1}^{s}(k_{\alpha}-2)\leq 6\sqrt{n}+7.
Definition 2.29.

A collection of prime divisors {D1,,Dr}\{D_{1},\ldots,D_{r}\} on a non-singular dd-fold ZZ is an asymptotic arrangement if Dred=D1++DrD_{red}=D_{1}+\ldots+D_{r} is SNC, and for prime numbers n0n\gg 0:

  1. (1)

    There are multiplicities 0<νj<n0<\nu_{j}<n, such that for any j<kj<k with DjDkD_{j}\cap D_{k}\neq\emptyset, we have qjkOnq_{jk}\in O_{n}, the unique integer such that νj+qjkνk0\nu_{j}+q_{jk}\nu_{k}\equiv 0 modulo nn.

  2. (2)

    There are line bundles {\mathcal{L}} such that 𝒪Z(j=1rνjDj)n{\mathcal{O}}_{Z}\left(\sum_{j=1}^{r}\nu_{j}D_{j}\right)\simeq{\mathcal{L}}^{\otimes n}.

Example 2.30.

Inside the proof of [Urz09, Th. 6.1], it was proved that for any large prime number nn there exist a partition

ν1++νr=n,\nu_{1}+\ldots+\nu_{r}=n,

with qjkOnq_{jk}\in O_{n} such that νj+qjkνk0\nu_{j}+q_{jk}\nu_{k}\equiv 0 modulo nn. We call it an asymptotic partition of nn. Indeed, the probability of a partition of nn to be asymptotic tends to 11 as nn grows to infinity. Thus, any collection of hyperplanes {H1,,Hr}\{H_{1},\ldots,H_{r}\} on d{\mathbb{P}}^{d} defining a SNC divisor, is itself an asymptotic arrangement with

D=ν1H1++νrHr=(ν1++νr)H=nH,D=\nu_{1}H_{1}+\ldots+\nu_{r}H_{r}=(\nu_{1}+\ldots+\nu_{r})H=nH,

where HH is a general hyperplane section on d{\mathbb{P}}^{d}.

3. Asymptoticity for a non-singular branch locus

3.1. Logarithmic asymptoticity

Let A(Z)=A(Z)A(Z)_{{\mathbb{R}}}=A(Z)\otimes_{{\mathbb{Z}}}{\mathbb{R}} be the extended Chow ring of a fixed non-singular variety ZZ. For each n1n\geq 1 assume the existence of finite log\log-morphisms hn:YnZh_{n}\colon Y_{n}\to Z between non-singular varieties of the same dimension d1d\geq 1 with deg(hn)=n\deg(h_{n})=n (Definition 2.10). We have a morphism of extended Chow rings hn:A(Z)A(Yn)h_{n}^{*}\colon A(Z)_{{\mathbb{R}}}\to A(Y_{n})_{{\mathbb{R}}} for each nn. Since hnh_{n} is flat, we have that hn(Ae(Z))Ae(Yn)h_{n}^{*}(A^{e}(Z))\subset A^{e}(Y_{n}) [Har77, III.9.6], then the same applies for the extended ring.

Theorem 3.1.

For n1n\geq 1 assume the existence of finite log\log-morphisms hn:YnZh_{n}\colon Y_{n}\to Z ramified at a non-singular divisor DD whose reduced form is a SNC divisor. Let D1,,DrD_{1},...,D_{r} be the components of DD, and DjD_{j}^{\prime} the reduced preimage of each DjD_{j}. Assume hnDj=nDjh_{n}^{*}D_{j}=nD^{\prime}_{j}. Then, we have,

limnci1cim(Yn)n=c¯i1c¯im(Z,D).\lim_{n\to\infty}\frac{c_{i_{1}}\ldots c_{i_{m}}(Y_{n})}{n}=\bar{c}_{i_{1}}\ldots\bar{c}_{i_{m}}(Z,D).
Proof.

The proof of the theorem is based on proving the following,

ce(Yn)=hn(c¯e(Z,D))+hn(D[1]c¯e1(Z,D))nhn(Ae(Z)),c_{e}(Y_{n})=h_{n}^{*}(\bar{c}_{e}(Z,D))+\frac{h_{n}^{*}(D^{[1]}\bar{c}_{e-1}(Z,D))}{n}\in h_{n}^{*}(A^{e}(Z)_{{\mathbb{R}}}), (5)

for all e0e\geq 0. Then, for any partition i1++im=di_{1}+...+i_{m}=d, we have

ci1cim(Yn)n=c¯i1c¯im(Z,D)+0e<m(j1,,jm)c¯j1,,je,je+11,,jm1(Z,D)D[me]nme,\frac{c_{i_{1}}...c_{i_{m}}(Y_{n})}{n}=\bar{c}_{i_{1}}...\bar{c}_{i_{m}}(Z,D)+\sum_{\begin{subarray}{c}0\leq e<m\\ (j_{1},...,j_{m})\end{subarray}}\frac{\bar{c}_{j_{1},...,j_{e},j_{e+1}-1,...,j_{m}-1}(Z,D)\,D^{[m-e]}}{n^{m-e}},

where the sum runs over each (j1,,jm)(j_{1},...,j_{m}) a permutation of {i1,,im}\{i_{1},...,i_{m}\}. The result follows directly since the combinatorial quantities obtained do not depend on nn. We proceed by induction on the dimension 0ed0\leq e\leq d. The trivial case for e=0e=0 is c0(Yn)=hc¯0(Z,D)=1,c_{0}(Y_{n})=h^{*}\bar{c}_{0}(Z,D)=1, assuming by convention c¯1(Z,D)=0.\bar{c}_{-1}(Z,D)=0. Since DD is non-singular, by Lemma 2.11 we have c¯e(Yn,D)=hn(c¯e(Z,D))\bar{c}_{e}(Y_{n},D)=h_{n}^{*}(\bar{c}_{e}(Z,D)). Thus, by Corollary 2.6 we get

ce(Yn)\displaystyle c_{e}(Y_{n}) =hn(c¯e(Z,D))Re(D).\displaystyle=h_{n}^{*}(\bar{c}_{e}(Z,D))-R_{e}(D^{\prime}).

Since hnDj=nDjh_{n}^{*}D_{j}=nD^{\prime}_{j}, we have,

ce(Yn)\displaystyle c_{e}(Y_{n}) =hn(c¯e(Z,D))k=0e1(1)ekhn(D[ek])nekck(Yn).\displaystyle=h_{n}^{*}(\bar{c}_{e}(Z,D))-\sum_{k=0}^{e-1}(-1)^{e-k}\frac{h_{n}^{*}(D^{[e-k]})}{n^{e-k}}c_{k}(Y_{n}).

Assuming the induction hypothesis (5) for ke1k\leq e-1 we get,

ce(Yn)\displaystyle c_{e}(Y_{n}) =hn(c¯e(Z,D))k=0e1(1)ekhn(D[ek])nek(h(c¯k(Z,D))+hn(D[1]c¯k1(Z,D))n)\displaystyle=h_{n}^{*}(\bar{c}_{e}(Z,D))-\sum_{k=0}^{e-1}(-1)^{e-k}\frac{h_{n}^{*}(D^{[e-k]})}{n^{e-k}}\left(h^{*}(\bar{c}_{k}(Z,D))+\frac{h_{n}^{*}(D^{[1]}\bar{c}_{k-1}(Z,D))}{n}\right)
=hn(c¯e(Z,D))k=0e1((1)ekhn(D[ek]c¯k(Z,D))nek(1)ek1hn(D[ek1]c¯k1(Z,D))nek1)\displaystyle=h_{n}^{*}(\bar{c}_{e}(Z,D))-\sum_{k=0}^{e-1}\left((-1)^{e-k}\frac{h_{n}^{*}(D^{[e-k]}\bar{c}_{k}(Z,D))}{n^{e-k}}-(-1)^{e-k-1}\frac{h_{n}^{*}(D^{[e-k-1]}\bar{c}_{k-1}(Z,D))}{n^{e-k-1}}\right)
=hn(c¯e(Z,D))+hn(D[1]c¯e1(Z,D))n1,\displaystyle=h_{n}^{*}(\bar{c}_{e}(Z,D))+\frac{h_{n}^{*}(D^{[1]}\bar{c}_{e-1}(Z,D))}{n^{1}},

where the last step is by a telescopic argument using c¯1(Z,D)=0\bar{c}_{-1}(Z,D)=0. ∎

The main situation to apply Theorem 3.1 is the case of nn-th root covers. In this case, take a non-singular SNC divisor D1++DrD_{1}+\ldots+D_{r} on ZZ, and we restrict our attention to prime numbers n2n\geq 2. For each nn, assume the existence of LL and 0<νj<n0<\nu_{j}<n such that D=jνjDjnLD=\sum_{j}\nu_{j}D_{j}\sim nL (Section 2.2). Construct the non-singular covers hn:YnZh_{n}\colon Y_{n}\to Z along each DD. Then hnDj=nDj,h_{n}^{*}D_{j}=nD_{j}^{\prime}, and we get.

Corollary 3.2.

Under the above hypothesis, the nn-th root covers YnY_{n} satisfy,

ci1cim(Yn)nc¯i1c¯im(Z,D),\frac{c_{i_{1}}\ldots c_{i_{m}}(Y_{n})}{n}\to\bar{c}_{i_{1}}\ldots\bar{c}_{i_{m}}(Z,D),

as nn\to\infty for prime numbers n0n\gg 0.

For our purposes in geography, this result has a disadvantage, the difficulty in finding good pairs (Z,D)(Z,D) whose covers YnY_{n} are minimal of general type. For example, from Theorem 2.13 would be enough KZK_{Z} big and nef, and DD ample with many components. However, at least the condition KZK_{Z} big seems difficult to assure since most of varieties with arbitrary collections of disjoint divisors appear to be a fiber space. In [BPS16] was proved the following: Assume that ZZ has a collection of disjoint divisors {Dj}jJ\{D_{j}\}_{j\in J}, if |J|0|J|\gg 0, then there exists a surjective morphism from ZZ to a curve such that every DjD_{j} is contained in a fiber. Thus, generically any variety ZZ of dimZ2\dim Z\geq 2 having collections of disjoint divisors is a fiber space ZVZ\to V over some variety VV.

Remark 3.3.

We can extend Corollary 3.2 to Abelian covers, i.e., to the case Gn=/n1/nkG_{n}={\mathbb{Z}}/n_{1}{\mathbb{Z}}\oplus\ldots\oplus{\mathbb{Z}}/n_{k}{\mathbb{Z}} a sequence of Abelian groups of order n=n1nkn=n_{1}\ldots n_{k} with each njn_{j} a prime number with njnkn_{j}\neq n_{k}. See [Par91] or [Gao11]. In this case, the Abelian covers YnXY_{n}\to X depend on a data DiniLiD^{i}\sim n_{i}L_{i} with DrediD^{i}_{red} a SNC divisor for i=1,,ki=1,\ldots,k. Then the SNC divisor to take is Dred=(D1++Dk)redD_{red}=(D^{1}+\ldots+D_{k})_{red}. The ramification numbers for each component DijD_{i}^{j} of DiD^{i} are given by hDji=ngcd(ni,νji(n))Dji=nDjih^{*}D_{j}^{i}=\frac{n}{\mbox{gcd}(n_{i},\nu_{j}^{i}(n))}{D^{\prime}_{j}}^{i}=n{D^{\prime}_{j}}^{i}, where νji(n)\nu_{j}^{i}(n) is the multiplicity of DjiD_{j}^{i} in DiD^{i}. From here, we leave to the reader the analog asymptotic result. However, we can ask: How can this argument be extended to any Galois cover?

By the above discussion, in the rest of this paper, we will study the above results for the case of 33-folds when DD has its components with non-empty intersections. Thus YnY_{n} will have singularities.

4. Asymptoticity of invariants

4.1. Asymptoticity of χ\chi for 33-folds

Consider a data (Z,D,n,)(Z,D,n,{\mathcal{L}}) as in Section 2.2, with ZZ a non-singular projective 33-fold. Let h:XnYnZh\colon X_{n}\to Y_{n}\to Z be any resolution of singularities of the branched nn-th root cover YnY_{n} along the effective divisor D=j=1rνjDjnLD=\sum_{j=1}^{r}\nu_{j}D_{j}\sim nL whose reduced form is SNC. We have,

L(i)=iLj=1riνjnDj=1n(iDj=1rniνjnDj)=1nj=1r{iνj}nDj.L^{(i)}=iL-\sum_{j=1}^{r}\left\lfloor\frac{i\nu_{j}}{n}\right\rfloor D_{j}=\frac{1}{n}\left(iD-\sum_{j=1}^{r}n\left\lfloor\frac{i\nu_{j}}{n}\right\rfloor D_{j}\right)=\frac{1}{n}\sum_{j=1}^{r}\{i\nu_{j}\}_{n}D_{j}.
Proposition 4.1.

We have,

χ(𝒪Xn)=nχ(𝒪Z)112(R1(n,D)+R2(n,D)+R3(n,D)),\chi({\mathcal{O}}_{X_{n}})=n\chi({\mathcal{O}}_{Z})-\frac{1}{12}(R_{1}(n,D)+R_{2}(n,D)+R_{3}(n,D)),

where

R1(n,D)=(n1)22nD[3]+(n1)(2n1)2n(D[1,2]+D[2,1])+3(n1)2D[1,1,1],R_{1}(n,D)=\frac{(n-1)^{2}}{2n}D^{[3]}+\frac{(n-1)(2n-1)}{2n}(D^{[1,2]}+D^{[2,1]})+\frac{3(n-1)}{2}D^{[1,1,1]},
R2(n,D)=(1n)2c1(Z)((2n1)nD[2]+3D[1,1])+(n1)2Dred(c12(Z)+c2(Z)),R_{2}(n,D)=\frac{(1-n)}{2}c_{1}(Z)\left(\frac{(2n-1)}{n}D^{[2]}+3D^{[1,1]}\right)+\frac{(n-1)}{2}D_{red}(c_{1}^{2}(Z)+c_{2}(Z)),
R3(n,D)\displaystyle R_{3}(n,D) =6(j<kd(νj,νk,n)DjDk(Dj+Dk+KZ)\displaystyle=6\left(\sum_{j<k}d(\nu_{j},\nu_{k},n)D_{j}D_{k}(D_{j}+D_{k}+K_{Z})\right.
+j<k<l(d(νj,νk,n)+d(νj,νl,n)+d(νk,νl,n))DjDkDl).\displaystyle\left.+\sum_{j<k<l}(d(\nu_{j},\nu_{k},n)+d(\nu_{j},\nu_{l},n)+d(\nu_{k},\nu_{l},n))D_{j}D_{k}D_{l}\right).
Proof.

By Corollary 2.16, and Hirzebruch-Riemann-Roch theorem for 33-folds we can compute

χ(𝒪Xn)\displaystyle\chi({\mathcal{O}}_{X_{n}}) =i=0n1χ(𝒪Z(L(i)))\displaystyle=\sum_{i=0}^{n-1}\chi({\mathcal{O}}_{Z}(-L^{(i)}))
=nχ(𝒪Z)112i=1n1(L(i)(L(i)+KZ)(2L(i)+KZ)+c2(Z).L(i))\displaystyle=n\chi({\mathcal{O}}_{Z})-\frac{1}{12}\sum_{i=1}^{n-1}\left(L^{(i)}(L^{(i)}+K_{Z})(2L^{(i)}+K_{Z})+c_{2}(Z).L^{(i)}\right)
=nχ(𝒪Z)112i=1n1(2(L(i))3+3(L(i))2KZ+L(i)KZ2+c2(Z).L(i))\displaystyle=n\chi({\mathcal{O}}_{Z})-\frac{1}{12}\sum_{i=1}^{n-1}\left(2(L^{(i)})^{3}+3(L^{(i)})^{2}K_{Z}+L^{(i)}K_{Z}^{2}+c_{2}(Z).L^{(i)}\right)
=nχ(𝒪Z)112R(n,D),\displaystyle=n\chi({\mathcal{O}}_{Z})-\frac{1}{12}R(n,D),

where R(n,D)R(n,D) is a quantity depending only on nn and DD. We have the following identities: i=1n1L(i)=(n1)2Dred\sum_{i=1}^{n-1}L^{(i)}=\frac{(n-1)}{2}D_{red}, and

i=1n1(L(i))2=(n1)(2n1)6nD[2]+(n1)2D[1,1]+2j<kd(νj,νk,n)DjDk.\sum_{i=1}^{n-1}(L^{(i)})^{2}=\frac{(n-1)(2n-1)}{6n}D^{[2]}+\frac{(n-1)}{2}D^{[1,1]}+2\sum_{j<k}d(\nu_{j},\nu_{k},n)D_{j}D_{k}.

The above is not difficult to deduce from the formulas in Lemma 2.24. To illustrate, we compute i(L(i))3\sum_{i}(L^{(i)})^{3} as follows. The first step, we compute explicitly:

(L(i))3\displaystyle(L^{(i)})^{3} =1n3(iDj=1rniνjnDj)3\displaystyle=\frac{1}{n^{3}}\left(iD-\sum_{j=1}^{r}n\left\lfloor\frac{i\nu_{j}}{n}\right\rfloor D_{j}\right)^{3}
=1n3(j=1r{iνj}nDj)3\displaystyle=\frac{1}{n^{3}}\left(\sum_{j=1}^{r}\{i\nu_{j}\}_{n}D_{j}\right)^{3}
=1n3(j=1r{iνj}n3Dj3+3j<k{iνj}n2{iνk}nDj2Dk+{iνj}n{iνk}n2DjDk2\displaystyle=\frac{1}{n^{3}}\left(\sum_{j=1}^{r}\{i\nu_{j}\}_{n}^{3}D_{j}^{3}+3\sum_{j<k}\{i\nu_{j}\}_{n}^{2}\{i\nu_{k}\}_{n}D_{j}^{2}D_{k}+\{i\nu_{j}\}_{n}\{i\nu_{k}\}_{n}^{2}D_{j}D_{k}^{2}\right.
+6j<k<l{iνj}n{iνk}n{iνl}nDjDkDl)\displaystyle\hskip 34.1433pt+\left.6\sum_{j<k<l}\{i\nu_{j}\}_{n}\{i\nu_{k}\}_{n}\{i\nu_{l}\}_{n}D_{j}D_{k}D_{l}\right)

Applying directly the formulas in Lemma 2.24, we get.

i=1n1(L(i))3\displaystyle\sum_{i=1}^{n-1}(L^{(i)})^{3} =(n1)24nD[3]+(n1)(2n1)4n(D[1,2]+D[2,1])+3(n1)4D[1,1,1]\displaystyle=\frac{(n-1)^{2}}{4n}D^{[3]}+\frac{(n-1)(2n-1)}{4n}(D^{[1,2]}+D^{[2,1]})+\frac{3(n-1)}{4}D^{[1,1,1]}
+3j<kd(νj,νk,n)(Dj2Dk+DjDk2)\displaystyle+3\sum_{j<k}d(\nu_{j},\nu_{k},n)(D_{j}^{2}D_{k}+D_{j}D_{k}^{2})
+3j<k<l(d(νj,νk,n)+d(νj,νl,n)+d(νk,νl,n))DjDkDl.\displaystyle+3\sum_{j<k<l}(d(\nu_{j},\nu_{k},n)+d(\nu_{j},\nu_{l},n)+d(\nu_{k},\nu_{l},n))D_{j}D_{k}D_{l}.

On the other hand,

i=1n1(L(i))2KZ=(1n)2c1(Z)((2n1)3nD[2]+D[1,1])+2j<kd(νj,νk,n)DjDkKZ,\sum_{i=1}^{n-1}(L^{(i)})^{2}K_{Z}=\frac{(1-n)}{2}c_{1}(Z)\left(\frac{(2n-1)}{3n}D^{[2]}+D^{[1,1]}\right)+2\sum_{j<k}d(\nu_{j},\nu_{k},n)D_{j}D_{k}K_{Z},
i=1n1L(i)KZ2=(n1)2KZ2Dred=(n1)2Dredc12(Z),\sum_{i=1}^{n-1}L^{(i)}K_{Z}^{2}=\frac{(n-1)}{2}K_{Z}^{2}D_{red}=\frac{(n-1)}{2}D_{red}c_{1}^{2}(Z),
i=1n1L(i)c2(Z)=(n1)2Dredc2(Z).\sum_{i=1}^{n-1}L^{(i)}c_{2}(Z)=\frac{(n-1)}{2}D_{red}c_{2}(Z).

From here it is not too difficult to note that we have R(n,D)=R1(n,D)+R2(n,D)+R3(n,D)R(n,D)=R_{1}(n,D)+R_{2}(n,D)+R_{3}(n,D), the quantities previously mentioned.

Theorem 4.2.

If {D1,,Dr}\{D_{1},\ldots,D_{r}\} is an asymptotic arrangement, then

χ(𝒪Xn)nc1c2¯(Z,D)24,\frac{\chi({\mathcal{O}}_{X_{n}})}{n}\to\frac{\overline{c_{1}c_{2}}(Z,D)}{24},

as nn\to\infty for prime numbers n0n\gg 0.

Proof.

First observe the following limits

limnR1(n,D)n=12D[3]+(D[1,2]+D[2,1])+32D[1,1,1]=12Dred(D[2]+D[1,1]),\lim_{n\to\infty}\frac{R_{1}(n,D)}{n}=\frac{1}{2}D^{[3]}+(D^{[1,2]}+D^{[2,1]})+\frac{3}{2}D^{[1,1,1]}=\frac{1}{2}D_{red}(D^{[2]}+D^{[1,1]}),
limnR2(n,D)n=12c1(Z)(2D[2]+3D[1,1])+12Dred(c12(Z)+c2(Z)).\lim_{n\to\infty}\frac{R_{2}(n,D)}{n}=-\frac{1}{2}c_{1}(Z)\left(2D^{[2]}+3D^{[1,1]}\right)+\frac{1}{2}D_{red}(c_{1}^{2}(Z)+c_{2}(Z)).

From formulas in Corollary 2.7, we get the identity

limnR1(n,D)+R2(n,D)n=12(c1c2(Z)c1c2¯(Z,D)).\lim_{n\to\infty}\frac{R_{1}(n,D)+R_{2}(n,D)}{n}=\frac{1}{2}(c_{1}c_{2}(Z)-\overline{c_{1}c_{2}}(Z,D)).

Since the collection of divisors is an asymptotic arrangement, we use Theorem 2.27 to get R3(n,D)/n0R_{3}(n,D)/n\approx 0 for prime numbers n0n\gg 0. Thus, in this case we have

χ(𝒪Xn)nc1c2(Z)24124(c1c2(Z)c1c2¯(Z,D))=c1c2¯(Z,D)24,\frac{\chi({\mathcal{O}}_{X_{n}})}{n}\approx\frac{c_{1}c_{2}(Z)}{24}-\frac{1}{24}(c_{1}c_{2}(Z)-\overline{c_{1}c_{2}}(Z,D))=\frac{\overline{c_{1}c_{2}}(Z,D)}{24},

for prime numbers n0n\gg 0.∎

Remark 4.3.

Observe that the formula for (L(i))3(L^{(i)})^{3} can be extended to higher dimensions. In the same way of Lemma 2.24, we can find formulas for (L(i))e(L^{(i)})^{e} depending only on the combinatorial aspects of DD and higher dimensional Dedekind sums. Thus, for asymptoticity of χ(𝒪Xn)\chi({\mathcal{O}}_{X_{n}}) in any dimension, we need asymptoticity of Dedekind sums, i.e., the higher dimensional analogs of Girstmair’s results (Theorem 2.27). For dimension d4d\geq 4 this is an open problem.

4.2. Toric local resolutions

In this section, we study the 33-fold singularity given by the normalization of tn=x1ν1x2ν2x3ν3t^{n}=x_{1}^{\nu_{1}}x_{2}^{\nu_{2}}x_{3}^{\nu_{3}}, with n2n\geq 2 a prime number and 0<νi<n0<\nu_{i}<n. The aim is to achieve good local resolutions of singularities in asymptotic terms with respect to nn. Since this singularity is toric, it can be resolved by subdivisions of its associated cone obtaining a refinement fan. To assure asymptotic properties, we have to pay attention to the combinatorial aspects of the refinement. Let 0<p,q<n0<p,q<n be integers such that ν1+pν30\nu_{1}+p\nu_{3}\equiv 0 and ν1+qν30\nu_{1}+q\nu_{3}\equiv 0 modulo nn. By Section 2.3, this 33-fold singularity is a toric variety Yp,q:=Spec(σ3)Y_{p,q}:=\mbox{Spec}(\sigma^{\vee}\cap{\mathbb{Z}}^{3}) defined by the cone σ=C(d1,d2,d3)3\sigma=C(d_{1},d_{2},d_{3})\subset{\mathbb{R}}^{3}, where d1=ne1pe3,d2=ne2qe3d_{1}=ne_{1}-pe_{3},d_{2}=ne_{2}-qe_{3}, and d3=e3d_{3}=e_{3} are the primitive ray generators. A transversal section of this cone is sketched in Figure 2.

Refer to caption
Figure 2. Transversal section of σ\sigma.

The fan defined by the cone σ\sigma has 2-dimensional faces (walls) given by τjk=C(dj,dk)\tau_{jk}=C(d_{j},d_{k}) for j<kj<k. By Section 2.4.1, each wall τjk\tau_{jk} can be resolved by Hirzebruch-Jung sequences (mjk,α,njk,α)α=0sjk+1(m_{jk,\alpha},n_{jk,\alpha})_{\alpha=0}^{s_{jk}+1} in direction djd_{j} to dkd_{k} with initial data

mjk,0=n,njk,0=0,njk,1=1,j<k,m_{jk,0}=n,\quad n_{jk,0}=0,\quad n_{jk,1}=1,\quad\forall j<k,
m13,1=p,m23,1=q,m12,1={pq}n,m_{13,1}=p^{\prime},\quad m_{23,1}=q^{\prime},\quad m_{12,1}=\{-p^{\prime}q\}_{n},

where pp^{\prime} and qq^{\prime} are the inverse modulo nn of pp and qq. Thus, there are integers kjk,α2k_{jk,\alpha}\geq 2, such that

nmjk,1=[kjk,1,,kjk,sjk].\frac{n}{m_{jk,1}}=[k_{jk,1},...,k_{jk,s_{jk}}].

Then, the walls τjk\tau_{jk} can be resolved by subdividing them in a sequence of steps by rays with generators (ejk,α)α=1sjk(e_{jk,\alpha})_{\alpha=1}^{s_{jk}} defined recursively as

ejk,α=mjk,αdj+njk,αdkn,1αsjk.e_{jk,\alpha}=\frac{m_{jk,\alpha}d_{j}+n_{jk,\alpha}d_{k}}{n},\quad 1\leq\alpha\leq s_{jk}.

If we fix j<kj<k, we denote by ekj,αe_{kj,\alpha} the exceptional divisors in direction dkd_{k} to djd_{j}. We have the relation ejk,α=ekj,s+1αe_{jk,\alpha}=e_{kj,s+1-\alpha}. In Figure 3 are illustrated the border generators.

Refer to caption
Figure 3. Lattice points that resolve each wall

In order to choose a good asymptotic local resolution of σ\sigma, imitating the 22-dimensional case, we can ask for a minimal local resolution, i.e., with nef canonical bundles. However, minimal varieties in higher dimensions may have terminal singularities, minimal singular models are not necessarily unique, and there are no efficient algorithms in the toric case. In this last, at least there exists a kind of optimal method. Minimal resolutions can be obtained by the canonical resolution of σ\sigma which is obtained by the canonical refinement of the cone [CLS11, Prop. 11.4.15]. However, the canonical refinement appears not to have a regular pattern for any p,qp,q. For example, if p=n1p=n-1, and 0<q<n0<q<n are free, the canonical and minimal resolutions look very simple but release a lot of exceptional data. See Figure 4. The resolution to the case p=qp=q takes the same form, just rotate the figure. And completely different from the above is the case p+q=np+q=n. See Figure 5. Thus, we do not consider these cases in the following, i.e., p,qn1p,q\neq n-1, p=qp=q or {p+q}n=0\{p+q\}_{n}=0. For more details see [TN23].

Refer to caption
Figure 4. Case p=n1p=n-1. We have cones generated by 4 generators. These cones define the canonical resolution of σ\sigma. The black dots, mark lattice points to blowing up and get a minimal non-singular resolution.
Refer to caption
Figure 5. Case p+q=np+q=n. The wall τ12\tau_{12} is of type (1,1)/n(1,1)/n. Then the minimal non-singular resolution depends only on the resolutions of the other walls.

To have a systematic way to construct resolutions, we choose the following way: We construct a cyclic resolution, i.e., a refinement of σ\sigma composed by 33-cones defining cyclic singularities. The advantage is that cyclic singularities have a well-known algorithm to resolve them [Fuj74]. Also, as we will see in the examples, we can construct some minimal resolutions. So, in future work, we expect that, under suitable conditions, these resolutions have good asymptotical behavior with respect to nn.

4.2.1. Cyclic resolution.

From (4) we know that every vPσNv\in P_{\sigma}\cap N can be written as

v=v1d1+v2d2+v3d3n,v3={pv1+qv2}n0vi<n.v=\frac{v_{1}d_{1}+v_{2}d_{2}+v_{3}d_{3}}{n},\quad v_{3}=\{pv_{1}+qv_{2}\}_{n}\quad 0\leq v_{i}<n.

Fix a vv, with v1+v2+v3nv_{1}+v_{2}+v_{3}\leq n and vi>0v_{i}>0.

Step 1: We refine by a star subdivision along vv obtaining a fan as illustrate Figure 6.

Refer to caption
Figure 6. Star subdivision along minimizer vv

Step 2: Now we refine each wall by doing toric blow-ups following the Hirzebruch-Jung algorithm. So we obtain a refinement σ\sigma^{*} and denote by XX the toric variety associated. This refinement gives us a birational projective morphism g:XYp,qg\colon X\to Y_{p,q} [CLS11, 11.1.6]. This refinement is sketched in Figure 7.

Refer to caption
Figure 7. Cyclic local resolution

Denote each 33-cone of σ\sigma^{*} by

σjk,α=C(v,ejk,α,ejk,α+1),0αsjk.\sigma_{jk,\alpha}=C(v,e_{jk,\alpha},e_{jk,\alpha+1}),\quad 0\leq\alpha\leq s_{jk}.

The 22-cones of σ\sigma^{*} are given in two types. The exterior walls τjk,α=C(ejk,α,ejk,α+1)\tau_{jk,\alpha}=C(e_{jk,\alpha},e_{jk,\alpha+1}), and the inner walls ρjk,α=C(v,ejk,α)\rho_{jk,\alpha}=C(v,e_{jk,\alpha}). For any permutation (vj,vk,vl)(v_{j},v_{k},v_{l}) with j<kj<k, using determinants and properties of Section 2.4.1, we have

mult(σjk,α)=vl,mult(ρjk,α)=gcd(vjnjk,αvkmjk,α,vl),mult(τjk,α)=1.\mbox{mult}(\sigma_{jk,\alpha})=v_{l},\quad\mbox{mult}(\rho_{jk,\alpha})=\mbox{gcd}(v_{j}n_{jk,\alpha}-v_{k}m_{jk,\alpha},v_{l}),\quad\mbox{mult}(\tau_{jk,\alpha})=1.
Lemma 4.4.

Each cone σjk,α\sigma_{jk,\alpha} is a cyclic singularity of type

(ajk,α,bjk,α,1)vl,ajk,α={mjk,α+1vknjk,α+1vj}vl,bjk,α={mjk,αvknjk,αvj}vl,\frac{(a_{jk,\alpha},b_{jk,\alpha},1)}{v_{l}},\quad a_{jk,\alpha}=\{m_{jk,\alpha+1}v_{k}-n_{jk,\alpha+1}v_{j}\}_{v_{l}},\quad b_{jk,\alpha}=\{m_{jk,\alpha}v_{k}-n_{jk,\alpha}v_{j}\}_{v_{l}},

where {}vl\{\cdot\}_{v_{l}} is the residue modulo vlv_{l}.

Proof.

Since τjk,α\tau_{jk,\alpha} is non-singular, then σjk,α\sigma_{jk,\alpha} is semi-unimodular respect to vv. By [Ash15, Prop. 1.2.3] if there is positive integer x,yx,y such that

xejk,α+yejk,α+1+vvl3,\frac{xe_{jk,\alpha}+ye_{jk,\alpha+1}+v}{v_{l}}\in{\mathbb{Z}}^{3},

then x,yx,y define the type of the cyclic singularity. Since gcd(n,vl)=1\gcd(n,v_{l})=1, we can solve the equations modulo vlv_{l} and the result follows.

Denote by h:XYp,q𝔸3h\colon X\to Y_{p,q}\to{\mathbb{A}}^{3} the composition with the natural projection to 𝔸3{\mathbb{A}}^{3}. Denote by DjD_{j} the divisor in 𝔸3{\mathbb{A}}^{3} defined by the coordinate xjx_{j}. Each ray on σ\sigma^{*} generated by dj,ejk,α,d_{j},e_{jk,\alpha}, or vv defines a toric divisor on XX given by

D~j=V(C(dj)),Ejk,α=V(C(ejk,α)),F=V(C(v)),\tilde{D}_{j}=V(C(d_{j})),\quad E_{jk,\alpha}=V(C(e_{jk,\alpha})),\quad F=V(C(v)),

where V()V(\cdot) denotes the orbit closure of a cone [CLS11, p. 121]. At the same time we fix notation for j<kj<k by Ekj,α:=Ejk,s+1αE_{kj,\alpha}:=E_{jk,s+1-\alpha}. we can compute

hDj=nD~j+kα=1sjkmjk,αEjk,α+vjF,h^{*}D_{j}=n\tilde{D}_{j}+\sum_{k}\sum_{\alpha=1}^{s_{jk}}m_{jk,\alpha}E_{jk,\alpha}+v_{j}F,
KX=gKYp,q+j<kα=1sjkNjk,αEjk,α+v1+v2+v3nnF,K_{X}=g^{*}K_{Y_{p,q}}+\sum_{j<k}\sum_{\alpha=1}^{s_{jk}}N_{jk,\alpha}E_{jk,\alpha}+\frac{v_{1}+v_{2}+v_{3}-n}{n}F, (6)

where

Njk,α:=mjk,α+njk,αn1.N_{jk,\alpha}:=\frac{m_{jk,\alpha}+n_{jk,\alpha}}{n}-1.

It is satisfied the relation kjk,αNjk,αNjk,α+1=Njk,α1(kjk,α2),k_{jk,\alpha}N_{jk,\alpha}-N_{jk,\alpha+1}=N_{jk,\alpha-1}-(k_{jk,\alpha}-2), which gives

α=1sjkNjk,α(kjk,α2)=(Njk,1+Nkj,1)α=1sjk(kjk,α2).\sum_{\alpha=1}^{s_{jk}}N_{jk,\alpha}(k_{jk,\alpha}-2)=-(N_{jk,1}+N_{kj,1})-\sum_{\alpha=1}^{s_{jk}}(k_{jk,\alpha}-2).
Proposition 4.5.

The {\mathbb{Q}}-divisor

h(n1n(D1+D2+D3))+j<kαNjk,αEjk,α+v1+v2+v3nnF,h^{*}\left(\frac{n-1}{n}(D_{1}+D_{2}+D_{3})\right)+\sum_{j<k}\sum_{\alpha}N_{jk,\alpha}E_{jk,\alpha}+\frac{v_{1}+v_{2}+v_{3}-n}{n}F,

is an effective {\mathbb{Z}}-divisor.

Proof.

The local pullback h(D1+D2+D3)h^{*}(D_{1}+D_{2}+D_{3}) equals

n(D1~+D2~+D3~)+j<k,α(mjk,α+njk,α)Ejk,α+(v1+v2+v3)Fjkl,p.n(\tilde{D_{1}}+\tilde{D_{2}}+\tilde{D_{3}})+\sum_{j<k,\alpha}(m_{jk,\alpha}+n_{jk,\alpha})E_{jk,\alpha}+(v_{1}+v_{2}+v_{3})F_{jkl,p}.

Thus, h(n1n(Dj+Dk+Dl))+Δh^{*}\left(\frac{n-1}{n}(D_{j}+D_{k}+D_{l})\right)+\Delta equals to

(n1)(D1~+D2~+D3~)+j<k,α(mjk,α+njk,α1)Ejk,α+(v1+v2+v31)F,(n-1)(\tilde{D_{1}}+\tilde{D_{2}}+\tilde{D_{3}})+\sum_{j<k,\alpha}\left(m_{jk,\alpha}+n_{jk,\alpha}-1\right)E_{jk,\alpha}+(v_{1}+v_{2}+v_{3}-1)F,

i.e., an effective {\mathbb{Z}}-divisor. ∎

Each inner wall defines a closed curve on XX given by

Cj=V(C(dj,v)),Cjk,α=V(C(ρjk,α)).C_{j}=V(C(d_{j},v)),\quad C_{jk,\alpha}=V(C(\rho_{jk,\alpha})).

The refinement σ\sigma^{*} is simplicial with cones of multiplicity one, and the canonical divisor KXK_{X} is a {\mathbb{Q}}-Cartier divisor. For any pair vj,vkv_{j},v_{k}, j<kj<k, let vlv_{l} be the another coordinate. The following relation among lattices generators,

ejk,α1+(kjk,α)ejk,α+0v+ejk,α+1=0,e_{jk,\alpha-1}+(-k_{jk,\alpha})e_{jk,\alpha}+0\cdot v+e_{jk,\alpha+1}=0,
vjelj,1+(vlmlj,1vjmlk,1vkn)dl+(1)v+vkelk,1=0,v_{j}e_{lj,1}+\left(\frac{v_{l}-m_{lj,1}v_{j}-m_{lk,1}v_{k}}{n}\right)d_{l}+(-1)v+v_{k}e_{lk,1}=0,

describe the intersection theory on XX [Ful93, Ch. V]. We have,

Ejk,αCjk,α±1=mult(ρjk,α)vl,Ejk,αCjk,α=kjk,αmult(ρjk,α)vl,FCjk,α=0,E_{jk,\alpha}C_{jk,\alpha\pm 1}=\frac{\mbox{mult}(\rho_{jk,\alpha})}{v_{l}},\quad E_{jk,\alpha}C_{jk,\alpha}=-\frac{k_{jk,\alpha}\mbox{mult}(\rho_{jk,\alpha})}{v_{l}},\quad FC_{jk,\alpha}=0,
D~lCl=gcd(vj,vk)(vlmlj,1vjmlk,1vk))nvjvk,FCl=gcd(vj,vk)vjvk,\tilde{D}_{l}\,C_{l}=\frac{\gcd(v_{j},v_{k})(v_{l}-m_{lj,1}v_{j}-m_{lk,1}v_{k}))}{nv_{j}v_{k}},\quad F\,C_{l}=-\frac{\gcd(v_{j},v_{k})}{v_{j}v_{k}},
KXCl=gcd(vj,vk)vjvk(vj+vk1+vlmlj,1vjmlk,1vkn),K_{X}\,C_{l}=-\frac{\gcd(v_{j},v_{k})}{v_{j}v_{k}}\left(v_{j}+v_{k}-1+\frac{v_{l}-m_{lj,1}v_{j}-m_{lk,1}v_{k}}{n}\right),
KXCjk,α=mult(ρjk,α)vl(kjk,α2),1αsjk.K_{X}\,C_{jk,\alpha}=\frac{\mbox{mult}(\rho_{jk,\alpha})}{v_{l}}(k_{jk,\alpha}-2),\quad 1\leq\alpha\leq s_{jk}. (7)
Lemma 4.6.

We have

F3=nv1v2v3.F^{3}=\frac{n}{v_{1}v_{2}v_{3}}.
Proof.

The divisor vjFv_{j}F is Cartier for any jj, then from the pullback identities above we have

v1v2v3F3=j=13(hDjnDjkαmjk,αEjk,α)=hD1hD2hD3=n,v_{1}v_{2}v_{3}F^{3}=\prod_{j=1}^{3}\left(h^{*}D_{j}-nD_{j}-\sum_{k}\sum_{\alpha}m_{jk,\alpha}E_{jk,\alpha}\right)=h^{*}D_{1}\,h^{*}D_{2}\,h^{*}D_{3}=n,

where the last is by projection formula.∎

As a consequence, we can compute,

KXF2\displaystyle K_{X}F^{2} =F2(D1+D2+D3+F)\displaystyle=-F^{2}\left(D_{1}+D_{2}+D_{3}+F\right)
=v1+v2+v3nv1v2v3.\displaystyle=\frac{v_{1}+v_{2}+v_{3}-n}{v_{1}v_{2}v_{3}}.

From the last one, we get

KX2F\displaystyle K_{X}^{2}\,F =lKXClgcd(vj,vk)j<k,αKXCjk,αmult(ρjk,α)KXF2.\displaystyle=-\sum_{l}\frac{K_{X}\,C_{l}}{\gcd(v_{j},v_{k})}-\sum_{j<k,\alpha}\frac{K_{X}\,C_{jk,\alpha}}{\mbox{mult}(\rho_{jk,\alpha})}-K_{X}F^{2}.
Example 4.7.

Case {p+q}n=1\{p+q\}_{n}=1: For v1=v2=1v_{1}=v_{2}=1, we have v3={p+q}n=1v_{3}=\{p+q\}_{n}=1. Thus, these coordinates define an interior lattice point vv, and it minimizes v1+v2+v3v_{1}+v_{2}+v_{3}. We have

mult(σjk,α)=mult(ρjk,α)=mult(τjk,α)=1,\mbox{mult}(\sigma_{jk,\alpha})=\mbox{mult}(\rho_{jk,\alpha})=\mbox{mult}(\tau_{jk,\alpha})=1,

i.e., XX is non-singular. On the other hand,

KXCjk,α=(kjk,α2),KXCl=0,K_{X}C_{jk,\alpha}=(k_{jk,\alpha}-2),\quad K_{X}C_{l}=0,

for all j,kj,k and ll. Thus, KXK_{X} is nef.

Example 4.8.

Case {p+q}n=2\{p+q\}_{n}=2: In this case, again v1=v2=1v_{1}=v_{2}=1 defines the minimizer interior lattice point vv. In this case, we have σ13,α\sigma_{13,\alpha} and σ23,α\sigma_{23,\alpha} as non-singular cones. On the other hand, each σ12,α\sigma_{12,\alpha} is cyclic singularity of order 22. Thus, they define canonical and terminal singularities. The first ones achieve a terminal resolution with one blow-up. Moreover,

KXC3=0,KXCj{0,12},j=1,2K_{X}C_{3}=0,\quad K_{X}C_{j}\in\left\{0,-\frac{1}{2}\right\},\quad j=1,2

so there are p,qp,q with canonical divisor KXK_{X} nef. In the worst case, i.e., KXCj<0K_{X}\,C_{j}<0 for j=1,2j=1,2, we can do toric flips a to get a nef toric variety given whose fan is sketched in Figure 8.

Refer to caption
Figure 8. Flipped fan for {p+q}n=2\{p+q\}_{n}=2.
Example 4.9.

If we drag the lattice point vv to one of the generators of the cone σ\sigma, we get a degenerated fan as in Figure 9. In this case, the singularities are of order nn, and as an advantage, we do not have a divisor FF.

Refer to caption
Figure 9. Degenerated cyclic resolution

As we see, having vj2v_{j}\leq 2 gives us good singularities to work. Indeed, if the vjsv_{j}^{\prime}s are small enough, the singularities are also good in asymptotic terms. The following arithmetic lemma will be useful in Section 4.3.

Lemma 4.10.

There exists vPσNv\in P_{\sigma}\cap N such that v1+v2+v3=nv_{1}+v_{2}+v_{3}=n, so KXK_{X} has multiplicity zero at FF. Moreover, for n0n\gg 0 we can choose vv such that the slopes vj/vk3v_{j}/v_{k}\leq 3.

Proof.

By (4) we have v3={v1p+v2q}nv_{3}=\{v_{1}p+v_{2}q\}_{n}. So v1+v2+v3=nv_{1}+v_{2}+v_{3}=n, implies that (v1,v2)(v_{1},v_{2}) are solutions (x,y)(x,y) of the Diophantine equation

ycxmodn,c={(p+1)(q+1)}n.y\equiv cx\mod n,\quad c=\{-(p+1)(q+1)^{\prime}\}_{n}.

Moreover, for any of those solutions with x+y<nx+y<n, we have v3=nxyv_{3}=n-x-y. A degenerate case is p+q=n2p+q=n-2, equivalently c=1c=1, thus the solution to the equation is the diagonal. Thus, we can choose x=y=n3x=y=\lfloor\frac{n}{3}\rfloor, and the result follows for this case. Now, we assume that c<n/2c<n/2, otherwise we do (x,y)(x,y)(x,y)\mapsto(-x,y). The integer points in the square [1,n1]2[1,n-1]^{2} solving the equation distribute in 2{\mathbb{R}}^{2} along the lines Lβ:y=cxβnL_{\beta}:y=cx-\beta n for 0βc10\leq\beta\leq c-1. Thus, over each LβL_{\beta} the integer solutions over the line are defined by those integers in the interval

Iβ=(βnc,(β+1)nc].I_{\beta}=\left(\left\lfloor\frac{\beta n}{c}\right\rfloor,\left\lfloor\frac{(\beta+1)n}{c}\right\rfloor\right].

For each 1knc1\leq k\leq\lfloor\frac{n}{c}\rfloor, we have

yk=y(βnc+k)=ckr,y_{k}=y\left(\left\lfloor\frac{\beta n}{c}\right\rfloor+k\right)=ck-r,

where 0r<c0\leq r<c is the residue of βn\beta n modulo cc. So, c(k1)ykckc(k-1)\leq y_{k}\leq ck. Let us choose β=c13,\beta=\lfloor\frac{c-1}{3}\rfloor, and k=n3ck=\lfloor\frac{n}{3c}\rfloor. In particular n3Iβ\lfloor\frac{n}{3}\rfloor\in I_{\beta}. So, as n0n\gg 0 we have ykn3y_{k}\approx\frac{n}{3}. By construction, we have xk=βnc+kx_{k}=\left\lfloor\frac{\beta n}{c}\right\rfloor+k. Let 0r20\leq r^{*}\leq 2 the residue of c1c-1 modulo 33, then xkn(cr)3cx_{k}\approx\frac{n(c-r^{*})}{3c}. Since 12crc1\frac{1}{2}\leq\frac{c-r^{*}}{c}\leq 1, the result follows choosing v1=xkv_{1}=x_{k} and v2=ykv_{2}=y_{k} as n0n\gg 0. ∎

4.3. Global resolution

Let {D1,,Dr}\{D_{1},...,D_{r}\} be an asymptotic arrangement on ZZ. Thus, for prime numbers n0n\gg 0 we have multiplicities 0<νj<n0<\nu_{j}<n depending on nn, with its respective qjkOnq_{jk}\in O_{n}. We have nn-th root covers fn:YnZf_{n}\colon Y_{n}\to Z branched at each D=jνjDjD=\sum_{j}\nu_{j}D_{j}. Let us fix a n0n\gg 0, so we drop the subscript nn of the morphisms, i.e., we are fixing a f:YnZf\colon Y_{n}\to Z.

For j<kj<k, the singularities of YnY_{n} over curves in Djk:=DjDkD_{jk}:=D_{j}\cap D_{k} are locally analytically isomorphic to Cqjk,n×C_{q_{jk},n}\times{\mathbb{C}} where Cqjk,nC_{q_{jk},n} is the surface cyclic quotient singularity of type 1n(qjk,1)\frac{1}{n}(q_{jk},1). For a triple j<k<lj<k<l, the singularity of YnY_{n} over a point in Djkl:=DjDkDlD_{jkl}:=D_{j}\cap D_{k}\cap D_{l} is locally analytically isomorphic to the normalization of Spec(σjklM)\mbox{Spec}(\sigma_{jkl}^{\vee}\cap M) where σjkl\sigma_{jkl} is a cone with walls of types 1n(qjk,1),1n(qjl,1),1n(qkl,1)\frac{1}{n}(q_{jk},1),\frac{1}{n}(q_{jl},1),\frac{1}{n}(q_{kl},1) as we see in Section 4.2. We will get the cyclic resolution XnYnX_{n}\to Y_{n} via weighted blow-ups in two steps.

Step 1: Since singularities over DjklD_{jkl} are isolated, for each σjkl\sigma_{jkl}, we do a weighted blow-up at a convenient interior lattice point vjklv^{jkl}. So, this refinement locally gives a projective morphism which is a blow-up over an isolated point [CLS11, 11.1.6]. In this case, we get a projective birational morphism h:XnYnh^{\prime}:X_{n}^{\prime}\to Y_{n}. We have exceptional divisors FjklF_{jkl} whose components are over the points of DjklD_{jkl} and they are isomorphic to weighted fake projective planes [Buc08]. For future computations, we fix the notation of vjklv^{jkl} and FjklF_{jkl} independent of the order of the triple j,k,lj,k,l. For example, vjkl=vkjlv^{jkl}=v^{kjl}.

Step 2: Since the centers of the above blow-ups are points, the singularity type over intersections DjkD_{jk} was not affected. For curves in DjkD_{jk}, locally by SNC property, we can assume that they are supported on a local equation xy=0xy=0 for local coordinates x,yx,y. Then over such curves, the singularities on XnX^{\prime}_{n} are locally analytically isomorphic to Cqjk,n×C_{q_{jk},n}\times{\mathbb{C}}, thus we use the Hirzebruch-Jung algorithm which is a weighted blow-up to resolve Cqjk,nC_{q_{jk},n}. The Hirzebruch-Jung resolution can be realized by a single blow-up Bl𝔪(Cqjk,n)Cqjk,Bl_{{\mathfrak{m}}}(C_{q_{jk,n}})\to C_{q_{jk,}} where 𝔪{\mathfrak{m}} is a maximal ideal determined explicitly in coordinates x,yx,y as we see at the end of Section 2.4.1 (also see [KM92, 10.5]). In terms of local resolutions, we need to follow an order compatible with the resolution, i.e., if we locally blow-up a curve in DjkD_{jk} then this operation must be reflected on the other local toric pictures following the centers to blowing-up. See Figure 10. Thus, this construction extends and we have resolved the curves DjkD_{jk}. Consequently, we get a projective morphism denoted by g:XnXnYng\colon X_{n}\to X^{\prime}_{n}\to Y_{n}, and denote by h:XnZh\colon X_{n}\to Z the composition. Since XnX_{n} was constructed by a sequence of weighted blow-ups with cyclic singularities, then, XnX_{n} is an embedded {\mathbb{Q}}-resolution of YnY_{n} [ABMMOG12, 2.1]. As we see in 2.17, the varieties XnX_{n} are irreducible. We summarize this in the following.

Proposition 4.11 (and Definition).

There exists a cyclic partial resolution g:XnYng:X_{n}\to Y_{n}, i.e. a projective, surjective, birational morphism such that XnX_{n} is irreducible and, it has at most isolated cyclic quotient singularities of order lower than nn.

Refer to caption
Figure 10. Assume r=4r=4 with Djkl=1D_{jkl}=1, then on YY the singularities over each DjklD_{jkl} can be sketched as in the figure. So the resolution process is in the following order: First the internals blow-ups, and then the walls in the following order D12,D13,D14,D23,D24,D34D_{12},D_{13},D_{14},D_{23},D_{24},D_{34}.

In the rest, we will abuse notation using DjkD_{jk} and DjklD_{jkl} for both, set-theoretic intersections DjDkD_{j}\cap D_{k} and DjDkDlD_{j}\cap D_{k}\cap D_{l}, and for intersections of cycles DjDkD_{j}D_{k} and DjDkDlD_{j}D_{k}D_{l}. Over each DjkD_{jk}, we get exceptional divisors Ejk,αE_{jk,\alpha}, 0αsjk0\leq\alpha\leq s_{jk}, where sjk=(n,qjk)s_{jk}=\ell(n,q_{jk}) and whose components are over those of DjkD_{jk}. From the local computations of the section above, we have

hDj=nD~j+DkDjα=1sjkmjk,αEjk,α+DklDjFkl,j,h^{*}D_{j}=n\tilde{D}_{j}+\sum_{D_{k}D_{j}\neq\emptyset}\sum_{\alpha=1}^{s_{jk}}m_{jk,\alpha}E_{jk,\alpha}+\sum_{D_{kl}D_{j}\neq\emptyset}{F_{kl,j}}, (8)

where Fkl,jF_{kl,j} is a divisor whose components are the exceptional divisors over points in DjklD_{jkl}.

Explicitly, for any triple of positive integers numbers j,k,lj,k,l, let ρkl(j){1,2,3}\rho_{kl}(j)\in\{1,2,3\} the position of jj if we order the triple. For example ρ23(1)=1\rho_{23}(1)=1, ρ57(6)=2\rho_{57}(6)=2 or ρ54(8)=3\rho_{54}(8)=3. Thus, if Fjkl,pF_{jkl,p} is the exceptional divisor over a point pDjklp\in D_{jkl}, then

Fjkl=pDjklFjkl,pF_{jkl}=\sum_{p\in D_{jkl}}F_{jkl,p}
Fkl,j=pDjklρkl(j)Fjkl,p.F_{kl,j}=\sum_{p\in D_{jkl}}\rho_{kl}(j)F_{jkl,p}. (9)

In terms of intersection theory, we have

Fjkl3=nv1jklv2jklv3jklDjkl,Dj~Dk~=0.F_{jkl}^{3}=\frac{n}{v^{jkl}_{1}v^{jkl}_{2}v^{jkl}_{3}}D_{jkl},\quad\tilde{D_{j}}\tilde{D_{k}}=0.

From Theorem 2.13 we have

KXnhK+Δ,K_{X_{n}}\sim_{{\mathbb{Q}}}h^{*}K+\Delta, (10)
K:=(KZ+n1nj=1rDj),Δ=j<kEjk+j<k<lVjklFjkl,K:=\left(K_{Z}+\frac{n-1}{n}\sum_{j=1}^{r}D_{j}\right),\quad\Delta=\sum_{j<k}E_{jk}+\sum_{j<k<l}V_{jkl}F_{jkl},
Ejk=α=1sjkNjk,αEjk,αNjk,α=mjk,α+njk,αnn,Vjkl=v1jkl+v2jkl+v3jklnn.E_{jk}=\sum_{\alpha=1}^{s_{jk}}N_{jk,\alpha}E_{jk,\alpha}\quad N_{jk,\alpha}=\frac{m_{jk,\alpha}+n_{jk,\alpha}-n}{n},\quad V_{jkl}=\frac{v_{1}^{jkl}+v_{2}^{jkl}+v_{3}^{jkl}-n}{n}.

Now we describe how the intersection theory on XnX_{n} behaves under pullbacks of divisors of ZZ. In what follows, we set Ejk,0=D~jE_{jk,0}=\tilde{D}_{j} and Ejk,sjk+1=D~kE_{jk,s_{jk}+1}=\tilde{D}_{k}.

Proposition 4.12.

Let G,GG,G^{\prime} any divisors on ZZ, then hGFkl,j=0h^{*}GF_{kl,j}=0 as 22-cycle, for any 1αsjk1\leq\alpha\leq s_{jk},

hGhGEjk,α=0,h^{*}Gh^{*}G^{\prime}E_{jk,\alpha}=0,
hGEjk,α2=kjk,αGDjk,hGEjk,αEjk,α±1=GDjk.h^{*}GE_{jk,\alpha}^{2}=-k_{jk,\alpha}GD_{jk},\quad h^{*}GE_{jk,\alpha}E_{jk,\alpha\pm 1}=GD_{jk}.
Proof.

We use the projection formula repeatedly. The first one is given by the fact that hFjk,lh_{*}F_{jk,l} has codimension 33. Now for a 1αsjk1\leq\alpha\leq s_{jk} we have hEjk,αh_{*}E_{jk,\alpha} supported in codimension 22, thus

hGhGEjk,α=GGhEjk,α=0.h^{*}Gh^{*}G^{\prime}E_{jk,\alpha}=GG^{\prime}h_{*}E_{jk,\alpha}=0.

Finally, for any α\alpha we have

hDjhGEjk,α=0=hG(mjk,α1Ejk,α1Ejk,α+mjk,αEjk,α2+mjk,α+1Ejk,α+1Ejk,α),h^{*}D_{j}h^{*}GE_{jk,\alpha}=0=h^{*}G(m_{jk,\alpha-1}E_{jk,\alpha-1}E_{jk,\alpha}+m_{jk,\alpha}E_{jk,\alpha}^{2}+m_{jk,\alpha+1}E_{jk,\alpha+1}E_{jk,\alpha}),
hDkhGEjk,α=0=hG(njk,α1Ejk,α1Ejk,α+njk,αEjk,α2+njk,α+1Ejk,α+1Ejk,α).h^{*}D_{k}h^{*}GE_{jk,\alpha}=0=h^{*}G(n_{jk,\alpha-1}E_{jk,\alpha-1}E_{jk,\alpha}+n_{jk,\alpha}E_{jk,\alpha}^{2}+n_{jk,\alpha+1}E_{jk,\alpha+1}E_{jk,\alpha}).

The recursive relations with kjk,αk_{jk,\alpha} give

[mjk,αmjk,α+1njk,αnjk,α+1][hG(Ejk,α2+kjk,αEjk,α1Ejk,α)hG(Ejk,α+1Ejk,αEjk,α1Ejk,α)]=[00].\left[\begin{array}[]{cc}m_{jk,\alpha}&m_{jk,\alpha+1}\\ n_{jk,\alpha}&n_{jk,\alpha+1}\\ \end{array}\right]\left[\begin{array}[]{c}h^{*}G(E_{jk,\alpha}^{2}+k_{jk,\alpha}E_{jk,\alpha-1}E_{jk,\alpha})\\ h^{*}G(E_{jk,\alpha+1}E_{jk,\alpha}-E_{jk,\alpha-1}E_{jk,\alpha})\end{array}\right]=\left[\begin{array}[]{c}0\\ 0\end{array}\right].

From Lemma 2.23 we have the determinant mjk,αnjk,α+1mjk,α+1njk,α=nm_{jk,\alpha}n_{jk,\alpha+1}-m_{jk,\alpha+1}n_{jk,\alpha}=n, thus

hG(Ejk,α2+kjk,αEjk,α1Ejk,α)=0h^{*}G(E_{jk,\alpha}^{2}+k_{jk,\alpha}E_{jk,\alpha-1}E_{jk,\alpha})=0
hG(Ejk,α+1Ejk,αEjk,α1Ejk,α)=0.h^{*}G(E_{jk,\alpha+1}E_{jk,\alpha}-E_{jk,\alpha-1}E_{jk,\alpha})=0.

In particular, we have

hGEjk,α+1Ejk,α=hGD~jEjk,1=GDjk,h^{*}GE_{jk,\alpha+1}E_{jk,\alpha}=h^{*}G\tilde{D}_{j}E_{jk,1}=GD_{jk},

and the result follows. ∎

Corollary 4.13.

For any divisor GG on ZZ we have

hGEjkKX=DjkG((Njk,1+Nkj,1)+α=1sjk(kjk,α2)).h^{*}GE_{jk}K_{X}=-D_{jk}G\left((N_{jk,1}+N_{kj,1})+\sum_{\alpha=1}^{s_{jk}}(k_{jk,\alpha}-2)\right).
Proof.

Since hGFh^{*}G\,F vanishes at top-dimensional intersections, and Ejk,α1Ejk,α2=0E_{jk,\alpha_{1}}E_{jk,\alpha_{2}}=0 for |α1α2|>1|\alpha_{1}-\alpha_{2}|>1, we have

hGEjkKX\displaystyle h^{*}GE_{jk}K_{X} =hG(Ejk)2=hGα=1sjkNjk,α2Ejk,α2+2Njk,αNjk,α+1Ejk,αEjk,α+1\displaystyle=h^{*}G(E_{jk})^{2}=h^{*}G\sum_{\alpha=1}^{s_{jk}}N_{jk,\alpha}^{2}E_{jk,\alpha}^{2}+2N_{jk,\alpha}N_{jk,\alpha+1}E_{jk,\alpha}E_{jk,\alpha+1}
=DjkGα=1sjkkjk,αNjk,α2+2Njk,αNjk,α+1\displaystyle=D_{jk}G\sum_{\alpha=1}^{s_{jk}}-k_{jk,\alpha}N_{jk,\alpha}^{2}+2N_{jk,\alpha}N_{jk,\alpha+1}
=DjkGα=1sjkNjk,αNjk,α+1Njk,αNjk,α1+Njk,α(kjk,α2)\displaystyle=D_{jk}G\sum_{\alpha=1}^{s_{jk}}N_{jk,\alpha}N_{jk,\alpha+1}-N_{jk,\alpha}N_{jk,\alpha-1}+N_{jk,\alpha}(k_{jk,\alpha}-2)
=DjkGα=1sjkNjk,α(kjk,α2)\displaystyle=D_{jk}G\sum_{\alpha=1}^{s_{jk}}N_{jk,\alpha}(k_{jk,\alpha}-2)
=DjkG((Njk,1+Nkj,1)+α=1sjk(kjk,α2)).\displaystyle=-D_{jk}G\left((N_{jk,1}+N_{kj,1})+\sum_{\alpha=1}^{s_{jk}}(k_{jk,\alpha}-2)\right).

where the last identity is by telescoping sum argument. ∎

Corollary 4.14.

If CC is a curve on XnX_{n} contained in H~j\tilde{H}_{j} and disjoint to any exceptional divisor FjklF_{jkl}, then

D~jC=hCn(Djjkmjk,1Dk).\tilde{D}_{j}C=\frac{h_{*}C}{n}\left(D_{j}-\sum_{j\neq k}m_{jk,1}D_{k}\right).
Proof.

From (8) we have

hDjC=DjhC=nD~j+kmjk,1Ejk,1C,h^{*}D_{j}C=D_{j}h_{*}C=n\tilde{D}_{j}+\sum_{k}m_{jk,1}E_{jk,1}C,
hDkC=DkhC=Ejk,1C,jk,h^{*}D_{k}C=D_{k}h_{*}C=E_{jk,1}C,\quad\forall j\neq k,

and the result follows. ∎

4.4. Asymptoticity of KXn3K_{X_{n}}^{3}.

For simplicity, let us denote KX=KXnK_{X}=K_{X_{n}}. Let us introduce the following notation

|D|jk:=|Dj|k+|Dk|j,|D|_{jk}:=|D_{j}|_{k}+|D_{k}|_{j},

where |Dj|k|D_{j}|_{k} satisfy

DllDj0Fll,jEjk,αKX=|Dj|k(kjk,α2).\sum_{D_{ll^{\prime}}D_{j}\neq 0}F_{ll^{\prime},j}E_{jk,\alpha}K_{X}=|D_{j}|_{k}(k_{jk,\alpha}-2).
Lemma 4.15.

We have

|Dj|k=lvρkl(j)jklvρkj(l)jklDjkl.|D_{j}|_{k}=\sum_{l}\dfrac{v^{jkl}_{\rho_{kl}(j)}}{{v^{jkl}_{\rho_{kj}(l)}}}D_{jkl}.
Proof.

Using equations from (7), observe that |Dj|k|D_{j}|_{k} depends on slopes of weights v1jkl,v2jkl,v3jklv^{jkl}_{1},v^{jkl}_{2},v^{jkl}_{3} of the lattice points vjklv^{jkl}. Using (9), we get the result.

We need this to compute the intersection of KXK_{X} with the external walls of the local toric resolution. Recursively we denote,

xjk,α=KXEjk,α1Ejk,α,1αsjk+1x_{jk,\alpha}=K_{X}E_{jk,\alpha-1}E_{jk,\alpha},\quad 1\leq\alpha\leq s_{jk}+1
yjk,α=KXEjk,α2,1αsjk.y_{jk,\alpha}=K_{X}E_{jk,\alpha}^{2},\quad 1\leq\alpha\leq s_{jk}.

Thus, we can write

hDjEjk,αKX=mjk,α1xjk,α+mjk,αyjk,α+mjk,α+1xjk,α+1+|Dj|k(kjk,α2)h^{*}D_{j}E_{jk,\alpha}K_{X}=m_{jk,\alpha-1}x_{jk,\alpha}+m_{jk,\alpha}y_{jk,\alpha}+m_{jk,\alpha+1}x_{jk,\alpha+1}+|D_{j}|_{k}(k_{jk,\alpha}-2) (11)

Using the {\mathbb{Q}}-numerical equivalence of KXK_{X} in 2.13, we compute

xjk,1=hDkDj~KX=Djk(K+ljNjl,1Dl),x_{jk,1}=h^{*}D_{k}\tilde{D_{j}}K_{X}=D_{jk}\left(K+\sum_{l\neq j}N_{jl,1}D_{l}\right),
xjk,s+1=hDjDk~KX=Djk(K+lkNkl,1Dl).x_{jk,s+1}=h^{*}D_{j}\tilde{D_{k}}K_{X}=D_{jk}\left(K+\sum_{l\neq k}N_{kl,1}D_{l}\right).
Proposition 4.16.

We have

xjk,α=xjk,1+1n(mjk,α(DjkDk|Dk|j)njk,α(DjkDj|Dj|k)),x_{jk,\alpha}=x_{jk,1}+\frac{1}{n}(m_{jk,\alpha}^{*}(D_{jk}D_{k}-|D_{k}|_{j})-n_{jk,\alpha}^{*}(D_{jk}D_{j}-|D_{j}|_{k})),
yjk,α=kjk,αxjk,α+(kjk,α2)n(njk,α+1(DjkDj|Dj|k)mjk,α+1(DjkDk|Dk|j)).y_{jk,\alpha}=-k_{jk,\alpha}x_{jk,\alpha}+\frac{(k_{jk,\alpha}-2)}{n}(n_{jk,\alpha+1}(D_{jk}D_{j}-|D_{j}|_{k})-m_{jk,\alpha+1}(D_{jk}D_{k}-|D_{k}|_{j})).

Where mjk,α=mjk,αmjk,α1mjk,1+mjk,0m_{jk,\alpha}^{*}=m_{jk,\alpha}-m_{jk,\alpha-1}-m_{jk,1}+m_{jk,0} and analogous for njk,αn_{jk,\alpha}^{*}.

Proof.

Using the recursion given by the kjk,αsk_{jk,\alpha}^{\prime}s, and formulas for hDjEjk,αKXh^{*}D_{j}\,E_{jk,\alpha}K_{X} and hDkEjk,αKXh^{*}D_{k}\,E_{jk,\alpha}K_{X} of (11), we have

[(Dj2Dk|Dj|k)(kjk,α2)(DjDk2|Dk|j)(kjk,α2)]=[mjk,αmjk,α+1njk,αnjk,α+1][kjk,αxjk,α+yjk,αxjk,α+1xjk,α]\left[\begin{array}[]{c}(D_{j}^{2}D_{k}-|D_{j}|_{k})(k_{jk,\alpha}-2)\\ (D_{j}D_{k}^{2}-|D_{k}|_{j})(k_{jk,\alpha}-2)\end{array}\right]=\left[\begin{array}[]{cc}m_{jk,\alpha}&m_{jk,\alpha+1}\\ n_{jk,\alpha}&n_{jk,\alpha+1}\end{array}\right]\left[\begin{array}[]{c}k_{jk,\alpha}x_{jk,\alpha}+y_{jk,\alpha}\\ x_{jk,\alpha+1}-x_{jk,\alpha}\end{array}\right]

The determinant mjk,αnjk,α+1mjk,α+1njk,α=nm_{jk,\alpha}n_{jk,\alpha+1}-m_{jk,\alpha+1}n_{jk,\alpha}=n, implies second relation for yjk,αy_{jk,\alpha}, and

xjk,α+1=xjk,α+(kjk,α2)n(mjk,α(DjkDk|Dk|j)njk,α(DjkDj|Dj|k))x_{jk,\alpha+1}=x_{jk,\alpha}+\frac{(k_{jk,\alpha}-2)}{n}(m_{jk,\alpha}(D_{jk}D_{k}-|D_{k}|_{j})-n_{jk,\alpha}(D_{jk}D_{j}-|D_{j}|_{k}))

The recurrence for xjk,αx_{jk,\alpha} with a telescopic sum argument give the result.

Proposition 4.17.

We have,

KX3=\displaystyle K_{X}^{3}= nK32j<k(DjkK+j<k<lVjklDjkl)(Njk,1+Nkj,1+(k2)jk)\displaystyle nK^{3}-2\sum_{j<k}\left(D_{jk}K+\sum_{j<k<l}V_{jkl}D_{jkl}\right)(N_{jk,1}+N_{kj,1}+(k-2)_{jk})
+j<k<lnVjkl3v1jklv2jklv3jklDjkl+j<kα=1sjkDjk(Dj+Dk)|D|jknNjk,α(kjk,α2)\displaystyle+\sum_{j<k<l}\frac{nV_{jkl}^{3}}{v_{1}^{jkl}v_{2}^{jkl}v_{3}^{jkl}}D_{jkl}+\sum_{j<k}\sum_{\alpha=1}^{s_{jk}}\frac{D_{jk}(D_{j}+D_{k})-|D|_{jk}}{n}N_{jk,\alpha}(k_{jk,\alpha}-2)
j<kα=1sjkNjk,α(xjk,α+yjk,α+xjk,α+1).\displaystyle-\sum_{j<k}\sum_{\alpha=1}^{s_{jk}}N_{jk,\alpha}(x_{jk,\alpha}+y_{jk,\alpha}+x_{jk,\alpha+1}).
Proof.

We will compute KX3K_{X}^{3} using the numerical equivalence of (10). Squaring we get

KX2hK2+(j<kEjk)2+j<k<lVjkl2Fjkl2+2j<kEjk(hK+j<k<lVjklFjkl).\displaystyle K_{X}^{2}\sim_{{\mathbb{Q}}}h^{*}K^{2}+\left(\sum_{j<k}E_{jk}\right)^{2}+\sum_{j<k<l}V_{jkl}^{2}F_{jkl}^{2}+2\sum_{j<k}E_{jk}\left(h^{*}K+\sum_{j<k<l}V_{jkl}F_{jkl}\right).

We have explicitly

(hK)2KX=(hK)3=nK3.(h^{*}K)^{2}K_{X}=(h^{*}K)^{3}=nK^{3}.
EjkFjklKX=Djklα=1sjkNjk,α(kjk,α2)=Djkl((Njk,1+Nkj,1)+α=1sjk(kjk,α2)).E_{jk}F_{jkl}K_{X}=D_{jkl}\sum_{\alpha=1}^{s_{jk}}N_{jk,\alpha}(k_{jk,\alpha}-2)=-D_{jkl}\left((N_{jk,1}+N_{kj,1})+\sum_{\alpha=1}^{s_{jk}}(k_{jk,\alpha}-2)\right).

In the rest, we will denote

(k2)jk:=α=1sjk(kjk,α2).(k-2)_{jk}:=\sum_{\alpha=1}^{s_{jk}}(k_{jk,\alpha}-2).

Using Corollary 4.13 for G=KG=K, we get

KX3=\displaystyle K_{X}^{3}= nK32j<k(DjkK+j<k<lVjklDjkl)(Njk,1+Nkj,1+(k2)jk)\displaystyle nK^{3}-2\sum_{j<k}\left(D_{jk}K+\sum_{j<k<l}V_{jkl}D_{jkl}\right)(N_{jk,1}+N_{kj,1}+(k-2)_{jk})
+j<k<lnVjkl3v1jklv2jklv3jklDjkl+KX(j<kEjk)2.\displaystyle+\sum_{j<k<l}\frac{nV_{jkl}^{3}}{v_{1}^{jkl}v_{2}^{jkl}v_{3}^{jkl}}D_{jkl}+K_{X}\left(\sum_{j<k}E_{jk}\right)^{2}.

Just rest to compute

KX(j<kEjk)2=j<kα=1sjkNjk,α(Njk,α1xjk,α+Njk,αyjk,α+Njk,α+1xjk,α+1).K_{X}\left(\sum_{j<k}E_{jk}\right)^{2}=\sum_{j<k}\sum_{\alpha=1}^{s_{jk}}N_{jk,\alpha}(N_{jk,\alpha-1}x_{jk,\alpha}+N_{jk,\alpha}y_{jk,\alpha}+N_{jk,\alpha+1}x_{jk,\alpha+1}).

From Corollary 4.13, we have

Djk(Dj+Dk)(kjk,α2)n\displaystyle\frac{D_{jk}(D_{j}+D_{k})(k_{jk,\alpha}-2)}{n} =h(Dj+Dk)Ejk,αKXn\displaystyle=\frac{h^{*}(D_{j}+D_{k})E_{jk,\alpha}K_{X}}{n}
=Njk,α1xjk,α+Njk,αyjk,α+Njk,α+1xjk,α+1\displaystyle=N_{jk,\alpha-1}x_{jk,\alpha}+N_{jk,\alpha}y_{jk,\alpha}+N_{jk,\alpha+1}x_{jk,\alpha+1}
+(xjk,α+yjk,α+xjk,α+1)+(kjk,α2)n|D|jk,\displaystyle+(x_{jk,\alpha}+y_{jk,\alpha}+x_{jk,\alpha+1})+\frac{(k_{jk,\alpha}-2)}{n}|D|_{jk},

So, we have explicitly

KX(j<kEjk)2\displaystyle K_{X}\left(\sum_{j<k}E_{jk}\right)^{2} =j<kα=1sjkDjk(Dj+Dk)|D|jknNjk,α(kjk,α2)\displaystyle=\sum_{j<k}\sum_{\alpha=1}^{s_{jk}}\frac{D_{jk}(D_{j}+D_{k})-|D|_{jk}}{n}N_{jk,\alpha}(k_{jk,\alpha}-2)
j<kα=1sjkNjk,α(xjk,α+yjk,α+xjk,α+1).\displaystyle-\sum_{j<k}\sum_{\alpha=1}^{s_{jk}}N_{jk,\alpha}(x_{jk,\alpha}+y_{jk,\alpha}+x_{jk,\alpha+1}).

Theorem 4.18.

If {D1,,Dr}\{D_{1},\ldots,D_{r}\} is an asymptotic arrangement, then

KXn3nc¯13(Z,D),\frac{K_{X_{n}}^{3}}{n}\approx-\bar{c}_{1}^{3}(Z,D),

for prime numbers n0n\gg 0.

Proof.

The first term of the sum contains αNjk,α(kjk2)\sum_{\alpha}N_{jk,\alpha}(k_{jk}-2), which is asymptotic respect to nn by previous discussion (Section 2.5). Thus, we just have to prove asymptoticity for

j<kα=1sjkNjk,α(xjk,α+yjk,α+xjk,α+1)\sum_{j<k}\sum_{\alpha=1}^{s_{jk}}N_{jk,\alpha}(x_{jk,\alpha}+y_{jk,\alpha}+x_{jk,\alpha+1}) (12)

Proceeding as above, it is not difficult to show the following identity,

α=1sjkDjk(Dj+Dk)|D|jkn(kjk,α2)\displaystyle\sum_{\alpha=1}^{s_{jk}}\frac{D_{jk}(D_{j}+D_{k})-|D|_{jk}}{n}(k_{jk,\alpha}-2) =α=1sjk(Njk,α+1)(xjk,α+yjk,α+xjk,α+1)\displaystyle=\sum_{\alpha=1}^{s_{jk}}(N_{jk,\alpha}+1)(x_{jk,\alpha}+y_{jk,\alpha}+x_{jk,\alpha+1})
Njk,1xjk,1Njk,sxjk,s+1.\displaystyle-N_{jk,1}x_{jk,1}-N_{jk,s}x_{jk,s+1}.

So, the asymptoticity of (12) depends only on the asymptoticity of

j<kα=1sjk(xjk,α+yjk,α+xjk,α+1).\sum_{j<k}\sum_{\alpha=1}^{s_{jk}}(x_{jk,\alpha}+y_{jk,\alpha}+x_{jk,\alpha+1}). (13)

By 4.16, xjk,α+yjk,α+xjk,α+1x_{jk,\alpha}+y_{jk,\alpha}+x_{jk,\alpha+1} equals to

kjk,α2n(mjk,α(DjkDk|Dk|j)njk,α(DjkDj|Dj|k)nxjk,1),\frac{k_{jk,\alpha}-2}{n}(m_{jk,\alpha}^{**}(D_{jk}D_{k}-|D_{k}|_{j})-n_{jk,\alpha}^{**}(D_{jk}D_{j}-|D_{j}|_{k})-nx_{jk,1}),

where mjk,α=mjk,α1mjk,α+1mjk,0+mjk,1m_{jk,\alpha}^{**}=m_{jk,\alpha-1}-m_{jk,\alpha+1}-m_{jk,0}+m_{jk,1}, and analogous for njk,αn_{jk,\alpha}^{**}. Observe that these terms are bounded by CnCn for some constant C>0C>0. On the other hand, the terms (DjkDj|Dj|k)(D_{jk}D_{j}-|D_{j}|_{k}) and (DjkDk|Dk|j)(D_{jk}D_{k}-|D_{k}|_{j}) asymptotically depend only on the slopes of coordinates of the chosen lattice points vjklv^{jkl} on each intersection DjklD_{jkl}. By Lemma 4.10, we can choose lattice points with slopes asymptotically bounded by 33 as nn grows, with KXK_{X} having Vjkl=0V_{jkl}=0 for all j<k<lj<k<l. So, we have

|D|jk6lDjkl.|D|_{jk}\leq 6\sum_{l}D_{jkl}.

Thus, as nn grows, all the terms in KX3/nK_{X}^{3}/n vanish except nK3nc¯13(Z,D)nK^{3}\approx-n\bar{c}_{1}^{3}(Z,D). ∎

4.5. Asymptoticity of e(Xn)e(X_{n}).

The topological characteristic can be computed from the topology of (Z,D)(Z,D) and the exceptional divisors Ejk,α,FjklE_{jk,\alpha},F_{jkl}. The divisors Fjkl=pDjklFjkl,pF_{jkl}=\sum_{p\in D_{jkl}}F_{jkl,p} where Fjkl,pF_{jkl,p} is the corresponding exceptional divisor over a pDjklp\in D_{jkl}. Thus, e(Fjkl)=Djkle(Fjkl,p)e(F_{jkl})=D_{jkl}e(F_{jkl,p}). In the toric picture (Section 4.2) of XnX_{n} over pp, let v=vjklv=v^{jkl} the ray generator defining Fjkl,pF_{jkl,p}.

Lemma 4.19.

We have e(Fjkl,p)=sjk+sjl+skl+3e(F_{jkl,p})=s_{jk}+s_{jl}+s_{kl}+3.

Proof.

It is well-known that Fjkl,pF_{jkl,p} is the toric variety associated to the star-fan Star(C(v))(Nv)\mbox{Star}(C(v))\subset(N_{v})_{{\mathbb{R}}}, i.e., the induced fan by the lattice quotient Nv=N/vNN_{v}=N/vN. In this case, the 22-cones of Star(C(v))\mbox{Star}(C(v)) are the induced by each C(ejk,α,ejk,α+1)C(e_{jk,\alpha},e_{jk,\alpha+1}). Since e(Fjkl,p)e(F_{jkl,p}) is the sum of its top-dimensional cones [CLS11, Thm. 12.3.9], we have the result.

The components of divisors Ejk,αE_{jk,\alpha} are determined locally as exterior divisors of the toric picture of XnX_{n}. They intersect FF at the rational curves Cjk,αC_{jk,\alpha}. Locally each component of Ejk,αE_{jk,\alpha} is isomorphic to 𝔸1×w1{\mathbb{A}}^{1}\times{\mathbb{P}}^{1}_{w} over DjkD_{jk}, where w1{\mathbb{P}}^{1}_{w} is a weighted projective line. This follows for the star-fan construction.

Lemma 4.20.

If Ejk,α=CDjkEjk,α,CE_{jk,\alpha}=\sum_{C\in D_{jk}}E_{jk,\alpha,C} is the decomposition in componenets, then we compute

e(Ejk,α)=4CDjk(1pg(C)).e(E_{jk,\alpha})=4\sum_{C\in D_{jk}}(1-p_{g}(C)).
Proof.

We have e(Ejk,α)=Ce(Ejk,α,C)e(E_{jk,\alpha})=\sum_{C}e(E_{jk,\alpha,C}). Since Ejk,α,CE_{jk,\alpha,C} is a fibration over CC with fiber w1{\mathbb{P}}^{1}_{w}. Since the topological characteristics of weighted and non-weighted projective spaces are the same, we have that e(Ejk,α,C)=e(w1)e(C)=4(1pg(C))e(E_{jk,\alpha,C})=e({\mathbb{P}}^{1}_{w})e(C)=4(1-p_{g}(C)). ∎

Lemma 4.21.

If XX is a complex algebraic variety, and AXA\subset X is a subvariety such that XAX\setminus A is smooth, then e(X)=e(XA)+e(A)e(X)=e(X\setminus A)+e(A).

Proof.

See [Ful93, p. 141].∎

Remark 4.22.

The above lemma implies the exclusion-inclusion principle, i.e., for subvarieties V,WXV,W\hookrightarrow X we have e(VW)+e(VW)=e(V)+e(W)e(V\cup W)+e(V\cap W)=e(V)+e(W).

Proposition 4.23.

We have,

e(Xn)\displaystyle e(X_{n}) =n(e(Y)e(D))+e(D)e(Sing(D))\displaystyle=n(e(Y)-e(D))+e(D)-e(\mbox{Sing}(D))
+j<kCDjk[sjk(34pg(C))1]j<k<l(sjk+sjl+skl3)Djkl\displaystyle+\sum_{j<k}\sum_{C\in D_{jk}}[s_{jk}(3-4p_{g}(C))-1]-\sum_{j<k<l}(s_{jk}+s_{jl}+s_{kl}-3)D_{jkl}
Proof.

Denote by RR the ramification divisor of h:XnZh\colon X_{n}\to Z. This is a n:1n:1 morphism which is an isomorphism outside RR we have

e(XR)=ne(YD)=n(e(Y)e(D)).e(X\setminus R)=ne(Y\setminus D)=n(e(Y)-e(D)).

On the other hand, R=jD~jExc(h)R=\bigcup_{j}\tilde{D}_{j}\cup\mbox{Exc}(h), where Exc(h)\mbox{Exc}(h) is the exceptional data of hh. Topologically is given by

Exc(h)=j<k<lFjklj<kαEkj,α.\mbox{Exc}(h)=\bigcup_{j<k<l}F_{jkl}\cup\bigcup_{j<k}\bigcup_{\alpha}E_{kj,\alpha}.

By the exclusion-inclusion observe that

e(jD~j)e(jD~jExc(h))=e(D)e(Sing(D)).e\left(\bigcup_{j}\tilde{D}_{j}\right)-e\left(\bigcup_{j}\tilde{D}_{j}\cap\mbox{Exc}(h)\right)=e(D)-e(\mbox{Sing}(D)).

So, we get

e(R)=e(D)e(Sing(D))+e(Exc(h)).e(R)=e(D)-e(\mbox{Sing}(D))+e(\mbox{Exc}(h)).

On the other hand, the components of Ejk,αEjk,αE_{jk,\alpha}E_{jk,\alpha} are curves over DjkD_{jk} isomorphic to their respective components. Also, each component of Ejk,αFjklE_{jk,\alpha}F_{jkl} is a rational curve over its corresponding point in DjklD_{jkl}. Thus, we have identities,

e(Ejk,αEjk,α+1)=e(Djk)e(E_{jk,\alpha}E_{jk,\alpha+1})=e(D_{jk})
e(Ejk,αFjkl)=2Djkl.e(E_{jk,\alpha}F_{jkl})=2D_{jkl}.

Using repeatedly the exclusion-inclusion principle we we compute

e(Exp(h))=j<kCDjk[sjk(34pg(C))1]j<k<l(sjk+sjl+skl3)Djkl.e(\mbox{Exp}(h))=\sum_{j<k}\sum_{C\in D_{jk}}[s_{jk}(3-4p_{g}(C))-1]-\sum_{j<k<l}(s_{jk}+s_{jl}+s_{kl}-3)D_{jkl}.

Theorem 4.24.

If {D1,,Dr}\{D_{1},\ldots,D_{r}\} is an asymptotic arrangement, then

e(Xn)nc¯3(Z,D),\frac{e(X_{n})}{n}\approx\bar{c}_{3}(Z,D),

for prime numbers n0n\gg 0.

Proof.

As a corollary of the previous identity for e(Xn)e(X_{n}), and by the previous discussion in Section 2.5, the lengths sjk/ns_{jk}/n are asymptotically zero as nn grows. Thus, e(Xn)n(e(Z)e(D))e(X_{n})\approx n(e(Z)-e(D)) as nn grows. ∎

5. Applications to the geography of 3-folds

5.1. The case of 33 hyperplanes.

Consider Z4Z\hookrightarrow{\mathbb{P}}^{4} of degree d=deg(Z)d=\deg(Z). In this case, we have explicitly

KZ=(d5)H|Z,c2(Z)=(10+d(d5))H2|Z,K_{Z}=(d-5)H|_{Z},\quad c_{2}(Z)=(10+d(d-5))H^{2}|_{Z},
c3(Z)=d(d2(d5)+10d10),c_{3}(Z)=-d\left(d^{2}(d-5)+10d-10\right),

for a generic hyperplane section HH. Take 3 hyperplane sections {H1,H2,H3}\{H_{1},H_{2},H_{3}\} in general position, and asymptotic partitions ν1+ν2+ν3=n\nu_{1}+\nu_{2}+\nu_{3}=n. Along D=jνjHjnHD=\sum_{j}\nu_{j}H_{j}\sim nH consider the respective nn-th root cover YnZY_{n}\to Z. Its singularities are over dd points in H1H2H3H_{1}H_{2}H_{3}. As we see in Example 4.7, these singularities admit a locally nef non-singular resolution. Unfortunately, in this resolution the lattice point vv does not satisfy the condition of Lemma 4.10, i.e., the volume KX3K_{X}^{3} will not be completely asymptotic to the logarithmic Chern number c¯13(Z,D)\bar{c}_{1}^{3}(Z,D). However, since the chosen vv satisfy vj=1v_{j}=1. So, following the methods of Theorem 4.18 to compute KX3K_{X}^{3}, we get

xjk,α=Ejk,αEjk,α+1KX=d(d3),1αsjk.x_{jk,\alpha}=E_{jk,\alpha}E_{jk,\alpha+1}K_{X}=d(d-3),\quad 1\leq\alpha\leq s_{jk}.

Now we compute

KX3\displaystyle K_{X}^{3} =d(d3)(nd23nd+3n9d+183j<k(k2)jk),\displaystyle=d(d-3)\left(nd^{2}-3nd+3n-9d+18-3\sum_{j<k}(k-2)_{jk}\right),

since

j<k(Njk,1+Nkj,1)=3n3n,\sum_{j<k}(N_{jk,1}+N_{kj,1})=-3\frac{n-3}{n},
KX(j<kEjk)2=d(d3)j<k(Njk,1+Njk,s+(k2)jk).\displaystyle K_{X}\,\left(\sum_{j<k}E_{jk}\right)^{2}=-d(d-3)\sum_{j<k}(N_{jk,1}+N_{jk,s}+(k-2)_{jk}).

In particular for prime numbers n0n\gg 0,

KX3n(KZ+H1+H2+H3)3d=d(d2)3d.\frac{K_{X}^{3}}{n}\approx(K_{Z}+H_{1}+H_{2}+H_{3})^{3}-d=d(d-2)^{3}-d.

On the other hand, from Section 4.1 we have

χ(𝒪X)=nχ(Z,𝒪Z)112(R1(n)+R2(n)+R3(n)),\chi({\mathcal{O}}_{X})=n\chi(Z,{\mathcal{O}}_{Z})-\frac{1}{12}(R_{1}(n)+R_{2}(n)+R_{3}(n)),

where

χ(Z,𝒪Z)=d(d5)(10+d(d5))24\chi(Z,{\mathcal{O}}_{Z})=-\frac{d(d-5)(10+d(d-5))}{24}
R1(n)=9d(n1)(2n1)2n,R_{1}(n)=\frac{9d(n-1)(2n-1)}{2n},
R2(n)=3d(d5)(n1)(5n1)2n+3d((d5)2+d(d5)+10)(n1)2R_{2}(n)=\frac{3d(d-5)(n-1)(5n-1)}{2n}+\frac{3d((d-5)^{2}+d(d-5)+10)(n-1)}{2}
R3(n)=6d(d2)(d(ν1,ν2,n)+d(ν1,ν3,n)+d(ν2,ν3,n)).R_{3}(n)=6d(d-2)(d(\nu_{1},\nu_{2},n)+d(\nu_{1},\nu_{3},n)+d(\nu_{2},\nu_{3},n)).

Since, the partition is asymptotic, for n0n\gg 0 we have

χ(𝒪X)nd(d2)(d1)224.\frac{\chi({\mathcal{O}}_{X})}{n}\approx-\frac{d(d-2)(d-1)^{2}}{24}.

For n0n\gg 0, the topological characteristic behaves as

e(X)nc3(3,Dred)=d(d5)(d2+2d+6).\frac{e(X)}{n}\approx c_{3}({\mathbb{P}}^{3},D_{red})=-d(d-5)(d^{2}+2d+6).

Following the proof of Theorem 5.2, we get KXK_{X} nef for n0n\gg 0. As a consequence of the above computations, we have.

Theorem 5.1.

For d5d\geq 5 and n0n\gg 0 there are minimal non-singular 33-folds XX of general type having degree nn over ZZ with slopes

c13c1c2(d2)31(d2)(d1)2,c3c1c2(d5)(d2+2d+6)(d2)(d1)2.\frac{c_{1}^{3}}{c_{1}c_{2}}\approx\frac{(d-2)^{3}-1}{(d-2)(d-1)^{2}},\quad\frac{c_{3}}{c_{1}c_{2}}\approx\frac{(d-5)(d^{2}+2d+6)}{(d-2)(d-1)^{2}}.

In particular, as the degree of ZZ grows, the slopes have limit point (1,1)(1,1).

5.2. Hyperplane sections arrangements.

The above partial resolution can be seen as a resolution of pairs

h:(Xn,D~red)(Z,Dred),h\colon(X_{n},\tilde{D}_{red})\to(Z,D_{red}),

where D~\tilde{D} is the inverse direct image of DD. The reduced divisor of DD^{\prime} is an SNC divisor. Indeed, in terms of log\log-resolutions [KM92, p. 5], we can see that our partial resolution has good behavior in logarithmic terms, i.e., they preserves the log\log-structure of the variety nn-th root cover YnY_{n}. The following result illustrates these ideas.

Theorem 5.2.

Let ZZ be a minimal non-singular projective 33-fold, and let {H1\{H_{1}, ,\ldots, Hr}H_{r}\} be a collection of hyperplane sections in general position. Then, for prime numbers n0n\gg 0 there are log\log-morphisms (Xn,D~red)(Z,Dred)(X_{n},\tilde{D}_{red})\to(Z,D_{red}) of degree nn such that:

  1. (1)

    XnX_{n} is of log\log-general type, i.e., KXn+D~redK_{X_{n}}+\tilde{D}_{red} is big and nef,

  2. (2)

    XnX_{n} has cyclic quotient singularities, and so log\log-terminal of order lower than nn, and

  3. (3)

    the slopes (K3/24χ,e/24χ)(-K^{3}/24\chi,e/24\chi) of XnX_{n} are arbitrarily near to (2,1/3)(2,1/3).

Proof.

We take D=j=1rνjHjD=\sum_{j=1}^{r}\nu_{j}H_{j}, where HjH_{j} are hyperplane sections on ZZ and j=1rνj=n\sum_{j=1}^{r}\nu_{j}=n an asymptotic partition. Recall that HjHkHl=Hj2Hk=deg(Z)H_{j}H_{k}H_{l}=H_{j}^{2}H_{k}=\deg(Z) for any j<k<lj<k<l. Take h:XnYnZh\colon X_{n}\to Y_{n}\to Z the asymptotic cyclic resolution constructed in Theorem 4.18. Again for simplicity let us denote KX=KXnK_{X}=K_{X_{n}}. From, the explicit description given in 4.5, we have

KX+Dred=KX+jD~j+j<k,αEjk,α+j<k<lFjkl=h(KZ+Dred).K_{X}+D^{\prime}_{red}=K_{X}+\sum_{j}\tilde{D}_{j}+\sum_{j<k,\alpha}E_{jk,\alpha}+\sum_{j<k<l}F_{jkl}=h^{*}(K_{Z}+D_{red}).

First observe that for any curve CC outside the exceptional data of hh, we have (KX+Dred)C0(K_{X}+D^{\prime}_{red})C\geq 0, by projection formula and since KZ+DredK_{Z}+D_{red} is ample. For every closed curve C=Cjk,αC=C_{jk,\alpha} of C=ClC=C_{l} of the local toric picture (Section 4.2) of the resolution, we have (KX+Dred)C=0(K_{X}+D^{\prime}_{red})\,C=0. For the remainder curves, we just need to concern about the positivity of its intersection with KXK_{X}. Since KZK_{Z} is a nef divisor, by Corollary 2.14 we must have any KXK_{X}-negative curve contained in the support of h(D)h^{*}(D). Thus, the rest of rational curves in Supp(hD)\mbox{Supp}(h^{*}D) are of the following types:

  1. (1)

    Curves defined by the closure of a wall Ejk,α1Ejk,αE_{jk,\alpha-1}E_{jk,\alpha} for 1αsjk.1\leq\alpha\leq s_{jk}.

  2. (2)

    A curve contained in Ejk,αE_{jk,\alpha} but not in Ejk,α±1E_{jk,\alpha\pm 1} for 1αsjk1\leq\alpha\leq s_{jk}.

  3. (3)

    A curve contained in H~j\tilde{H}_{j}.

If CC is of type (1), from 4.12 we have,

(KX+Dred)Ejk,αEjk,α+1=h(KZ+D)Ejk,αEjk,α+1=(KZ+Dred)Hjk>0,(K_{X}+D^{\prime}_{red})E_{jk,\alpha}E_{jk,\alpha+1}=h^{*}(K_{Z}+D)E_{jk,\alpha}E_{jk,\alpha+1}=(K_{Z}+D_{red})H_{jk}>0,

for any α\alpha. If CC is of type (2), then CC must be a fiber of the ruled surface Ejk,αE_{jk,\alpha},i.e., is in the class of Cjk,αC_{jk,\alpha}. But, by (6) we have KXC>0K_{X}C>0. Finally, if CC is of type (3), we assume that it does not intersect interior divisors FjklF_{jkl}. If does it, then by the toric local description CC must be of the form Ejk,1H~jE_{jk,1}\tilde{H}_{j} for some kk. Again by projection formula, we have (KX+Dred)C=(KZ+D)hC0(K_{X}+D^{\prime}_{red})C=(K_{Z}+D)h_{*}C\geq 0. Then, KX+DredK_{X}+D^{\prime}_{red} is a nef divisor, and moreover (KX+Dred)3=(KZ+D)3>0(K_{X}+D^{\prime}_{red})^{3}=(K_{Z}+D)^{3}>0.Thus, by [Laz04, Th. 2.2.16.], the divisor KX+DredK_{X}+D^{\prime}_{red} is big. Now, from Theorem 4.18 we know that for n0,n\gg 0,

KX3n\displaystyle\frac{K_{X}^{3}}{n} c13(Z,D)=(KZ+rH)3\displaystyle\approx-c_{1}^{3}(Z,D)=\left(K_{Z}+rH\right)^{3}
=KZ3+r3deg(Z)+3rKZH2+3KZ2H,\displaystyle=K_{Z}^{3}+r^{3}\deg(Z)+3rK_{Z}H^{2}+3K_{Z}^{2}H,

where HH is a generic hyperplane section on ZZ. Thus, if we choose rr depending on nn with r(n)/n0r(n)/n\to 0 as nn grows, then the numbers |D|jk|D|_{jk} goes to zero respect with nn. Then, we have KX3>0K_{X}^{3}>0. Moreover, from Example 2.9 we have (K3/24χ,e/24χ)(X)(-K^{3}/24\chi,e/24\chi)(X) arbitrarily near to (2,1/3)(2,1/3). ∎

6. Discussion

In this section, we will briefly discuss some possible future paths in order to extend this work.

6.1. Asymptoticity through minimal models

One of the main horizons of this research is to achieve the asymptoticity of invariants through minimal models. This means that the invariants of XnX_{n}, with respect to nn, could be asymptotically equal to the respective invariants of its minimal model. Thus, we will be in a very nice position to do geography, i.e., the study of arrangements of hypersurfaces is identified through the slopes of Chern numbers with a "region" of minimal projective varieties. As we see in Theorem 5.2 and Theorem 5.1, if the basis pair (Z,D)(Z,D) has ZZ minimal of general type and DD composed by ample divisors, then our constructions preserve important features in terms of minimal models. However, this in general is not something easy to work on. For the future of this work, 33 aspects are important.

  1. (1)

    Asymptotic study of (partial) desingularization of cyclic quotient singularities of dimension 3\geq 3.

  2. (2)

    Hirzebruch-Riemann-Roch for singular varieties with terminal and log\log - terminal singularities with their asymptotic analogs. In particular, this requires to establish what invariants will be the correct version of the Chern numbers. As we did in the case of dimension 33.

  3. (3)

    The behavior of the invariants after applying the MMP to our constructed varieties.

In the next section, we discuss (1). If we achieve our goal we will be, able to construct good partial resolutions XnYnX_{n}\to Y_{n}, i.e., the Chern numbers, with respect to nn, are asymptotically equal to the logarithmic Chern numbers of the basis (Z,D)(Z,D). We expect that we can improve the singularities to the terminal ones, so we will be able to run the MMP, i.e. we want to construct a terminal good partial resolution. For (2), we have results of [Rei87] and [BS05] which are a kind of starting point for future work. These contain versions of the Hirzebruch-Riemann-Roch theorem for varieties with canonical and cyclic quotient singularities. For (3), we think that the answer could be hidden in all the massive previous work done around the minimal model program [BCHM10], [KM92]. We expect, that the involved invariants do not suffer dramatic changes after flipping or contractions operations as occur in the case of surfaces. Then, asymptotically with respect to nn, the invariants remain unchanged. We state the above discussion as conjecture.

Conjecture 6.1.

Let XnYn(Z,D)X_{n}\to Y_{n}\to(Z,D) be a terminal good partial resolution of singularities of the nn-th root cover construction. Assume that KYnK_{Y_{n}} is nef, and let XnX^{\prime}_{n} a minimal model of XnX_{n}. Then, for any partition i1++im=di_{1}+\ldots+i_{m}=d we have

ci1cim(Xn)nc¯i1c¯im(Z,D),\frac{c_{i_{1}}\ldots c_{i_{m}}(X^{\prime}_{n})}{n}\approx\bar{c}_{i_{1}}\ldots\bar{c}_{i_{m}}(Z,D),

for prime numbers n0n\gg 0.

6.2. What about the length of resolution of 33-fold c.q.s

Cyclic quotient singularities of dimension 33 can be desingularized using a generalization of the Hirzebruch-Jung algorithm, which is the Fujiki-Oka algorithm. See [Ash19] for a modern treatment. After the local cyclic toric resolution of Section 4.2, instinctively we want to desingularize each one with the Fujiki-Oka process. However, since we want asymptoticy of invariants in our resolutions, we ask for the topological length and the intersection number behavior of such an algorithm. For the first, we mean the amount of new topological data, i.e., how the Betti numbers grow for the chosen resolutions. For last, we mean how the new curves and divisors on the exceptional data affect the volume KX3K_{X}^{3}. As we discussed in Section 2.5, the algorithm in dimension 22 has both aspects behaving as n\sim\sqrt{n} for a suitable class of integer numbers.

Let us assume that we choose a partial resolution for the local cyclic resolution, so the amount of new topological data will behave approximately as

3nj<k,α(vl,ajk,α,bjk,α),\sim 3\sqrt{n}\sum_{j<k,\alpha}\ell(v_{l},a_{jk,\alpha},b_{jk,\alpha}),

where (vl,ajk,α,bjk,α)\ell(v_{l},a_{jk,\alpha},b_{jk,\alpha}) is a length number depending on each cyclic singularities given in Lemma 4.4. Thus, asymptotically respect with nn, we require that (vl,ajk,α,bjk,α)n1/c\ell(v_{l},a_{jk,\alpha},b_{jk,\alpha})\sim n^{1/c} for c<1/2c<1/2. In particular, Fujiki-Oka algorithm for a cyclic quotient singularity of type 1n(a,b,1)\frac{1}{n}(a,b,1) contains the processes for those of dimension two 1n(a,1)\frac{1}{n}(a,1) and 1n(b,1)\frac{1}{n}(b,1). Thus, in the best case, we will have (vl,ajk,α,bjk,α)vl\ell(v_{l},a_{jk,\alpha},b_{jk,\alpha})\sim\sqrt{v_{l}}. To assure asymptoticity in Theorem 4.18 we must have vln/3v_{l}\sim n/3, thus after resolve we lose the asymptoticity on the topological side. On the other hand, if we admit all vlsv_{l}^{\prime}s small as we see in Theorem 5.1, then after resolve we lose the asymptoticity of the volume. These observations lead us to the following problem: the existence of a well-behaved algorithmic terminal resolution for cyclic quotient singularities, i.e., having only terminal singularities.

The terminalization of a toric singularity is a well-known process [CLS11, Sec. 11.4]. Indeed, assume that our toric singularity has associated cone σd\sigma\subset{\mathbb{R}}^{d}. First, we have to compute the convex hull of σd{0}\sigma\cap{\mathbb{Z}}^{d}-\{0\}. This will give us a refinement of σ\sigma, which is a canonical resolution, i.e. having at most canonical singularities with ample canonical bundle. Finally, each canonical toric singularity defined by a cone can be terminalizated by blowing-up each lattice point on the plane generated by the primitive generators of the cone. However, we do not know the growing behavior of this algorithm. In fact, it is known that the best convex-hull algorithm behaves as nlogn\sim n\log n when the number of lattice points is nn [Gre90]. This not seems like a good algorithm to choose.

Question 6.2.

How can we construct a terminal algorithm for cyclic quotient singularities with the desired asymptotic properties? Is it possible?

As we see in Figure 4 and Figure 5, to achieve a well-behaved resolution it is probable that we will have to impose different conditions on the integer aa and bb. In Theorem 5.2 we see that there are 33-folds with cyclic quotient singularities accumulating in the well-known point of the map (2,1/3)(2,1/3). We are curious if after applying the process proposed in Section 6.1, the minimal 33-folds expected will preserve the accumulating point or they move out. Finally, the principal motivation for all this work is the connection between arrangements of hypersurfaces and the geography of invariants of minimal varieties. For us will be interesting to explore the geography through arbitrary arrangements of planes on 3{\mathbb{P}}^{3}.

Question 6.3.

What is the region covered by minimal models of nn-th root cover YnY_{n} along arrangements of planes in 3{\mathbb{P}}^{3}?

References

  • [ABMMOG12] E. Artal-Bartolo, J. Martín-Morales, and J. Ortigas-Galindo. Intersection theory on abelian-quotient V-surfaces and \mathbb{Q}-resolutions. Eleventh international conference Zaragoza-Pau on applied mathematics and statistics, Jaca, Spain, September 15–17, 2010., pages 13–23, 2012.
  • [Ash15] T. Ashikaga. Toric modifications of cyclic orbifolds and an extended Zagier reciprocity for Dedekind sums. Tôhoku Math. J. (2), 67(3):323–347, 2015.
  • [Ash19] T. Ashikaga. Multidimensional continued fractions for cyclic quotient singularities and Dedekind sums. Kyoto J. Math., 59(4):993–1039, 2019.
  • [Bar77] P. Barkan. Sur les sommes de Dedekind et les fractions continues finies. C. R. Acad. Sci. Paris Sér. A-B, 284(16):A923–A926, 1977.
  • [BCHM10] C. Birkar, P. Cascini, C. Hacon, and J. McKernan. Existence of minimal models for varieties of log general type. J. Am. Math. Soc., 23(2):405–468, 2010.
  • [BPS16] F. Bogomolov, A. Pirutka, and A. Silberstein. Families of disjoint divisors on varieties. Eur. J. Math., 2(4):917–928, 2016.
  • [BS05] A. Buckley and B. Szendrői. Orbifold Riemann-Roch for threefolds with an application to Calabi-Yau geometry. J. Algebr. Geom., 14(4):601–622, 2005.
  • [Buc08] Weronika Buczynska. Fake weighted projective spaces. arXiv: math.AG.0805.1211, 2008.
  • [CCJ20] J. Chen, M. Chen, and C. Jiang. The Noether inequality for algebraic threefold (Appendix by Kollar, J.). Math. Ann. 202, Springer-Verlag, 2020.
  • [CH06] M. Chen and C. Hacon. On the geography of Gorenstein minimal 3-folds of general type. Asian J. Math., 10(4):757–763, 2006.
  • [CL01] M. Chang and A. Lopez. A linear bound on the Euler number of threefolds of Calabi-Yau and of general type. Manuscr. Math., 105(1):47–67, 2001.
  • [CLS11] T. Cox, J. Little, and H. Schenck. Toric Varieties, volume 124. AMS, Graduate Studies in Mathematics, 2011.
  • [DS22] R. Du and H. Sun. Inequalities of Chern classes on nonsingular projective nn-folds with ample canonical or anti-canonical line bundles. J. Differ. Geom., 122(3):377–398, 2022.
  • [EFU22] S. Eterović, F. Figueroa, and G. Urzúa. On the geography of line arrangements. Adv. Geom., 22(2):269–276, 2022.
  • [EV92] H. Esnault and E. Viehweg. Lectures on Vanishing Theorems, volume OWS 20. Birkhause Basel, 1992.
  • [Fuj74] A. Fujiki. On resolutions of cyclic quotient singularities. Publ. Res. Inst. Math. Sci., 10:293–328, 1974.
  • [Ful93] W. Fulton. Introduction to toric varieties, volume 131. Princeton University Press, 1993.
  • [Gao11] Y. Gao. A note on finite abelian covers. Sci. China, Math., 54(7):1333–1342, 2011.
  • [Gir03] K. Girstmair. Zones of large and small values for Dedekind sums. Acta Arith., 109(3):299–308, 2003.
  • [Gir06] K. Girstmair. Continued fractions and Dedekind sums: three-term relations and distribution. J. Number Theory, 119(1):66–85, 2006.
  • [GKPT19] D. Greb, S. Kebekus, T. Peternell, and B. Taji. The Miyaoka-Yau inequality and uniformisation of canonical models. Ann. Sci. Éc. Norm. Supér. (4), 52(6):1487–1535, 2019.
  • [Gre90] J. Greenfield. A proof for a quickhull algorithm. Electrical Engineering and Computer Science - Technical Reports. 65., 1990.
  • [Har77] R. Hartshorne. Algebraic geometry, volume GTM 52. Springer, 1977.
  • [Hir83] F. Hirzebruch. Arrangement of lines and algebraic surfaces. Progress in Math. Vol.II, 36:113–240, 1983.
  • [Hol88] R.-P. Holzapfel. Chern number relations for locally abelian Galois coverings of algebraic surfaces. Math. Nachr., 138:263–292, 1988.
  • [Hu18] Y. Hu. Inequality for Gorenstein minimal 3-folds of general type. Commun. Anal. Geom., 26(2):347–359, 2018.
  • [Hun89] B. Hunt. Complex manifolds geography in dimension 2 and 3. J. Differential Geometry, 30:51–153, 1989.
  • [Iit77] S. Iitaka. On logarithmic Kodaira dimension of algebraic varieties. Complex An.and Alg. Geom., pages 175–190, 1977.
  • [Iit78] S. Iitaka. Geometry on complements of lines in 2\mathbb{P}^{2}. Tokyo J. Math., 1:1–19, 1978.
  • [KM92] J. Kollár and S. Mori. Classification of three-dimensional flips. J. Am. Math. Soc., 5(3):533–703, 1992.
  • [Kol91] J. Kollár. Flips, flops, minimal models, etc. Surveys in differential geometry. Vol. I, pages 113–199, 1991.
  • [Laz04] R. Lazarsfeld. Positivity in algebraic geometry. I. Classical setting: line bundles and linear series, volume 48 of Ergeb. Math. Grenzgeb., 3. Folge. Berlin: Springer, 2004.
  • [Mat10] K. Matsuki. Introduction to the Mori program. Universitext. Springer-Verlag, New York, 2010.
  • [Miy87] Y. Miyaoka. The pseudo-effectivity of 3c2c123c_{2}-c^{2}_{1} for varieties with numerically effective canonical classes. Adv. Stud. Pure Math., 10:449––476, 1987.
  • [Par91] R. Pardini. Abelian covers of algebraic varieties. J. Reine Angew. Math., 417:191–213, 1991.
  • [Rei87] M. Reid. Young person’s guide to canonical singularities. Algebraic geometry, Proc. Summer Res. Inst., Brunswick/Maine 1985, part 1, Proc. Symp. Pure Math. 46, 345-414 (1987)., 1987.
  • [Som84] A. Sommese. On the density of ratios of chern numbers of algebraic surfaces. Math. Ann., 268:207–221, 1984.
  • [Sta18] The Stacks Project Authors. Stacks Project. https://stacks.math.columbia.edu, 2018.
  • [TN23] Y. Torres-Nova. On geography of 33-folds via asymptotic behavior of its invariants. Ph.D. Thesis, Pontificia Universidad Católica de Chile, https://repositorio.uc.cl/handle/11534/74443, 2023.
  • [Urz09] G. Urzúa. Algebraic surfaces and arrangements of curves. Journal of Algebraic Geometry, 19:169 – 189, 2009.
  • [Urz16] G. Urzúa. Chern slopes of surfaces of general type in positive characteristic. Duke Math. J., 166:975–1004, 2016.
  • [Vie77] E. Viehweg. Rational singularities of higher dimensional schemes. Proc. Am. Math. Soc., 63:6–8, 1977.
  • [Vie82] E. Viehweg. Vanishing theorems. J. Reine Angew. Math., 335:1–8, 1982.
  • [Zag73] D. Zagier. Higher dimensional Dedekind sums. Math. Ann., 202:149–172, 1973.