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

\newaliascnt

lemmathm \aliascntresetthelemma \newaliascntpropthm \aliascntresettheprop \newaliascntcorthm \aliascntresetthecor \newaliascntwarningthm \aliascntresetthewarning \newaliascntremthm \aliascntresettherem \newaliascntexamplethm \aliascntresettheexample \newaliascntdefinitionthm \aliascntresetthedefinition \newaliascntwarthm \aliascntresetthewar \newaliascntblockthm \aliascntresettheblock \newaliascntconvthm \aliascntresettheconv

Local Newton nondegenerate Weil divisors in toric varieties

András Némethi Alfréd Rényi Institute of Mathematics, ELKH
     Reáltanoda utca 13-15, H-1053, Budapest, Hungary
     ELTE - University of Budapest, Dept. of Geometry, Budapest, Hungary
     BCAM - Basque Center for Applied Math., Mazarredo, 14 E48009 Bilbao, Basque Country, Spain
nemethi.andras@renyi.hu
 and  Baldur Sigurðsson Institute of Mathematics,
     18 Đường Hoàng Quốc Việt, Quận Cầu Giấy, 10307, Hanoi, Vietnam
baldursigurds@gmail.com
Abstract.

We introduce and develop the theory of Newton nondegenerate local Weil divisors (X,0)(X,0) in toric affine varieties. We characterize in terms of the toric combinatorics of the Newton diagram different properties of such singular germs: normality, Gorenstein property, or being an Cartier divisor in the ambient space. We discuss certain properties of their (canonical) resolution X~X\widetilde{X}\to X and the corresponding canonical divisor. We provide combinatorial formulae for the delta–invariant δ(X,0)\delta(X,0) and for the cohomology groups Hi(X~,𝒪X~)H^{i}(\widetilde{X},\mathcal{O}_{\widetilde{X}}) for i>0i>0. In the case dim(X,0)=2\dim(X,0)=2, we provide the (canonical) resolution graph from the Newton diagram and we also prove that if such a Weil divisor is normal and Gorenstein, and the link is a rational homology sphere, then the geometric genus is given by the minimal path cohomology, a topological invariant of the link.

The first author was partially supported by NKFIH Grant “Élvonal (Frontier)” KKP 126683.

1. Introduction

\theblock.

Hypersurface (or complete intersection) germs with nondegenerate Newton principal part constitute a very important family of singularities. They provide a bridge between toric geometry and the combinatorics of polytopes. The computation of their analytic and topological invariants serve as guiding models for the general cases, and also as testing ground for different general conjectures and ideas.

On the other hand, from the point of view of the general classification theorems in algebraic/analytic geometry and singularity theory, these hypersurface germs are rather restrictive. In particular, it is highly desired to extend such germs to a more general setting. Besides the algebraic/analytic motivations there are also several topological ones too: one has to create a flexible family, which is able to follow at analytic level different inductive (cutting and pasting procedures) of the topology. For example, the link of a surface singularity is an oriented plumbed 3–manifold associated with a graph. In inductive proofs and constructions it is very efficient to consider their splice or JSJ decomposition. This would correspond to cutting the Newton diagram by linear planes though their 1–faces, in this way creating non-regular cones as well, as completely general toric 3–folds as ambient spaces for our germs.

The first goal of the present work is to introduce and develop the theory of Weil divisors in general affine toric varieties with additional Newton nondegeneracy condition. By such extensions we wish to cover non–Gorenstein singularities as well, or germs which are not necessarily Cartier divisors in their canonical ambient toric spaces. In the toric presentation two combinatorial/geometrical packages are needed: the fan and geometry of the ambient toric variety, and the ‘dual fan’ (as a subdivision of the previous one) together with the Newton polytope associated with the equations of the Weil divisor.

In fact, we will focus on three level of invariants.

The first level is the analytic geometry of the abstract or embedded singular germs, e.g. normality, or being Gorenstein or isolated singularity, or being Cartier (or {\mathbb{Q}}–Cartier) in the ambient toric variety. Furthermore, at this level we wish to understand/determine several numerical sheaf–cohomological invarints as well.

The second level is the toric combinatorics. In terms of this we wish to characterize the above analytic properties and provide formulae for the numerical invariants.

The third level appears explicitly in the case of curve and surface singularities. In the case of surfaces we construct the resolution graph (as the plumbing graph of the link, hence as a complete topological invariant). It is always a very interesting and difficult task to decide whether the numerical analytic invariants can be recovered from the resolution graph. (This is much harder than the formulae via the toric combinatorics: the Newton polytope preserves considerably more information from the structure of the equations than the resolution graph.) In the last part we prove that the geometric genus of the resolution can be recovered from the graph. This is a new substantial step in a project which aims to provide topological interpretations for sheaf–cohomological invariants, see [23, 25, 21, 22]

\theblock.

Next we provide some additional concrete comments and the detailed presentation of the sections.

After recalling some notation and results from toric geometry, we generalize the notion of a Newton nondegenerate hypersurface in r\mathbb{C}^{r} to an arbitrary Weil divisor in an affine toric variety in section 3. These Newton nondegenerate Weil divisors can be resolved using toric geometry similarly as in the classical case [27], or in a different generalization [4]. In section 4 we consider Newton nondegenerate curves. In section 5 we provide conditions for Newton nondegenerate surface singularities to be isolated, and in section 6 we generalize Oka’s algorithm [27] to construct a resolution of a Newton nondegenerate Weil divisor, along with an explicit description of its resolution graph.

In section 7, we give a formula for the δ\delta-invariant and dimensions of cohomologies of the structure sheaf on a resolution of a Newton nondegenerate germ in terms of the Newton polyhedron, see theorem 7.1, whose statement should have independent interest. In particular, this yields a formula for the geometric genus. In the classical case, this formula was given by Merle and Teissier in [19, Théorème 2.1.1].

In section 8, we give a formula for a canonical divisor on a resolution of a Newton nondegenerate Weil divisor, as well as the canonical cycle in the surface case, in terms of the Newton diagram, see section 8. This formula generalizes results of Oka [27, §9]. In the surface case, we also prove in section 9 that the Gorenstein property is identified by the Newton polyhedron, theorem 9.1. A similar, but weaker, condition implies that the singularity is \mathbb{Q}-Gorenstein, but is not sufficient, as shown by an example in section 9.

Using the above results, and a technical result verfied in section 11, we generalize a previous result [25] for the classical case of Newton nondegenerate hypersurface singularities in 3\mathbb{C}^{3}, namely that the geometric genus is determined by a computation sequence, and is therefore topologically determined:

1.1 Theorem.

Let (X,0)(Y,0)(X,0)\subset(Y,0) be a two-dimensional Newton nondegenerate Weil divisor in the affine toric ambients space YY. Assume that (X,0)(X,0) is normal and Gorenstein, and that its link is a rational homology sphere. Then the geometric genus pg(X,0)p_{g}(X,0) equals the minimal path lattice cohomology associated with the link of (X,0)(X,0). In particular, the geometric genus is determined by the topology of (X,0)(X,0).

2. Toric preliminaries

In this section, we will recall some definitions and statements from toric geometry. For an introduction, see e.g. [13] and [10].

\theblock.

Let NN be a free Abelian group of rank rr\in\mathbb{N} and set M=N=Hom(N,)M=N^{\vee}=\mathop{\rm Hom}\nolimits(N,\mathbb{Z}), as well as M=MM_{\mathbb{R}}=M\otimes\mathbb{R} and N=NN_{\mathbb{R}}=N\otimes\mathbb{R}. If σN\sigma\subset N_{\mathbb{R}} is a cone, the dual cone is defined as

σ={uM|vσ:u,v0}.\sigma^{\vee}=\left\{u\in M_{\mathbb{R}}\,\middle|\,\forall v\in\sigma:\,\langle u,v\rangle\geq 0\right\}.

We also set

σ={uM|vσ:u,v=0}.\sigma^{\perp}=\left\{u\in M_{\mathbb{R}}\,\middle|\,\forall v\in\sigma:\,\langle u,v\rangle=0\right\}.

We will always assume cones to be finitely generated and rational. To a cone σN\sigma\subset N_{\mathbb{R}} we associate the semigroup SσS_{\sigma}, the algebra AσA_{\sigma} and the affine variety UσU_{\sigma} by setting

Sσ=σM,Aσ=[Sσ],Uσ=Spec(Aσ).S_{\sigma}=\sigma^{\vee}\cap M,\quad A_{\sigma}=\mathbb{C}[S_{\sigma}],\quad U_{\sigma}=\mathop{\rm Spec}\nolimits(A_{\sigma}).

A variety of the form UσU_{\sigma} is called an affine toric variety. It has a canonical action of the rr-torus 𝕋r=()r\mathbb{T}^{r}=(\mathbb{C}^{*})^{r}.

\theblock.

A fan \triangle in NN is a collection of cones in NN_{\mathbb{R}} satisfying the following two conditions.

  1. (i)

    Any face of a cone in \triangle is in \triangle.

  2. (ii)

    The intersection of two cones in \triangle is a face of each of them.

The support of a fan \triangle is defined as ||=σσ|\triangle|=\cup_{\sigma\in\triangle}\sigma. If τ,σ\tau,\sigma\in\triangle and τσ\tau\subset\sigma, then we get a morphism UτUσU_{\tau}\to U_{\sigma}. These morphisms form a direct system, whose limit is denoted by YY_{\triangle} and called the associated toric variety. The actions of 𝕋r\mathbb{T}^{r} on the affine varieties UσU_{\sigma} for σ\sigma\in\triangle glue together to form an action on YY_{\triangle}. Note that the canonical maps UσYU_{\sigma}\to Y_{\triangle} are open inclusions (note also that the notation YY_{\triangle} differs from [13]).

Let ~\widetilde{\triangle} be another fan in a lattice N~\widetilde{N} and let ϕ:N~N\phi:\widetilde{N}\to N be a linear map. Assume that for any σ~~\widetilde{\sigma}\in\widetilde{\triangle} there is a σ\sigma\in\triangle so that ϕ(σ~)σ\phi(\widetilde{\sigma})\subset\sigma. This induces maps Uσ~UσYU_{\widetilde{\sigma}}\to U_{\sigma}\to Y_{\triangle}, which glue together to form a map Y~YY_{\widetilde{\triangle}}\to Y_{\triangle\vphantom{\widetilde{\triangle}}}.

\thelemma Lemma (Proposition, p. 39, [13]).

Let ~\widetilde{\triangle} and \triangle be fans as above. The induced map Y~YY_{\widetilde{\triangle}}\to Y_{\triangle\vphantom{\widetilde{\triangle}}} is proper if and only if ϕ1(||)=|~|\phi^{-1}(|\triangle|)=|\widetilde{\triangle}|.

\theblock.

For any pMp\in M, there is an associated rational function on UσU_{\sigma}. These glue together to form a rational function xpx^{p} on YY_{\triangle}. We refer to these functions as monomials. A monomial xpx^{p} is a regular function on YY_{\triangle} if p||=σσp\in|\triangle|^{\vee}=\cap_{\sigma\in\triangle}\sigma^{\vee}. A map ϕ:N~N\phi:\widetilde{N}\to N as above induces ϕ:MM~\phi^{*}:M\to\widetilde{M}. The monomial xpx^{p} on YY_{\triangle} then pulls back to xϕ(p)x^{\phi^{*}(p)}.

\theblock.

For σ\sigma\in\triangle, let OσO_{\sigma} be the closed subset of UσU_{\sigma} defined by the ideal generated by monomials xpx^{p} where p(σσ)Mp\in(\sigma^{\vee}\setminus\sigma^{\perp})\cap M. We identify this set with its image in YY_{\triangle}. The closure of OσO_{\sigma} in YY_{\triangle} is denoted by V(σ)V(\sigma). In the case when σ\sigma is a ray, V(σ)V(\sigma) is a Weil divisor and we write Dσ=V(σ)D_{\sigma}=V(\sigma). The orbits of the 𝕋r\mathbb{T}^{r} action on YY_{\triangle} are precisely the sets OσO_{\sigma} for σ\sigma\in\triangle. Furthermore, we have (as sets)

Uσ=τσOτ,V(σ)=στOτ,Oσ=V(σ)στV(τ).U_{\sigma}=\coprod_{\tau\subset\sigma}O_{\tau},\quad V(\sigma)=\coprod_{\sigma\subset\tau}O_{\tau},\quad O_{\sigma}=V(\sigma)\setminus\bigcup_{\sigma\subsetneq\tau}V(\tau).

Let NσN_{\sigma} be the subgroup of NN generated by σN\sigma\cap N and define

N(σ)=N/Nσ,M(σ)=σM,Mσ=M/M(σ).N(\sigma)=N/N_{\sigma},\quad M(\sigma)=\sigma^{\perp}\cap M,\quad M_{\sigma}=M/M(\sigma).

Note that this way we have MσNσM_{\sigma}^{\phantom{\vee}}\cong N_{\sigma}^{\vee} and M(σ)N(σ)M(\sigma)\cong N(\sigma)^{\vee}. Let πσ:NN(σ)\pi_{\sigma}:N_{\mathbb{R}}\to N_{\mathbb{R}}(\sigma) be the canonical projection and set

Star(σ)={πσ(τ)|στ}.\mathop{\rm Star}\nolimits(\sigma)=\left\{\pi_{\sigma}(\tau)\,\middle|\,\sigma\subset\tau\in\triangle\right\}.

This set is a fan in N(σ)N(\sigma), whose associated toric variety is identified canonically with the orbit closure V(σ)V(\sigma). Similarly, let ϖσ:MMσ\varpi_{\sigma}:M\to M_{\sigma} be the canonical projection. Assuming σ\sigma\in\triangle has dimension ss, we have (Uσ,Oσ)(Yσ×()rs,({0}×()rs)(U_{\sigma},O_{\sigma})\cong(Y_{\sigma}\times(\mathbb{C}^{*})^{r-s},(\{0\}\times(\mathbb{C}^{*})^{r-s}). In particular, OσYO_{\sigma}\subset Y_{\triangle} has YσY_{\sigma} as a transverse type.

\thedefinition Definition.
  • (i)

    For a cone ΣN\Sigma\subset N_{\mathbb{R}}, let Σ\triangle_{\Sigma} denote the fanfan consisting of all the faces of Σ\Sigma. We also write YΣY_{\Sigma} instead of YΣY_{\triangle_{\Sigma}}.

  • (ii)

    If \triangle is a fan and ii\in\mathbb{N}, define

    (i)={σ|dimσ=i}.\triangle^{(i)}=\left\{\sigma\in\triangle\,\middle|\,\dim\sigma=i\right\}.
  • (iii)

    A regular cone (resp. simplicial cone) is a cone generated by a subset of an integral (resp. rational) basis of NN.

  • (iv)

    A subdivision of a fan \triangle is a fan ~\widetilde{\triangle} so that |~|=|||\widetilde{\triangle}|=|\triangle| and each cone in \triangle is a union of cones in ~\widetilde{\triangle}. A regular subdivision is a subdivision consisting of regular cones.

  • (v)

    If ΣN\Sigma\subset N_{\mathbb{R}} is a cone and \triangle is a subdivision of Σ\triangle_{\Sigma}, denote by \triangle^{*} the fan consisting of σ\sigma\in\triangle for which σΣ\sigma\subset\partial\Sigma. Here we see Σ\partial\Sigma as the union of the proper faces of Σ\Sigma. As a result, \triangle^{*} is a subdivision of the fan Σ{Σ}\triangle_{\Sigma}\setminus\{\Sigma\}.

  • (vi)

    Let 1,2\triangle_{1},\triangle_{2} be subdivisions of a fan \triangle. We say that 2\triangle_{2} refines 1\triangle_{1} if 2\triangle_{2} is a subdivision of 1\triangle_{1}, or that 2\triangle_{2} is a refinement of 1\triangle_{1}.

  • (vii)

    Let \triangle be a fan with a subdivision 1\triangle_{1} and let σ\sigma\in\triangle. The restriction of 1\triangle_{1} to σ\sigma is defined as

    1|σ={τ1|τσ}.\triangle_{1}|_{\sigma}=\left\{\tau\in\triangle_{1}\,\middle|\,\tau\subset\sigma\right\}.

3. Analytic Weil divisors in affine toric varieties

\theblock.

Throughout this section, as well as the following sections, we will assume that NN has rank rr and that Σ\Sigma is an rr-dimensional, rational, finitely generated, strictly convex cone in NN_{\mathbb{R}}. This means that ΣN\Sigma\subset N_{\mathbb{R}} is generated over 0\mathbb{R}_{\geq 0} by a finite set of elements from NN, which generate NN as a vectorspace, and that Σ={0}\Sigma^{\perp}=\{0\}. In particular, the orbit OΣO_{\Sigma} consists of a single point, which we denote by 0, and refer to as the origin. Let YΣY_{\Sigma} be the affine toric variety associated with Σ\Sigma.

Any subdivision \triangle of Σ\triangle_{\Sigma} induces a modification π:YYΣ\pi_{\triangle}:Y_{\triangle}\to Y_{\Sigma}.

In the sequel we denote by (YΣ,0)(Y_{\Sigma},0) the analytic germ of YΣY_{\Sigma} at 0, and usually we will denote by YY a (small Stein) representative of (YΣ,0)(Y_{\Sigma},0). (Hence (Y,0)=(YΣ,0)(Y,0)=(Y_{\Sigma},0).) If πΔ\pi_{\Delta} is a toric modification, in the discussions regarding the local analytic germ (Y,0)(Y,0), we will use the same notation YΔY_{\Delta} for πΔ1(Y)\pi_{\Delta}^{-1}(Y) and DσD_{\sigma} for DσπΔ1(Y)D_{\sigma}\cap\pi_{\Delta}^{-1}(Y). Similarly, OσO_{\sigma} might stay for OσYYO_{\sigma}\cap Y\subset Y as well. If in some argument we really wish to stress the differences, we write YΔlocY^{loc}_{\Delta}, DσlocD^{loc}_{\sigma}, OσlocO^{loc}_{\sigma} for the local objects.

Assume that f𝒪Y,0f\in\mathcal{O}_{Y,0} is the germ of a holomorphic function at the origin, which has an expansion

(3.1) f(x)=pSΣapxp,ap.f(x)=\sum_{p\in S_{\Sigma}}a_{p}x^{p},\quad a_{p}\in\mathbb{C}.

Then ({f=0},0)(Y,0)(\{f=0\},0)\subset(Y,0) is the germ of an analytic space. We set supp(f)={pSΣ|ap0}\mathop{\rm supp}\nolimits(f)=\left\{p\in S_{\Sigma}\,\middle|\,a_{p}\neq 0\right\} too.

\thedefinition Definition.

The Newton polyhedron of ff with respect to Σ\Sigma is the polyhedron

Γ+(f)=conv(supp(f)+Σ),\Gamma_{+}(f)=\mathop{\rm conv}\nolimits(\mathop{\rm supp}\nolimits(f)+\Sigma^{\vee}),

where conv\mathop{\rm conv}\nolimits denotes the convex closure in MM_{\mathbb{R}}. The union of compact faces of Γ+(f)\Gamma_{+}(f) is denoted by Γ(f)\Gamma(f) and is called the Newton diagram of ff with respect to Σ\Sigma.

\theblock.

The fan f\triangle_{f} and some combinatorial properties. It follows from definition that Σ\Sigma is precisely the set of those linear functions on MM_{\mathbb{R}} having a minimal value on Γ+(f)\Gamma_{+}(f). Denote by F()F(\ell) the minimal set of Σ\ell\in\Sigma on Γ+(f)\Gamma_{+}(f). For 1,2Σ\ell_{1},\ell_{2}\in\Sigma, say that 12\ell_{1}\sim\ell_{2} if and only if F(1)=F(2)F(\ell_{1})=F(\ell_{2}). Then \sim is an equivalence relation on Σ\Sigma having finitely many equivalence classes, each of whose closure is a finitely generated rational strictly convex cone. These cones form a fan, which we will denote by f\triangle_{f}. We refer to f\triangle_{f} as the dual fan associated with ff and Σ\Sigma. Note that f\triangle_{f} refines Σ\triangle_{\Sigma}.

For any σf\sigma\in\triangle_{f}, the face F()F(\ell) is independent of the choice of σ\ell\in\sigma^{\circ}, where σσ\sigma^{\circ}\subset\sigma is the relative interior, that is, the topological interior of σ\sigma as a subset of Nσ,N_{\sigma,\mathbb{R}}. For σf(1)\sigma\in\triangle_{f}^{(1)}, the set σN\sigma\cap N is a semigroup generated by a unique element, which we denote by σ\ell_{\sigma}. For a series

g𝒪Y,0[xM]={xph|pM,h𝒪Y,0},g\in\mathcal{O}_{Y,0}[x^{M}]=\left\{x^{p}h\,\middle|\,p\in M,\,h\in\mathcal{O}_{Y,0}\right\},

the support supp(g)\mathop{\rm supp}\nolimits(g) is defined similarly as above, and for σf(1)\sigma\in\triangle_{f}^{(1)} we set

wtσ(g)=min{σ(p)|psupp(g)}.\mathop{\rm wt}\nolimits_{\sigma}(g)=\min\left\{\ell_{\sigma}(p)\,\middle|\,p\in\mathop{\rm supp}\nolimits(g)\right\}.

One verifies that for any such gg

(3.2) the vanishing order of g along DσYf is exactly wtσ(g).\mbox{the vanishing order of $g$ along $D_{\sigma}\subset Y_{\triangle_{f}}$ is exactly $\mathop{\rm wt}\nolimits_{\sigma}(g)$}.
\thedefinition Definition.

Let σf\sigma\in\triangle_{f} and σ\ell\in\sigma^{\circ}. Define

Fσ=F(),fσ=pFσapxp.F_{\sigma}=F(\ell),\quad f_{\sigma}=\sum_{p\in F_{\sigma}}a_{p}x^{p}.

If σN\sigma^{\prime}\subset N_{\mathbb{R}} is a cone, and σσ\sigma^{\prime\circ}\subset\sigma^{\circ} (for example, if σ\sigma^{\prime} is an element of a refinement of f\triangle_{f}), then we set Fσ=FσF_{\sigma^{\prime}}=F_{\sigma}.

If σΣ\sigma\subset\Sigma is one dimensional, set mσ=wtσ(f)m_{\sigma}=\mathop{\rm wt}\nolimits_{\sigma}(f). Thus, σ|Fσmσ\ell_{\sigma}|_{F_{\sigma}}\equiv m_{\sigma}. Note that we have

Γ+(f)={uM|σf(1):σ(u)mσ}.\Gamma_{+}(f)=\left\{u\in M_{\mathbb{R}}\,\middle|\,\forall\sigma\in\triangle_{f}^{(1)}:\,\ell_{\sigma}(u)\geq m_{\sigma}\right\}.

This can be compared with the following set.

\thedefinition Definition.

Let

Γ+(f)={uM|σf(1):σ(u)mσ},\Gamma_{+}^{*}(f)=\left\{u\in M_{\mathbb{R}}\,\middle|\,\forall\sigma\in\triangle_{f}^{*(1)}:\,\ell_{\sigma}(u)\geq m_{\sigma}\right\},

where f(1)\triangle_{f}^{*(1)} denotes the set of rays in f\triangle_{f} contained in the boundary of Σ\Sigma.

\thedefinition Definition.

Denote by (X,0)(Y,0)(X,0)\subset(Y,0) the union of those local primary components of the germ defined by ff (with their non-reduced structure), which are not invariant by the torus action. If ff is reduced along the non-invariant components, this means the following. Let UYU\subset Y be a neighbourhood of the origin on which ff converges and let XUX^{\prime}\subset U be defined by f=0f=0. Then XX is the closure of X{Dσ|σΣ(1)}X^{\prime}\setminus\cup\left\{D_{\sigma}\,\middle|\,\sigma\in\triangle_{\Sigma}^{(1)}\right\} in UU.

\therem Remark.

(i) For any pMp\in M, the function xpfx^{p}f defines the same germ (X,0)(X,0). Thus, we may allow f𝒪Y,0[xM]={xpg|pM,g𝒪Y,0}f\in\mathcal{O}_{Y,0}[x^{M}]=\left\{x^{p}g\,\middle|\,p\in M,\,g\in\mathcal{O}_{Y,0}\right\} as well.

(ii) Since the divisors {Dσ:σΣ(1)}\{D_{\sigma}\,:\,\sigma\in\triangle_{\Sigma}^{(1)}\} are torus-invariant, the divisor of ff in YΣY_{\Sigma} is X+σmσDσX+\sum_{\sigma}m_{\sigma}D_{\sigma}.

\theprop Proposition.
  • (i)

    We have Γ+(f)=p+Σ\Gamma_{+}(f)=p+\Sigma^{\vee} for some pMp\in M if and only if f=Σ\triangle_{f}=\triangle_{\Sigma} if and only if the germ XX at 0 is the empty germ.

  • (ii)

    For a σΣ\sigma\in\triangle_{\Sigma}, we have OσXO_{\sigma}\subset X if and only if the normal fan f\triangle_{f} subdivides σ\sigma into smaller cones, i.e. f|σσ\triangle_{f}|_{\sigma}\neq\triangle_{\sigma}.

  • (iii)

    The ideal IX𝒪Y,0I_{X}\subset\mathcal{O}_{Y,0} which defines (X,0)(X,0) in (Y,0)(Y,0), is generated by the functions xpfx^{p}f for pMp\in M satisfying σ(p)+mσ0\ell_{\sigma}(p)+m_{\sigma}\geq 0 for all σΣ(1)\sigma\in\triangle^{(1)}_{\Sigma}.

Proof.

Statement i is clear, since Γ+(f)\Gamma_{+}(f) is of the form p+Σp+\Sigma^{\vee} if and only if ff is a product of a monomial and a unit in 𝒪Y,0\mathcal{O}_{Y,0}.

Statement ii follows from i, and the fact that the intersection of XX and a generic transverse space YσY_{\sigma} to OσO_{\sigma} has Newton polygon ϖσ(Γ(f))\varpi_{\sigma}(\Gamma(f)), cf. section 2.

iii Assuming the given conditions on pp, the function xpfx^{p}f is meromorphic and has no poles. Since YY is normal, xpfx^{p}f is analytic and vanishes on XX. As a result, xpfIXx^{p}f\in I_{X}.

To show that these generate IXI_{X}, take gIXg\in I_{X}. We must show that g=hfg=hf, with h𝒪Y,0[xM]h\in\mathcal{O}_{Y,0}[x^{M}] and σ(p)+mσ0\ell_{\sigma}(p)+m_{\sigma}\geq 0 for psupp(h)p\in\mathop{\rm supp}\nolimits(h).

Let IX,MI_{X,M} be the localization of IXI_{X} along the invariant divisors, that is, the ideal of meromorphic function germs on (Y,0)(Y,0), regular on the open torus and vanishing on XX. It follows that IX,M=f𝒪Y,0[xM]I_{X,M}=f\cdot\mathcal{O}_{Y,0}[x^{M}] and IX=IX,M𝒪Y,0I_{X}=I_{X,M}\cap\mathcal{O}_{Y,0}.

Thus, g=xrhfg=x^{r}hf for some h𝒪Y,0h\in\mathcal{O}_{Y,0} and rMr\in M. Then, there exist finite families (hi)i(h_{i})_{i} of units in 𝒪Y,0\mathcal{O}_{Y,0} and exponents (pi)i(p_{i})_{i} in MM so that xrh=ixpihix^{r}h=\sum_{i}x^{p_{i}}h_{i} and the support of xrhx^{r}h is the disjoint union of the supports of xpihix^{p_{i}}h_{i}. Let us take any σΣ(1)\sigma\in\triangle^{(1)}_{\Sigma}. The condition on disjointness of supports gives

miniwtσxpihif=wtσxrhf=wtσg0.\min_{i}\mathop{\rm wt}\nolimits_{\sigma}x^{p_{i}}h_{i}f=\mathop{\rm wt}\nolimits_{\sigma}x^{r}hf=\mathop{\rm wt}\nolimits_{\sigma}g\geq 0.

As a result, we have σ(pi)+mσ0\ell_{\sigma}(p_{i})+m_{\sigma}\geq 0 for all ii. The result follows. ∎

\thedefinition Definition.

Let ff and f\triangle_{f} be as above. We say that Γ+(f)\Gamma_{+}(f), or ff, is (\mathbb{Q}-)pointed if there exists a pMp\in M (pM)p\in M_{\mathbb{Q}}) such that σ(p)=mσ\ell_{\sigma}(p)=m_{\sigma} for all σΣ(1)\sigma\in\triangle_{\Sigma}^{(1)}.

\theprop Proposition.
  • (i)

    If Σ\Sigma is regular (resp. simplicial), then any Newton polyhedron (w.r.t. Σ\Sigma) is pointed at some pMp\in M (resp. pMp\in M_{\mathbb{Q}}).

  • (ii)

    ff is pointed at pMp\in M if and only if (X,0)(X,0) in (Y,0)(Y,0) is defined by a single equation xpfx^{-p}f (cf. section 3). In other words, ff is pointed if and only if (X,0)(X,0) is a Cartier divisor in (Y,0)(Y,0).

  • (iii)

    ff is pointed at pMp\in M_{\mathbb{Q}} if and only if (X,0)(X,0) is a \mathbb{Q}-Cartier divisor in (Y,0)(Y,0).

Proof.

(i) Use the fact that {σ:σΣ(1)}\{\ell_{\sigma}\,:\,\sigma\in\triangle_{\Sigma}^{(1)}\} is an integral (resp. rational) basis.

(ii) If ff is pointed at pMp\in M then by section 3, xpfIXx^{-p}f\in I_{X}. Moreover, if xqfIXx^{-q}f\in I_{X} for some qMq\in M, then σ(pq)0\ell_{\sigma}(p-q)\geq 0 for any σΣ(1)\sigma\in\triangle_{\Sigma}^{(1)}, hence pqΣMp-q\in\Sigma\cap M and xpq𝒪Y,0x^{p-q}\in\mathcal{O}_{Y,0}.

Conversely, assume that (X,0)(Y,0)(X,0)\subset(Y,0) is an (analytic) Cartier divisor. Let ~f\widetilde{\triangle}_{f} be a smooth subdivision of f\triangle_{f}, and set Y~=Y~f\widetilde{Y}=Y_{\widetilde{\triangle}_{f}}. This is a smooth variety, and the map π:Y~YΣ\pi:\widetilde{Y}\to Y_{\Sigma} is a resolution of YY. Take a small Stein representative YlocYY^{\rm loc}\subset Y, and set Y~loc=π1(Yloc)\widetilde{Y}^{\rm loc}=\pi^{-1}(Y^{\rm loc}). Then we have the vanishing H1(Y~,𝒪Y~)=0H^{\geq 1}(\widetilde{Y},\mathcal{O}_{\widetilde{Y}})=0 (see e.g. [13, Corrollary, p. 74] or [10, §8.5]), and also its local analogue H1(Y~loc,𝒪Y~loc)=0H^{\geq 1}(\widetilde{Y}^{\rm loc},\mathcal{O}_{\widetilde{Y}^{\rm loc}})=0 (since the local analytic germ (Y,0)(Y,0) is rational too). Thus, from the exponential exact sequence, Pic(Y~)=H2(Y~,)\mathop{\rm Pic}\nolimits(\widetilde{Y})=H^{2}(\widetilde{Y},\mathbb{Z}) and Pic(Y~loc)=H2(Y~loc,)\mathop{\rm Pic}\nolimits(\widetilde{Y}^{\rm loc})=H^{2}(\widetilde{Y}^{\rm loc},\mathbb{Z}). On the other hand, YY is weighted homogeneous (as any affine toric variety), hence H2(Y~,)=H2(Y~loc,)H^{2}(\widetilde{Y},\mathbb{Z})=H^{2}(\widetilde{Y}^{\rm loc},\mathbb{Z}). In particular, Pic(Y~)Pic(Y~loc)\mathop{\rm Pic}\nolimits(\widetilde{Y})\cong\mathop{\rm Pic}\nolimits(\widetilde{Y}^{\rm loc}). Here the first group is the Picard group of the algebraic variety, while the second one is the Picard group of the analytic manifold.

Next, consider the Chow group Ar1(Y)A_{r-1}(Y) of codimension one, i.e. the group freely generated by Weil divisors, modulo linear equivalence. Note that since Y~\widetilde{Y} is smooth, we have Ar1(Y~)Pic(Y~)A_{r-1}(\widetilde{Y})\cong\mathop{\rm Pic}\nolimits(\widetilde{Y}) and Ar1(Y~loc)Pic(Y~loc)A_{r-1}(\widetilde{Y}^{\rm loc})\cong\mathop{\rm Pic}\nolimits(\widetilde{Y}^{\rm loc}). If we factor these isomorphic groups by the subgroups generated by the exceptional divisors, we find that the restriction induces an isomorphism Ar1(Y)Ar1(Yloc)A_{r-1}(Y)\cong A_{r-1}(Y^{\rm loc}).

Denote by DσlocD^{\rm loc}_{\sigma} the restriction image of DσD_{\sigma} under the above isomorphism. Since (X,0)(Y,0)(X,0)\subset(Y,0) is local analytic Cartier, and the local divisor of ff in YY is X+DflocX+D^{\rm loc}_{f}, where Dfloc=σΣ(1)mσDσlocD^{\rm loc}_{f}=\sum_{\sigma\in\triangle_{\Sigma}^{(1)}}m_{\sigma}D^{\rm loc}_{\sigma}, we find that the class of DflocD^{\rm loc}_{f} is zero in Ar1(Yloc)A_{r-1}(Y^{\rm loc}). But then, by the above isomorphisms, the class of Df=σΣ(1)mσDσD_{f}=\sum_{\sigma\in\triangle_{\Sigma}^{(1)}}m_{\sigma}D_{\sigma} is zero in Ar1(Y)A_{r-1}(Y).

Finally note that Ar1(Y)A_{r-1}(Y) can be computed as follows [13, 3.4]. Consider the group Div𝕋(Y)=Dσ|σΣ\mathop{\rm Div}\nolimits_{\mathbb{T}}(Y)=\mathbb{Z}\left\langle D_{\sigma}\,\middle|\,\sigma\in\triangle_{\Sigma}\right\rangle of invariant divisors and the inclusion MDiv𝕋(Y)M\hookrightarrow\mathop{\rm Div}\nolimits_{\mathbb{T}}(Y) sending pMp\in M to σσ(p)Dσ\sum_{\sigma}\ell_{\sigma}(p)D_{\sigma}. Along with the map Div𝕋Ar1(Y)\mathop{\rm Div}\nolimits_{\mathbb{T}}\to A_{r-1}(Y), this gives a short exact sequence

0MDiv𝕋(Y)Ar1(Y)0.0\to M\to\mathop{\rm Div}\nolimits_{\mathbb{T}}(Y)\to A_{r-1}(Y)\to 0.

Since DfAr1(Y)D_{f}\in A_{r-1}(Y) maps to zero in Ar1(Yloc)A_{r-1}(Y^{\rm loc}) under the above isomorphism, and DfDiv𝕋(Y)D_{f}\in\mathop{\rm Div}\nolimits_{\mathbb{T}}(Y), we find that DfD_{f} is in the image of MM. But this means exactly that there exists pMp\in M such that σ(p)=mσ\ell_{\sigma}(p)=m_{\sigma} for all σΣ(1)\sigma\in\triangle^{(1)}_{\Sigma}.

(iii) Use part (ii) for a certain power of ff. ∎

\thedefinition Definition.

We say that ff has Newton nondegenerate principal part with respect to Σ\Sigma (or simply that ff or (X,0)(X,0) is Newton nondegenerate) if for every σf\sigma\in\triangle_{f} with FσF_{\sigma} compact, the variety Spec([M]/(fσ))\mathop{\rm Spec}\nolimits(\mathbb{C}[M]/(f_{\sigma})) (that is, {x𝕋r|fσ(x)=0}\left\{x\in\mathbb{T}^{r}\,\middle|\,f_{\sigma}(x)=0\right\} with its non-reduced structure) is smooth. Note that fσf_{\sigma} is a polynomial since FσF_{\sigma} is compact.

\thelemma Lemma.

Assume that (X,0)(Y,0)(X,0)\subset(Y,0) is Newton nondegenerate and let σΣ\sigma\in\triangle_{\Sigma}. If OσXO_{\sigma}\subset X, then the generic transverse type of XX to OσO_{\sigma} is a Newton nondegenerate singularity with Newton polyhedron ϖσ(Γ+(f))Mσ\varpi_{\sigma}(\Gamma_{+}(f))\subset M_{\sigma}.

Proof.

The statement follows by restricting ff to a toric subspace transverse to OσO_{\sigma}, see section 2. ∎

\theblock.

The fan ~f\widetilde{\triangle}_{f} and the associated resolution. Assume that ff is Newton nondegenerate. Let ~f\widetilde{\triangle}_{f} be a regular subdivision of f\triangle_{f}. Then Y~=Y~f\widetilde{Y}=Y_{\widetilde{\triangle}_{f}} is a smooth variety, and we have a modification π:Y~Y\pi:\widetilde{Y}\to Y. As a result of the nondegeneracy of ff, the strict transform X~\widetilde{X} of XX in Y~\widetilde{Y} intersects all orbits in Y~\widetilde{Y} smoothly. In particular, X~\widetilde{X} is smooth, and π\pi is an embedded resolution of (X,0)(Y,0)(X,0)\subset(Y,0).

\thelemma Lemma.

Assume (X,0)(Y,0)(X,0)\subset(Y,0) is a Newton nondegenerate Weil divisor. Then, the singular locus of the germ (X,0)(X,0) is contained in the union of codimension 2\geq 2 orbits in (Y,0)(Y,0).

Proof.

Let Y(r2)Y^{(\leq r-2)} be the union of orbits of dimension r2\leq r-2, that is, codimension 2\geq 2, in YY. Let π:Y~Y\pi:\widetilde{Y}\to Y be as in section 3. The restriction π1(YY(r2))YY(r2)\pi^{-1}(Y\setminus Y^{(\leq r-2)})\to Y\setminus Y^{(\leq r-2)} is an isomorphism, and X~\widetilde{X} is smooth. Therefore, XY(r2)X\setminus Y^{(\leq r-2)} is smooth. ∎

4. Newton nondegenerate curve singularities

In this section, we will assume that rkN=2\mathop{\mathrm{rk}}N=2 and that ΣN\Sigma\subset N_{\mathbb{R}} is a two dimensional finitely generated strictly convex rational cone. Nondegenerate rank 2 singularities appear naturally in the r=3r=3 case as transversal types of certain orbits.

We will introduce the canonical subdivision and we establish criterions for irreducibility and smoothness. They will be used in the context of rank r=3r=3 cones in the definition of their canonical subdivision and in the characterization of Newton nondegenerate isolated surface singularities.

\theblock.

Canonical primitive sequence. Assume first that Σ\Sigma is nonregular. Then there exists a sequence of vectors 0,,s+1ΣN\ell_{0},\ldots,\ell_{s+1}\in\Sigma\cap N, called the canonical primitive sequence [27] and integers b1,,bs2b_{1},\ldots,b_{s}\geq 2, called the associated selfintersection numbers, so that:

  1. (i)

    If 0js0\leq j\leq s, then j,j+1\ell_{j},\ell_{j+1} form an integral basis for NN.

  2. (ii)

    If 0<js0<j\leq s, then bjj=j1+j+1b_{j}\ell_{j}=\ell_{j-1}+\ell_{j+1}.

  3. (iii)

    The set {0,,s+1}\{\ell_{0},\ldots,\ell_{s+1}\} is a minimal set of generators for the semigroup ΣN\Sigma\cap N.

This data is uniquely determined up to reversing the order of (j)j(\ell_{j})_{j} and (bj)j(b_{j})_{j}. It can, in fact, be determined as follows. Let α\alpha be the absolute value of the determinant of the 2×22\times 2 matrix whose columns ,\ell,\ell^{\prime} are the primitive generators of the one dimensional faces of Σ\Sigma, given in any integral basis. Then, there exists a unique integer 0β<α0\leq\beta<\alpha so that β+αN\beta\ell+\ell^{\prime}\in\alpha N. The selfintersection numbers are determined as the negative continued fraction expansion

αβ=b11b21bs.\frac{\alpha}{\beta}=b_{1}-\frac{1}{b_{2}-\ddots-\frac{1}{b_{s}}}.

We use the notation [b1,,bs][b_{1},\ldots,b_{s}] for the right hand side above. We have

0=,1=β+α.\ell_{0}=\ell,\quad\ell_{1}=\frac{\beta\ell+\ell^{\prime}}{\alpha}.

Along with condition (ii), this determines the canonical primitive sequence recursively and we have s+1=\ell_{s+1}=\ell^{\prime}.

Refer to caption
4=(5,3)\ell_{4}=(5,3)2=(3,2)\ell_{2}=(3,2)1=(1,1)\ell_{1}=(1,1)0=(0,1)\ell_{0}=(0,1)
Figure 1. In this example, Σ\Sigma is generated by (0,1)(0,1) and (5,3)(5,3). The canonical primitive sequence consists of four elements, including the generators of the cone.

Alternatively, the vectors 0,1,,s+1\ell_{0},\ell_{1},\ldots,\ell_{s+1} are the integral points lying on compact faces of the convex closure of the set ΣN{0}\Sigma\cap N\setminus\{0\}. For a detailed discussion of this construction, see [26, 1.6].

If Σ\Sigma is regular, then we prefer to modify the minimality of the resolution considered above, and set s=1s=1, 1=\ell_{1}=\ell and 2=\ell_{2}=\ell^{\prime} and 1=0+2\ell_{1}=\ell_{0}+\ell_{2}. Accordingly, in item (ii), we will have b1=1-b_{1}=-1. In particular, the set {0,1,2}\{\ell_{0},\ell_{1},\ell_{2}\} is not a minimal set of generators of the semigroup ΣN\Sigma\cap N. We make this choice here mostly for technical reasons (directed by properties of the induced reslution), which will appear in section 10. The same choice is made in [27], Definition (3.5).

\thedefinition Definition.

Let Σ\Sigma be a two dimensional rational strictly convex cone with a canonical primitive sequence 0,1,,s+1\ell_{0},\ell_{1},\ldots,\ell_{s+1}. The canonical subdivision of Σ\triangle_{\Sigma} is the unique subdivision ~Σ\widetilde{\triangle}_{\Sigma} for which

~Σ(1)={0i| 0is+1}.\widetilde{\triangle}_{\Sigma}^{(1)}=\left\{\mathbb{R}_{\geq 0}\langle\ell_{i}\rangle\,\middle|\,0\leq i\leq s+1\right\}.

For each i=1,,si=1,\ldots,s, there is a unique number bi1-b_{i}\in\mathbb{Z}_{\leq-1} satisfying i1bii+i+1=0\ell_{i-1}-b_{i}\ell_{i}+\ell_{i+1}=0. We define α(0,s+1)\alpha(\ell_{0},\ell_{s+1}) and β(0,s+1)\beta(\ell_{0},\ell_{s+1}) as the numerator and denominator, respectively, of the negative continued fraction

[b1,,bs]=b11b21,[b_{1},\ldots,b_{s}]=b_{1}-\frac{1}{b_{2}-\frac{1}{\cdots}},

(we require gcd(α(0,s+1),β(0,s+1))=1\gcd(\alpha(\ell_{0},\ell_{s+1}),\beta(\ell_{0},\ell_{s+1}))=1, and β(0,s+1)0\beta(\ell_{0},\ell_{s+1})\geq 0, so that these numbers are well defined). The number α(0,s+1)\alpha(\ell_{0},\ell_{s+1}) is referred to as the determinant of Σ\Sigma.

\therem Remark.

Let 1,2N\ell_{1},\ell_{2}\in N be two linearly independent elements. Then we have α(1,2)=1\alpha(\ell_{1},\ell_{2})=1 if and only if 1,2\ell_{1},\ell_{2} form part of an integral basis of NN. In general, α=α(1,2)\alpha=\alpha(\ell_{1},\ell_{2}) can be computed as the content of the restriction of 2\ell_{2} to the kernel of 1\ell_{1}. In other words, let KNK\subset N be the kernel of 1\ell_{1}. Then 2|K\ell_{2}|_{K} is divisible by α\alpha, and (2|K)/α(\ell_{2}|_{K})/\alpha is primitive.

\thelemma Lemma.

If Σ\Sigma is not a regular cone, then YΣY_{\Sigma} has a cyclic quotient singularity at the origin and the map Y~ΣYΣY_{\widetilde{\triangle}_{\Sigma}}\to Y_{\Sigma} induced by the identity map on NN is the minimal resolution.

Proof.

See Proposition 1.19 and Proposition 1.24 of [26]. ∎

\theprop Proposition.

Assume that rkN=2\mathop{\mathrm{rk}}N=2, and that ff is Newton nondegenerate with respect to ΣN\Sigma\subset N_{\mathbb{R}} defining a germ (X,0)(X,0).

  1. (i)

    The germ (X,0)(X,0) is irreducible if and only if Γ(f)\Gamma(f) is a single interval with no integral interior points. In fact, in general, the number of components in (X,0)(X,0) is precisely the combinatorial length of Γ(f)\Gamma(f).

  2. (ii)

    Assume that (X,0)(X,0) is irreducible and let σf(1)\sigma\in\triangle_{f}^{(1)} so that Γ(f)=Fσ\Gamma(f)=F_{\sigma}. Then (X,0)(X,0) is smooth if and only if σ\ell_{\sigma} lies on the boundary of the convex hull of the set ΣN\Sigma^{\circ}\cap N. In other words, let 0,,s+1\ell_{0},\ldots,\ell_{s+1} be the canonical primitive sequence of Σ\Sigma. Then either σ\ell_{\sigma} is one of 1,,s\ell_{1},\ldots,\ell_{s}, or there is an a>0a\in\mathbb{Z}_{>0} such that either

    σ=a0+1orσ=as+1+s.\ell_{\sigma}=a\ell_{0}+\ell_{1}\quad or\quad\ell_{\sigma}=a\ell_{s+1}+\ell_{s}.
  3. (iii)

    The curve (X,0)(X,0) is smooth if and only if the following condition holds: If pMp\in M and σ(p)>mσ\ell_{\sigma}(p)>m_{\sigma} for all σΣ(1)\sigma\in\triangle_{\Sigma}^{(1)}, then σ(p)>mσ\ell_{\sigma}(p)>m_{\sigma} for all σf(1)\sigma\in\triangle_{f}^{(1)}.

\therem Remark.

One can ask why the vectors 0\ell_{0} and s+1\ell_{s+1} do not appear in the list of (ii). The answer is that the corresponding divisors DσD_{\sigma}, though they intersect EE transversaly, they are 𝕋\mathbb{T}–invariant, hence they are eliminated by the convention of the definition 3.6.

Proof of section 4.

We start with the following observations. Write σi=0i\sigma_{i}=\mathbb{R}_{\geq 0}\langle\ell_{i}\rangle. Let \triangle^{\prime} be a regular subdivision of Σ\triangle_{\Sigma} which refines both f\triangle_{f} and the canonical subdivision of Σ\triangle_{\Sigma}. The map YYΣY_{\triangle^{\prime}}\to Y_{\Sigma} is then a resolution of YΣY_{\Sigma} with exceptional divisor EE^{\prime}. We can write E=i=1sEiE^{\prime}=\cup_{i=1}^{s^{\prime}}E^{\prime}_{i}, where each EiE_{i}^{\prime} is a rational curve. Furthermore, if iji\neq j, then EiE_{i}^{\prime} and EjE_{j}^{\prime} intersect if and only if |ij|=1|i-j|=1. In fact, we can write

(1)={σ1,,σs}{σ,τ}\triangle^{\prime(1)}=\{\sigma_{1}^{\prime},\ldots,\sigma_{s^{\prime}}^{\prime}\}\cup\{\sigma,\tau\}

where σ,τ\sigma,\tau are the two faces of Σ\Sigma and Ei=V(σi)E_{i}^{\prime}=V(\sigma_{i}^{\prime}).

Similarly as in [27], we see that YY_{\triangle^{\prime}} resolves (X,0)(X,0) and that the strict transform XX^{\prime} of XX in YY_{\triangle^{\prime}} intersects the exceptional divisor EE^{\prime} transversally in smooth points of EE^{\prime}. In fact, these intersection points lie in the open orbit OσiEiO_{\sigma_{i}^{\prime}}\subset E^{\prime}_{i}. Therefore, we have (see [27, Theorem 5.1])

|XEi|=χ(XOσi)=Vol1(F(i))|X^{\prime}\cap E^{\prime}_{i}|=\chi(X^{\prime}\cap O_{\sigma^{\prime}_{i}})=\mathop{\rm Vol}\nolimits_{1}(F(\ell_{i}^{\prime}))

where i\ell_{i}^{\prime} is the primitive generator of σi\sigma_{i}^{\prime}. Now, the components of (X,0)(X,0) are in bijection with the intersection points XEX^{\prime}\cap E^{\prime}, which proves (i).

For (ii), let ~Σ\widetilde{\triangle}_{\Sigma} be the canonical subdivision, and π:Y~Y\pi:\widetilde{Y}\to Y the associated modification, which is a resolution of YY. Let X~Y~\widetilde{X}\subset\widetilde{Y} be the strict transform of XX. The minimal cycle of the resolution Y~Y\widetilde{Y}\to Y is the reduced exceptional divisor EY~E\subset\widetilde{Y} and (Y,0)(Y,0) is rational. By [5], the pullback of the maximal ideal of 0Y0\in Y is the reduced exceptional divisor in Y~\widetilde{Y}, and the maximal ideal has no base points in Y~\widetilde{Y}. It follows that the multiplicity of (X,0)(X,0) is the intersection number between X~\widetilde{X} and EE. In particular, (X,0)(X,0) is smooth if and only if EX~E\cup\widetilde{X} is a normal crossing divisor. If σ=σi\sigma=\sigma_{i} for some 1is1\leq i\leq s, then this is indeed the case. Otherwise, there is an 0is0\leq i\leq s so that σ=ai+bi+1\ell_{\sigma}=a\ell_{i}+b\ell_{i+1}. In a neighbourhood of EiEi+1E_{i}\cap E_{i+1} we have coordinates u,vu,v so that Ei={x=0}E_{i}=\{x=0\}, Ei+1={y=0}E_{i+1}=\{y=0\} and we have some generic coefficients c,dc,d so that the strict transform of XX is defined by cxb+dyacx^{b}+dy^{a}. Thus, (X,0)(X,0) is not smooth if 1<i<s1<i<s. In the case i=0i=0 (the case i=si=s is similar), X~\widetilde{X} is smooth and transverse to E1E_{1} if and only if b=1b=1.

The condition in item (iii) is equivalent with the equality

(4.1) (Γ+(f)Γ+(f))M=.(\Gamma^{*}_{+}(f)\setminus\Gamma_{+}(f))\cap M=\emptyset.

Choose a basis for NN, inducing an isomorphism NMN\cong M via the dual basis, as well as an inner product on NMN_{\mathbb{R}}\cong M_{\mathbb{R}}. If we rotate the segment Γ(f)\Gamma(f) by π/2\pi/2 and translate it, then it can be identified with the vector i\ell_{i} (segment tit\ell_{i}, t[0,1]t\in[0,1]). Consider the parallelogram P(i)P(\ell_{i}) whose sides are parallel to 0\ell_{0} and s+1\ell_{s+1}, and it has i\ell_{i} as diagonal. It is divided by i\ell_{i} into two triangles, each of them can be identified by Γ+(f)Γ+(f)\Gamma^{*}_{+}(f)\setminus\Gamma_{+}(f). Hence, eq. 4.1 holds if and only if P(i)N=P(\ell_{i})^{\circ}\cap N=\emptyset.

Clearly, P(i)NP(\ell_{i})^{\circ}\cap N is empty if σconv(ΣN)\ell_{\sigma}\in\partial\mathop{\rm conv}\nolimits(\Sigma^{\circ}\cap N). The converse can be seen as follows. Let (ib)i(\ell^{\mathrm{b}}_{i})_{i\in\mathbb{Z}} be a family consisting of integral points on conv(ΣN)\partial\mathop{\rm conv}\nolimits(\Sigma^{\circ}\cap N), ordered according one of the orientation of this boundary. Two consecutive elements of this family form a basis of NN, and

ΣN=i0ib,i+1b{0}.\Sigma^{\circ}\cap N=\bigcup_{i\in\mathbb{Z}}\mathbb{Z}_{\geq 0}\langle\ell^{\mathrm{b}}_{i},\ell^{\mathrm{b}}_{i+1}\rangle\setminus\{0\}.

It follows that the set of irreducible elements in the semigroup ΣN\Sigma^{\circ}\cap N are presicely the elements on the boundary conv(ΣN)\partial\mathop{\rm conv}\nolimits(\Sigma^{\circ}\cap N). In particular, if σ(conv(ΣN))\ell_{\sigma}\in(\mathop{\rm conv}\nolimits(\Sigma^{\circ}\cap N))^{\circ}, then σ=+′′\ell_{\sigma}=\ell^{\prime}+\ell^{\prime\prime} for some ,′′ΣN\ell^{\prime},\ell^{\prime\prime}\in\Sigma^{\circ}\cap N. It follows that ,′′P(i)\ell^{\prime},\ell^{\prime\prime}\in P(\ell_{i})^{\circ}. ∎

Refer to caption
conv(ΣN)Σ\mathop{\rm conv}\nolimits(\Sigma^{\circ}\cap N)\subset\SigmaΓ+(f)Σ\Gamma_{+}(f)\subset\Sigma^{\vee}
Figure 2. The integral points in the interior of the parallellogram P(σ)P(\ell_{\sigma}).
\thecor Corollary.

Consider the notation from the proof of section 4(ii), that is, (X,0)(X,0) irreducible and σ=ai+bi+1\ell_{\sigma}=a\ell_{i}+b\ell_{i+1} with gcd(a,b)=1\gcd(a,b)=1. Then the multiplicity of (X,0)(X,0) is

mult(X,0)={bi=0,a+b0<i<s,ai=s.\mathop{\rm mult}\nolimits(X,0)=\begin{cases}b&i=0,\\ a+b&0<i<s,\\ a&i=s.\end{cases}\qed
\therem Remark.

Let ,\ell,\ell^{\prime} be any two linearly independent integral vectors in any free \mathbb{Z} module, and let NN be the free \mathbb{Z} module generated by them. Then the definitions from 4 and 4 can be repeated in NN. Then the determinant of two such vectors can be seen as the greatest common divisor of the maximal minors of the matrix having the coordinate vectors of ,\ell,\ell^{\prime} as rows, see [27]. Note that α(,)=α(,)\alpha(\ell,\ell^{\prime})=\alpha(\ell^{\prime},\ell). Moreover, β(0,s+1)β(s+1,0)1(modα(0,s+1))\beta(\ell_{0},\ell_{s+1})\beta(\ell_{s+1},\ell_{0})\equiv 1\,(\mathop{\rm mod}\nolimits\alpha(\ell_{0},\ell_{s+1})), cf. [29, Proposotion 5.6].

5. Isolated surface singularities

In the next theorem we give necessary and sufficient conditions for a Newton nondegenerate surface singularity to be isolated, in terms of the Newton polyhedron. In particular, we assume that r=3r=3 in this section. This is a (non-direct) generalization of a result of Kouchnirenko valid in the classical case [15].

5.1 Theorem.

Let (X,0)(X,0) be a Newton nondegenerate singularity and assume rkN=3\mathop{\mathrm{rk}}N=3. The following are equivalent

  1. (i)

    (X,0)(X,0) has an isolated singularity.

  2. (ii)

    If pMp\in M satisfies σ(p)>mσ\ell_{\sigma}(p)>m_{\sigma} for all σΣ(1)\sigma\in\triangle_{\Sigma}^{(1)}, then σ(p)>mσ\ell_{\sigma}(p)>m_{\sigma} for all σf(1)\sigma\in\triangle_{f}^{*(1)}.

  3. (iii)

    Let σ1,σ2Σ(1)\sigma_{1},\sigma_{2}\in\triangle_{\Sigma}^{(1)} and σ=0σ1,σ2Σ(2)\sigma=\mathbb{R}_{\geq 0}\langle\sigma_{1},\sigma_{2}\rangle\in\triangle_{\Sigma}^{(2)} and assume that τf(1)\tau\in\triangle_{f}^{(1)} with τσ\tau\subset\sigma. If pMp\in M so that σ1(p)>mσ1\ell_{\sigma_{1}}(p)>m_{\sigma_{1}} and σ2(p)>mσ2\ell_{\sigma_{2}}(p)>m_{\sigma_{2}}, then τ(p)>mτ\ell_{\tau}(p)>m_{\tau}.

  4. (iv)

    Let σ1,σ2Σ(1)\sigma_{1},\sigma_{2}\in\triangle_{\Sigma}^{(1)} and σ=0σ1,σ2Σ(2)\sigma=\mathbb{R}_{\geq 0}\langle\sigma_{1},\sigma_{2}\rangle\in\triangle_{\Sigma}^{(2)}. Then there is at most one τf(1)\tau\in\triangle_{f}^{(1)} with τσ\tau\subset\sigma and σ1τσ2\sigma_{1}\neq\tau\neq\sigma_{2}. If such a τ\tau exists, then τ\ell_{\tau} is one of the following

    (5.1) 1,,s,a0+1,s+as+1,a0\ell_{1},\ldots,\ell_{s},\quad a\ell_{0}+\ell_{1},\quad\ell_{s}+a\ell_{s+1},\qquad a\in\mathbb{Z}_{\geq 0}

    and, furthermore, there exists an ee\in\mathbb{Q} so that

    (5.2) eτ+σ1α(τ,σ1)+σ2α(τ,σ2)=0,e\ell_{\tau}+\frac{\ell_{\sigma_{1}}}{\alpha(\ell_{\tau},\ell_{\sigma_{1}})}+\frac{\ell_{\sigma_{2}}}{\alpha(\ell_{\tau},\ell_{\sigma_{2}})}=0,

    (see section 4 for α(,)\alpha(\cdot,\cdot)) and

    (5.3) emτ+mσ1α(τ,σ1)+mσ2α(τ,σ2)=1.em_{\tau}+\frac{m_{\sigma_{1}}}{\alpha(\ell_{\tau},\ell_{\sigma_{1}})}+\frac{m_{\sigma_{2}}}{\alpha(\ell_{\tau},\ell_{\sigma_{2}})}=-1.
Proof.

By section 3, the singular locus of the punctured germ X{0}X\setminus\{0\} is a union of orbits OσO_{\sigma} for some σΣ(2)\sigma\in\triangle_{\Sigma}^{(2)}. For such a σ\sigma, we have (V(σ),0)(X,0)(V(\sigma),0)\subset(X,0) if and only if the projection of Γ+(f)\Gamma_{+}(f) in MσM_{\sigma} is nontrivial, by section 3. By the same lemma, if (V(σ),0)(X,0)(V(\sigma),0)\subset(X,0), then the generic transverse type to V(σ)V(\sigma) in (X,0)(X,0) is a Newton nondegenerate curve with Newton polyhedron the projection of Γ+(f)\Gamma_{+}(f) to MσM_{\sigma}. Therefore item (i)\Leftrightarrowitem (iii) follows from section 4. The equivalence of item (ii) and item (iii) is an exercise.

The generic transverse type to (V(σ),0)(V(\sigma),0) in (X,0)(X,0) is smooth if and only if its diagram has a single face corresponding to a τ\tau as in eq. 5.1, and this face has length one. item (i)\Leftrightarrowitem (iv) follows, once we show that given such a τ\tau, an ee\in\mathbb{Q} satisfying eq. 5.2 exists and is unique, and that, furthermore, the left hand side of eq. 5.3 is minus the combinatorial length of the face FF of the Newton diagram corresponding to τ\tau.

Take a smooth subdivision of σ\sigma containing τ\tau as a ray, and let τi\tau_{i} be the ray adjacent to τ\tau between τ\tau and σi\sigma_{i}. Then there exists a b-b\in\mathbb{Z} so that

(5.4) bτ+τ1+τ2=0.-b\ell_{\tau}+\ell_{\tau_{1}}+\ell_{\tau_{2}}=0.

Furthermore, for i=1,2i=1,2, we may assume that

τi=βiτ+σiαi\ell_{\tau_{i}}=\frac{\beta_{i}\ell_{\tau}+\ell_{\sigma_{i}}}{\alpha_{i}}

where αi/βi\alpha_{i}/\beta_{i} is the continued fraction associated with τ\ell_{\tau} and σi\ell_{\sigma_{i}}. As a result, eq. 5.4 can be rewritten as (5.2) with e=b+β1/α1+β2/α2e=-b+\beta_{1}/\alpha_{1}+\beta_{2}/\alpha_{2}. Let p1,p2p_{1},p_{2} be the endpoints of FF so that τ1(p2p1)>0\ell_{\tau_{1}}(p_{2}-p_{1})>0 and τ2(p1p2)>0\ell_{\tau_{2}}(p_{1}-p_{2})>0. Since τi\ell_{\tau_{i}} is a primitive function on the affine hull of the face of FF, τ1(p2p1)=τ2(p1p2)=\ell_{\tau_{1}}(p_{2}-p_{1})=\ell_{\tau_{2}}(p_{1}-p_{2})= the length of FF. We find

emτ+mσ1α1+mσ2α2=eτ(p1)+σ1(p1)α1+σ2(p2)α2=bτ(p1)+τ1(p1)+τ2(p2)=τ2(p2p1).\begin{split}em_{\tau}+\frac{m_{\sigma_{1}}}{\alpha_{1}}+\frac{m_{\sigma_{2}}}{\alpha_{2}}&=e\ell_{\tau}(p_{1})+\frac{\ell_{\sigma_{1}}(p_{1})}{\alpha_{1}}+\frac{\ell_{\sigma_{2}}(p_{2})}{\alpha_{2}}\\ &=-b\ell_{\tau}(p_{1})+\ell_{\tau_{1}}(p_{1})+\ell_{\tau_{2}}(p_{2})=\ell_{\tau_{2}}(p_{2}-p_{1}).\qed\end{split}

6. Resolution of Newton nondegenerate surface singularities

In this section, we retain the notation introduced in section 3, with the assumption that rkN=3\mathop{\mathrm{rk}}N=3. We describe Oka’s algorithm which describes explicitly the graph of a resolution of a Newton nondegenerate Weil divisor of dimension 22. This algorithm was originally described by Oka [27] for Newton nondegenerate hypersurface singularities in (3,0)(\mathbb{C}^{3},0). The general methods for resolving Newton nondegenerate hypersurface singularities have been used in e.g. [32] and [3, Chapter 8].

\thedefinition Definition.

A canonical subdivision of f\triangle_{f} is a subdivision ~f\widetilde{\triangle}_{f} satisfying the following.

  1. (i)

    ~f\widetilde{\triangle}_{f} is a regular subdivision of f\triangle_{f}.

  2. (ii)

    If σf(2)f\sigma\in\triangle^{(2)}_{f}\setminus\triangle_{f}^{*}, then ~f|σ\widetilde{\triangle}_{f}|_{\sigma} is the canonical subdivision ~σ\widetilde{\triangle}_{\sigma} of σ\triangle_{\sigma} given in section 4.

\theblock.

The existence of a canonical subdivision is proved in [27, §3]. We fix such a subdivision ~f\widetilde{\triangle}_{f}. We will denote by Y~\widetilde{Y} the toric variety associated with ~f\widetilde{\triangle}_{f}. The map Y~Y\widetilde{Y}\to Y is denoted by π\pi, and the strict transform of XX under this map is denoted by X~\widetilde{X}. We denote by πX\pi_{X} the restriction π|X~:X~X\pi|_{\widetilde{X}}:\widetilde{X}\to X. By section 2, the map Y~Y\widetilde{Y}\to Y is proper, hence X~X\widetilde{X}\to X is proper as well.

\thedefinition Definition.

For i,di,d\in\mathbb{N}, define

~f(i,d)={σ~f(i)|dim(FσΓ(f))=d}~f(i,d)=~f(i,d)~f.\begin{split}\widetilde{\triangle}_{f}^{(i,d)}&=\left\{\sigma\in\widetilde{\triangle}_{f}^{(i)}\,\middle|\,\dim(F_{\sigma}\cap\Gamma(f))=d\right\}\\ \widetilde{\triangle}_{f}^{*(i,d)}&=\widetilde{\triangle}_{f}^{(i,d)}\cap\widetilde{\triangle}_{f}^{*}.\end{split}
\thedefinition Definition.

We start by defining a graph GG^{*} as follows. Index the set ~f(1,2)\widetilde{\triangle}_{f}^{(1,2)} by a set 𝒩\mathcal{N}, i.e. write ~f(1,2)={σn|n𝒩}\widetilde{\triangle}_{f}^{(1,2)}=\left\{\sigma_{n}\,\middle|\,n\in\mathcal{N}\right\} in such a way that the map 𝒩~f(1,2)\mathcal{N}\to\widetilde{\triangle}_{f}^{(1,2)}, nσnn\mapsto\sigma_{n} is bijective. Similarly, index the set Δ~f(1,2)Δ~f(1,1)\widetilde{\Delta}_{f}^{(1,2)}\cup\widetilde{\Delta}_{f}^{*(1,1)} by 𝒩\mathcal{N}^{*}. Hence 𝒩𝒩\mathcal{N}\subset\mathcal{N}^{*}. The elements of 𝒩\mathcal{N}^{*} are referred to as extended nodes, while 𝒩{\mathcal{N}} as nodes.

Denote by FnF_{n} the face of Γ+(f)\Gamma_{+}(f) corresponding to σn\sigma_{n} and by n\ell_{n} the primitive integral generator of σn\sigma_{n}. Note that n𝒩n\in\mathcal{N} if and only if FnF_{n} is bounded. For n,n𝒩n,n^{\prime}\in\mathcal{N}^{*}, let tn,nt_{n,n^{\prime}} be the length of the segment FnFnF_{n}\cap F_{n^{\prime}} if this is a bounded segment of dimension 1. If FnFnF_{n}\cap F_{n^{\prime}} is unbounded, or has dimension 0, then we set tn,n=0t_{n,n^{\prime}}=0. Now, for every pair nn and n𝒩n^{\prime}\in\mathcal{N}^{*}, we join n,nn,n^{\prime} by tn,nt_{n,n^{\prime}} bamboos of type α(n,n)/β(n,n)\alpha(\ell_{n},\ell_{n^{\prime}})/\beta(\ell_{n},\ell_{n^{\prime}}), as in fig. 3. This finishes the construction of the graph GG^{*}. Denote its set of vertices 𝒱\mathcal{V}^{*}.

Define the graph GG as the induced full subgraph of GG^{*} on the set of vertices 𝒱=𝒱(𝒩𝒩)\mathcal{V}=\mathcal{V}^{*}\setminus(\mathcal{N}^{*}\setminus\mathcal{N}).

Refer to caption
b1-b_{1}b1-b_{1}\cdotsb2-b_{2}b1-b_{1}b2-b_{2}b2-b_{2}\cdots\cdotsbs1-b_{s-1}bs1-b_{s-1}bs1-b_{s-1}bs-b_{s}bs-b_{s}bs-b_{s}nnnn^{\prime}\vdots\vdots
Figure 3. We join n,n𝒩n,n^{\prime}\in\mathcal{N} by tn,nt_{n,n^{\prime}} bamboos of the above form, where the sequence b1,,bsb_{1},\ldots,b_{s} is defined as b1=1b_{1}=1 if α(n,n)=1\alpha(\ell_{n},\ell_{n^{\prime}})=1, and by a negative continued fraction expansion α(n,n)/β(n,n)=[b1,,bs]\alpha(\ell_{n},\ell_{n^{\prime}})/\beta(\ell_{n},\ell_{n^{\prime}})=[b_{1},\ldots,b_{s}] otherwise.

In order to have a plumbing graph structure on GG, we must specify an Euler number and a genus for each vertex, as well as a sign for each edge. All edges are positive. Vertices appearing on bamboos have genus zero, whereas the genus gng_{n} associated with n𝒩n\in\mathcal{N} is defined as the number of integral interior points in the polygon FnF_{n}.

To every extended node n𝒩n\in\mathcal{N}^{*} we have associated the cone σn\sigma_{n} and its primitive integral generator n\ell_{n}. If v1,,vsv_{1},\ldots,v_{s} are the vertices appearing on a bamboo, in this order, from nn to n𝒩n^{\prime}\in\mathcal{N}^{*}, let 0,1,,s+1\ell_{0},\ell_{1},\ldots,\ell_{s+1} be the canonical primitive sequence associated with n,n\ell_{n},\ell_{n^{\prime}}. We then set v=i\ell_{v}=\ell_{i} for v=viv=v_{i}, i=1,,si=1,\ldots,s, and σv=0i\sigma_{v}=\mathbb{R}_{\geq 0}\langle\ell_{i}\rangle. This induces a map γ:𝒱~f(1)\gamma:\mathcal{V}\to\widetilde{\triangle}_{f}^{(1)} with the property that γ(n)=σn\gamma(n)=\sigma_{n} for n𝒩n\in\mathcal{N}^{*}, and v,w\ell_{v},\ell_{w} generate an element of ~f(2)\widetilde{\triangle}_{f}^{(2)} if v,wv,w are adjacent in GG^{*}.

For any v𝒱v\in\mathcal{V}, let 𝒱v\mathcal{V}_{v} and𝒱v\mathcal{V}_{v}^{*} be the set of neighbours of vv in GG and GG^{*}, respectively. Then there exists a unique bv1-b_{v}\in\mathbb{Z}_{\leq-1} satisfying

bvv+u𝒱vu=0inN,-b_{v}\ell_{v}+\sum_{u\in\mathcal{V}_{v}^{*}}\ell_{u}=0\ \ \mbox{in}\ \ N,

The number bv-b_{v} is the Euler number associated with v𝒱v\in\mathcal{V}. We note that if vv lies on a bamboo, with the notation of the previous paragraph, v=viv=v_{i}, then bv=bi-b_{v}=-b_{i} and bi2-b_{i}\leq-2 unless α(n,n)=1\alpha(\ell_{n},\ell_{n^{\prime}})=1.

\therem Remark.

The link of an isolated surface singularity is a rational homology sphere if and only if it has a resolution whose graph is a tree and all vertices have genus zero, see e.g. [20]. The above construction produces such a graph if and only if all integral points on Γ(f)\Gamma(f) lie on its boundary Γ(f)\partial\Gamma(f).

Indeed, if PΓ(f)P\subset\Gamma(f) is a vertex which is not on the boundary, then the nodes corresponding to faces of Γ(f)\Gamma(f) containing PP lie on an embedded cycle. Similarly, if SΓ(f)S\subset\Gamma(f) is a face of dimension 11 which is not a subset of the boundary, and SS contains integral interior points, then the nodes corresponding to the two faces containing SS are joined by more than one bamboo, inducing an embedded cycle in GG. Finally, if FΓ(f)F\subset\Gamma(f) is a two dimensional face containing interior integral interior points, then the corresponding node has nonzero genus. The converse is not difficult.

The classical case Y=3Y=\mathbb{C}^{3} is discussed in details in [7].

\theexample Example.

Let Σ=03\Sigma=\mathbb{R}^{3}_{\geq 0}, and consider standard coordinates x,y,zx,y,z on Y=3Y=\mathbb{C}^{3}, and the function

f(x,y,z)=x5+x2y2+y7+z10.f(x,y,z)=x^{5}+x^{2}y^{2}+y^{7}+z^{10}.

The Newton diagram Γ(f)\Gamma(f) consists of two triangular faces, whose intersection is a segment of length two. The diagram, as well as the graph obtained by Oka’s algorithm can be seen in fig. 4.

Refer to caption
1-1

(0,1,0)(0,1,0)

(2,4,1)(2,4,1)

4-44-43-3[2][2]1-12-2

(2,3,1)(2,3,1)

(3,2,1)(3,2,1)

(7,3,2)(7,3,2)

(25,10,7)(25,10,7)

(26,10,7)(26,10,7)

(1,0,0)(1,0,0)

(0,0,1)(0,0,1)(2,3,2)(2,3,2)1-13-32-2(10,4,3)(10,4,3)(5,2,2)(5,2,2)3-3x5x^{5}x2y2x^{2}y^{2}y7y^{7}z10z^{10}9-9
Figure 4. A Newton diagram, and the graph GG^{*}, with the subgraph GG in black.
\theprop Proposition.

Let (X,0)(X,0) be a Newton nondegenerate surface singularity. Then the map X~X\widetilde{X}\to X is a resolution of (X,0)(X,0) whose resolution graph is GG.

More precisely, X~\widetilde{X} is smooth and the exceptional set EX~E\subset\widetilde{X} is a normal crossing divisor. For each σ~f(1)\sigma\in\widetilde{\triangle}_{f}^{(1)}, we can enumerate the irreducible components of EσE_{\sigma} by γ1(σ)\gamma^{-1}(\sigma) so that Eσ=vγ1(σ)EvE_{\sigma}=\amalg_{v\in\gamma^{-1}(\sigma)}E_{v}, where EvE_{v} is a smooth curve.

If γ(v)~f(1)~\gamma(v)\in\widetilde{\triangle}_{f}^{(1)}\setminus\widetilde{\triangle}^{*}, then EvE_{v} is compact, has genus gvg_{v}, and its normal bundle in X~\widetilde{X} has Euler number bv-b_{v}. If γ(v)~(1)\gamma(v)\in\widetilde{\triangle}^{(1)*}, then EvE_{v} is a smooth germ, transverse to a smooth point of the exceptional divisor.

Furthermore, if v,w𝒱v,w\in\mathcal{V}, then the number of intersection points |EvEw||E_{v}\cap E_{w}| equals the number of edges between vv and ww in GG.

Proof.

The proof goes exactly as in [27]

\thedefinition Definition.

For v𝒱v\in\mathcal{V}^{*}, (recall section 3 and section 3) let

Fv=Fγ(v),v=γ(v),mv=mγ(v).F_{v}=F_{\gamma(v)},\quad\ell_{v}=\ell_{\gamma(v)},\quad m_{v}=m_{\gamma(v)}.
\thelemma Lemma.

For v𝒱v\in\mathcal{V}, we have

bvv+u𝒱vu=0,bvmv+u𝒱vmu=2Vol2(Fv).-b_{v}\ell_{v}+\sum_{u\in\mathcal{V}^{*}_{v}}\ell_{u}=0,\quad-b_{v}m_{v}+\sum_{u\in\mathcal{V}^{*}_{v}}m_{u}=-2\mathop{\rm Vol}\nolimits_{2}(F_{v}).
Proof.

The first equality follows from construction, see also [27, §6]. The second equality follows from [7, Prop. 4.4.4] and the formula α1=β0+s+1\alpha\ell_{1}=\beta\ell_{0}+\ell_{s+1}, where 0,1,,s+1\ell_{0},\ell_{1},\ldots,\ell_{s+1} is a primitive sequence. ∎

\therem Remark.
  • (i)

    The exceptional divisor EE is the union of EσE_{\sigma} for which σ~f(1)\sigma\in\widetilde{\triangle}_{f}^{(1)} is a cone which is not contained in Σ\partial\Sigma, or, equivalently, FσF_{\sigma} is compact.

  • (ii)

    If σ~f(1,2)\sigma\in\widetilde{\triangle}_{f}^{(1,2)}, then EσE_{\sigma} is a compact smooth irreducible curve. If σ~f(1,1)~f\sigma\in\widetilde{\triangle}_{f}^{(1,1)}\setminus\widetilde{\triangle}^{*}_{f}, then EσE_{\sigma} is the union of tσt_{\sigma} disjoint smooth compact rational curves. For σ~f(1,1)\sigma\in\widetilde{\triangle}_{f}^{*(1,1)}, the intersection Eσ=V(σ)X~E_{\sigma}=V(\sigma)\cap\widetilde{X} is the disjoint union of tt smooth curve germs, where tt is the length of the segment FσΓ(f)F_{\sigma}\cap\Gamma(f). If σ~f(1,0)\sigma\in\widetilde{\triangle}_{f}^{(1,0)}, then Eσ=E_{\sigma}=\emptyset (the global divisor DσD_{\sigma} does not intersect X~\widetilde{X}).

\thedefinition Definition.

We denote by L=Ev|v𝒱L=\mathbb{Z}\left\langle E_{v}\,\middle|\,v\in\mathcal{V}\right\rangle the lattice of integral cycles in X~\widetilde{X} supported on the exceptional divisor EE.

\thedefinition Definition.

Let g𝒪Y,0g\in\mathcal{O}_{Y,0} and denote its restriction by g¯𝒪X,0\overline{g}\in\mathcal{O}_{X,0}. For any v𝒱v\in\mathcal{V}^{*}, we define

wtv(g)=min{v(p)|psupp(g)},wt(g)=v𝒱wtv(g)EvL,wtv(g¯)=max{wtv(g+h)|hIX},wt(g¯)=v𝒱wtv(g¯)EvL.\begin{split}\mathop{\rm wt}\nolimits_{v}(g)=\min\left\{\ell_{v}(p)\,\middle|\,p\in\mathop{\rm supp}\nolimits(g)\right\},\quad&\mathop{\rm wt}\nolimits(g)=\sum_{v\in\mathcal{V}}\mathop{\rm wt}\nolimits_{v}(g)E_{v}\in L,\\ \mathop{\rm wt}\nolimits_{v}(\overline{g})=\max\left\{\mathop{\rm wt}\nolimits_{v}(g+h)\,\middle|\,h\in I_{X}\right\},\quad&\mathop{\rm wt}\nolimits(\overline{g})=\sum_{v\in\mathcal{V}}\mathop{\rm wt}\nolimits_{v}(\overline{g})E_{v}\in L.\end{split}

For σ=γ(v)\sigma=\gamma(v), we also write wtσ\mathop{\rm wt}\nolimits_{\sigma} instead of wtv\mathop{\rm wt}\nolimits_{v}, as this is independent of vγ1(σ)v\in\gamma^{-1}(\sigma).

Similarly, for any v𝒱v\in\mathcal{V}, let divv\mathop{\rm div}\nolimits_{v} be the valuation on 𝒪X,0\mathcal{O}_{X,0} associated with the divisor EvE_{v}, that is, for g¯𝒪X,0\overline{g}\in\mathcal{O}_{X,0}, denote by divv(g¯)\mathop{\rm div}\nolimits_{v}(\overline{g}) the order of vanishing of the function πX(g¯)\pi_{X}^{*}(\overline{g}) along EvE_{v}. Set also

div(g¯)=v𝒱divv(g¯)EvL.\mathop{\rm div}\nolimits(\overline{g})=\sum_{v\in\mathcal{V}}\mathop{\rm div}\nolimits_{v}(\overline{g})E_{v}\in L.
\therem Remark.
  • (i)

    If σ=γ(v)\sigma=\gamma(v) and |γ1(σ)|>1|\gamma^{-1}(\sigma)|>1, then divv\mathop{\rm div}\nolimits_{v} is not independent of the choice of vγ1(σ)v\in\gamma^{-1}(\sigma).

  • (ii)

    For σ~f(1)\sigma\in\widetilde{\triangle}_{f}^{(1)}, the function wtσ:𝒪Y,0{0}\mathop{\rm wt}\nolimits_{\sigma}:\mathcal{O}_{Y,0}\setminus\{0\}\to\mathbb{Z} is the valuation on 𝒪Y,0\mathcal{O}_{Y,0} associated with the irreducible divisor V(σ)Y~V(\sigma)\subset\widetilde{Y}, cf. eq. 3.2.

  • (iii)

    In general, the functions wtv\mathop{\rm wt}\nolimits_{v} and divv\mathop{\rm div}\nolimits_{v} do not coincide on 𝒪X,0\mathcal{O}_{X,0}. However, wtv(g¯)divv(g¯)\mathop{\rm wt}\nolimits_{v}(\overline{g})\leq\mathop{\rm div}\nolimits_{v}(\overline{g}) for any g¯𝒪X,0\overline{g}\in\mathcal{O}_{X,0} and v𝒱v\in\mathcal{V}. Furthermore, if pMp\in M and γ(v)~f(1,>0)~f\gamma(v)\in\widetilde{\triangle}_{f}^{(1,>0)}\setminus\widetilde{\triangle}_{f}^{*}, then divv(xp)=wtv(xp)=v(p)\mathop{\rm div}\nolimits_{v}(x^{p})=\mathop{\rm wt}\nolimits_{v}(x^{p})=\ell_{v}(p). In particular, this defines a group homomorphism MLM\to L, pwt(xp)p\to\mathop{\rm wt}\nolimits(x^{p}).

7. The geometric genus

In this section we provide a formula for the delta invariant and geometric genus for an arbitrary generalized Newton nondegenerate singularity in terms of its Newton polyhedron. In this section, the rank rr of NN is under no restriction. Recall that we say that ff (or Γ+(f)\Gamma_{+}(f)) is pointed at pMp\in M_{\mathbb{Q}}, if for any σΣ(1)\sigma\in\triangle_{\Sigma}^{(1)} we have mσ=σ(p)m_{\sigma}=\ell_{\sigma}(p), see section 3.

\therem Remark.

In the proof of theorem 7.1, one of the main steps consists of computing the cohomology of a line bundle on a toric variety. To do this, we build on classical methods [13, 10]. A more general method to compute such cohomology has been described by Altmann and Ploog in [2].

\thedefinition Definition.

For a point xx in an analytic variety XX, denote by 𝒪¯X,x\overline{\mathcal{O}}_{X,x} the normalization of its local ring 𝒪X,x\mathcal{O}_{X,x}. The delta invariant associated with xXx\in X is defined as

δ(X,x)=dim𝒪¯X,x/𝒪X,x.\delta(X,x)=\dim_{\mathbb{C}}\overline{\mathcal{O}}_{X,x}/\mathcal{O}_{X,x}.

Let X~X\widetilde{X}\to X be a resolution of the singularity xXx\in X and assume that XX has dimension dd. Assume, furthermore, that δ(X,x)<\delta(X,x)<\infty, and that the higher direct image sheaves Riπ𝒪X~R^{i}\pi_{*}\mathcal{O}_{\widetilde{X}}, i>0i>0, are concentrated at xx. The geometric genus pg=pg(X,0)p_{g}=p_{g}(X,0) is defined as

(1)d1pg(X,x)=δ(X,x)+i=1d1(1)ihi(X~,𝒪X~).(-1)^{d-1}p_{g}(X,x)=\delta(X,x)+\sum_{i=1}^{d-1}(-1)^{i}h^{i}(\widetilde{X},\mathcal{O}_{\widetilde{X}}).

We say that (X,x)(X,x) is rational if δ(X,x)=0\delta(X,x)=0 and hi(X~,𝒪X~)=0h^{i}(\widetilde{X},\mathcal{O}_{\widetilde{X}})=0 for i>0i>0.

7.1 Theorem.

Let (X,0)(Y,0)(X,0)\subset(Y,0) be a Newton nondegenerate Weil divisor of dimension d=r1d=r-1.

  1. (i)

    We have the following canonical identifications

    𝒪¯X,0/𝒪X,0pMH~0(Γ+(xpf)Σ,),Hi(X~,𝒪X~)pMH~i(Γ+(xpf)Σ,),i>0.\begin{split}\overline{\mathcal{O}}_{X,0}/\mathcal{O}_{X,0}&\cong\bigoplus_{p\in M}\widetilde{H}^{0}(\Gamma_{+}(x^{p}f)\setminus\Sigma^{\vee},\mathbb{C}),\\ H^{i}(\widetilde{X},\mathcal{O}_{\widetilde{X}})&\cong\bigoplus_{p\in M}\widetilde{H}^{i}(\Gamma_{+}(x^{p}f)\setminus\Sigma^{\vee},\mathbb{C}),\quad i>0.\end{split}

    In particular, if these vector spaces have finite dimension, then

    δ(X,0)=pMh~0(Γ+(xpf)Σ,),pg(X,0)=(1)d1pMχ~(Γ+(xpf)Σ,),\begin{split}\delta(X,0)&=\sum_{p\in M}\widetilde{h}^{0}(\Gamma_{+}(x^{p}f)\setminus\Sigma^{\vee},\mathbb{C}),\\ p_{g}(X,0)&=(-1)^{d-1}\sum_{p\in M}\widetilde{\chi}(\Gamma_{+}(x^{p}f)\setminus\Sigma^{\vee},\mathbb{C}),\end{split}

    where χ~\widetilde{\chi} denotes the reduced Euler characteristic, that is, the alternating sum of ranks of reduced singular cohomology groups.

  2. (ii)

    We have

    h~d1(Γ+(xpf)Σ,)={1if0Γ+(xpf)Γ+(xpf),0else.\widetilde{h}^{d-1}(\Gamma_{+}(x^{p}f)\setminus\Sigma^{\vee},\mathbb{C})=\begin{cases}1&\textrm{if}\quad 0\in\Gamma_{+}^{*}(x^{p}f)^{\circ}\setminus\Gamma_{+}(x^{p}f)^{\circ},\\ 0&\textrm{else}.\end{cases}

    In particular, hd1(X~,𝒪X~)=|MΓ+(f)Γ+(f)|h^{d-1}(\widetilde{X},\mathcal{O}_{\widetilde{X}})=|M\cap\Gamma_{+}^{*}(f)^{\circ}\setminus\Gamma_{+}(f)^{\circ}| (recall section 3).

  3. (iii)

    Assume that ff is \mathbb{Q}-pointed, that d2d\geq 2, and that (X,0)(X,0) has only rational singularities outside the origin. Then (X,0)(X,0) is normal and hi(X~,𝒪X~)=0h^{i}(\widetilde{X},\mathcal{O}_{\widetilde{X}})=0 for 1i<d11\leq i<d-1.

\thecor Corollary.

Assume that d=2d=2 and (X,0)(X,0) is normal. Then

pg(X,0)=|MΓ+(f)Γ+(f)|.p_{g}(X,0)=|M\cap\Gamma_{+}^{*}(f)^{\circ}\setminus\Gamma_{+}(f)^{\circ}|.\qed

This generalizes a result of Merle and Teissier [19] valid for the classical case Σ=03\Sigma={\mathbb{R}}_{\geq 0}^{3}.

\thecor Corollary.

Assume that d=1d=1 and (X,0)(X,0) is an irreducible germ of a curve, and that σ~f(1)\sigma\in\widetilde{\triangle}^{(1)}_{f} satisfies Fσ=Γ(f)F_{\sigma}=\Gamma(f) (cf. section 4item (i)). Then δ(X,0)\delta(X,0) is the number of unordered pairs ,′′ΣN\ell^{\prime},\ell^{\prime\prime}\in\Sigma^{\circ}\cap N satisfying +′′=σ\ell^{\prime}+\ell^{\prime\prime}=\ell_{\sigma}.

Proof.

Let P(σ)P(\ell_{\sigma}) be the parallelogram introduced in the proof of section 4. The diagonal splits P(σ)P(\ell_{\sigma}) into two triangles, T1T_{1} and T2T_{2}, say. If T1\ell^{\prime}\in T_{1}^{\circ}, then σT2\ell_{\sigma}-\ell^{\prime}\in T^{\circ}_{2}. This induces a bijection between elements T1N\ell^{\prime}\in T^{\circ}_{1}\cap N and unordered pairs {,′′}ΣN\{\ell^{\prime},\ell^{\prime\prime}\}\subset\Sigma^{\circ}\cap N adding up to σ\ell_{\sigma}. By rotating by π/2\pi/2 as in the proof of section 4, T1NT_{1}^{\circ}\cap N is in bijection with MΓ+(f)Γ+(f)M\cap\Gamma_{+}^{*}(f)^{\circ}\setminus\Gamma_{+}(f)^{\circ}. ∎

\therem Remark.

Assume that d2d\geq 2, and that XX is rational outside {0}\{0\}. Then, for 0<i<d10<i<d-1, we have

Hi(X~,𝒪X~)Hi(X~E,𝒪X~)Hi(X{0},𝒪X)H{0}i+1(X,𝒪X).H^{i}(\widetilde{X},\mathcal{O}_{\widetilde{X}})\cong H^{i}(\widetilde{X}\setminus E,\mathcal{O}_{\widetilde{X}})\cong H^{i}(X\setminus\{0\},\mathcal{O}_{X})\cong H^{i+1}_{\{0\}}(X,\mathcal{O}_{X}).

Here, the first isomorphism comes from the long exact sequence for cohomology with support in EE, and the vanishing HEi(X~,𝒪X~)=0H^{i}_{E}(\widetilde{X},\mathcal{O}_{\widetilde{X}})=0, for i<di<d [14, Corollary 3.3]. The second isomorphism follows from the rationality assumption, and the Leray spectral sequence. The third isomorphism comes from the similar long exact sequence for cohomology with support in {0}\{0\}, and the fact that Hj(X,𝒪X)=0H^{j}(X,\mathcal{O}_{X})=0 for j>0j>0, if we choose a Stein representative XX of the germ (X,0)(X,0). This last long exact sequence furthermore gives

H{0}1(X,𝒪X)H0(X{0},𝒪X)H0(X,𝒪X)𝒪¯X,0/𝒪X,0.H^{1}_{\{0\}}(X,\mathcal{O}_{X})\cong\frac{H^{0}(X\setminus\{0\},\mathcal{O}_{X})}{H^{0}(X,\mathcal{O}_{X})}\cong\overline{\mathcal{O}}_{X,0}/\mathcal{O}_{X,0}.

Therefore, in this case, the groups described in theorem 7.1 are closely related with the depth of 𝒪X,0\mathcal{O}_{X,0}. In particular, the conclusion of theorem 7.1item (iii) is that (X,0)(X,0) is a Cohen–Macaulay ring.

If ff is pointed at pMp\in M, then this statement can be proved as follows. Since (X,0)(X,0) is a Cartier divisor in (Y,0)(Y,0), cf. section 3ii, and (Y,0)(Y,0) is Cohen–Macaulay [9, Theorem 6.3.5] so is (X,0)(X,0) [9, Theorem 2.1.3].

Proof of theorem 7.1.

To prove item (i), we use results and notation from [10, §7], see also [13, 3.5]. Define

Dm={mσDσ|σ~f(1)}.D_{m}=\sum\left\{m_{\sigma}D_{\sigma}\,\middle|\,\sigma\in\widetilde{\triangle}_{f}^{(1)}\right\}.

Then Dm+X~D_{m}+\widetilde{X} is the divisor of the pullback of ff to Y~\widetilde{Y}, and we have a short exact sequence

0𝒪Y~(Dm)f𝒪Y~𝒪X~0.0\to\mathcal{O}_{\widetilde{Y}}(D_{m})\stackrel{{\scriptstyle\cdot f}}{{\to}}\mathcal{O}_{\widetilde{Y}}\to\mathcal{O}_{\widetilde{X}}\to 0.

By [10, Corollary 7.4], we have Hi(Y~,𝒪Y~)=0H^{i}(\widetilde{Y},\mathcal{O}_{\widetilde{Y}})=0 for all i>0i>0. Furthermore, H0(X~,𝒪X~)𝒪¯X,0H^{0}(\widetilde{X},\mathcal{O}_{\widetilde{X}})\cong\overline{\mathcal{O}}_{X,0}, and the image of H0(Y~,𝒪Y~)=𝒪Y,0H^{0}(\widetilde{Y},\mathcal{O}_{\widetilde{Y}})=\mathcal{O}_{Y,0} in 𝒪¯X,0\overline{\mathcal{O}}_{X,0} is 𝒪X,0\mathcal{O}_{X,0}. Therefore,

𝒪¯X,0/𝒪X,0H1(Y~,𝒪Y~(Dm)),andHi(X~,𝒪X~)Hi+1(Y~,𝒪Y~(Dm)),i>0.\begin{split}\overline{\mathcal{O}}_{X,0}/\mathcal{O}_{X,0}&\cong H^{1}(\widetilde{Y},\mathcal{O}_{\widetilde{Y}}(D_{m})),\ \ \mbox{and}\\ H^{i}(\widetilde{X},\mathcal{O}_{\widetilde{X}})&\cong H^{i+1}(\widetilde{Y},\mathcal{O}_{\widetilde{Y}}(D_{m})),\quad i>0.\end{split}

Denote by gg the order function defined in [10, §6] (using the natural trivialization of 𝒪Y~(Dm)\mathcal{O}_{\widetilde{Y}}(D_{m}) on the open torus)

g:|~f|,g()=min{(q)|qΓ+(f)}g:|\widetilde{\triangle}_{f}|\to\mathbb{R},\quad g(\ell)=-\min\left\{\ell(q)\,\middle|\,q\in\Gamma_{+}(f)\right\}

and define the sets

Zp={|~f||(p)g()},pM.Z_{p}=\left\{\ell\in|\widetilde{\triangle}_{f}|\,\middle|\,\ell(p)\geq g(\ell)\right\},\quad p\in M.

We note that ZpZ_{p} is a convex cone and that 0Zp0\in Z_{p} for all pMp\in M. By [10, Theorem 7.2], we have isomorphisms

Hi+1(Y~,𝒪Y~(Dm))pMHZpi+1(|~f|,).H^{i+1}(\widetilde{Y},\mathcal{O}_{\widetilde{Y}}(D_{m}))\cong\bigoplus_{p\in M}H^{i+1}_{Z_{p}}(|\widetilde{\triangle}_{f}|,\mathbb{C}).

Since |~f|=Σ|\widetilde{\triangle}_{f}|=\Sigma is a convex set, the long exact sequence associated with cohomology with supports provides, for any pMp\in M

0H~i(|~f|,)H~i(|~f|Zp,)HZpi+1(|~f|,)H~i+1(|~f|,)0.0\cong\widetilde{H}^{i}(|\widetilde{\triangle}_{f}|,\mathbb{C})\to\widetilde{H}^{i}(|\widetilde{\triangle}_{f}|\setminus Z_{p},\mathbb{C})\cong H^{i+1}_{Z_{p}}(|\widetilde{\triangle}_{f}|,\mathbb{C})\to\widetilde{H}^{i+1}(|\widetilde{\triangle}_{f}|,\mathbb{C})\cong 0.

To finish the proof of (i), we will show that for any pMp\in M, the spaces |~f|Zp=ΣZp|\widetilde{\triangle}_{f}|\setminus Z_{p}=\Sigma\setminus Z_{p} and Γ+(xpf)Σ\Gamma_{+}(x^{p}f)\setminus\Sigma^{\vee} are in fact homotopically equivalent. We start by noting that the the condition Zp|~f|Z_{p}\subset\partial|\widetilde{\triangle}_{f}| (including the case when Zp=Z_{p}=\emptyset) is equivalent to 0Γ+(xpf)Γ(xpf)0\in\Gamma_{+}(x^{p}f)\setminus\Gamma(x^{p}f). If this happens then we can choose a qΣq\in\Sigma^{\vee} small so that qΓ+(xpf)Γ(xpf)-q\in\Gamma_{+}(x^{p}f)\setminus\Gamma(x^{p}f) as well, and so Γ+(xpf)Σ\Gamma_{+}(x^{p}f)\setminus\Sigma^{\vee} is star-shaped with center q-q. In particular, in this case,

ΣZp{a point}Γ+(xpf)Σ,\Sigma\setminus Z_{p}\sim\{\mbox{a point}\}\sim\Gamma_{+}(x^{p}f)\setminus\Sigma^{\vee},

where \sim denotes the homotopy equivalence. Thus, in what follows, we assume that ZpZ_{p} contains an interior point in Σ\Sigma, equivalently, 0Γ+(xpf)Γ(xpf)0\notin\Gamma_{+}(x^{p}f)\setminus\Gamma(x^{p}f).

Choose 0Σ\ell_{0}\in\Sigma^{\circ} and q0(Σ)q_{0}\in(\Sigma^{\vee})^{\circ} satisfying 0(q0)=1\ell_{0}(q_{0})=1 and define the hyperplanes

H={N|(q0)=1},H={qM|0(q)=1}.H=\left\{\ell\in N_{\mathbb{R}}\,\middle|\,\ell(q_{0})=1\right\},\quad H^{\vee}=\left\{q\in M_{\mathbb{R}}\,\middle|\,\ell_{0}(q)=1\right\}.

Then, seeing HH and HH^{\vee} as linear spaces by choosing origins 0,q0\ell_{0},q_{0}, the pairing H×H(,q)(q)1H\times H^{\vee}\ni(\ell,q)\mapsto\ell(q)-1 is nondegenerate and the polyhedrons HΣH\cap\Sigma and HΣH^{\vee}\cap\Sigma^{\vee} are each others polar sets as in [13, 1.5].

Since 0Zp0\in Z_{p}, we have

ΣZp(HΣZp)×HΣZp.\Sigma\setminus Z_{p}\,\sim\,(H\cap\Sigma\setminus Z_{p})\times\mathbb{R}\,\sim\,H\cap\Sigma\setminus Z_{p}.

By the assumptions made above, there is an ZpΣ\ell\in Z_{p}\cap\Sigma^{\circ}. Both ΣH\Sigma\cap H and ZpHZ_{p}\cap H are compact convex polyhedrons in HH. Projection away from \ell onto (HΣ)\partial(H\cap\Sigma) then induces a homotopy equivalence

HΣZpHΣZp.H\cap\Sigma\setminus Z_{p}\sim H\cap\partial\Sigma\setminus Z_{p}.

By projection, we mean that any element in a ray r=+>0Hr=\ell+\mathbb{R}_{>0}\ell^{\prime}\subset H maps to the unique element in r(HΣ)r\cap\partial(H\cap\Sigma). By section 7item (i), this has the subset

{Hσ|σf,HσZp=}\cup\left\{H\cap\sigma\,\middle|\,\sigma\in\triangle_{f}^{*},\,H\cap\sigma\cap Z_{p}=\emptyset\right\}

as a strong deformation retract. All this yields

(7.1) ΣZp{Hσ|σf,σZp={0}}.\Sigma\setminus Z_{p}\sim\cup\left\{H\cap\sigma\,\middle|\,\sigma\in\triangle_{f}^{*},\,\sigma\cap Z_{p}=\{0\}\right\}.

Using a projection, this time onto Σ\partial\Sigma^{\vee} in MM, having as center any element in (ΣΓ+(xpf))(\Sigma^{\vee}\cap\Gamma_{+}(x^{p}f))^{\circ}, we get a homotopy equivalence

Γ+(xpf)ΣΓ+(xpf)Σ.\Gamma_{+}(x^{p}f)\setminus\Sigma^{\vee}\sim\Gamma_{+}(x^{p}f)^{\circ}\cap\partial\Sigma^{\vee}.

By section 7item (ii), we have a homotopy equivalence

Γ+(xpf)Σ{(σΣ)|σΣ,σ{0},(σΣ)Γ+(xpf)}.\Gamma_{+}(x^{p}f)^{\circ}\cap\partial\Sigma^{\vee}\sim\cup\left\{(\sigma^{\perp}\cap\Sigma^{\vee})^{\circ}\,\middle|\,\sigma\in\triangle_{\Sigma},\,\sigma\neq\{0\},\,(\sigma^{\perp}\cap\Sigma^{\vee})^{\circ}\cap\Gamma_{+}(x^{p}f)^{\circ}\neq\emptyset\right\}.

Since, by assumption made above, 0Γ+(xpf)0\notin\Gamma_{+}(x^{p}f)^{\circ}, and so the right hand side above has a free action by >0\mathbb{R}_{>0} which has a section given by intersection with HH^{\vee}. Furthermore, one checks that if σf\sigma\in\triangle_{f}^{*}, then

(σΣ)Γ+(xpf)σ{0}:(p)+m<0.(\sigma^{\perp}\cap\Sigma^{\vee})^{\circ}\cap\Gamma_{+}(x^{p}f)^{\circ}\neq\emptyset\quad\Leftrightarrow\quad\forall\ell\in\sigma\setminus\{0\}:\,\ell(p)+m_{\ell}<0.

Here, the condition on the left is equivalent to σZp={0}\sigma\cap Z_{p}=\{0\}, so

(7.2) Γ+(xpf)Σ{H(σΣ)|σf,σZp={0}}.\Gamma_{+}(x^{p}f)\setminus\Sigma^{\vee}\sim\cup\left\{H\cap(\sigma^{\perp}\cap\Sigma^{\vee})^{\circ}\,\middle|\,\sigma\in\triangle_{f}^{*},\,\sigma\cap Z_{p}=\{0\}\right\}.

Now, consider the CW structure KK given by the cells HσH\cap\sigma in HΣH\cap\partial\Sigma and KK^{\prime} given by cells H(σΣ)H^{\vee}\cap(\sigma^{\perp}\cap\Sigma^{\vee}) in HΣH^{\vee}\cap\partial\Sigma^{\vee}. Using barycentric subdivision, one obtains a homeomorphism ϕ:HΣHΣ\phi:H\cap\partial\Sigma\to H^{\vee}\cap\partial\Sigma^{\vee}, sending the center of a cell HσH\cap\sigma to the center of the dual cell HσH\cap\sigma^{\vee}, thus identifying KK with the dual of KK^{\prime}. By this identification, the left hand side of eq. 7.2 is a regular neighbourhood around the image under ϕ\phi of the left hand side of eq. 7.1. This concludes (i).

Next, we prove (ii). By the above discussion, the result is clear in the cases when Zp=Z_{p}=\emptyset or ZpΣZ_{p}\subset\partial\Sigma. Assuming that this is not the case, the complex, say, AA, on the right hand side of eq. 7.1 is a closed subset of HΣSd1H\cap\partial\Sigma\sim S^{d-1}. Then hd1(A,)=0h^{d-1}(A,\mathbb{C})=0, unless A=HΣA=H\cap\partial\Sigma, in which case hd1(A,)=1h^{d-1}(A,\mathbb{C})=1. But this is equivalent to (p)+m<0\ell(p)+m_{\ell}<0 for all Σ0\ell\in\partial\Sigma\setminus 0, that is, 0Γ+(xpf)0\in\Gamma_{+}^{*}(x^{p}f)^{\circ}.

For (iii), we will show that Γ+(xpf)Σ\Gamma_{+}(x^{p}f)\setminus\Sigma^{\vee} has trivial homology in degrees i<d1i<d-1 for all pMp\in M. By assumption, there is a qMq\in M_{\mathbb{Q}} so that for σΣ(1)\sigma\in\triangle_{\Sigma}^{(1)} we have mσ=σ(q)m_{\sigma}=\ell_{\sigma}(q). We can again assume that 0Γ+(xpf)Γ(xpf)0\in\Gamma_{+}(x^{p}f)\setminus\Gamma(x^{p}f). We must show that h~i(A,)=0\widetilde{h}^{i}(A,\mathbb{C})=0 for i<d1i<d-1, where AA is the right hand side of eq. 7.1. We note that by definition, AA consists of cells HσH\cap\sigma for σf\sigma\in\triangle_{f}^{*} satisfying Hσ:(p)<m\forall\ell\in H\cap\sigma:\,\ell(p)<-m_{\ell}. Define similarly

AΣ={Hσ|σΣ,Hσ:(p)<(q)}.A_{\Sigma}=\cup\left\{H\cap\sigma\,\middle|\,\sigma\in\triangle_{\Sigma}^{*},\forall\ell\in H\cap\sigma:\,\ell(p)<-\ell(q)\right\}.

Define

Aq={HΣ|(p)<(q)}.A_{q}=\left\{\ell\in H\cap\partial\Sigma\,\middle|\,\ell(p)<-\ell(q)\right\}.

This space can be either Sd1S^{d-1}, an d1d-1 dimensional ball, or empty. In each case, H~i(Aq,)=0\widetilde{H}^{i}(A_{q},\mathbb{C})=0 for i<d1i<d-1. We will show that AqAΣAA_{q}\supset A_{\Sigma}\subset A, and that these inclusions are homotopy equivalences. For the first one, in fact, this is clear by definition and section 7(i).

For the second one, denote by AΣiA_{\Sigma}^{i} the ii-skeleton of the complex AΣA_{\Sigma}, and define similarly

Ai=A{σ|σΣ(i+2)}.A^{i}=A\setminus\cup\left\{\sigma^{\circ}\,\middle|\,\sigma\in\triangle_{\Sigma}^{*(\geq i+2)}\right\}.

We will prove by induction on ii that AΣiAiA^{i}_{\Sigma}\subset A^{i} and that this is a homotopy equivalence. The case i=0i=0 follows from the pointed condition: assuming σΣ(1)\sigma\in\triangle_{\Sigma}^{*(1)} is a ray, there is a t>0t>0 so that Hσ={tσ}H\cap\sigma=\{t\ell_{\sigma}\}. By assumption, we have mσ=σ(q)m_{\ell_{\sigma}}=\ell_{\sigma}(q), so that HσAΣH\cap\sigma\subset A_{\Sigma} if and only if HσAH\cap\sigma\subset A. Since A0A^{0} consists only of such zero-cells, we get AΣ0=A0A_{\Sigma}^{0}=A^{0}.

Next, assume that for some i>0i>0 we have an inclusion AΣi1Ai1A^{i-1}_{\Sigma}\subset A^{i-1} which is a homotopy equivalence. Let σΣ(i+1)\sigma\in\triangle_{\Sigma}^{*(i+1)} provide an ii-cell HσH\cap\sigma in AΣA_{\Sigma}. In this case, we want to show that HσAiH\cap\sigma\subset A^{i}. In fact, we have (Hσ)AΣi1\partial(H\cap\sigma)\subset A^{i-1}_{\Sigma}, hence (Hσ)Ai1\partial(H\cap\sigma)\subset A^{i-1}, by induction. But by the rationality assumption on the transverse type, it follows from (ii) and section 3 that we must have σAi\sigma\subset A^{i}, thus AΣiAiA_{\Sigma}^{i}\subset A^{i}.

To show that this inclusion is a homotopy equivalence, let σf(i+1)\sigma\in\triangle_{f}^{*(i+1)} provide an ii-cell HσH\cap\sigma which is not in AΣiA_{\Sigma}^{i}. By definition, we see that σAi\sigma\not\subset A^{i} as well. In fact, similarly as in the proof of (i), the inclusion (Hσ)Ai(Hσ)Ai\partial(H\cap\sigma)\cap A^{i}\subset(H\cap\sigma)\cap A^{i} is a strong deformation retract. Since these cells, along with AΣiA_{\Sigma}^{i} provide a finite closed covering, these glue together to form a strong deformation retract AiAΣiA^{i}\to A_{\Sigma}^{i}. ∎

\thelemma Lemma.

Let K,LNK,L\subset\mathbb{R}^{N}. Assume that KK is given as a finite disjoint union K=αIKαK=\cup_{\alpha\in I}K_{\alpha} of relatively open convex polyhedrons KαK_{\alpha}, i.e. each KαK_{\alpha} is given by a finite number of affine equations and strict inequalities. Furthermore, assume the following two conditions:

  • If FF is the face of K¯α\overline{K}_{\alpha} for some α\alpha, then F=K¯βF=\overline{K}_{\beta} for some β\beta.

  • For any α,β\alpha,\beta, the intersection K¯αK¯β\overline{K}_{\alpha}\cap\overline{K}_{\beta} is a face of both K¯α\overline{K}_{\alpha} and K¯β\overline{K}_{\beta}.

Note that the polyhedrons KαK_{\alpha} may be unbounded. In this case

  1. (i)

    Assume that KK is compact and LL is convex. Then the inclusion

    (7.3) αI{K¯α|K¯αL=}KL\bigcup_{\alpha\in I}\left\{\overline{K}_{\alpha}\,\middle|\,\overline{K}_{\alpha}\cap L=\emptyset\right\}\subset K\setminus L

    is a strong deformation retract.

  2. (ii)

    Assume that LL is convex. Then the inclusion

    αI{Kα|KαL}KL\bigcup_{\alpha\in I}\left\{K_{\alpha}\,\middle|\,K_{\alpha}\cap L\neq\emptyset\right\}\subset K\cap L

    is a strong deformation retract.

Proof.

We prove item (i), similar arguments work for (ii). We use induction on the number of α\alpha with KαLK_{\alpha}\cap L\neq\emptyset. Indeed, if this number is zero, then the inclusion in eq. 7.3 is an equality.

Otherwise, there is an α0\alpha_{0} with Kα0LK_{\alpha_{0}}\cap L\neq\emptyset. Define

I={αI|K¯αK¯α0}I,K=αIKα.I^{\prime}=\left\{\alpha\in I\,\middle|\,\overline{K}_{\alpha}\not\supset\overline{K}_{\alpha_{0}}\right\}\subsetneq I,\quad K^{\prime}=\cup_{\alpha\in I^{\prime}}K_{\alpha}.

Then the left hand side of eq. 7.3 does not change if we replace II by II^{\prime}. Therefore, using the induction hypothesis, it is enough to show that the inclusion KLKLK^{\prime}\setminus L\subset K\setminus L is a homotopy equivalence. We do this by constructing a deformation retract h:KL×[0,1]KLh:K\setminus L\times[0,1]\to K\setminus L. For this, we use the finite closed covering K¯αL\overline{K}_{\alpha}\setminus L, αI\alpha\in I of KLK\setminus L. It is then enough to define the restriction hαh_{\alpha} of hh to (K¯αL)×[0,1](\overline{K}_{\alpha}\setminus L)\times[0,1] for αI\alpha\in I in such a way that these definitions coincide on intersections.

For any αI\alpha\in I^{\prime}, we define hα(x,t)=xh_{\alpha}(x,t)=x. Let qKα0Lq\in K_{\alpha_{0}}\cap L. If αII\alpha\in I\setminus I^{\prime}, then qK¯αq\in\overline{K}_{\alpha}, and we define hαh_{\alpha} by projecting away from qq, that is, for any xKαx\in K_{\alpha} there is a unique yy in the intersection of K¯αKα0\partial\overline{K}_{\alpha}\setminus K_{\alpha_{0}} and they ray starting at qq passing through xx. We define hα(x,t)=(1t)x+tyh_{\alpha}(x,t)=(1-t)x+ty. One readily verifies that these functions are continuous, agree on intersections of their domains and define a strong deformation retract. ∎

8. Canonical divisors and cycle

In this section we describe possible canonical divisors for Y~=Y~f\widetilde{Y}=Y_{\widetilde{\triangle}_{f}} and X~\widetilde{X}. Furthermore, in the case d=2d=2, we give a formula for the canonical cycle.

\thedefinition Definition.

Let X~X\widetilde{X}\to X be a resolution of singularities of an (r1)(r-1)-dimensional singularity. A canonical divisor KX~K_{\widetilde{X}} on X~\widetilde{X} is any divisor satisfying 𝒪X~(KX~)ΩX~r1\mathcal{O}_{\widetilde{X}}(K_{\widetilde{X}})\cong\Omega_{\widetilde{X}}^{r-1}.

If r=3r=3 then let E=v𝒱EvE=\cup_{v\in\mathcal{V}}E_{v} be the exceptional divisor of a resolution X~X\widetilde{X}\to X, where EvE_{v} are the irreducible components of EE. Recall that we denoted by LL the lattice of integral cycles in X~\widetilde{X} supported on the exceptional divisor EE: that is, L=Ev|v𝒱L=\mathbb{Z}\left\langle E_{v}\,\middle|\,v\in\mathcal{V}\right\rangle. We also set L=LL_{\mathbb{Q}}=L\otimes\mathbb{Q} and

L=Hom(L,){lL|lL:(l,l)},L^{\prime}=\mathop{\rm Hom}\nolimits(L,\mathbb{Z})\cong\left\{l^{\prime}\in L_{\mathbb{Q}}\,\middle|\,\forall l\in L:\,(l^{\prime},l)\in\mathbb{Z}\right\},

where (,)(\cdot,\cdot) denotes the intersection form, extended linearly to LL_{\mathbb{Q}}. Moreover, set EvLE^{*}_{v}\in L^{\prime} for the unique rational cycle satisfying (Ev,Ev)=1(E_{v},E_{v}^{*})=-1 and (Ew,Ev)=0(E_{w},E_{v}^{*})=0 for wvw\neq v.

In this surface singularity case the canonical cycle ZKLZ_{K}\in L^{\prime} is the unique rational cycle on X~\widetilde{X} supported on the exceptional divisor, satisfying the adjunction formula

(Ev,ZK)=bv+22gv(E_{v},Z_{K})=-b_{v}+2-2g_{v}

for any irreducible component EvE_{v} of the exceptional divisor, where bv-b_{v} is the Euler number of the normal bundle of EvX~E_{v}\subset\widetilde{X}, and gvg_{v} is the genus of EvE_{v} (we assume here that the components EvE_{v} of the exceptional divisor are smooth).

\therem Remark.

The cycles ZKZ_{K} and EvE_{v}^{*} are well defined, since the intersection matrix, with entries (Ev,Ew)(E_{v},E_{w}), associated with any resolution is negative definite. Notice also that any two canonical divisors are linearly equivalent, and that any canonical divisor KK is numerically equivalent to ZK-Z_{K}. However, it can happen that 𝒪X~(KX~+ZK)\mathcal{O}_{\widetilde{X}}(K_{\widetilde{X}}+Z_{K}) has infinite order in the Picard group.

\theprop Proposition.

Fix any rr. Let (X,0)(Y,0)(X,0)\subset(Y,0) be a Newton nondegenerate Weil divisor, and ~f\widetilde{\triangle}_{f} a subdivision of the normal fan f\triangle_{f} so that Y~Y\widetilde{Y}\to Y is an embedded resolution. Then the divisors

(8.1) KY~=σ~f(1)DσDiv(Y~),KX~=σ~f(1)(1+mσ)EσDiv(X~)K_{\widetilde{Y}}=-\sum_{\sigma\in\widetilde{\triangle}^{(1)}_{f}}D_{\sigma}\in\mathop{\rm Div}\nolimits(\widetilde{Y}),\quad K_{\widetilde{X}}=-\sum_{\sigma\in\widetilde{\triangle}^{(1)}_{f}}(1+m_{\sigma})E_{\sigma}\in\mathop{\rm Div}\nolimits(\widetilde{X})

are possible canonical divisors for Y~\widetilde{Y} and X~\widetilde{X}, respectively.

Furthermore, in the surface case (r=3r=3), the canonical cycle on X~\widetilde{X} is given by the formula

(8.2) ZKE=wt(f)(mn+1)Ev,Z_{K}-E=\mathop{\rm wt}\nolimits(f)-\sum(m_{n}+1)E_{v}^{*},

where the sum to the right runs through edges {n,v}\{n,v\} in the graph GG^{*} so that n𝒩𝒩n\in\mathcal{N}^{*}\setminus\mathcal{N} and v𝒱v\in\mathcal{V} (and the identity is in LL).

Proof.

For KY~K_{\widetilde{Y}}, see e.g. 4.3 of [13]. Since the divisor X~+σ~f(1)mσDσ=(πf)\widetilde{X}+\sum_{\sigma\in\widetilde{\triangle}_{f}^{(1)}}m_{\sigma}D_{\sigma}=(\pi^{*}f) is principal in Y~\widetilde{Y} (and Dσ|X~=EσD_{\sigma}|_{\widetilde{X}}=E_{\sigma}), the adjunction formula gives

KX~=(KY~+X~)|X~=σ~f(1)(mσ+1)Eσ,K_{\widetilde{X}}=\left.\left(K_{\widetilde{Y}}+\widetilde{X}\right)\right|_{\widetilde{X}}=-\sum_{\sigma\in\widetilde{\triangle}_{f}^{(1)}}(m_{\sigma}+1)E_{\sigma},

which proves eq. 8.1. To prove eq. 8.2, it is enough to show that in LL for all v𝒱v\in\mathcal{V},

(8.3) (ZKE,Ev)=(wt(f)(mn+1)Ev,Ev),(Z_{K}-E,E_{v})=\left(\mathop{\rm wt}\nolimits(f)-\sum(m_{n}+1)E_{v}^{*},E_{v}\right),

where the sum is as in eq. 8.2. Recall that wt(f)=v𝒱mvEv\mathop{\rm wt}\nolimits(f)=\sum_{v\in\mathcal{V}}m_{v}E_{v}. We note that the adjunction formula gives (ZKE,Ev)=22gvδv(Z_{K}-E,E_{v})=2-2g_{v}-\delta_{v} for all v𝒱v\in\mathcal{V}, where δv\delta_{v} is the valency of the vertex vv in GG, and gvg_{v} is the genus of EvE_{v}. Furthermore, it follows from section 6 that if v𝒱v\in\mathcal{V}, then

  • δv=1\delta_{v}=1 if and only vv is on the end of a bamboo joining a node n𝒩n\in\mathcal{N} and an extended node n𝒩𝒩n^{\prime}\in\mathcal{N}^{*}\setminus\mathcal{N}. In this case, vv has exactly one neighbour in 𝒱𝒱\mathcal{V}^{*}\setminus\mathcal{V} in the graph GG^{*}.

  • δv=2\delta_{v}=2 if and only vv is on a bamboo joining two extended nodes, and is not of the form described in the previous item.

  • δv3\delta_{v}\geq 3 if and only if vv is a node.

Consider first the case δv=1\delta_{v}=1, and let nn be the unique neighbour of vv in 𝒩𝒩\mathcal{N}^{*}\setminus\mathcal{N}. It follows from section 6 that (wt(f),Ev)=mn(\mathop{\rm wt}\nolimits(f),E_{v})=-m_{n}, since FvF_{v} is a segment, and so has area zero. As a result, the right hand side of eq. 8.3 is 1=(ZKE,Ev)1=(Z_{K}-E,E_{v}).

Next, assume that δv=2\delta_{v}=2. Then both sides of eq. 8.3 vanish (use again section 6).

Assume finally that v𝒩v\in\mathcal{N}. Then, vv has no neighbours in 𝒩𝒩\mathcal{N}^{*}\setminus\mathcal{N}. Furthermore, δv\delta_{v} coincides with the number of integral points on the boundary of FvF_{v}, since each edge adjacent to vv can be seen to correspond to a primitive segment of the boundary. By using Pick’s theorem and section 6, we therefore get

(ZKE,Ev)=22gvδv=2Vol2(Fv)=(Ev,wt(f)),(Z_{K}-E,E_{v})=2-2g_{v}-\delta_{v}=-2\mathop{\rm Vol}\nolimits_{2}(F_{v})=(E_{v},\mathop{\rm wt}\nolimits(f)),

which finishes the proof. ∎

\therem Remark.

As mσm_{\sigma} depends on the choice of ff up to a xpx^{p} multiplication, the right hand side of the second formula from eq. 8.1 depends on this choice too. In fact, the monomial rational function xpx^{p} realizes the linear equivalence between the two divisors KX~K_{\widetilde{X}} associated with two such choices.

9. Gorenstein surface singularities

In this section we prove theorem 9.1, which characterizes nondegenerate normal surface Gorenstein singularities by their Newton polyhedron. The key technical sections 9 and 9 provide the tools for the proof. They are proved using vanishing of certain cohomology groups calculated by toric methods. In the first lemma, the restriction r=3r=3 is not needed. However, the second lemma relies on the negative definiteness of the intersection form, restricting our result to the surface case.

\thedefinition Definition.

Let ff and f\triangle_{f} be as above. We say that Γ+(f)\Gamma_{+}(f), or ff, is (\mathbb{Q}-)Gorenstein-pointed if there exists a pMp\in M (pMp\in M_{\mathbb{Q}}) such that σ(p)=mσ+1\ell_{\sigma}(p)=m_{\sigma}+1 for all σf(1,1)\sigma\in\triangle^{*(1,1)}_{f}.

\theexample Example.

Recall that (Y,0)(Y,0) is Gorenstein if and only if there is a pMp\in M satisfying σ(p)=1\ell_{\sigma}(p)=1 for all σΣ(1)\sigma\in\triangle_{\Sigma}^{(1)}, see e.g. [8], Theorem 6.32. Therefore, if (X,0)(X,0) is Cartier, and f=Σ\triangle_{f}^{*}=\triangle_{\Sigma}^{*}, then ff is Gorenstein pointed (since mσ=0m_{\sigma}=0 for σΣ(1)\sigma\in\triangle_{\Sigma}^{(1)}). Furthermore, (X,0)(X,0) is Gorenstein since (Y,0)(Y,0) is Gorenstein and ff forms a regular sequence.

Similarly, (Y,0)(Y,0) is \mathbb{Q}-Gorenstein if there is a pMp\in M_{\mathbb{Q}} satisfying σ(p)=1\ell_{\sigma}(p)=1 for all σΣ(1)\sigma\in\triangle_{\Sigma}^{(1)}, see e.g. [1]. Therefore, if (X,0)(X,0) is Cartier, and f=Σ\triangle_{f}^{*}=\triangle_{\Sigma}^{*}, then ff is \mathbb{Q}-Gorenstein pointed.

\therem Remark.

Though the two combinatorial conditions in definitions 3 and 9 look very similar, they codify two rather different geometrical properties. Being ‘pointed’ codifies an embedding property, namely that (X,0)(Y,0)(X,0)\subset(Y,0) is Cartier, see section 3. However, being ‘Gorenstein pointed’ codifies an abstract property of the germ (X,0)(X,0), namely its Gorenstein property, see theorem 9.1 below.

\theblock.

Recall also that (X,0)(X,0) is Gorenstein if it admits a Gorenstein form. A Gorenstein form is a nowhere vanishing section in H0(X0,ΩX02)=H0(X~E,ΩX~E2)H^{0}(X\setminus 0,\Omega^{2}_{X\setminus 0})=H^{0}(\widetilde{X}\setminus E,\Omega^{2}_{\widetilde{X}\setminus E}). A Gorenstein pluri-form is a nowhere vanishing section in H0(X~E,(ΩX~E2)k)H^{0}(\widetilde{X}\setminus E,(\Omega^{2}_{\widetilde{X}\setminus E})^{\otimes k}) for some k>0k\in\mathbb{Z}_{>0}.

In this section KY~K_{\widetilde{Y}} and KX~K_{\widetilde{X}} are canonical divisors with a choice as in eq. 8.1.

\thedefinition Definition.

Let ωf\omega_{f} be some meromorphic 2-form on X~\widetilde{X} whose divisor (ωf)(\omega_{f}) is KX~K_{\widetilde{X}}.

9.1 Theorem.

Assume that (X,0)(Y,0)(X,0)\subset(Y,0) is a normal Newton nondegenerate surface singularity (i.e. r=3r=3). The following conditions are equivalent:

  1. (i)

    ff is Gorenstein-pointed at some pMp\in M.

  2. (ii)

    There exists a pMp\in M so that for all v𝒱𝒱v\in\mathcal{V}^{*}\setminus\mathcal{V} we have v(p)=mv+1\ell_{v}(p)=m_{v}+1.

  3. (iii)

    There exists a pMp\in M so that for all v𝒱v\in\mathcal{V} we have v(p)=mv+1mv(ZK)\ell_{v}(p)=m_{v}+1-m_{v}(Z_{K}).

  4. (iv)

    There exists a pMp\in M so that xpωfx^{p}\omega_{f} is a Gorenstein form.

  5. (v)

    (X,0)(X,0) is Gorenstein.

When these conditions hold, (i), (ii), (iii) and (iv) uniquely identify the same point pp.

In fact, the analogues of parts item (i)item (iv) are equivalent over rational points pMp\in M_{\mathbb{Q}} as well.

\theprop Proposition.

Under the assumption of theorem 9.1, the following conditions are equivalent, and imply that (X,0)(X,0) is \mathbb{Q}-Gorenstein:

  1. (i)

    ff is \mathbb{Q}-Gorenstein-pointed at some pMp\in M_{\mathbb{Q}}.

  2. (ii)

    There exists a pMp\in M_{\mathbb{Q}} so that for all v𝒱𝒱v\in\mathcal{V}^{*}\setminus\mathcal{V} we have v(p)=mv+1\ell_{v}(p)=m_{v}+1.

  3. (iii)

    There exists a pMp\in M_{\mathbb{Q}} so that for all v𝒱v\in\mathcal{V} we have v(p)=mv+1mv(ZK)\ell_{v}(p)=m_{v}+1-m_{v}(Z_{K}).

  4. (iv)

    There exists a pMp\in M_{\mathbb{Q}} so that xkp(ωf)kx^{kp}(\omega_{f})^{\otimes k} is a Gorenstein pluri-form for some k>0k\in\mathbb{Z}_{>0}.

Furthermore, all these these conditions identify the very same pp uniquely.

Proof.

(ii) is a rephrasing of (i), since f(1,1)=𝒱𝒱\triangle^{*(1,1)}_{f}=\mathcal{V}^{*}\setminus\mathcal{V}.

(ii)\Rightarrow(iii) For any pMp\in M_{\mathbb{Q}} consider the cycles

Z1:=v𝒱v(p)EvL,Z2:=(mn+1)EvL,Z_{1}:=\sum_{v\in\mathcal{V}}\ell_{v}(p)E_{v}\in L_{\mathbb{Q}},\qquad Z_{2}:=\sum(m_{n}+1)E_{v}^{*}\in L_{\mathbb{Q}},

where the sum runs over edges {n,v}\{n,v\} in GG^{*} so that n𝒱n\in\mathcal{V}^{*} and v𝒱v\in\mathcal{V} (as in eq. 8.2), and Z:=n𝒱𝒱n(p)EnZ^{*}:=\sum_{n\in\mathcal{V}^{*}\setminus\mathcal{V}}\ell_{n}(p)E_{n} (where all these EnE_{n}’s are the noncompact curves in X~\widetilde{X}).

If {n,v}\{n,v\} is an edge as above, then (Z2,Ev)=(mn+1)(Z_{2},E_{v})=-(m_{n}+1). Moreover, (Z,Ev)X~=n(p)(Z^{*},E_{v})_{\widetilde{X}}=\ell_{n}(p). Therefore, by assumption (ii), (Z+Z2,Eu)X~=0(Z^{*}+Z_{2},E_{u})_{\widetilde{X}}=0 for any u𝒱u\in\mathcal{V}. On the other hand, by section 6, (Z+Z1,Eu)X~=0(Z^{*}+Z_{1},E_{u})_{\widetilde{X}}=0 for any u𝒱u\in\mathcal{V} as well. Hence Z1=Z2Z_{1}=Z_{2}. But by eq. 8.2 mu(Z2)=mv+1mv(ZK)m_{u}(Z_{2})=m_{v}+1-m_{v}(Z_{K}).

(iii)\Rightarrow(ii) With the above notations, (iii) shows that Z1=Z2Z_{1}=Z_{2}. Let {n,v}\{n,v\} be an edge as above, let w𝒱w\in\mathcal{V} be the other neighbour of vv, and note that Ev=bvEvEwE_{v}=b_{v}E^{*}_{v}-E^{*}_{w} in LL^{\prime}. Then,

mn+1=(Z2,Ev)=(Z1,bvEv+Ew)=v(p)bvw(p)=n(p)m_{n}+1=(Z_{2},-E_{v})=(Z_{1},-b_{v}E^{*}_{v}+E^{*}_{w})=\ell_{v}(p)b_{v}-\ell_{w}(p)=\ell_{n}(p)

(in the last equality use section 6).

For (ii)\Leftrightarrow(iv) use the second identity of eq. 8.1. ∎

\therem Remark.

Similarly as in theorem 9.1, one may ask whether the equivalent cases in 9 are equivalent with the property that (X,0)(X,0) is \mathbb{Q}-Gorenstein. If ff is \mathbb{Q}-Gorenstein-pointed at pMp\in M_{\mathbb{Q}}, then (iv) implies that (X,0)(X,0) is \mathbb{Q}-Gorenstein. The converse does not hold, as seen by the following example.

Let N=3N=\mathbb{Z}^{3} and

Σ=0(1,0,0),(0,1,0),(1,0,1),(0,1,1),f(x)=x(0,0,2)+x(1,0,1)+x(0,2,0)+2x(1,2,1).\Sigma=\mathbb{R}_{\geq 0}\langle(1,0,0),(0,1,0),(1,0,1),(0,1,1)\rangle,\quad f(x)=x^{(0,0,2)}+x^{(1,0,1)}+x^{(0,2,0)}+2x^{(1,2,-1)}.

Write σi\sigma_{i}, i=1,2,3,4i=1,2,3,4 for the rays generated by the vector specified above and denote by mim_{i} the corresponding multiplicities. We find m1=m2=m3=0m_{1}=m_{2}=m_{3}=0 and m4=1m_{4}=1. As a result, since the linear equation

(100010101011)p=(1112)\left(\begin{matrix}1&0&0\\ 0&1&0\\ 1&0&1\\ 0&1&1\end{matrix}\right)\cdot p=\left(\begin{matrix}1\\ 1\\ 1\\ 2\end{matrix}\right)

has no solution, hence ff is not \mathbb{Q}-Gorenstein pointed.

Refer to caption
(0,2,0)(0,2,0)(0,0,2)(0,0,2)(1,0,1)(1,0,1)(1,2,1)(1,2,-1)(1,0,0)(1,0,0)(0,1,0)(0,1,0)(0,1,1)(0,1,1)(1,0,1)(1,0,1)(1,1,1)(1,1,1)3-3GG^{*}GGΓ(f)\Gamma(f)
Figure 5. A Newton diagram, and the output of Oka’s algorithm. The dotted line shows the intersection of the affine hull of the only face of the diagram intersected with Σ\partial\Sigma^{\vee}. For simplicity, here in GG^{*} we have blown down the (1)(-1)-vertices constructed in the last paragraph of section 4.

On the other hand, one verifies that the Weil divisor defined by ff is normal using theorem 7.1. Furthermore, Oka’s algorithm shows that this singularity has a resolution with an exceptional divisor consisting of a single rational curve with Euler number 3-3. Such a singularity is a cyclic quotient singularity. In particular, it is \mathbb{Q}-Gorenstein.

\theblock.

Next, we focus on the proof of theorem 9.1. The equivalences of the first four cases follow from (or, as) section 9. For item (i)\Rightarrowitem (v) note that if ff is Gorenstein-pointed at pMp\in M then xpωfx^{p}\omega_{f} trivializes the canonical bundle. The implication item (v)\Rightarrowitem (i) will be proved below based on two lemmas.

\thelemma Lemma.

Let g¯H0(X~E,𝒪X~(KX~))\,\overline{g}\in H^{0}(\widetilde{X}\setminus E,\mathcal{O}_{\widetilde{X}}(K_{\widetilde{X}})), that is, g¯\overline{g} is a meromorphic function on the complement of the exceptional divisor in X~\widetilde{X} satisfying

(9.1) (g¯)KX~|X~E=v𝒱𝒱(mv+1)Ev.(\overline{g})\geq-K_{\widetilde{X}}|_{\widetilde{X}\setminus E}=\sum_{v\in\mathcal{V}^{*}\setminus\mathcal{V}}(m_{v}+1)E_{v}.

Then, there exists a Laurent series g𝒪Y,0[xM]g\in\mathcal{O}_{Y,0}[x^{M}] satisfying (πg)|X~E=g¯(\pi^{*}g)|_{\widetilde{X}\setminus E}=\overline{g} and

(9.2) σ~f(1):wtσgmσ+1.\forall\sigma\in\widetilde{\triangle}_{f}^{*(1)}:\,\mathop{\rm wt}\nolimits_{\sigma}g\geq m_{\sigma}+1.
Proof.

Let I=H0(X~E,𝒪X~(KX~))I=H^{0}(\widetilde{X}\setminus E,\mathcal{O}_{\widetilde{X}}(K_{\widetilde{X}})) and let JJ be the set of meromorphic functions obtained as a restriction of Laurent series satisfying eq. 9.2. We want to show that I=JI=J.

We immediately see JIJ\subset I. In fact, this inclusion fits into an exact sequence as follows. Recall the notation Dm=σ~f(1)mσDσD_{m}=\sum_{\sigma\in\widetilde{\triangle}_{f}^{(1)}}m_{\sigma}D_{\sigma} from the proof of theorem 7.1, and KY~=σ~f(1)DσK_{\widetilde{Y}}=-\sum_{\sigma\in\widetilde{\triangle}_{f}^{(1)}}D_{\sigma}. Also, define DcD_{\mathrm{c}} as the union of compact divisors in Y~\widetilde{Y}, that is, σDσ\cup_{\sigma}D_{\sigma} for σ~f(1)\sigma\not\in\widetilde{\triangle}_{f}^{*(1)}. Since (πf)=X~+Dm(\pi^{*}f)=\widetilde{X}+D_{m}, we have a short exact sequence of sheaves

0𝒪Y~Dc(KY~)f𝒪Y~Dc(Dm+KY~)𝒪X~E(Dm+KY~)00\to\mathcal{O}_{\widetilde{Y}\setminus D_{\mathrm{c}}}(K_{\widetilde{Y}})\stackrel{{\scriptstyle\cdot f}}{{\to}}\mathcal{O}_{\widetilde{Y}\setminus D_{\mathrm{c}}}(-D_{m}+K_{\widetilde{Y}})\to\mathcal{O}_{\widetilde{X}\setminus E}(-D_{m}+K_{\widetilde{Y}})\to 0

yielding a long exact sequence of cohomology groups. We have

I=H0(X~E,𝒪X~E(Dm+KY~)),I=H^{0}(\widetilde{X}\setminus E,\mathcal{O}_{\widetilde{X}\setminus E}(-D_{m}+K_{\widetilde{Y}})),

since KX~=(Dm+KY~)|X~K_{\widetilde{X}}=(-D_{m}+K_{\widetilde{Y}})|_{\widetilde{X}}. Furthermore, since Y~\widetilde{Y} is normal, H0(Y~Dc,𝒪Y~Dc(Dm+KY~))H^{0}(\widetilde{Y}\setminus D_{\mathrm{c}},\mathcal{O}_{\widetilde{Y}\setminus D_{\mathrm{c}}}(-D_{m}+K_{\widetilde{Y}})) is the set of Laurent series satisfying eq. 9.2. Thus, its image in II is JJ. Therefore, the quotient I/JI/J injects into H1(Y~Dc,𝒪Y~Dc(KY~))H^{1}(\widetilde{Y}\setminus D_{\mathrm{c}},\mathcal{O}_{\widetilde{Y}\setminus D_{\mathrm{c}}}(K_{\widetilde{Y}})). On the other hand,

(9.3) H1(Y~Dc,𝒪Y~Dc(KY~))pMHZ(p)1(Σ,),H^{1}(\widetilde{Y}\setminus D_{\mathrm{c}},\mathcal{O}_{\widetilde{Y}\setminus D_{\mathrm{c}}}(K_{\widetilde{Y}}))\cong\bigoplus_{p\in M}H^{1}_{Z(p)}(\partial\Sigma,\mathbb{C}),

where, following Fulton [13], ψK:Σ\psi_{K}:\partial\Sigma\to\mathbb{R} is the unique function restricting to linear function on all σ~f\sigma\in\widetilde{\triangle}_{f}^{*}, and satisfying ψK(σ)=1\psi_{K}(\ell_{\sigma})=1 for σ~f(1)\sigma\in\widetilde{\triangle}_{f}^{(1)*}, and for pMp\in M we set

Z(p)={Σ|(p)ψK()}.Z(p)=\left\{\ell\in\partial\Sigma\,\middle|\,\ell(p)\geq\psi_{K}(\ell)\right\}.

Firstly, since Σ\partial\Sigma is contractible, we find

HZ(p)1(Σ,)H~0(ΣZ(p),).H^{1}_{Z(p)}(\partial\Sigma,\mathbb{C})\cong\widetilde{H}^{0}(\partial\Sigma\setminus Z(p),\mathbb{C}).

Secondly, define Z(p)Z^{\prime}(p) as the union of those cones σ~f\sigma\in\widetilde{\triangle}_{f}^{*} satisfying p|σ0p|_{\sigma}\geq 0 (i.e. (p)0\ell(p)\geq 0 for all σ\ell\in\sigma), and let Z′′(p)Z^{\prime\prime}(p) be the set of Σ\ell\in\partial\Sigma satisfying (p)0\ell(p)\geq 0. By section 7, the inclusions

ΣZ(p)ΣZ(p)ΣZ′′(p)\partial\Sigma\setminus Z(p)\subset\partial\Sigma\setminus Z^{\prime}(p)\supset\partial\Sigma\setminus Z^{\prime\prime}(p)

are strong deformation retracts. But the right hand side above is either a contractible set, or it has the homotopy of Sr2S^{r-2}. In particular, it is connected, by our assumption r>2r>2, and so eq. 9.3 vanishes. ∎

\thelemma Lemma.

Assume that (X,0)(X,0) is a Gorenstein normal surface singularity, i.e. r=3r=3, and that we have a Gorenstein form ω\omega on X~E\widetilde{X}\setminus E. Thus, KX~ZK-K_{\widetilde{X}}-Z_{K} is linearly trivial, and there exists

g¯H0(X~,𝒪X~(KX~+ZK)),(g¯)=(ω)(ωf)=ZKKX~.\overline{g}\in H^{0}(\widetilde{X},\mathcal{O}_{\widetilde{X}}(K_{\widetilde{X}}+Z_{K})),\quad(\overline{g})=(\omega)-(\omega_{f})=-Z_{K}-K_{\widetilde{X}}.

Then there is a g𝒪Y,0[xM]g\in\mathcal{O}_{Y,0}[x^{M}] satisfying

(9.4) (πg)|X~=g¯andv𝒱:wtv(g)=divv(g¯).(\pi^{*}g)|_{\widetilde{X}}=\overline{g}\quad\mathrm{and}\quad\forall v\in\mathcal{V}^{*}:\,\mathop{\rm wt}\nolimits_{v}(g)=\mathop{\rm div}\nolimits_{v}(\overline{g}).
Proof.

By the previous section 9, we can find a gg satisfying g|X~E=g¯g|_{\widetilde{X}\setminus E}=\overline{g} and eq. 9.2. Let A=(g¯)A=(\overline{g}) and B=v𝒱wtv(g)EvB=\sum_{v\in\mathcal{V}^{*}}\mathop{\rm wt}\nolimits_{v}(g)E_{v}. We want to prove that A=BA=B. Both AA and BB are supported in the exceptional divisor and the noncompact curves EvE_{v} for v𝒱𝒱v\in\mathcal{V}^{*}\setminus\mathcal{V}, and by our assumptions, they have the same multiplicity along this noncompact part. Thus, ABA-B is supported on the exceptional divisor. Furthermore, we have wtv(g)divv(g¯)\mathop{\rm wt}\nolimits_{v}(g)\leq\mathop{\rm div}\nolimits_{v}(\overline{g}) for v𝒱v\in\mathcal{V}, thus BA0B-A\leq 0.

For the reverse inequality, note first that (A,Ev)=0(A,E_{v})=0 for all v𝒱v\in\mathcal{V} since AA is principal. For any v𝒱v\in\mathcal{V}, let qMq\in M be an element of the support of the principal part of gg with respect to v\ell_{v}, i.e. qsupp(g)q\in\mathop{\rm supp}\nolimits(g) and v(q)=wtv(g)\ell_{v}(q)=\mathop{\rm wt}\nolimits_{v}(g). By definition, we also have u(q)wtu(g)\ell_{u}(q)\geq\mathop{\rm wt}\nolimits_{u}(g) for all u𝒱vu\in\mathcal{V}^{*}_{v}. Therefore,

(B,Ev)=bvwtv(g)+{wtu(g)|u𝒱v}bvv(q)+{u(q)|u𝒱v}=0.(B,E_{v})=-b_{v}\mathop{\rm wt}\nolimits_{v}(g)+\sum\left\{\mathop{\rm wt}\nolimits_{u}(g)\,\middle|\,u\in\mathcal{V}^{*}_{v}\right\}\leq-b_{v}\ell_{v}(q)+\sum\left\{\ell_{u}(q)\,\middle|\,u\in\mathcal{V}^{*}_{v}\right\}=0.

As a result, BAB-A is in the Lipman cone, and so, BA0B-A\geq 0, proving eq. 9.4. ∎

Proof of theorem 9.1.

The first four conditions are equivalent by section 9, and item (iv) clearly implies item (v).

Assuming that (X,0)(X,0) is Gorenstein, let ω\omega be a Gorenstein form. Then there is meromorphic g¯\overline{g} so that g¯ωf=ω\overline{g}\omega_{f}=\omega on X~E\widetilde{X}\setminus E. By section 9, g¯\overline{g} is the restriction of a Laurent series g𝒪Y,0[xM]g\in\mathcal{O}_{Y,0}[x^{M}] satisfying eq. 9.4.

For any v𝒱v\in\mathcal{V}, denote by gvg_{v} the principal part of gg with respect to the weight v\ell_{v}. We make the

Claims:

  1. (a)

    For any n𝒩n\in\mathcal{N}, gng_{n} is a monomial, that is, there is a pnMp_{n}\in M so that gn=anxpng_{n}=a_{n}x^{p_{n}} for some ana_{n}\in\mathbb{C}^{*}.

  2. (b)

    If vv is a vertex on a bamboo connecting n𝒩n\in\mathcal{N} and some other node in 𝒩\mathcal{N}^{*}, then gv=anxpng_{v}=a_{n}x^{p_{n}}.

By (b), the exponent p=pnp=p_{n} does not depend on nn, finishing the proof since hence xpωfx^{p}\omega_{f} is a Gorenstein form.

(a) is proved as follows. Set qsupp(gn)q\in\mathop{\rm supp}\nolimits(g_{n}) arbitrarily. We then have wtn(g)=n(q)\mathop{\rm wt}\nolimits_{n}(g)=\ell_{n}(q), and also wtu(g)u(q)\mathop{\rm wt}\nolimits_{u}(g)\leq\ell_{u}(q), for any other uu, since supp(gn)supp(g)\mathop{\rm supp}\nolimits(g_{n})\subset\mathop{\rm supp}\nolimits(g). In particular,

bnwtn(g)+u𝒱nwtu(g)bnn(q)+u𝒱nu(q).-b_{n}\mathop{\rm wt}\nolimits_{n}(g)+\sum_{u\in\mathcal{V}_{n}}\mathop{\rm wt}\nolimits_{u}(g)\leq-b_{n}\ell_{n}(q)+\sum_{u\in\mathcal{V}_{n}}\ell_{u}(q).

The right hand side is sero since n+u𝒱nu=0\ell_{n}+\sum_{u\in\mathcal{V}_{n}}\ell_{u}=0 for n𝒩n\in\mathcal{N}. On the other hand, by the section 9, we have wtv(g)=divv(g¯)\mathop{\rm wt}\nolimits_{v}(g)=\mathop{\rm div}\nolimits_{v}(\overline{g}) for all vv, thus, the left hand side above equals (div(g),En)(\mathop{\rm div}\nolimits(g),E_{n}). Furthermore, since (g¯)=(ω)(ωf)(\overline{g})=(\omega)-(\omega_{f}), gg does not have any zeroes or poles outside the exceptional divisor, in a neighbourhood around EnE_{n}, hence (div(g¯),En)=((g),En)=0(\mathop{\rm div}\nolimits(\overline{g}),E_{n})=((g),E_{n})=0. Therefore, the inequality above is an equality, and we have wtu(g)=u(q)\mathop{\rm wt}\nolimits_{u}(g)=\ell_{u}(q) for u𝒱nu\in\mathcal{V}_{n}.

This fact is true for any choice of qq, therefore, u(q)=wtu(g)=u(q)\ell_{u}(q^{\prime})=\mathop{\rm wt}\nolimits_{u}(g)=\ell_{u}(q) for any u𝒱nu\in\mathcal{V}_{n} and for any other choice qq^{\prime}. But the vectors {u}u𝒱n\{\ell_{u}\}_{u\in\mathcal{V}_{n}} form a generator set, hence necessarily q=qq=q^{\prime}.

For (b), assume that nn and n𝒩n^{\prime}\in\mathcal{N}^{*} are joined by a bamboo, consisting of vertices v1,,vsv_{1},\ldots,v_{s}, with v1𝒱nv_{1}\in\mathcal{V}_{n} and vs𝒱nv_{s}\in\mathcal{V}_{n^{\prime}}, and vi,vi+1v_{i},v_{i+1} neighbours for i=1,,s1i=1,\ldots,s-1. For convenience, we set v0=nv_{0}=n and vs+1=nv_{s+1}=n^{\prime}. We start by showing that wti(g)=i(pn)\mathop{\rm wt}\nolimits_{i}(g)=\ell_{i}(p_{n}) using induction (we replace the subscript viv_{i} by just ii for legibility). Indeed, for i=0i=0 this is clear, and we showed in the proof of (a) that this holds for i=1i=1. For the induction step we use the recursive formulas

i+1bii+i1=0,wti+1(g)biwti(g)+wti1(g)=0.\ell_{i+1}-b_{i}\ell_{i}+\ell_{i-1}=0,\quad\mathop{\rm wt}\nolimits_{i+1}(g)-b_{i}\mathop{\rm wt}\nolimits_{i}(g)+\mathop{\rm wt}\nolimits_{i-1}(g)=0.

The first one holds by section 6, and the second one follows from wti(g)=divi(g)\mathop{\rm wt}\nolimits_{i}(g)=\mathop{\rm div}\nolimits_{i}(g) similarly as above, although for the case i=si=s, we may have to use a component of the noncompact curve EnE_{n^{\prime}}.

We now see that for any 1is1\leq i\leq s, the support of gig_{i} consists of points qMq\in M for which i(q)=i(pn)\ell_{i}(q)=\ell_{i}(p_{n}) and i±1(q)i±1(pn)\ell_{i\pm 1}(q)\geq\ell_{i\pm 1}(p_{n}). But these equations are equivalent to n(q)=n(pn)\ell_{n}(q)=\ell_{n}(p_{n}) and n(q)=n(pn)\ell_{n^{\prime}}(q)=\ell_{n^{\prime}}(p_{n}). Therefore, supp(gi)=supp(gn)\mathop{\rm supp}\nolimits(g_{i})=\mathop{\rm supp}\nolimits(g_{n}) for these ii. ∎

10. The geometric genus and the diagonal computation sequence

In this section we construct the diagonal computation sequence, and show that it computes the geometric genus of any Newton nondegenerate, \mathbb{Q}-Gorenstein pointed, normal surface singularity having a rational homology sphere link. Any computation sequence provides an upper bound for the geometric genus. The smallest such bound is a topological invariant, and we show that this is realized by this diagonal sequence. This is done by showing that the diagonal computation sequence counts the lattice points “under the diagram”, whose number is precisely the geometric genus, according to section 7.

\theblock.

Discussions regarding general normal surface singularities. Throughout this section, when not mentioned specifically, π:(X~,E)(X,0)\pi:(\widetilde{X},E)\to(X,0) denotes a resolution of a normal surface singularity (X,0)(X,0) with exceptional divisor EE, whose irreducible decomposition is E=v𝒱EvE=\cup_{v\in\mathcal{V}}E_{v}.

We assume that (X,0)(X,0) has a rational homology sphere link; thus Ev1E_{v}\cong\mathbb{CP}^{1} for all v𝒱v\in\mathcal{V}.

We use the notations LL, LL^{\prime} and EvE^{*}_{v} as in section 8. For Z=vrvEvZ=\sum_{v}r_{v}E_{v} with rvr_{v}\in\mathbb{Q} we write Z=vrvEv\lfloor Z\rfloor=\sum_{v}\lfloor r_{v}\rfloor E_{v}. ZKZ_{K} denotes the canonical cycle. Note that ZK=0Z_{K}=0 if and only if (X,0)(X,0) is an ADEADE germ. Otherwise, it is known that in the minimal resolution, or, even in the minimal good resolution, all the coefficients of ZKZ_{K} are strictly positive. However, usually this is not the case in non-minimal resolutions, i.e. in our GG it is not automatically guaranteed.

\thelemma Lemma.

In any resolution X~X\widetilde{X}\to X of a normal surface singularity with ZK0\lfloor Z_{K}\rfloor\geq 0 we have

(10.1) pg=dimH0(X~,𝒪X~(KX~+ZK))H0(X~,𝒪X~(KX~)).p_{g}=\dim_{\mathbb{C}}\frac{H^{0}(\widetilde{X},\mathcal{O}_{\widetilde{X}}(K_{\widetilde{X}}+\lfloor Z_{K}\rfloor))}{H^{0}(\widetilde{X},\mathcal{O}_{\widetilde{X}}(K_{\widetilde{X}}))}.
Proof.

By the generalized version of Grauert–Riemenschneider vanishing we have the two vanishings

(10.2) H1(X~,𝒪X~(KX~))=0,H1(X~,𝒪X~(ZK))=0.H^{1}(\widetilde{X},\mathcal{O}_{\widetilde{X}}(K_{\widetilde{X}}))=0,\ \ \ H^{1}(\widetilde{X},\mathcal{O}_{\widetilde{X}}(-\lfloor Z_{K}\rfloor))=0.

Hence, if ZK=0\lfloor Z_{K}\rfloor=0 then pg=0p_{g}=0 too. Otherwise, from the long exact sequence of cohomology groups associated with

0𝒪X~(KX~)𝒪X~(KX~+ZK)𝒪ZK(KX~+ZK)0,0\to\mathcal{O}_{\widetilde{X}}(K_{\widetilde{X}})\to\mathcal{O}_{\widetilde{X}}(K_{\widetilde{X}}+\lfloor Z_{K}\rfloor)\to\mathcal{O}_{\lfloor Z_{K}\rfloor}(K_{\widetilde{X}}+\lfloor Z_{K}\rfloor)\to 0,

we obtain that the right hand side of eq. 10.1 equals dimH0(ZK,𝒪ZK(KX~+ZK))\dim\,H^{0}(\lfloor Z_{K}\rfloor,\mathcal{O}_{\lfloor Z_{K}\rfloor}(K_{\widetilde{X}}+\lfloor Z_{K}\rfloor)). By Serre duality, this equals H1(ZK,𝒪ZK)H^{1}(\lfloor Z_{K}\rfloor,\mathcal{O}_{\lfloor Z_{K}\rfloor}). Now, the short exact sequence

0𝒪X~(ZK)𝒪X~𝒪ZK0,0\to\mathcal{O}_{\widetilde{X}}(-\lfloor Z_{K}\rfloor)\to\mathcal{O}_{\widetilde{X}}\to\mathcal{O}_{\lfloor Z_{K}\rfloor}\to 0,

with the above vanishing eq. 10.2 gives H1(ZK,𝒪ZK)H1(X~,𝒪X~)pgH^{1}(\lfloor Z_{K}\rfloor,\mathcal{O}_{\lfloor Z_{K}\rfloor})\cong H^{1}(\widetilde{X},\mathcal{O}_{\widetilde{X}})\cong\mathbb{C}^{p_{g}}. ∎

\thedefinition Definition.

A computation sequence is a sequence of cycles (Zi)i=0k(Z_{i})_{i=0}^{k} from ZK+LZ_{K}+L,

ZKZK=Z0<<ZkZ_{K}-\lfloor Z_{K}\rfloor=Z_{0}<\ldots<Z_{k}

such that

  • (i)

    for all 0i<k0\leq i<k there is a v(i)𝒱v(i)\in\mathcal{V} so that Zi+1=Zi+Ev(i)Z_{i+1}=Z_{i}+E_{v(i)}, and

  • (ii)

    ZkZKZ_{k}\geq Z_{K} and ZkZKZ_{k}-Z_{K} is the union of some reduced and non-intersecting rational (1)(-1)-curves

Given such a sequence (Zi)i=0k(Z_{i})_{i=0}^{k}, we define

i=𝒪X~(KX~+ZKZi),𝒬i=i/i+1.\mathcal{L}_{i}=\mathcal{O}_{\widetilde{X}}(K_{\widetilde{X}}+Z_{K}-Z_{i}),\quad\mathcal{Q}_{i}=\mathcal{L}_{i}/\mathcal{L}_{i+1}.

Then 𝒬i\mathcal{Q}_{i} is a line bundle on Ev(i)E_{v(i)}. Denote by did_{i} its degree. Since KX~+ZKK_{\widetilde{X}}+Z_{K} is numerically equivalent to zero, we have di=(Zi,Ev(i))d_{i}=(-Z_{i},E_{v(i)}). In particular, since Ev(i)1E_{v(i)}\cong\mathbb{CP}^{1}, we get 𝒬i=𝒪Ev(i)(di)\mathcal{Q}_{i}=\mathcal{O}_{E_{v(i)}}(-d_{i}) and

h0(Ev(i),𝒬i)=max{0,(Zi,Ev(i))+1}.h^{0}(E_{v(i)},\mathcal{Q}_{i})=\max\{0,(-Z_{i},E_{v(i)})+1\}.
\theblock.

Given a computation sequence (Zi)i(Z_{i})_{i}, the inclusion 𝒪X~(KX~+ZKZk)𝒪X~(KX~)\mathcal{O}_{\widetilde{X}}(K_{\widetilde{X}}+Z_{K}-Z_{k})\hookrightarrow\mathcal{O}_{\widetilde{X}}(K_{\widetilde{X}}) induces an isomorphism

H0(X~,𝒪X~(KX~+ZKZk)H0(X~,𝒪X~(KX~)).H^{0}(\widetilde{X},\mathcal{O}_{\widetilde{X}}(K_{\widetilde{X}}+Z_{K}-Z_{k})\stackrel{{\scriptstyle\cong}}{{\longrightarrow}}H^{0}(\widetilde{X},\mathcal{O}_{\widetilde{X}}(K_{\widetilde{X}})).

Indeed, let 𝒰𝒱\mathcal{U}\subset\mathcal{V} be such that ZkZK=u𝒰EuZ_{k}-Z_{K}=\sum_{u\in\mathcal{U}}E_{u}. Then we have a short exact sequence

0𝒪X~(KX~E𝒰)𝒪X~(KX~)𝒪Eu(KX~)0,0\to\mathcal{O}_{\widetilde{X}}(K_{\widetilde{X}}-E_{\mathcal{U}})\to\mathcal{O}_{\widetilde{X}}(K_{\widetilde{X}})\to\bigoplus\mathcal{O}_{E_{u}}(K_{\widetilde{X}})\to 0,

which induces an exact sequence

0H0(X~,𝒪X~(KX~+ZKZk)H0(X~,𝒪X~(KX~))H0(Eu,𝒪Eu(KX~)),0\to H^{0}(\widetilde{X},\mathcal{O}_{\widetilde{X}}(K_{\widetilde{X}}+Z_{K}-Z_{k})\to H^{0}(\widetilde{X},\mathcal{O}_{\widetilde{X}}(K_{\widetilde{X}}))\to\bigoplus H^{0}(E_{u},\mathcal{O}_{E_{u}}(K_{\widetilde{X}})),

and the right hand side vanishes, since (Eu,KX~)=22gu+bu=1(E_{u},K_{\widetilde{X}})=-2-2g_{u}+b_{u}=-1.

\thecor Corollary.

Let (Zi)i=0k(Z_{i})_{i=0}^{k} be a computation sequence. Then

(10.3) pg=i=0k1dimH0(X~,i)H0(X~,i+1)i=0k1max{0,di+1}.p_{g}=\sum_{i=0}^{k-1}\dim\,\frac{H^{0}(\widetilde{X},\mathcal{L}_{i})}{H^{0}(\widetilde{X},\mathcal{L}_{i+1})}\leq\sum_{i=0}^{k-1}\max\{0,d_{i}+1\}.

with equality if and only if the map H0(X~,i)H0(Ev(i),𝒬i)H^{0}(\widetilde{X},\mathcal{L}_{i})\to H^{0}(E_{v(i)},\mathcal{Q}_{i}) is surjective for all 0i<k0\leq i<k. ∎

\therem Remark.
  • (i)

    We note in particular that if there exists a computation sequence (Zi)i=0k(Z_{i})_{i=0}^{k} so that (Zi,Ev(i))>0(Z_{i},E_{v(i)})>0 for all ii, then pg=0p_{g}=0, that is, (X,0)(X,0) is rational. In general, if (Zi,Ev(i))>0(Z_{i},E_{v(i)})>0 for some ii, then the inequality between the ithi^{\textrm{th}} terms in the sums eq. 10.3 is an equality.

  • (ii)

    Let S(Zi)S(Z_{i}) be the sum imax{0,di+1}\sum_{i}\max\{0,d_{i}+1\} from the right hand side of eq. 10.3 associated with (Zi)(Z_{i}). Then we have

    (10.4) pgmin(Zi)S(Zi),p_{g}\leq\min_{(Z_{i})}S(Z_{i}),

    where the minimum is taken over all computation sequences. Note that min(Zi)S(Zi)\min_{(Z_{i})}S(Z_{i}) is an invariant associated with the topological type (graph), hence in this way we get a topological upper bound for the geometric genus of all possible analytic types supported on a fixed topological type.

    On the other hand we emphasize the following facts. In general it is hard to identify a sequence which minimizes {S(Zi)}\{S(Z_{i})\}. Also, for an arbitrary fixed topological type, it is not even true that there exists an analytic type supported on the fixed topological type for which eq. 10.4 holds. Furthermore, it is even harder to identify those analytic structures which maximize pgp_{g}, e.g., if eq. 10.4 holds for some analytic structure, then which are these maximizing analytic structures, see e.g. [24].

    In the sequel our aim is the following: in our toric Newton nondegenerate case we construct combinatorially a sequence (it will be called ‘diagonal sequence’), which satisfies eq. 10.3 with equality (in particular it minimizes {S(Zi)}\{S(Z_{i})\} as well). This also shows that if a topological type is realized by a Newton nondegenerate Weil divisor, then this germ maximizes the geometric genus of analytic types supported by that topological type.

\theblock.

We recall the construction of the Laufer operator and generalized Laufer sequences with respect to 𝒩𝒱\mathcal{N}\subset\mathcal{V}. We claim that for any cycle ZLZ\in L^{\prime}, there is a smallest cycle x(Z)Z+Lx(Z)\in Z+L satisfying

(10.5) {n𝒩:mn(x(Z))=mn(Z),v𝒱𝒩:(x(Z),Ev)0.\left\{\begin{array}[]{l}\forall n\in\mathcal{N}:\,m_{n}(x(Z))=m_{n}(Z),\\ \forall v\in\mathcal{V}\setminus\mathcal{N}:\,(x(Z),E_{v})\leq 0.\end{array}\right.

The existence and uniqueness of such an element is explained in [21] in the case when |𝒩|=1|\mathcal{N}|=1 and in general in [16, 25, 31]. The name comes from a construction of Laufer in [17, Proposition 4.1]. Note that x(Z)x(Z) only depends on the multiplicities mn(Z)m_{n}(Z) of ZZ for n𝒩n\in\mathcal{N} and the class [Z]H=L/L[Z]\in H=L^{\prime}/L.

The following properties hold for the operator xx, assuming Z1Z2LZ_{1}-Z_{2}\in L:

Monotonicity: If Z1Z2Z_{1}\leq Z_{2} then x(Z1)x(Z2)x(Z_{1})\leq x(Z_{2}).

Idempotency: We have x(x(Z))=x(Z)x(x(Z))=x(Z) for any ZLZ\in L^{\prime}.

Lower bound by intersection numbers: If ZLZ\in L^{\prime} and ZLZ^{\prime}\in L_{\mathbb{Q}} so that mn(Z)=mn(Z)m_{n}(Z)=m_{n}(Z^{\prime}) for n𝒩n\in\mathcal{N} and (Z,Ev)0(Z^{\prime},E_{v})\geq 0 for all v𝒱𝒩v\in\mathcal{V}\setminus\mathcal{N}, then x(Z)Zx(Z)\geq Z^{\prime}.

Generalized Laufer sequence: Assume that Zx(Z)Z\leq x(Z). First note that if (Z,Ev)>0(Z,E_{v})>0 for some v𝒱𝒩v\in\mathcal{V}\setminus\mathcal{N}, then we have Z+Evx(Z)Z+E_{v}\leq x(Z) as well, similarly as in the proof of Proposition 4.1 [17]. We claim that there exists a generalized Laufer sequence which connects ZZ with x(Z)x(Z). It is determined recursively as follows. Start by setting Z0=ZZ_{0}=Z. Assume that we have constructed ZiZ_{i}. By induction, we then have Zix(Z)Z_{i}\leq x(Z). If (Z,Ev)0(Z,E_{v})\leq 0 for all v𝒱𝒩v\in\mathcal{V}\setminus\mathcal{N} then by the minimality of x(Z)x(Z) we get Zi=x(Z)Z_{i}=x(Z); hence the construction is finished and we stop. Otherwise, there is a v𝒱𝒩v\in\mathcal{V}\setminus\mathcal{N} so that (Z,Ev)>0(Z,E_{v})>0. We then define Zi+1=Zi+EvZ_{i+1}=Z_{i}+E_{v} (for some choice of such vv).

\therem Remark.

The computation sequence (Zi)i=0k(Z_{i})_{i=0}^{k} (as in corollary 10), what we will construct, will have several intermediate parts formed by generalized Laufer sequences as above. Note that if ZiZ_{i} and Zi+1=Zi+EvZ_{i+1}=Z_{i}+E_{v} are two consecutive elements in a Laufer sequence, then di=(Zi,Ev)>0-d_{i}=(Z_{i},E_{v})>0, hence max{0,di+1}=0\max\{0,d_{i}+1\}=0, and the comment from section 10 applies: this step does not contribute in the sum on the right hand side of eq. 10.3. Informally, we say that parts given by Laufer sequences “do not contribute to the geometric genus”.

\theblock.

The Newton nondegenerate case. Let us consider again the resolution X~X\widetilde{X}\to X of Newton nondegenerate Weil divisor as in section 6. Let KX~K_{\widetilde{X}} denote a canonical divisor as in section 8. In this section we will assume that in the dual resolution graph GG we have mn(ZK)1m_{n}(Z_{K})\geq 1 for any node nn. This assumption will be justified in section 11.

From the assumption mn(ZK)1m_{n}(Z_{K})\geq 1, valid for any node nn, an immediate application of the construction of GG from section 6 gives that ZK0Z_{K}\geq 0. Thus ZK>0\lfloor Z_{K}\rfloor>0.

\thelemma Lemma.
  • (i)

    x(ZKZK)ZKZKx(Z_{K}-\lfloor Z_{K}\rfloor)\geq Z_{K}-\lfloor Z_{K}\rfloor.

  • (ii)

    Let 𝒰𝒱\mathcal{U}\subset\mathcal{V} be the set of (1)(-1)-vertices appearing on bamboos joining n,n𝒩n,n^{\prime}\in\mathcal{N}^{*} with α(n,n)=1\alpha(\ell_{n},\ell_{n^{\prime}})=1 in section 6. Then x(ZK)=ZK+u𝒰Eux(Z_{K})=Z_{K}+\sum_{u\in\mathcal{U}}E_{u}. In particular, the sequence constructed in section 10 satisfies ii in section 10.

Proof.

i Since x(Z)ZLx(Z)-Z\in L for any ZLZ\in L^{\prime}, it is enough to show that x(ZKZK)0x(Z_{K}-\lfloor Z_{K}\rfloor)\geq 0. We can analyse each component of G𝒩G\setminus\mathcal{N} independently, let GBG_{B} be such a bamboo formed from E1,,EsE_{1},\ldots,E_{s}, with dual vectors in GBG_{B} denoted by EiE_{i}^{*}. If a0a\geq 0 and b0b\geq 0 are the multiplicities of ZKZKZ_{K}-\lfloor Z_{K}\rfloor along the neighboring nodes of GBG_{B} in GG (with convention that a=0a=0 if there is only one such node), we search for a cycle xx with (x,Ei)(aE1+bEs,Ei)(x,E_{i})\leq(aE_{1}^{*}+bE_{s}^{*},E_{i}) for all ii. Thus, x(aE1+bEs)x-(aE_{1}^{*}+bE_{s}^{*}) is in the Lipman cone of GBG_{B}, hence xaE1+bEs0x\geq aE_{1}^{*}+bE_{s}^{*}\geq 0.

ii Using the lower bound by intersection numbers, we find that x(ZK)ZKE+n𝒩Enx(Z_{K})\geq Z_{K}-E+\sum_{n\in\mathcal{N}}E_{n}. Since x(ZK)=x(ZKE+n𝒩En)x(Z_{K})=x(Z_{K}-E+\sum_{n\in\mathcal{N}}E_{n}), there exists a Laufer sequence from ZKE+n𝒩EnZ_{K}-E+\sum_{n\in\mathcal{N}}E_{n} to x(ZK)x(Z_{K}). Now, one verifies that the construction/algorithm of this sequence chooses each vertex v𝒱(𝒩𝒰)v\in\mathcal{V}\setminus(\mathcal{N}\cup\mathcal{U}) once, and each vertex in 𝒰\mathcal{U} twice. ∎

\thedefinition Definition.

A (coarse) diagonal computation sequence (Z¯i)i=0k¯(\bar{Z}_{i})_{i=0}^{\bar{k}} with respect to 𝒩\mathcal{N} is defined as follows. Start with Z0=ZKZKZ_{0}=Z_{K}-\lfloor Z_{K}\rfloor, and define Z¯0=x(ZKZK)\bar{Z}_{0}=x(Z_{K}-\lfloor Z_{K}\rfloor). Assuming Z¯i\bar{Z}_{i} (i0i\geq 0) has been defined, and that Z¯i|𝒩<ZK|𝒩\bar{Z}_{i}|_{\mathcal{N}}<Z_{K}|_{\mathcal{N}}, choose a v¯(i)𝒩\bar{v}(i)\in\mathcal{N} minimizing the ratio

(10.6) nr(n):=mn(Z¯i)mn(ZKE),n𝒩.n\mapsto r(n):=\frac{m_{n}(\bar{Z}_{i})}{m_{n}(Z_{K}-E)},\ \ \ n\in\mathcal{N}.

Then set Z¯i+1=x(Z¯i+Ev¯(i))\bar{Z}_{i+1}=x(\bar{Z}_{i}+E_{\bar{v}(i)}). If Z¯i|𝒩=(ZKE)|𝒩\bar{Z}_{i}|_{\mathcal{N}}=(Z_{K}-E)|_{\mathcal{N}}, then we record k¯=i\bar{k}^{\prime}=i. If Z¯i|𝒩=ZK|𝒩\bar{Z}_{i}|_{\mathcal{N}}=Z_{K}|_{\mathcal{N}}, then we stop, and set k¯=i\bar{k}=i.

We refine the above choice as follows. Choose some node n0𝒩n_{0}\in\mathcal{N} and define a partial order \leq on the set 𝒩\mathcal{N}: for n,n𝒩n,n^{\prime}\in\mathcal{N}, define nnn\leq n^{\prime} if nn lies on the geodesic joining nn^{\prime} and n0n_{0} (here we make use of the assumption that the link is a rational homology sphere, in particular, GG is a tree). When choosing v¯(i)\bar{v}(i), if given a choice of several nodes minimizing {r(n)}n\{r(n)\}_{n}, and minn{r(n)}<1\min_{n}\{r(n)\}<1, then, we choose v¯(i)\bar{v}(i) minimal of those with respect to (𝒩,)(\mathcal{N},\leq). If minn{r(n)}=1\min_{n}\{r(n)\}=1, let 𝒩𝒩\mathcal{N}^{\prime}\subset\mathcal{N} be the set of nodes nn for which r(n)=1r(n)=1. If 𝒩\mathcal{N}^{\prime} has one element we have to chose that one. Otherwise, let GG^{\prime} be the minimal connected subgraph of GG containing 𝒩\mathcal{N}^{\prime}, and we choose v¯(i)\bar{v}(i) as a leaf of GG^{\prime}.

Note that by section 10i, Z0=ZKZKx(ZKZK)=Z¯0Z_{0}=Z_{K}-\lfloor Z_{K}\rfloor\leq x(Z_{K}-\lfloor Z_{K}\rfloor)=\bar{Z}_{0}, hence there exists a Laufer sequence connecting Z0Z_{0} with Z¯0\bar{Z}_{0}. Furthermore, using idempotency and monotonicity of the Laufer operator section 10, we find

Z¯i+Ev¯(i)=x(Z¯i)+Ev¯(i)x(Z¯i+Ev¯(i))=Z¯i+1.\bar{Z}_{i}+E_{\bar{v}(i)}=x(\bar{Z}_{i})+E_{\bar{v}(i)}\leq x(\bar{Z}_{i}+E_{\bar{v}(i)})=\bar{Z}_{i+1}.

As a result, we can join Z¯i+Ev¯(i)\bar{Z}_{i}+E_{\bar{v}(i)} and Z¯i+1\bar{Z}_{i+1} by a Laufer sequence. This way, we obtain a computation sequence (Zi)i(Z_{i})_{i}, connecting ZKZKZ_{K}-\lfloor Z_{K}\rfloor with x(ZK)x(Z_{K}). Finally, by section 10ii, x(Zk)x(Z_{k}) satisfies the requirement section 10ii too, hence section 10 applies.

\theblock.

For a diagonal computation sequence as above at each step, except for the step from Z¯i\bar{Z}_{i} to Z¯i+Ev¯(i)\bar{Z}_{i}+E_{\bar{v}(i)}, we have di<0d_{i}<0, we find, using sections 10 and 10

(10.7) pgi=0k¯1max{0,(Z¯i,Ev¯(i))+1}.p_{g}\leq\sum_{i=0}^{\bar{k}-1}\max\{0,(-\bar{Z}_{i},E_{\bar{v}(i)})+1\}.
10.1 Theorem.

Let (X,0)(X,0) be a normal Newton nondegenerate Weil divisor given by a function ff, with a rational homology sphere link, and assume that the polyhedron Γ+(f)\Gamma_{+}(f) is \mathbb{Q}-Gorenstein pointed at pMp\in M_{\mathbb{Q}}. Then, a diagonal computation sequence (Zi)i(Z_{i})_{i} constructed above computes the geometric genus, that is, equality holds in eq. 10.7.

In this sequel we prove the theorem under the assumption 10.9 regarding the multiplicities of ZKZ_{K}, by the results of the next section this assumption can be removed.

\thedefinition Definition.

Let n𝒩n\in\mathcal{N}, corresponding to the face FnΓ(f)F_{n}\subset\Gamma(f). Denote by CnC_{n} the convex hull of the union of FnF_{n} and {p}\{p\}. Set also

Cn=CnnnCn,C^{-}_{n}=C_{n}\setminus\bigcup_{n^{\prime}\geq n}C_{n^{\prime}},

where we use the partial ordering \leq on 𝒩\mathcal{N} defined in section 10. For i=0,,k¯1i=0,\ldots,\bar{k}-1, let HiH_{i} be the hyperplane in MM_{\mathbb{R}} defined as the set of points qMq\in M_{\mathbb{R}} satisfying n(qp)=mv¯(i)(Z¯i)\ell_{n}(q-p)=m_{\bar{v}(i)}(\bar{Z}_{i}). For i=0,,k¯1i=0,\ldots,\bar{k}-1, we set

Fi=Cv¯(i)Hi,Fi=Cv¯(i)Hi.F_{i}=C_{\bar{v}(i)}\cap H_{i},\qquad F^{-}_{i}=C_{\bar{v}(i)}^{-}\cap H_{i}.
\therem Remark.

The affine plane HiH_{i} contains an affine lattice MHM\cap H, that is there is an affine isomorphism H2H\to\mathbb{R}^{2}, inducing a bijection HM2H\cap M\to\mathbb{Z}^{2}. The polyhedron FiF_{i} is then the image of a lattice polyhedron with no integral integer points under a homothety with ratio in [0,1[[0,1[ if i<k¯i<\bar{k}^{\prime}. These properties allow us to apply section 10 in the proof of theorem 10.1. Furthermore, the polygon FiF_{i} is always nonempty, even if FiF_{i}^{-} may be empty.

\theblock.

The sets CnC^{-}_{n} form a partitioning of the union of segments starting at pp and ending in points on Γ(f)\Gamma(f), that is, n𝒩Cn\cup_{n\in\mathcal{N}}C_{n}. This follows from the construction as follows. The partially ordered set (𝒩,)(\mathcal{N},\leq) is an lower semilattice, i.e. any subset has a largest lower bound. If qn𝒩Cnq\in\cup_{n\in\mathcal{N}}C_{n}, and 𝒩\mathcal{I}\subset\mathcal{N} is the set of nodes nn for which qCnq\in C_{n}, then qCnqq\in C_{n_{q}}^{-}, and pCnp\notin C^{-}_{n} for nnqn\neq n_{q}, where nqn_{q} is the largerst lower bound of \mathcal{I}.

The integral points qq in the union of the sets CnFnC_{n}^{-}\setminus F_{n} are presicely the integral points satisfying σ(q)>mσ\ell_{\sigma}(q)>m_{\sigma} for all σ~f(1,1)\sigma\in\widetilde{\triangle}_{f}^{*(1,1)} and σ(q)mσ\ell_{\sigma}(q)\leq m_{\sigma} for some σff\sigma\in\triangle_{f}\setminus\triangle_{f}^{*}. Indeed, by the rational homology sphere assumption, any integral point on the Newton diagram Γ(f)\Gamma(f) must lie on the boundary Γ(f)\partial\Gamma(f), see section 6. These are the points “under the Newton diagram”; by theorem 7.1, the number of these points is pgp_{g}. It follows from construction that the family (FiM)i=0k¯1(F^{-}_{i}\cap M)_{i=0}^{\bar{k}-1} forms a partition of these points. We conclude:

(10.8) pg=i=0k¯1|FiM|.p_{g}=\sum_{i=0}^{\bar{k}^{\prime}-1}|F^{-}_{i}\cap M|.
\thedefinition Definition.

For r,xr,x\in\mathbb{R}, denote by rx\lceil r\rceil_{x} the smallest real number larger or equal to rr and congruent to xx modulo \mathbb{Z}. That is,

rx=min{a|ar,ax(mod)}\lceil r\rceil_{x}=\min\left\{a\in\mathbb{R}\,\middle|\,a\geq r,\;a\equiv x\;(\mathop{\rm mod}\nolimits\mathbb{Z})\right\}
\therem Remark.

The number rx\lceil r\rceil_{x} depends on xx only up to an integer. For all ii, we have Z¯iZK(modL)\bar{Z}_{i}\equiv Z_{K}\;(\mathop{\rm mod}\nolimits L). In particular, given an n𝒩n\in\mathcal{N}, we have mn(Z¯i)mn(ZKE)(mod)m_{n}(\bar{Z}_{i})\equiv m_{n}(Z_{K}-E)\;(\mathop{\rm mod}\nolimits\mathbb{Z}).

\thelemma Lemma.

Let ZLZ\in L^{\prime} and take n,n𝒩n,n^{\prime}\in\mathcal{N}^{*} connected by a bamboo, and u𝒱u\in\mathcal{V} a neighbour of nn on this bamboo. Then

(10.9) mu(x(Z))=βmn(Z)+mn(Z)αmu(Z)m_{u}(x(Z))=\left\lceil\frac{\beta m_{n}(Z)+m_{n^{\prime}}(Z)}{\alpha}\right\rceil_{m_{u}(Z)}

where α=α(n,n)\alpha=\alpha(\ell_{n},\ell_{n^{\prime}}) and β=β(n,n)\beta=\beta(\ell_{n},\ell_{n^{\prime}}) (see section 4 and section 4). Furthermore, if for all v𝒱v\in\mathcal{V} lying on the bamboo joining n,nn,n^{\prime}, we have (Z,Ev)=0(Z,E_{v})=0, then x(Z)=Zx(Z)=Z along the bamboo and

(10.10) mu(x(Z))=βmn(Z)+mn(Z)α.m_{u}(x(Z))=\frac{\beta m_{n}(Z)+m_{n^{\prime}}(Z)}{\alpha}.
Proof.

We prove eq. 10.9, eq. 10.10 follows similarly. Let u=u1,,usu=u_{1},\ldots,u_{s} be the vertices on the bamboo with Euler numbers b1,,bs-b_{1},\ldots,-b_{s} as in fig. 3. Set m~0=m0=mn(Z)\widetilde{m}_{0}=m_{0}=m_{n}(Z) and m~s+1=ms+1=mn(Z)\widetilde{m}_{s+1}=m_{s+1}=m_{n^{\prime}}(Z). There exists a unique set of numbers m~1,,m~s\widetilde{m}_{1},\ldots,\widetilde{m}_{s}\in\mathbb{Q} so that the equations

(10.11) m~i1bim~i+m~i+1=0,i=1,,s\widetilde{m}_{i-1}-b_{i}\widetilde{m}_{i}+\widetilde{m}_{i+1}=0,\quad i=1,\ldots,s

are satisfied. This follows from the fact that the intersection matrix of the bamboo is invertible over \mathbb{Q}. In fact, it follows from [12, Lemma 20.2] that in fact,

m~1=βm0+ms+1α.\widetilde{m}_{1}=\frac{\beta m_{0}+m_{s+1}}{\alpha}.

This, and the lower bound by intersection numbers from section 10, implies that mu(x(Z))m~1m_{u}(x(Z))\geq\widetilde{m}_{1}, and therefore mu(x(Z))m~1mu(Z)m_{u}(x(Z))\geq\lceil\widetilde{m}_{1}\rceil_{m_{u}(Z)}, since x(Z)ZLx(Z)-Z\in L.

For the inverse inequality, we must show that there exist numbers m1,,msm_{1},\ldots,m_{s} satisfying

(10.12) mi1bimi+mi+10,mimi(Z)(mod),m_{i-1}-b_{i}m_{i}+m_{i+1}\leq 0,\qquad m_{i}\equiv m_{i}(Z)\;(\mathop{\rm mod}\nolimits\mathbb{Z}),

for i=1,,si=1,\ldots,s, and so that m1m_{1} is the right hand side of eq. 10.9. Let n=0,1,,s,s+1=n\ell_{n}=\ell_{0},\ell_{1},\ldots,\ell_{s},\ell_{s+1}=\ell_{n^{\prime}} be the canoncial primitive sequence as in section 6, and note that β=α(1,s+1)\beta=\alpha(\ell_{1},\ell_{s+1}). Set recursively

mi=α(i,s+1)mi1+ms+1α(i1,s+1)mui(Z)i=1,,s.m_{i}=\left\lceil\frac{\alpha(\ell_{i},\ell_{s+1})m_{i-1}+m_{s+1}}{\alpha(\ell_{i-1},\ell_{s+1})}\right\rceil_{m_{u_{i}}(Z)}\qquad i=1,\ldots,s.

Note that, by definition, mimi(Z)m_{i}\equiv m_{i}(Z). The assumption ZLZ\in L^{\prime} therefore implies that the left hand side of eq. 10.12 is an integer. It is then enough to prove eq. 10.12 for i=1i=1. This equation is clear if s=1s=1, so we assume that s>1s>1. Setting γ=α(2,s)\gamma=\alpha(\ell_{2},\ell_{s}), we find

m2m~2=γm1+ms+1βmu2(Z)γm~1+m~s+1β=γβ(m1m~1)+rm_{2}-\widetilde{m}_{2}=\left\lceil\frac{\gamma m_{1}+m_{s+1}}{\beta}\right\rceil_{m_{u_{2}}(Z)}-\frac{\gamma\widetilde{m}_{1}+\widetilde{m}_{s+1}}{\beta}=\frac{\gamma}{\beta}(m_{1}-\widetilde{m}_{1})+r

where 0r<10\leq r<1. In order to prove eq. 10.12, we start by subtracting zero, i.e. the left hand side of eq. 10.11. The left hand side of eq. 10.12 equals

m0m~0b1(m1m~1)+m2m~2=(b1+γβ)(m1m~1)+r<1,m_{0}-\widetilde{m}_{0}-b_{1}(m_{1}-\widetilde{m}_{1})+m_{2}-\widetilde{m}_{2}=\left(-b_{1}+\frac{\gamma}{\beta}\right)(m_{1}-\widetilde{m}_{1})+r<1,

since γ/β<1\gamma/\beta<1. Since the left hand side is an integer, eq. 10.12 follows. ∎

\thelemma Lemma.

If k¯i<k¯\bar{k}^{\prime}\leq i<\bar{k}, then (Z¯i,Ev¯(i))>0(\bar{Z}_{i},E_{\bar{v}(i)})>0. As a result, the corresponding terms in eq. 10.7 vanish.

Proof.

Let u𝒱nu\in\mathcal{V}_{n} be a neighbour of v¯(i)\bar{v}(i). Assume first that uu lies on a bamboo connecting v¯(i)\bar{v}(i) and n𝒩n\in\mathcal{N}. We then have mv¯(i)(Z¯i)=mv¯(i)(ZKE)m_{\bar{v}(i)}(\bar{Z}_{i})=m_{\bar{v}(i)}(Z_{K}-E). Furthermore, mn(Z¯i)=mn(ZKE)+εm_{n}(\bar{Z}_{i})=m_{n}(Z_{K}-E)+\varepsilon, where ε\varepsilon equals 0 or 11. By the previous lemma, we find

mu(Z¯i)=βmv¯(i)(Z¯i)+mn(ZKE)+εαmu(ZK)=mu(ZKE)+ε.m_{u}(\bar{Z}_{i})=\left\lceil\frac{\beta m_{\bar{v}(i)}(\bar{Z}_{i})+m_{n}(Z_{K}-E)+\varepsilon}{\alpha}\right\rceil_{m_{u}(Z_{K})}=m_{u}(Z_{K}-E)+\varepsilon.

with α,β\alpha,\beta as in the lemma.

Next, assume that uu lies on a bamboo connecting v¯(i)\bar{v}(i) and n𝒩𝒩n^{\prime}\in\mathcal{N}^{*}\setminus\mathcal{N}. Name the vertices on the bamboo u1,,usu_{1},\ldots,u_{s} as in the proof of the previous lemma. We then have (ZKE,Euj)=0(Z_{K}-E,E_{u_{j}})=0 for j=1,,s1j=1,\ldots,s-1, and (ZKE,Eus)=1(Z_{K}-E,E_{u_{s}})=1. By the lower bound on intersection numbers, we find x(ZKE)ZKEx(Z_{K}-E)\geq Z_{K}-E. A Laufer sequence which computes x(ZKE)x(Z_{K}-E) from ZKEZ_{K}-E may start with Eus,Eus1,,Eu1E_{u_{s}},E_{u_{s-1}},\ldots,E_{u_{1}}. This shows that mu(x(ZKE))mu(ZKE)+1m_{u}(x(Z_{K}-E))\geq m_{u}(Z_{K}-E)+1 in this case.

As a result, for every u𝒱v¯(i)u\in\mathcal{V}_{\bar{v}(i)}, we have mu(x(ZKE))mu(ZKE)m_{u}(x(Z_{K}-E))\geq m_{u}(Z_{K}-E), with an equality for at most one neighbour. As a result, since (ZKE,Ev)=2δv(Z_{K}-E,E_{v})=2-\delta_{v} we find

(Z¯i,Ev¯(i))(ZKE,Ev¯(i))+(δv¯(i)1)=1.(\bar{Z}_{i},E_{\bar{v}(i)})\geq(Z_{K}-E,E_{\bar{v}(i)})+(\delta_{\bar{v}(i)}-1)=1.

The final statement of the lemma is now clear. ∎

\thelemma Lemma.

Let F2F\subset\mathbb{R}^{2} be an integral polygon with no internal integral points. Let S1,,SrS_{1},\ldots,S_{r} be the faces of FF and let cjc_{j} be the integral lenght of SjS_{j}. Let 0ρ<10\leq\rho<1, J{1,,r}J\subset\{1,\ldots,r\}. Then let ai:2a_{i}:\mathbb{R}^{2}\to\mathbb{R} be the unique integral affine function whose minimal set on ρF\rho F is ρSj\rho S_{j} and this minimal value is λj]1,0]\lambda_{j}\in]-1,0] if jJj\notin J and λj[1,0[\lambda_{j}\in[-1,0[ if jJj\in J. Set Fρ=ρFjJρSjF_{\rho}^{-}=\rho F\setminus\cup_{j\in J}\rho S_{j}. Then there exists an aa\in\mathbb{Z} satifying

j=1scjaja,|Fρ2|=max{0,a+1}.\sum_{j=1}^{s}c_{j}a_{j}\equiv a,\qquad|F_{\rho}^{-}\cap\mathbb{Z}^{2}|=\max\{0,a+1\}.
Proof.

This is [31, Theroem 4.2.2]. ∎

Proof of theorem 10.1.

Recall the order \leq on the set 𝒩\mathcal{N} defined in section 10. We extend this order in the obvious way to all of 𝒱\mathcal{V}. Also, by assumption, ff is \mathbb{Q}-pointed at the point pMp\in M_{\mathbb{Q}}. Fix an 0ik¯0\leq i\leq\bar{k}^{\prime} and set H=HiH=H_{i}. For u𝒱v¯(i)u\in\mathcal{V}_{\bar{v}(i)}, define

λu=inf{u(q)|qFi}\lambda_{u}=\inf\left\{\ell_{u}(q)\,\middle|\,q\in F_{i}\right\}

(recall that FiF_{i} is nonempty, see section 10) and

νu={λu+1ifuv¯(i)andλu,λuelse.\nu_{u}=\begin{cases}\lambda_{u}+1&\mathrm{if}\;u\leq\bar{v}(i)\;\mathrm{and}\;\lambda_{u}\in\mathbb{Z},\\ \lceil\lambda_{u}\rceil&\mathrm{else.}\end{cases}

Define the affine functions au:Ha_{u}:H\to\mathbb{R}, au=u|Hνua_{u}=\ell_{u}|_{H}-\nu_{u}. By construction, these are primitive integral functions on HH with respect to the affine lattice HMH\cap M. It now follows from section 10 that there is an aa\in\mathbb{Z} so that uaua\sum_{u}a_{u}\equiv a and |FiM|=max{0,a+1}|F^{-}_{i}\cap M|=\max\{0,a+1\}.

On the other hand, we claim that νuu(p)mu(Z¯i)\nu_{u}-\ell_{u}(p)\leq m_{u}(\bar{Z}_{i}) for u𝒱v¯(i)u\in\mathcal{V}_{\bar{v}(i)}. Using section 6, and the definition of HiH_{i}, i.e. v¯(i)(qp)|H=mv¯(i)(Z¯i)\ell_{\bar{v}(i)}(q-p)|_{H}=m_{\bar{v}(i)}(\bar{Z}_{i}) for qHq\in H, it follows that

a=uau(q)=uu(qp)(νuu(p))bv¯(i)v¯(i)(qp)umu(Z¯i)=(Z¯i,Ev¯(i)).a=\sum_{u}a_{u}(q)=\sum_{u}\ell_{u}(q-p)-(\nu_{u}-\ell_{u}(p))\geq b_{\bar{v}(i)}\ell_{\bar{v}(i)}(q-p)-\sum_{u}m_{u}(\bar{Z}_{i})=(-\bar{Z}_{i},E_{\bar{v}(i)}).

where qq is any element of HH. As a result, using eq. 10.7 and section 10, as well as eq. 10.8, we have

pg=i=0k¯1|FiM|i=0k¯1max{0,(Z¯i,Ev¯(i))+1}pg,p_{g}=\sum_{i=0}^{\bar{k}^{\prime}-1}|F^{-}_{i}\cap M|\geq\sum_{i=0}^{\bar{k}-1}\max\{0,(-\bar{Z}_{i},E_{\bar{v}(i)})+1\}\geq p_{g},

and so these inequalities are in fact equalities.

We are left with proving the claim νumu(Z¯i)+u(p)\nu_{u}\leq m_{u}(\bar{Z}_{i})+\ell_{u}(p) for u𝒱v¯(i)u\in\mathcal{V}_{\bar{v}(i)}. Fix uu, and let n𝒩n\in\mathcal{N}^{*} so that uu lies on a bamboo connecting v¯(i)\bar{v}(i) and nn. Let S=Fv¯(i)FnS=F_{\bar{v}(i)}\cap F_{n}. Then SS is the minimal set of u\ell_{u} on Fv¯(i)F_{\bar{v}(i)}, i.e., S=FuS=F_{u}. Let AA be the affine hull of S{p}S\cup\{p\}. Since the two affine functions

v¯(i)v¯(i)(p)mv¯(i)(ZKE),nn(p)mn(ZKE),\frac{\ell_{\bar{v}(i)}-\ell_{\bar{v}(i)}(p)}{m_{\bar{v}(i)}(Z_{K}-E)},\quad\frac{\ell_{n}-\ell_{n}(p)}{m_{n}(Z_{K}-E)},

both take value 0 on pp and 11 on SS, by theorem 9.1item (iii), they conincide on AA. Let

r=mv¯(i)(Z¯i)mv¯(i)(ZKE)r=\frac{m_{\bar{v}(i)}(\bar{Z}_{i})}{m_{\bar{v}(i)}(Z_{K}-E)}

Using the minimality of eq. 10.6, we get for any qp+r(Sp)HAq\in p+r(S-p)\subset H\cap A

(10.13) n(qp)mn(ZKE)=v¯(i)(qp)mv¯(i)(ZKE)=mv¯(i)(Z¯i)mv¯(i)(ZKE)mn(Z¯i)mn(ZKE),\frac{\ell_{n}(q-p)}{m_{n}(Z_{K}-E)}=\frac{\ell_{\bar{v}(i)}(q-p)}{m_{\bar{v}(i)}(Z_{K}-E)}=\frac{m_{\bar{v}(i)}(\bar{Z}_{i})}{m_{\bar{v}(i)}(Z_{K}-E)}\leq\frac{m_{n}(\bar{Z}_{i})}{m_{n}(Z_{K}-E)},

and so n(qp)mn(Z¯i)\ell_{n}(q-p)\leq m_{n}(\bar{Z}_{i}). In the case when nv¯(i)n\leq\bar{v}(i), or equivalently, uv¯(i)u\leq\bar{v}(i), this inequality is strict. It follows, using section 10, that

(10.14) mu(Z¯i)=β(v¯(i),n)mv¯(i)(Z¯i)+mn(Z¯i)α(v¯(i),n)mu(Z¯i)β(v¯(i),n)v¯(i)(qp)+n(qp)α(v¯(i),n)=u(qp)=λuu(p).\begin{split}m_{u}(\bar{Z}_{i})&=\left\lceil\frac{\beta(\ell_{\bar{v}(i)},\ell_{n})m_{\bar{v}(i)}(\bar{Z}_{i})+m_{n}(\bar{Z}_{i})}{\alpha(\ell_{\bar{v}(i)},\ell_{n})}\right\rceil_{m_{u}(\bar{Z}_{i})}\\ &\geq\frac{\beta(\ell_{\bar{v}(i)},\ell_{n})\ell_{\bar{v}(i)}(q-p)+\ell_{n}(q-p)}{\alpha(\ell_{\bar{v}(i)},\ell_{n})}\\ &=\ell_{u}(q-p)\\ &=\lambda_{u}-\ell_{u}(p).\end{split}

Therefore, since mu(Z¯i)mu(ZK)u(p)m_{u}(\bar{Z}_{i})\equiv m_{u}(Z_{K})\equiv-\ell_{u}(p) (mod)(\mathop{\rm mod}\nolimits\mathbb{Z}), we find

mu(Z¯i)λuu(p).m_{u}(\bar{Z}_{i})\geq\lceil\lambda_{u}\rceil-\ell_{u}(p).

This proves the claim, unless uv¯(i)u\leq\bar{v}(i) and λu\lambda_{u}\in\mathbb{Z}. In that case, the numbers v¯(i)(q)\ell_{\bar{v}(i)}(q) and u(q)=λu\ell_{u}(q)=\lambda_{u} are both integers. Since n,u\ell_{n},\ell_{u} form a part of an integral basis of N=MN=M^{\vee}, we can assume that qMq\in M, hence,

β(v¯(i),n)v¯(i)(qp)+n(qp)α(v¯(i),n)=u(qp)u(p)mu(ZK)mu(Z¯i)(mod).\frac{\beta(\ell_{\bar{v}(i)},\ell_{n})\ell_{\bar{v}(i)}(q-p)+\ell_{n}(q-p)}{\alpha(\ell_{\bar{v}(i)},\ell_{n})}\\ =\ell_{u}(q-p)\equiv-\ell_{u}(p)\equiv m_{u}(Z_{K})\equiv m_{u}(\bar{Z}_{i})\quad(\mathop{\rm mod}\nolimits\mathbb{Z}).

As a result, since we have a strict inequality mn(Z¯i)>n(qp)m_{n}(\bar{Z}_{i})>\ell_{n}(q-p) we get a strict inequality in eq. 10.14 as well. Therefore, we have

mu(Z¯i)>λuu(p)andmu(Z¯i)λuu(p)(mod),m_{u}(\bar{Z}_{i})>\lambda_{u}-\ell_{u}(p)\qquad\mathrm{and}\qquad m_{u}(\bar{Z}_{i})\equiv\lambda_{u}-\ell_{u}(p)\quad(\mathop{\rm mod}\nolimits\mathbb{Z}),

and so mu(Z¯i)λuu(p)+1=νuu(p)m_{u}(\bar{Z}_{i})\geq\lambda_{u}-\ell_{u}(p)+1=\nu_{u}-\ell_{u}(p), which finishes the proof of the claim. ∎

11. Removing B1B_{1}-facets

In this section we consider only surface singularities, i.e. we assume that r=3r=3. We consider removable B1B_{1}-facets of two dimensional Newton diagrams and show that they can be removed without affecting certain invariants of nondegenerate Weil divisors. This is stated in section 11. In parallel we also prove section 11, which allows us to assume that the divisor ZKEZ_{K}-E on the resolution provided by Oka’s algorithm has nonnegative multiplicities on nodes, cf. section 10 and the sentence after theorem 10.1. Similar computations are given in [7], providing a stronger result in the case of a hypersurface singularity in 3\mathbb{C}^{3} with rational homology sphere link.

The concept of a B1B_{1}-facet appears in [11] in the case of hypersurfaces in KrK^{r}, where KK is a pp-adic field, and is further studied in [18, 6].

\thedefinition Definition.

Let FΓ(f)F\subset\Gamma(f) be a compact facet, i.e. of dimension 22. Then FF is a B1B_{1}-facet if FF has exactly 33 vertices p1,p2,p3p_{1},p_{2},p_{3} so that there is a σΣ(1)\sigma\in\triangle^{(1)}_{\Sigma} so that mσ=σ(p1)=σ(p2)=σ(p3)1m_{\sigma}=\ell_{\sigma}(p_{1})=\ell_{\sigma}(p_{2})=\ell_{\sigma}(p_{3})-1. A B1B_{1}-facet FF is removable if furthermore, the segment [p2,p3][p_{2},p_{3}] is contained in the boundary Γ(f)\partial\Gamma(f) of Γ(f)\Gamma(f).

Refer to caption
p3p_{3}p2p_{2}p1p_{1}σ\sigmaFFσF\sigma_{F}σ+\sigma_{+}
Figure 6. On the left we have a Newton diagram in 03\mathbb{R}^{3}_{\geq 0} with a removable B1B_{1} facet FF. To the left, we see the 22-skeleton of the dual fan, and an intersection with a hyperplane. In this example we have σ(p1)=σ(p2)=0\ell_{\sigma}(p_{1})=\ell_{\sigma}(p_{2})=0 and σ(p3)=1\ell_{\sigma}(p_{3})=1.
\thedefinition Definition.

Let T(f)T(f) be closure in NN_{\mathbb{R}} of the union of cones in f\triangle_{f} which correspond to compact facets of Γ+(f)\Gamma_{+}(f) which have dimension >0>0. This is the tropicalization of ff. We say that Σ\Sigma is generated by the tropicalization of ff, if Σ\Sigma is generated as a cone by the set T(f)T(f).

Let Σ\Sigma^{\prime} be the cone generated by T(f)T(f). This is a finitely generated rational strictly convex cone, and if (X,0)(X,0) is not rational, then Σ\Sigma^{\prime} has dimension r=3r=3. This cone induces an affine toric variety Y=YΣY^{\prime}=Y_{\Sigma^{\prime}}, and the function ff defines a Weil divisor (X,0)(Y,0)(X^{\prime},0)\subset(Y^{\prime},0). Furthermore, the inclusion ΣΣ\Sigma^{\prime}\subset\Sigma induces a morphism YYY^{\prime}\to Y, which restricts to a morphism (X,0)(X,0)(X^{\prime},0)\to(X,0).

\therem Remark.

The closure of T(f)T(f) in a certain partial compactification of NN_{\mathbb{R}} is called the local tropicalization of (X,0)(X,0) [30].

Refer to caption
x3x^{3}xy3xy^{3}z2z^{2}T(f)T(f)MM_{\mathbb{R}}NN_{\mathbb{R}}
Figure 7. Here, Σ=03\Sigma=\mathbb{R}^{3}_{\geq 0} is the positive octant, and f(x,y,z)=x3+xy3+z2f(x,y,z)=x^{3}+xy^{3}+z^{2} is the E7E_{7} singularity in normal form. In this case, T(f)T(f) does not generate Σ\Sigma, but the cone generated by (2,0,1)(2,0,1), (0,1,0)(0,1,0) and (0,0,1)(0,0,1).
\thelemma Lemma.

Let σ~f\sigma\in\widetilde{\triangle}_{f}. Then the orbit OσO_{\sigma} intersects X~\widetilde{X} if and only if σT(f)\sigma\subset T(f).

Proof.

The orbit OσO_{\sigma} is an affine variety Oσ=Spec([M(σ)])O_{\sigma}=\mathop{\rm Spec}\nolimits(\mathbb{C}[M(\sigma)]) (recall M(σ)=Mσ)M(\sigma)=M\cap\sigma^{\perp}), and if pσp_{\sigma} is an element of the affine hull of FσF_{\sigma}, then xpσfσ[M(σ)]x^{-p_{\sigma}}f_{\sigma}\in\mathbb{C}[M(\sigma)] and

X~OσSpec([M(σ)](xpσfσ)).\widetilde{X}\cap O_{\sigma}\cong\mathop{\rm Spec}\nolimits\left(\frac{\mathbb{C}[M(\sigma)]}{(x^{-p_{\sigma}}f_{\sigma})}\right).

Therefore, X~Oσ\widetilde{X}\cap O_{\sigma} is empty if and only if xpσfσx^{-p_{\sigma}}f_{\sigma} is a unit in [M(σ)]\mathbb{C}[M(\sigma)], which is equivalent to fσf_{\sigma} being a monomial, i.e. dimFσ=0\dim F_{\sigma}=0. ∎

\thelemma Lemma.

Let (X,0)(X,0) and (X,0)(X^{\prime},0) be as in section 11. If (X,0)(X,0) is normal, then the morphism (X,0)(X,0)(X^{\prime},0)\to(X,0) is an isomorphism.

Proof.

We can assume that the smooth subdivision ~f\widetilde{\triangle}_{f} subdivides the cone Σ\Sigma^{\prime}, so that we get a subdivision ~f=~f|Σ\widetilde{\triangle}_{f}^{\prime}=\widetilde{\triangle}_{f}|_{\Sigma^{\prime}} of the cone Σ\Sigma^{\prime}. Let Y~\widetilde{Y}^{\prime} be the corresponding toric variety. Let T(f)\triangle_{T(f)} be the fan consisting of cones σ~f\sigma\in\widetilde{\triangle}_{f} which are contained in T(f)T(f). We then get open inclusions

YT(f)Y~Y~Y_{T(f)}\subset\widetilde{Y}^{\prime}\subset\widetilde{Y}

where YT(f)Y_{T(f)} is the toric variety associated with the fan T(f)\triangle_{T(f)}.

It follows from section 11 that the strict transforms X~\widetilde{X} and X~\widetilde{X}^{\prime} of XX and XX^{\prime}, respectively, are contained in YT(f)Y_{T(f)}, and so X~=X~\widetilde{X}^{\prime}=\widetilde{X}. As a result, X{0}X~π1(0)=X~π1(0)X{0}X^{\prime}\setminus\{0\}\cong\widetilde{X}^{\prime}\setminus\pi^{-1}(0)=\widetilde{X}\setminus\pi^{-1}(0)\cong X\setminus\{0\}. Since (X,0)(X,0) is normal, the morphism (X,0)(X,0)(X^{\prime},0)\to(X,0) is an isomorphism. ∎

\theblock.

Assume that FΓ(f)F\subset\Gamma(f) is a removable B1B_{1}-face, and let σΣ(1)\sigma\in\triangle^{(1)}_{\Sigma} and pip_{i} be as in section 11. If FF is the only facet of Γ(f)\Gamma(f), then we leave as an exercise to show that the graph GG is equivalent to a string of rational curves, and so (X,0)(X,0) is rational. We will always assume that FF is not the only facet of Γ(f)\Gamma(f). There exists an element of Σ\Sigma^{\circ} which is constant on the segment [p1,p3][p_{1},p_{3}] (e.g. the normal vector to FF). As a result, the boundary Σ\partial\Sigma intersects the hyperplane of elements N\ell\in N_{\mathbb{R}} which are constant on [p1,p3][p_{1},p_{3}] in two rays, σ+\sigma_{+} and σ\sigma_{-}, where σ+\ell\in\sigma_{+} satisfies |[p1,p3]maxF\ell|_{[p_{1},p_{3}]}\equiv\max_{F}\ell, and σ\ell\in\sigma_{-} satisfies |[p1,p3]minF\ell|_{[p_{1},p_{3}]}\equiv\min_{F}\ell.

Let +N\ell_{+}\in N be a primitive generator of σ+\sigma_{+}, set m+=maxFm_{+}=\max_{F}\ell and define

f¯(x)={apxp|pM,+(p)m+},\bar{f}(x)=\sum\left\{a_{p}x^{p}\,\middle|\,p\in M,\,\ell_{+}(p)\geq m_{+}\right\},

where apa_{p} are the coefficients of ff as in eq. 3.1. Let (X¯,0)(\bar{X},0) be the Weil divisor defined by f¯\bar{f}. We get a Newton polyhedron Γ+(f¯)\Gamma_{+}(\bar{f}), from which we calculate invariants of (X¯,0)(\bar{X},0) as described in previous sections. It follows from this construction that Γ(f¯)=Γ(f)F¯\Gamma(\bar{f})=\overline{\Gamma(f)\setminus F}, and that f¯\bar{f} is Newton nondegenerate.

Now, assume that Σ\Sigma is generated by the tropicalization of ff. Let σ1\sigma_{1} and σ3f(1)\sigma_{3}\in\triangle_{f}^{(1)} be the rays corresponding to the noncompact faces of Γ+(f)\Gamma_{+}(f) containing the segments [p2,p3][p_{2},p_{3}] and [p1,p2][p_{1},p_{2}], respectively. Let 1,3\ell_{1},\ell_{3} be primitive generators of σ1,σ3\sigma_{1},\sigma_{3}. By construction, and the above assumption that Σ\Sigma is generated by T(f)T(f), we have 01,3Σ\mathbb{R}_{\geq 0}\langle\ell_{1},\ell_{3}\rangle\subset\partial\Sigma, and so +01,3f\ell_{+}\in\mathbb{R}_{\geq 0}\langle\ell_{1},\ell_{3}\rangle\in\triangle_{f}.

In fact, we have +=1+t3\ell_{+}=\ell_{1}+t\ell_{3} where t=1(p1p2)t=\ell_{1}(p_{1}-p_{2}). Indeed, +\ell_{+} is the unique positive linear combination of 1\ell_{1} and 3\ell_{3} which vanishes on p1p3p_{1}-p_{3}, and is primitive. Since 3=σ\ell_{3}=\ell_{\sigma}, by definition of FF, and since 1(p3)=1(p2)\ell_{1}(p_{3})=\ell_{1}(p_{2}), we have

(1+t3)(p1p3)=1(p1p3)+1(p1p2)3(p1p3)=1(p1p2)1(p1p2)=0.(\ell_{1}+t\ell_{3})(p_{1}-p_{3})=\ell_{1}(p_{1}-p_{3})+\ell_{1}(p_{1}-p_{2})\cdot\ell_{3}(p_{1}-p_{3})=\ell_{1}(p_{1}-p_{2})-\ell_{1}(p_{1}-p_{2})=0.

Furthermore, we have 1(p3p2)=0\ell_{1}(p_{3}-p_{2})=0 and 3(p3p2)=1\ell_{3}(p_{3}-p_{2})=1, and so by section 4, 1,3\ell_{1},\ell_{3} form a part of an integral basis, which implies that 1+t3\ell_{1}+t\ell_{3} is primitive.

Now, define tt^{\prime} as the combinatorial length of the segment [p1,p2][p_{1},p_{2}]. We have t|tt^{\prime}|t and via Oka’s algorithm (section 6), this segment corresponds to tt^{\prime} bamboos in GG, each consisting of a single (1)(-1)-curve, whereas [p2,p3][p_{2},p_{3}] corresponds to one bamboo with determinant t/tt/t^{\prime}.

\theprop Proposition.

Let ff, FF and f¯\bar{f} be as above, and assume that ff is Newton nondegenerate. Assume also that Σ\Sigma is generated by the tropicalization T(f)T(f) as described in section 11. Then

  1. (i)

    f¯\bar{f} is Newton nondegenerate.

  2. (ii)

    Γ(f¯)=Γ(f)F¯\Gamma(\bar{f})=\overline{\Gamma(f)\setminus F}.

  3. (iii)

    The singularities (X,0)(X,0) and (X¯,0)(\bar{X},0) have diffeomorphic links.

  4. (iv)

    The singularities (X,0)(X,0) and (X¯,0)(\bar{X},0) have equal geometric genera and δ\delta-invariants.

  5. (v)

    If (X,0)(X,0) is normal, then (X¯,0)(\bar{X},0) is normal.

  6. (vi)

    If ff is \mathbb{Q}-Gorenstein-pointed at pMp\in M_{\mathbb{Q}}, then so is f¯\bar{f}. In particular, if (X,0)(X,0) is Gorenstein, then (X¯,0)(\bar{X},0) is also Gorenstein.

Proof.

item (i) and item (ii) follow from definition.

We now prove item (iii). We have GG, the output of Oka’s algorithm for the Newton polyhedron Γ+(f)\Gamma_{+}(f), and G¯\bar{G}, the output of Oka’s algorithm for Γ+(f¯)\Gamma_{+}(\bar{f}). Let σFf\sigma_{F}\in\triangle_{f} be the ray dual to FF and let FF^{\prime} be the unique face of Γ+(f)\Gamma_{+}(f) adjacent to FF, i.e. FF=[p1,p3]F^{\prime}\cap F=[p_{1},p_{3}]. Then σF0F,+f¯\sigma_{F}\subset\mathbb{R}_{\geq 0}\langle\ell_{F^{\prime}},\ell_{+}\rangle\in\triangle_{\bar{f}}, and we can subdivide the canonical subdivision of 0F,+\mathbb{R}_{\geq 0}\langle\ell_{F}^{\prime},\ell_{+}\rangle so that we can assume that σF~f¯\sigma_{F}\in\widetilde{\triangle}_{\bar{f}}. We can therefore identify vertices vFv_{F} of GG and G¯\bar{G} corresponding to the same ray σF~f(1)\sigma_{F}\in\widetilde{\triangle}^{(1)}_{f} and σF~f¯(1)\sigma_{F}\in\widetilde{\triangle}^{(1)}_{\bar{f}}. It is then clear from construction that the components of GvFG\setminus v_{F} and G¯vF\bar{G}\setminus v_{F} in the direction of vFv_{F^{\prime}} are isomorphic. After blowing down the (1)(-1)-curves corresponding to the segment [p1,p2][p_{1},p_{2}], we must show

  • The two bamboos joining F\ell_{F} with +\ell_{+} on one hand, and with 1\ell_{1} on the other, are isomorphic.

  • The vertex vFv_{F} has the same Euler number in GG and in G¯\bar{G}.

Refer to caption
1-1v3v_{3}v1v_{1}vFv_{F}vFv_{F^{\prime}}vFv_{F^{\prime}}v+v_{+}GG^{*}G¯\bar{G}^{*}v¯F\bar{v}_{F}uuuuuu^{\prime}u¯\bar{u}^{\prime}
Figure 8. The (1)(-1)-curve to the left is blown down, so that the two graphs GG and G¯\bar{G}, obtained by deleting v3,v1,v+v_{3},v_{1},v_{+} and their adjacent edges, look topologically the same. To the right, the bamboo connecting v¯F\bar{v}_{F^{\prime}} and v+v_{+} corresponds to a subdivision of the cone generated by +\ell_{+} and F\ell_{F^{\prime}} which contains the ray generated by F\ell_{F}.

For the first of these, we prove that

α(F,+)=α(F,1),β(F,+)=β(F,1).\alpha(\ell_{F},\ell_{+})=\alpha(\ell_{F},\ell_{1}),\quad\beta(\ell_{F},\ell_{+})=\beta(\ell_{F},\ell_{1}).

We calculate α(F,+)\alpha(\ell_{F},\ell_{+}) as the greatest common divisor of maximal minors of the matrix having coordinate vectors for F\ell_{F} and +=1+t3\ell_{+}=\ell_{1}+t\ell_{3} as rows. But α(F,1)=t/t\alpha(\ell_{F},\ell_{1})=t/t^{\prime}, and so adding a multiple of tt to 1\ell_{1} does not modify the greatest common divisor of these determinants, hence α(F,+)=α(F,1+t3)=α(F,1)\alpha(\ell_{F},\ell_{+})=\alpha(\ell_{F},\ell_{1}+t\ell_{3})=\alpha(\ell_{F},\ell_{1}).

The invariant β(F,+)\beta(\ell_{F},\ell_{+}) can be calculated as the unique number 0β<α(F,+)0\leq\beta<\alpha(\ell_{F},\ell_{+}) so that βF++\beta\ell_{F}+\ell_{+} is a multiple of α(F,+)\alpha(\ell_{F},\ell_{+}). On the other hand, we find, setting β=β(F,1)\beta=\beta(\ell_{F},\ell_{1}) and α=α(F,+)=α(F,1)=t/t\alpha=\alpha(\ell_{F},\ell_{+})=\alpha(\ell_{F},\ell_{1})=t/t^{\prime},

βF++α=βF+1+t3α=βF+1α+t3N.\frac{\beta\ell_{F}+\ell_{+}}{\alpha}=\frac{\beta\ell_{F}+\ell_{1}+t\ell_{3}}{\alpha}=\frac{\beta\ell_{F}+\ell_{1}}{\alpha}+t^{\prime}\ell_{3}\in N.

Finally, we show that vFv_{F} has the same Euler number in the graphs GG and G¯\bar{G}. Denote these by bF-b_{F} and b¯F-\bar{b}_{F}. After blowing down the (1)(-1) curves associated with the segment [p1,p2][p_{1},p_{2}], the vertex vFv_{F} has two neighbors in either graph GG or G¯\bar{G}. Denote by v1v_{-1} and v¯1\bar{v}_{-1} the neighbor of vFv_{F} contained in the same component of GvFG\setminus v_{F} and G¯vF\bar{G}_{\setminus}v_{F} as vFv_{F^{\prime}}. It is then clear that v1=v¯1\ell_{v_{-1}}=\ell_{\bar{v}_{-1}}.

Denote by u,v¯u,\bar{v} the neighbours of vF,v¯Fv_{F},\bar{v}_{F} in the direction of v1,v+v_{1},v_{+}, respectively, and u,u¯u^{\prime},\bar{u}^{\prime} the other neighbours, as in fig. 8. Then we have u=u¯\ell_{u^{\prime}}=\ell_{\bar{u}^{\prime}} and

u=βF+1α,u¯=βF++α=u+t3,\ell_{u}=\frac{\beta\ell_{F}+\ell_{1}}{\alpha},\quad\ell_{\bar{u}}=\frac{\beta\ell_{F}+\ell_{+}}{\alpha}=\ell_{u}+t^{\prime}\ell_{3},

where α,β\alpha,\beta are as above. The two numbers bF-b_{F} and b¯F-\bar{b}_{F} are identified by section 6

bFF+u+u+t3=0,b¯FF+u¯+u¯=0,-b_{F}\ell_{F}+\ell_{u}+\ell_{u^{\prime}}+t^{\prime}\ell_{3}=0,\quad-\bar{b}_{F}\ell_{F}+\ell_{\bar{u}}+\ell_{\bar{u}^{\prime}}=0,

which leads to their equality.

Next, we prove item (iv) and item (v). By theorem 7.1, it suffices to show that

Γ+(f)(Σ+q),Γ+(f¯)(Σ+q)\Gamma_{+}(f)\setminus(\Sigma^{\vee}+q),\qquad\Gamma_{+}(\bar{f})\setminus(\Sigma^{\vee}+q)

have the same cohomology for all qMq\in M. By shifting Γ+(f)\Gamma_{+}(f), we simplify the following proof by assuming q=0q=0. The inclusion

Γ(f)ΣΓ+(f)Σ\Gamma(f)\setminus\Sigma^{\vee}\subset\Gamma_{+}(f)\setminus\Sigma^{\vee}

is a homotopy equivalence. Indeed, one can construct a suitable vectorfied on Γ+(f)Σ\Gamma_{+}(f)\setminus\Sigma^{\vee} pointing in the direction of Σ-\Sigma^{\vee}, whose trajectories end up in Γ(f)Σ\Gamma(f)\setminus\Sigma, thus giving a homotopy inverse to the above inclusion.

Now, let KK be the union of faces of Γ(f)\Gamma(f) which do not intersect Σ\Sigma^{\vee}. By section 7, the inclusion KΓ(f)ΣK\subset\Gamma(f)\setminus\Sigma is a homotopy equivalence. Define K¯\bar{K} similarly, using f¯\bar{f}. Thus it suffices to prove that H~i(K,K¯;)\widetilde{H}^{i}(K,\bar{K};\mathbb{Z}) vanish for all ii. By excision, this is equivalent to showing

(11.1) i0:H~i(KF,K¯F;)=0.\forall i\in\mathbb{Z}_{\geq 0}:\,\widetilde{H}^{i}(K\cap F,\bar{K}\cap F;\mathbb{Z})=0.

If Σ\Sigma^{\vee} does not intersect the face FF, then KF=F=K¯FK\cap F=F=\bar{K}\cap F. Also, if p2Σp_{2}\in\Sigma^{\vee}, then KF=K¯FK\cap F=\bar{K}\cap F. In either case, eq. 11.1 holds. We can therefore assume that p2Kp_{2}\in K and FKF\not\subset K. With these assumptions at hand, it is then enough to prove that excactly one of the segements [p1,p2][p_{1},p_{2}] and [p2,p3][p_{2},p_{3}] is contained in KK, i.e. it cannot happen that either both or neither is contained in KK.

Let AA be the affine hull of FnF_{n}, i.e. the hyperplane in MM_{\mathbb{R}} defined by n=mn\ell_{n}=m_{n}, and let C=ΣAC=\Sigma^{\vee}\cap A. Define a point rAr\in A by

3(r)=0,1(r)=0,n(r)=mn.\ell_{3}(r)=0,\quad\ell_{1}(r)=0,\quad\ell_{n}(r)=m_{n}.

This is well defined, since the functions 1,3,n\ell_{1},\ell_{3},\ell_{n} are linearly independent. Then CC is a convex polygon in AA, and rr is a vertex of CC. Furthermore, rr is the unique point in CC where both functions 1|C\ell_{1}|_{C} and 3|C\ell_{3}|_{C} take their minimal values.

If neither of the segments [p1,p2][p_{1},p_{2}], [p2,p3][p_{2},p_{3}] are contained in KK, i.e. both intersect Σ\Sigma^{\vee}, then we can choose r1C[p1,p2]r_{1}\in C\cap[p_{1},p_{2}] and r2C[p2,p3]r_{2}\in C\cap[p_{2},p_{3}]. Furthermore, we have 3(r)3(r1)=3(p2)\ell_{3}(r)\leq\ell_{3}(r_{1})=\ell_{3}(p_{2}), and 1(r)1(r2)=1(p2)\ell_{1}(r)\leq\ell_{1}(r_{2})=\ell_{1}(p_{2}). Therefore, p2p_{2} is in the convex hull of r,r1,r2r,r_{1},r_{2}, and so p2Cp_{2}\in C, contrary to the assumption p2Kp_{2}\in K.

Refer to caption
FnF_{n}rrrr^{\prime}3<m3\ell_{3}<m_{\ell_{3}}3>m3\ell_{3}>m_{\ell_{3}}p1p_{1}p3p_{3}p2p_{2}1>m1\ell_{1}>m_{\ell_{1}}1<m1\ell_{1}<m_{\ell_{1}}AA
Figure 9. The segment [r,r][r^{\prime},r] intersects neither [p1,p2][p_{1},p_{2}] nor [p2,p3][p_{2},p_{3}].

Next, assume that both segments [p1,p2][p_{1},p_{2}], [p2,p3][p_{2},p_{3}] are contained in KK. We start by showing that in this case, we have rFnr\in F_{n}. By assumption, we can choose rCFnr^{\prime}\in C\cap F_{n}.

We have 3(r)>m3\ell_{3}(r^{\prime})>m_{\ell_{3}}. One verifies (see fig. 9) that if 3(r)m3\ell_{3}(r)\leq m_{\ell_{3}}, then we would have 1(r)<1(r)\ell_{1}(r^{\prime})<\ell_{1}(r), but rr is a minimum for 1|C\ell_{1}|_{C}. Therefore, we can assume that 3(r)>m3\ell_{3}(r)>m_{\ell_{3}}, similarly, 1(r)>m1\ell_{1}(r)>m_{\ell_{1}}. It follows, since CFC\cap F\neq\emptyset, that rFr\in F, so we can assume that r=rr^{\prime}=r. But, since r[p1,p2][p2,p3]r\notin[p_{1},p_{2}]\cup[p_{2},p_{3}], we find

3(p2)<3(r)<3(p3)=3(p2)+1,\ell_{3}(p_{2})<\ell_{3}(r)<\ell_{3}(p_{3})=\ell_{3}(p_{2})+1,

and so 3(r)\ell_{3}(r)\notin\mathbb{Z}. But this is a contradiction, since 3(r)=3(q)\ell_{3}(r)=\ell_{3}(q)\in\mathbb{Z}.

Next we prove item (vi). Assume that Γ+(f)\Gamma_{+}(f) is \mathbb{Q}-Gorenstein pointed at pMp\in M_{\mathbb{Q}}. It suffices to show that +(p)=m¯++1\ell_{+}(p)=\bar{m}_{\ell_{+}}+1, where m¯+\bar{m}_{\ell_{+}} is the minimal value of +\ell_{+} on Γ+(f¯)\Gamma_{+}(\bar{f}). We immediately find

m¯+=+(p3)=1(p3)+t3(p3)=m1+t(m3+1)=1(p)1+t3(p)=+(p)1.\bar{m}_{\ell_{+}}=\ell_{+}(p_{3})=\ell_{1}(p_{3})+t\ell_{3}(p_{3})=m_{\ell_{1}}+t(m_{\ell_{3}}+1)=\ell_{1}(p)-1+t\ell_{3}(p)=\ell_{+}(p)-1.\qed
\theexample Example.

Consider the cone Σ=03\Sigma=\mathbb{R}^{3}_{\geq 0} and the function

f(x,y,z)=x3+xy3+z5+y10z,f(x,y,z)=x^{3}+xy^{3}+z^{5}+y^{10}z,

which defines a nonrational singularity (X,0)(X,0). In this case, Γ(f)\Gamma(f) has a B1B_{1}-facet

F=conv{(1,3,0),(0,10,1),(0,0,5)},F=\mathop{\rm conv}\nolimits\{(1,3,0),(0,10,1),(0,0,5)\},

corresponding to a node n𝒩n\in\mathcal{N}. The normal vector to FF is (19,2,5)(19,2,5) and eq. 8.2 gives mn(ZKE)=1m_{n}(Z_{K}-E)=-1. By the above computations, removing the monomial y10zy^{10}z from ff gives another singularity with the same link and geometric genus, but ZKEZ_{K}-E is nonnegative on the other node. After removing FF we find

f¯(x,y,z)=x3+xy3+z5.\bar{f}(x,y,z)=x^{3}+xy^{3}+z^{5}.

Note that Σ\Sigma is generated by the tropicalization of ff, but the tropicalization of f¯\bar{f} generates the cone 0(5,0,1),(0,1,0),(0,0,1)\mathbb{R}_{\geq 0}\langle(5,0,1),(0,1,0),(0,0,1)\rangle.

Refer to caption
x3x^{3}xy3xy^{3}(1,0,1)(1,0,1)(1,0,0)(1,0,0)(0,0,1)(0,0,1)(0,1,0)(0,1,0)(15,10,9)(15,10,9)z5z^{5}(5,0,1)(5,0,1)FF(19,2,5)(19,2,5)y5z3y^{5}z^{3}y10zy^{10}z
Figure 10. A diagram with a B1B_{1}-facet FF and its dual. The dotted line to the right replaces its two neighbouring segments if the B1B_{1}-facet is removed.
\theblock.

In what follows, we connect the above construction with the coefficients of ZKEZ_{K}-E. We introduce a simplified graph, whose vertices are the nodes of GG. whose vertices are the nodes of GG, and a bamboo of GG connecting two nodes of GG is replaced in G𝒩G_{\mathcal{N}} by an edge. Then G𝒩G_{\mathcal{N}} is a tree, with an edge connecting n,nn,n^{\prime} if and only if FnF_{n} and FnF^{\prime}_{n} intersect in a segment (of length 11). Recall that a leaf of a tree is a vertex with exactly one neighbour. If we assume that |𝒩|>1|\mathcal{N}|>1, then we see that the following are equivalent, since G𝒩G_{\mathcal{N}} is a tree:

  • n𝒩n\in\mathcal{N} is a leaf in G𝒩G_{\mathcal{N}},

  • Γ(f)Fn\Gamma(f)\setminus F_{n} is connected,

  • all edges of FnF_{n}, except for one, lie on the boundary Γ(f)\partial\Gamma(f) of the Newton diagram.

If |𝒩|=1|\mathcal{N}|=1, then there is a unique n𝒩n\in\mathcal{N}, and Γ(f)=F\Gamma(f)=F, in particular, Γ(f)=Fn\partial\Gamma(f)=\partial F_{n}. Finally, if |𝒩|=0|\mathcal{N}|=0, and if we assume that (X,0)(X,0) is normal, then (X,0)(X,0) is rational.

Refer to caption
length = ttlength = ttlength = ss
Figure 11. A big triangle, a small triangle of type t=3t=3, and a trapezoid of type (t,s)=(4,2)(t,s)=(4,2).

The following lemma is elementary:

\thelemma Lemma.

Let FF be an integral polyhedron in 2\mathbb{R}^{2}, having no integral interior points. Then, up to an integral affine automorphism of 2\mathbb{R}^{2}, FF is one one the following:

  • Big triangle The convex hull of (0,0)(0,0), (2,0)(2,0), (0,2)(0,2).

  • Small triangle of type tt The convex hull of (0,0)(0,0), (t,0)(t,0), (0,1)(0,1).

  • Trapezoid of type (t,s)(t,s) The convex hull of (0,0)(0,0), (t,0)(t,0), (0,1)(0,1), (s,1)(s,1), where t,st,s\in\mathbb{Z}, ts>0t\geq s>0 and t>0t>0. ∎

\thelemma Lemma.

Assume that (X,0)(X,0) is normal, Gorenstein-pointed at pMp\in M, and not rational. If n𝒩n\in\mathcal{N} is a leaf in G𝒩G_{\mathcal{N}} and mn(ZKE)<0m_{n}(Z_{K}-E)<0, then FnF_{n} is a removable B1B_{1}-facet of Γ(f)\Gamma(f) (See section 11 for the definition of G𝒩G_{\mathcal{N}}).

Proof.

By assumption, FnF_{n} has two adjacent edges contained in Γ(f)\partial\Gamma(f), say [q1,q2][q_{1},q_{2}] and [q2,q3][q_{2},q_{3}]. Let F1,F2F_{1},F_{2} be the noncompact faces of Γ+(f)\Gamma_{+}(f) containing the segments [q1,q2][q_{1},q_{2}] and [q2,q3][q_{2},q_{3}], respectively, and let 1,2Σ\ell_{1},\ell_{2}\in\partial\Sigma be the primitive functions having F1,F2F_{1},F_{2} as minimal sets on Γ+(f)\Gamma_{+}(f), denote these minimal values by m1,m2m_{\ell_{1}},m_{\ell_{2}}.

Let l1=length([q2,q3])l_{1}=\mathop{\mathrm{length}}([q_{2},q_{3}]) and α1=1(q3q2)/l1\alpha_{1}=\ell_{1}(q_{3}-q_{2})/l_{1} and l2=length([q2,q3])l_{2}=\mathop{\mathrm{length}}([q_{2},q_{3}]) and α2=2(q1q2)/l2\alpha_{2}=\ell_{2}(q_{1}-q_{2})/l_{2}. Then, the bamboos corresponding to the segements [q1,q2][q_{1},q_{2}] and [q2,q3][q_{2},q_{3}] have determininats α1,α2\alpha_{1},\alpha_{2}, see section 4.

Assume first that FnF_{n} is a small triangle of type tt, that the segment [q1,q2][q_{1},q_{2}] has length tt, and that α1=1\alpha_{1}=1. This implies that FnF_{n} is a removable B1B_{1}-facet.

Otherwise, let AA be the affine hull of FnF_{n}. If FnF_{n} is a big triangle, a trapezoid, or a small triangle as above, but with α1>1\alpha_{1}>1, then the square

{qA|m11(q)m1+1,m22(q)m2+1}\left\{q\in A\,\middle|\,m_{\ell_{1}}\leq\ell_{1}(q)\leq m_{\ell_{1}}+1,\,m_{\ell_{2}}\leq\ell_{2}(q)\leq m_{\ell_{2}}+1\right\}

is contained in FnF_{n}. In particular, its vertex q0q_{0}, the unique point in AA satisfying i(q0)=mi+1\ell_{i}(q_{0})=m_{\ell_{i}}+1 for i=1,2i=1,2, is contained in FnF_{n}. The set

R={qΣ|i(q)=0,i=1,2}R=\left\{q\in\Sigma^{\vee}\,\middle|\,\ell_{i}(q)=0,\;i=1,2\right\}

is a one dimensional face of Σ\Sigma^{\vee} (here we use the condition that Σ\Sigma is generated by the tropicalization of (X,0)(X,0)). By our assumption mnn(p)m_{n}\leq\ell_{n}(p) we have pq0+RΓ+(f)p\in q_{0}+R^{\circ}\subset\Gamma_{+}(f)^{\circ}, contradicting the assumption that (X,0)(X,0) is not rational. ∎

\theprop Proposition.

Assume that (X,0)(X,0) is normal, Gorenstein-pointed at pMp\in M, and not rational. If there is an n𝒩n\in\mathcal{N} so that mn(ZKE)<0m_{n}(Z_{K}-E)<0, then Γ(f)\Gamma(f) has a removable B1B_{1}-facet.

Proof.

If nn is a leaf in G𝒩G_{\mathcal{N}} (see section 11), then FnF_{n} is removable by section 11. So let us assume that nn is not a leaf in G𝒩G_{\mathcal{N}}, i.e. that Γ(f)Fn\Gamma(f)\setminus F_{n} is disconnected. The inclusion

Γ(f)FnΓ+(f)({nmn}Γ+(f))\Gamma(f)^{\circ}\setminus F_{n}\subset\Gamma_{+}^{*}(f)^{\circ}\setminus(\{\ell_{n}\leq m_{n}\}\cup\Gamma_{+}(f)^{\circ})

is a strong homotopy retract (here we set Γ(f)=Γ(f)Γ(f)\Gamma(f)^{\circ}=\Gamma(f)\setminus\partial\Gamma(f)). In particular, the right hand side is disconnected as well. But it follows from our assumptions that the point pp is in the right hand side above. Let CC be a component of Γ(f)Fn\Gamma(f)\setminus F_{n} contained in a component of the right hand side which does not contain pp. Then, for any nn^{\prime} so that FnC¯F_{n^{\prime}}\subset\overline{C} we have n(p)>mn\ell_{n^{\prime}}(p)>m_{n^{\prime}}, i.e. mn(ZKE)<0m_{n^{\prime}}(Z_{K}-E)<0. Let GCG_{C} be the induced subgraph of G𝒩G_{\mathcal{N}} having vertices nn^{\prime} for FnC¯F_{n^{\prime}}\subset\overline{C}. This graph is a nonempty tree, and so has either exactly one vertex, or at has least two leaves. In the first case, the unique vertex nn^{\prime} of GCG_{C} is a leaf of GG. In the second case, GCG_{C} has at least two leaves, so we can choose a leaf nn^{\prime} of GCG_{C} which is not adjacent to nn in GG. In either case, FnF_{n^{\prime}} is a removable B1B_{1}-facet by section 11. ∎

\theprop Proposition.

Assume that ff defines a normal Newton nondegenerate Weil divisor (X,0)(X,0), which is not rational. Then there exists a normal Newton nondegenerate Weil divisor (X¯,0)(\bar{X},0), defined by a function f¯\bar{f} and a cone Σ\Sigma^{\prime} (possibly different than Σ\Sigma) satisfying the following conditions:

  • (X¯,0)(\bar{X},0) and (X,0)(X,0) have diffeomorphic links.

  • pg(X¯,0)=pg(X,0)p_{g}(\bar{X},0)=p_{g}(X,0).

  • If (X,0)(X,0) is Gorenstein or pointed at pMp\in M_{\mathbb{Q}}, then so is (X¯,0)(\bar{X},0).

  • If FnΓ+(f¯)F_{n}\subset\Gamma_{+}(\bar{f}) is a compact facet, then mn(ZKE)0m_{n}(Z_{K}-E)\geq 0.

In fact, Γ(f¯)\Gamma(\bar{f}) is the union of those facets FnF_{n} of Γ(f)\Gamma(f) for which mn(ZKE)0m_{n}(Z_{K}-E)\geq 0.

Proof.

By section 11, we can assume that Σ\Sigma is generated by T(f)T(f), since (X,0)(X,0) is normal (see section 11). The result therefore follows, using induction on the number of facets of Γ(f)\Gamma(f), and sections 11 and 11 below. ∎

12. Examples

\theexample Example.

Let N=M=3N=M=\mathbb{Z}^{3} and let a,b,ca,b,c\in\mathbb{N} be natural numbers with no common factor, and let 0r<s0\leq r<s\in\mathbb{N} be coprime with srcs\leq rc. Take

Σ=0(ra,0,s)(0,rb,s)(0,0,1),f=x1a+x2b+x3c.\Sigma^{\vee}=\mathbb{R}_{\geq 0}\left\langle\begin{array}[]{@{(}rrr@{)}}ra,&0,&-s\\ 0,&rb,&-s\\ 0,&0,&1\end{array}\right\rangle,\quad f=x_{1}^{a}+x_{2}^{b}+x_{3}^{c}.

The cone Σ\Sigma is then generated by

1=(1,0,0),2=(0,1,0),3=1gcd(ab,s)(bs,as,abr).\ell_{1}=(1,0,0),\quad\ell_{2}=(0,1,0),\quad\ell_{3}=\frac{1}{\gcd(ab,s)}(bs,as,abr).

Corresponding to these, we have irreducible invariant divisors D1,D2,D3YD_{1},D_{2},D_{3}\subset Y and multiplicities

m1=0,m2=0,m3=absgcd(ab,s).m_{1}=0,\quad m_{2}=0,\quad m_{3}=\frac{abs}{\gcd(ab,s)}.

The Newton diagram Γ(f)\Gamma(f) consists of a single face with normal vector 0=(bc,ac,ab)\ell_{0}=(bc,ac,ab) and m0=abcm_{0}=abc. Fulton shows in 3.4 of [13] that the group of Weil divisors modulo linear equivalence on YY is generated by D1,D2,D3D_{1},D_{2},D_{3}, and that j=13aiDi\sum_{j=1}^{3}a_{i}D_{i} is Cartier if and only if there is a p=(p1,p2,p3)M=3p=(p_{1},p_{2},p_{3})\in M=\mathbb{Z}^{3} so that aj=j(p)a_{j}=\ell_{j}(p) for j=1,2,3j=1,2,3.

In our case, XX is equivalent to i=13miDi=m3D3-\sum_{i=1}^{3}m_{i}D_{i}=-m_{3}D_{3}. Therefore, if XX is Cartier, then there is a p=(p1,p2,p3)Mp=(p_{1},p_{2},p_{3})\in M so that i(p)=mi\ell_{i}(p)=m_{i}. Therefore, we find p1=p2=0p_{1}=p_{2}=0, and

abrgcd(ab,s)p3=absgcd(ab,s).\frac{abr}{\gcd(ab,s)}p_{3}=\frac{abs}{\gcd(ab,s)}.

Therefore, XX is Cartier if and only if r|sr|s, i.e. r=1r=1.

Refer to caption
MM_{\mathbb{R}}NN_{\mathbb{R}}(5,3,10)(1,0,0)(0,1,0)(35,21,15)y5y^{5}z7z^{7}x3x^{3}
Figure 12. In the above examples, we have a=3a=3, b=5b=5, c=7c=7, r=2r=2 and s=3s=3. The cone Σ\Sigma is generated by the vectors (1,0,0)(1,0,0), (0,1,0)(0,1,0) and (5,3,10)(5,3,10). Furthermore, (35,21,15)(35,21,15) is the normal vector to the unique facet of Γ(f)\Gamma(f).
\theexample Example.

In [24], Némethi and Okuma analyse upper and lower bounds for the geometric genus of singularities with a specific topological type, namely, whose link is given by the plumbing graph in fig. 13.

Refer to caption
3-33-32-22-21-11-113-13
Figure 13. A resolution graph

They show that for this graph, the path lattice cohomology is 44, but that the maximal geometric genus among analytic structures with this topological type is 33. As a result, this graph is not the topological type of a Newton nondegenerate Weil divisor in a toric affine space.

On the other hand, this topological type is realized by the complete intersection given by the splice equations

X={z4|z12z2+z32+z43=z13+z22+z42z3=0}.X=\left\{z\in\mathbb{C}^{4}\,\middle|\,z_{1}^{2}z_{2}+z_{3}^{2}+z_{4}^{3}=z_{1}^{3}+z_{2}^{2}+z_{4}^{2}z_{3}=0\right\}.

This singularity is in fact a Newton nondegenerate isolated complete intersection [28]. As a result, the methods of section 10 do not generalize in the most straightforward way to Newton nondegenerate complete intersections.

References

  • [1] Altmann, Klaus. Toric \mathbb{Q}-Gorenstein singularities. arXiv:alg-geom/9403003.
  • [2] Altmann, Klaus; Ploog, David. Displaying the cohomology of toric line bundles. Izv. Math., 84(4):683–693, 2020.
  • [3] Arnold, Vladimir I.; Gusein-Zade, Sabir M.; Varchenko, Alexander N. Singularities of differentiable maps. Volume 2. Modern Birkhäuser Classics. Birkhäuser/Springer, New York, 2012.
  • [4] Aroca, Fuensanta; Gómez-Morales, Mirna; Shabbir, Khurram. Torical modification of Newton non-degenerate ideals. Rev. R. Acad. Cienc. Exactas Fís. Nat., Ser. A Mat., RACSAM, 107(1):221–239, 2013.
  • [5] Artin, Michael. On isolated rational singularities of surfaces. Am. J. Math., 88:129–136, 1966.
  • [6] Bories, Bart; Veys, Willem. Igusa’s pp-adic local zeta function and the monodromy conjecture for non-degenerate surface singularities., volume 1145. Providence, RI: American Mathematical Society (AMS), 2016.
  • [7] Braun, Gábor; Némethi, András. Invariants of Newton non-degenerate surface singularities. Compos. Math., 143(4):1003–1036, 2007.
  • [8] Bruns, Winfried; Gubeladze, Joseph. Polytopes, rings, and K-theory. New York, NY: Springer, 2009.
  • [9] Bruns, Winfried; Herzog, Jürgen. Cohen-Macaulay rings. Rev. ed, volume 39. Cambridge: Cambridge University Press, rev. ed. edition, 1998.
  • [10] Danilov, V. I. The geometry of toric varieties. Uspekhi Mat. Nauk, 33(2(200)):85–134, 247, 1978.
  • [11] Denef, Jan. Poles of pp-adic complex powers and Newton polyhedra. Nieuw Arch. Wiskd., IV. Ser., 13(3):289–295, 1995.
  • [12] Eisenbud, David; Neumann, Walter D. Three-dimensional link theory and invariants of plane curve singularities., volume 110 of Annals of Mathematics Studies. Princeton University Press, 1985.
  • [13] Fulton, William. Introduction to toric varieties, volume 131 of Annals of Mathematics Studies. Princeton University Press, Princeton, NJ, 1993. The William H. Roever Lectures in Geometry.
  • [14] Karras, Ulrich. Local cohomology along exceptional sets. Math. Ann., 275:673–682, 1986.
  • [15] Kouchnirenko, A. G. Polyèdres de Newton et nombres de Milnor. Invent. Math., 32:1–31, 1976.
  • [16] László, Tamás. Lattice cohomology and Seiberg–Witten invariants of normal surface singularities. Ph.D. thesis, Central European University, 2013. ArXiv:1310.3682.
  • [17] Laufer, Henry B. On rational singularities. Am. J. Math., 94:597–608, 1972.
  • [18] Lemahieu, Ann; Van Proeyen, Lise. Monodromy conjecture for nondegenerate surface singularities. Trans. Am. Math. Soc., 363(9):4801–4829, 2011.
  • [19] Merle, Michel; Teissier, Bernard. Conditions d’adjonction. (D’après Du Val). In Séminaire sur les singularités des surfaces, volume 777 of Lecture Notes in Mathematics. Springer-Verlag, Berlin Heiderberg New York, 1980.
  • [20] Némethi, András. Resolution graphs of some surface singularities. I: Cyclic coverings. In Singularities in algebraic and analytic geometry. Papers from the AMS special session, San Antonio, TX, USA, January 13-14, 1999, 89–128. Providence, RI: American Mathematical Society (AMS), 2000.
  • [21] Némethi, András. On the Ozsváth-Szabó invariant of negative definite plumbed 3-manifolds. Geom. Topol., 9:991–1042, 2005.
  • [22] Némethi, András. Pairs of invariants of surface singularities. In Proceedings of the international congress of mathematicians, ICM 2018, Rio de Janeiro, Brazil, August 1–9, 2018. Volume II. Invited lectures, 749–779. Hackensack, NJ: World Scientific; Rio de Janeiro: Sociedade Brasileira de Matemática (SBM), 2018.
  • [23] Némethi, András; Nicolaescu, Liviu I. Seiberg–Witten invariants and surface singularities. Geom. Topol., 6:269–328, 2002.
  • [24] Némethi, András; Okuma, Tomohiro. Analytic singularities supported by a specific integral homology sphere link. Methods Appl. Anal., 24(2):303–320, 2017.
  • [25] Némethi, András; Sigurðsson, Baldur. The geometric genus of hypersurface singularities. J. Eur. Math. Soc. (JEMS), 18(4):825–851, 2016.
  • [26] Oda, Tadao. Convex bodies and algebraic geometry. An introduction to the theory of toric varieties. Berlin etc.: Springer-Verlag, 1988.
  • [27] Oka, Mutsuo. On the resolution of the hypersurface singularities. In Complex analytic singularities, volume 8 of Adv. Stud. Pure Math., 405–436. North-Holland, Amsterdam, 1987.
  • [28] Oka, Mutsuo. Principal zeta-function of non-degenerate complete intersection singularity. J. Fac. Sci., Univ. Tokyo, Sect. I A, 37(1):11–32, 1990.
  • [29] Popescu-Pampu, Patrick. The geometry of continued fractions and the topology of surface singularities. In Singularities in geometry and topology 2004, volume 46 of Adv. Stud. Pure Math., 119–195. Math. Soc. Japan, Tokyo, 2007.
  • [30] Popescu-Pampu, Patrick; Stepanov, Dmitry. Local tropicalization. In Algebraic and combinatorial aspects of tropical geometry. Proceedings based on the CIEM workshop on tropical geometry, International Centre for Mathematical Meetings (CIEM), Castro Urdiales, Spain, December 12–16, 2011, 253–316. Providence, RI: American Mathematical Society (AMS), 2013.
  • [31] Sigurðsson, Baldur. The geometric genus and the Seiberg–Witten invariant of Newton nondegenerate surface singularities. Ph.D. thesis, Central European University, 2015. ArXiv:1601.02572.
  • [32] Varchenko, Alexander N. Zeta-function of monodromy and Newton’s diagram. Invent. Math., 37:253–262, 1976.