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
Abstract.
We introduce and develop the theory of Newton nondegenerate local Weil divisors 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 and the corresponding canonical divisor. We provide combinatorial formulae for the delta–invariant and for the cohomology groups for . In the case , 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.
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 –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 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 -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 -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 , namely that the geometric genus is determined by a computation sequence, and is therefore topologically determined:
1.1 Theorem.
Let be a two-dimensional Newton nondegenerate Weil divisor in the affine toric ambients space . Assume that is normal and Gorenstein, and that its link is a rational homology sphere. Then the geometric genus equals the minimal path lattice cohomology associated with the link of . In particular, the geometric genus is determined by the topology of .
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 be a free Abelian group of rank and set , as well as and . If is a cone, the dual cone is defined as
We also set
We will always assume cones to be finitely generated and rational. To a cone we associate the semigroup , the algebra and the affine variety by setting
A variety of the form is called an affine toric variety. It has a canonical action of the -torus .
\theblock.
A fan in is a collection of cones in satisfying the following two conditions.
-
(i)
Any face of a cone in is in .
-
(ii)
The intersection of two cones in is a face of each of them.
The support of a fan is defined as . If and , then we get a morphism . These morphisms form a direct system, whose limit is denoted by and called the associated toric variety. The actions of on the affine varieties for glue together to form an action on . Note that the canonical maps are open inclusions (note also that the notation differs from [13]).
Let be another fan in a lattice and let be a linear map. Assume that for any there is a so that . This induces maps , which glue together to form a map .
\thelemma Lemma (Proposition, p. 39, [13]).
Let and be fans as above. The induced map is proper if and only if .
\theblock.
For any , there is an associated rational function on . These glue together to form a rational function on . We refer to these functions as monomials. A monomial is a regular function on if . A map as above induces . The monomial on then pulls back to .
\theblock.
For , let be the closed subset of defined by the ideal generated by monomials where . We identify this set with its image in . The closure of in is denoted by . In the case when is a ray, is a Weil divisor and we write . The orbits of the action on are precisely the sets for . Furthermore, we have (as sets)
Let be the subgroup of generated by and define
Note that this way we have and . Let be the canonical projection and set
This set is a fan in , whose associated toric variety is identified canonically with the orbit closure . Similarly, let be the canonical projection. Assuming has dimension , we have . In particular, has as a transverse type.
\thedefinition Definition.
-
(i)
For a cone , let denote the fanfan consisting of all the faces of . We also write instead of .
-
(ii)
If is a fan and , define
-
(iii)
A regular cone (resp. simplicial cone) is a cone generated by a subset of an integral (resp. rational) basis of .
-
(iv)
A subdivision of a fan is a fan so that and each cone in is a union of cones in . A regular subdivision is a subdivision consisting of regular cones.
-
(v)
If is a cone and is a subdivision of , denote by the fan consisting of for which . Here we see as the union of the proper faces of . As a result, is a subdivision of the fan .
-
(vi)
Let be subdivisions of a fan . We say that refines if is a subdivision of , or that is a refinement of .
-
(vii)
Let be a fan with a subdivision and let . The restriction of to is defined as
3. Analytic Weil divisors in affine toric varieties
\theblock.
Throughout this section, as well as the following sections, we will assume that has rank and that is an -dimensional, rational, finitely generated, strictly convex cone in . This means that is generated over by a finite set of elements from , which generate as a vectorspace, and that . In particular, the orbit consists of a single point, which we denote by , and refer to as the origin. Let be the affine toric variety associated with .
Any subdivision of induces a modification .
In the sequel we denote by the analytic germ of at 0, and usually we will denote by a (small Stein) representative of . (Hence .) If is a toric modification, in the discussions regarding the local analytic germ , we will use the same notation for and for . Similarly, might stay for as well. If in some argument we really wish to stress the differences, we write , , for the local objects.
Assume that is the germ of a holomorphic function at the origin, which has an expansion
(3.1) |
Then is the germ of an analytic space. We set too.
\thedefinition Definition.
The Newton polyhedron of with respect to is the polyhedron
where denotes the convex closure in . The union of compact faces of is denoted by and is called the Newton diagram of with respect to .
\theblock.
The fan and some combinatorial properties. It follows from definition that is precisely the set of those linear functions on having a minimal value on . Denote by the minimal set of on . For , say that if and only if . Then is an equivalence relation on 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 . We refer to as the dual fan associated with and . Note that refines .
For any , the face is independent of the choice of , where is the relative interior, that is, the topological interior of as a subset of . For , the set is a semigroup generated by a unique element, which we denote by . For a series
the support is defined similarly as above, and for we set
One verifies that for any such
(3.2) |
\thedefinition Definition.
Let and . Define
If is a cone, and (for example, if is an element of a refinement of ), then we set .
If is one dimensional, set . Thus, . Note that we have
This can be compared with the following set.
\thedefinition Definition.
Let
where denotes the set of rays in contained in the boundary of .
\thedefinition Definition.
Denote by the union of those local primary components of the germ defined by (with their non-reduced structure), which are not invariant by the torus action. If is reduced along the non-invariant components, this means the following. Let be a neighbourhood of the origin on which converges and let be defined by . Then is the closure of in .
\therem Remark.
(i) For any , the function defines the same germ . Thus, we may allow as well.
(ii) Since the divisors are torus-invariant, the divisor of in is .
\theprop Proposition.
-
(i)
We have for some if and only if if and only if the germ at is the empty germ.
-
(ii)
For a , we have if and only if the normal fan subdivides into smaller cones, i.e. .
-
(iii)
The ideal which defines in , is generated by the functions for satisfying for all .
Proof.
Statement i is clear, since is of the form if and only if is a product of a monomial and a unit in .
Statement ii follows from i, and the fact that the intersection of and a generic transverse space to has Newton polygon , cf. section 2.
iii Assuming the given conditions on , the function is meromorphic and has no poles. Since is normal, is analytic and vanishes on . As a result, .
To show that these generate , take . We must show that , with and for .
Let be the localization of along the invariant divisors, that is, the ideal of meromorphic function germs on , regular on the open torus and vanishing on . It follows that and .
Thus, for some and . Then, there exist finite families of units in and exponents in so that and the support of is the disjoint union of the supports of . Let us take any . The condition on disjointness of supports gives
As a result, we have for all . The result follows. ∎
\thedefinition Definition.
Let and be as above. We say that , or , is (-)pointed if there exists a ( such that for all .
\theprop Proposition.
-
(i)
If is regular (resp. simplicial), then any Newton polyhedron (w.r.t. ) is pointed at some (resp. ).
-
(ii)
is pointed at if and only if in is defined by a single equation (cf. section 3). In other words, is pointed if and only if is a Cartier divisor in .
-
(iii)
is pointed at if and only if is a -Cartier divisor in .
Proof.
(i) Use the fact that is an integral (resp. rational) basis.
(ii) If is pointed at then by section 3, . Moreover, if for some , then for any , hence and .
Conversely, assume that is an (analytic) Cartier divisor. Let be a smooth subdivision of , and set . This is a smooth variety, and the map is a resolution of . Take a small Stein representative , and set . Then we have the vanishing (see e.g. [13, Corrollary, p. 74] or [10, §8.5]), and also its local analogue (since the local analytic germ is rational too). Thus, from the exponential exact sequence, and . On the other hand, is weighted homogeneous (as any affine toric variety), hence . In particular, . 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 of codimension one, i.e. the group freely generated by Weil divisors, modulo linear equivalence. Note that since is smooth, we have and . If we factor these isomorphic groups by the subgroups generated by the exceptional divisors, we find that the restriction induces an isomorphism .
Denote by the restriction image of under the above isomorphism. Since is local analytic Cartier, and the local divisor of in is , where , we find that the class of is zero in . But then, by the above isomorphisms, the class of is zero in .
Finally note that can be computed as follows [13, 3.4]. Consider the group of invariant divisors and the inclusion sending to . Along with the map , this gives a short exact sequence
Since maps to zero in under the above isomorphism, and , we find that is in the image of . But this means exactly that there exists such that for all .
(iii) Use part (ii) for a certain power of . ∎
\thedefinition Definition.
We say that has Newton nondegenerate principal part with respect to (or simply that or is Newton nondegenerate) if for every with compact, the variety (that is, with its non-reduced structure) is smooth. Note that is a polynomial since is compact.
\thelemma Lemma.
Assume that is Newton nondegenerate and let . If , then the generic transverse type of to is a Newton nondegenerate singularity with Newton polyhedron .
Proof.
The statement follows by restricting to a toric subspace transverse to , see section 2. ∎
\theblock.
The fan and the associated resolution. Assume that is Newton nondegenerate. Let be a regular subdivision of . Then is a smooth variety, and we have a modification . As a result of the nondegeneracy of , the strict transform of in intersects all orbits in smoothly. In particular, is smooth, and is an embedded resolution of .
\thelemma Lemma.
Assume is a Newton nondegenerate Weil divisor. Then, the singular locus of the germ is contained in the union of codimension orbits in .
Proof.
Let be the union of orbits of dimension , that is, codimension , in . Let be as in section 3. The restriction is an isomorphism, and is smooth. Therefore, is smooth. ∎
4. Newton nondegenerate curve singularities
In this section, we will assume that and that is a two dimensional finitely generated strictly convex rational cone. Nondegenerate rank 2 singularities appear naturally in the 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 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 is nonregular. Then there exists a sequence of vectors , called the canonical primitive sequence [27] and integers , called the associated selfintersection numbers, so that:
-
(i)
If , then form an integral basis for .
-
(ii)
If , then .
-
(iii)
The set is a minimal set of generators for the semigroup .
This data is uniquely determined up to reversing the order of and . It can, in fact, be determined as follows. Let be the absolute value of the determinant of the matrix whose columns are the primitive generators of the one dimensional faces of , given in any integral basis. Then, there exists a unique integer so that . The selfintersection numbers are determined as the negative continued fraction expansion
We use the notation for the right hand side above. We have
Along with condition (ii), this determines the canonical primitive sequence recursively and we have .
Alternatively, the vectors are the integral points lying on compact faces of the convex closure of the set . For a detailed discussion of this construction, see [26, 1.6].
If is regular, then we prefer to modify the minimality of the resolution considered above, and set , and and . Accordingly, in item (ii), we will have . In particular, the set is not a minimal set of generators of the semigroup . 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 be a two dimensional rational strictly convex cone with a canonical primitive sequence . The canonical subdivision of is the unique subdivision for which
For each , there is a unique number satisfying . We define and as the numerator and denominator, respectively, of the negative continued fraction
(we require , and , so that these numbers are well defined). The number is referred to as the determinant of .
\therem Remark.
Let be two linearly independent elements. Then we have if and only if form part of an integral basis of . In general, can be computed as the content of the restriction of to the kernel of . In other words, let be the kernel of . Then is divisible by , and is primitive.
\thelemma Lemma.
If is not a regular cone, then has a cyclic quotient singularity at the origin and the map induced by the identity map on is the minimal resolution.
Proof.
See Proposition 1.19 and Proposition 1.24 of [26]. ∎
\theprop Proposition.
Assume that , and that is Newton nondegenerate with respect to defining a germ .
-
(i)
The germ is irreducible if and only if is a single interval with no integral interior points. In fact, in general, the number of components in is precisely the combinatorial length of .
-
(ii)
Assume that is irreducible and let so that . Then is smooth if and only if lies on the boundary of the convex hull of the set . In other words, let be the canonical primitive sequence of . Then either is one of , or there is an such that either
-
(iii)
The curve is smooth if and only if the following condition holds: If and for all , then for all .
\therem Remark.
One can ask why the vectors and do not appear in the list of (ii). The answer is that the corresponding divisors , though they intersect transversaly, they are –invariant, hence they are eliminated by the convention of the definition 3.6.
Proof of section 4.
We start with the following observations. Write . Let be a regular subdivision of which refines both and the canonical subdivision of . The map is then a resolution of with exceptional divisor . We can write , where each is a rational curve. Furthermore, if , then and intersect if and only if . In fact, we can write
where are the two faces of and .
Similarly as in [27], we see that resolves and that the strict transform of in intersects the exceptional divisor transversally in smooth points of . In fact, these intersection points lie in the open orbit . Therefore, we have (see [27, Theorem 5.1])
where is the primitive generator of . Now, the components of are in bijection with the intersection points , which proves (i).
For (ii), let be the canonical subdivision, and the associated modification, which is a resolution of . Let be the strict transform of . The minimal cycle of the resolution is the reduced exceptional divisor and is rational. By [5], the pullback of the maximal ideal of is the reduced exceptional divisor in , and the maximal ideal has no base points in . It follows that the multiplicity of is the intersection number between and . In particular, is smooth if and only if is a normal crossing divisor. If for some , then this is indeed the case. Otherwise, there is an so that . In a neighbourhood of we have coordinates so that , and we have some generic coefficients so that the strict transform of is defined by . Thus, is not smooth if . In the case (the case is similar), is smooth and transverse to if and only if .
The condition in item (iii) is equivalent with the equality
(4.1) |
Choose a basis for , inducing an isomorphism via the dual basis, as well as an inner product on . If we rotate the segment by and translate it, then it can be identified with the vector (segment , ). Consider the parallelogram whose sides are parallel to and , and it has as diagonal. It is divided by into two triangles, each of them can be identified by . Hence, eq. 4.1 holds if and only if .
Clearly, is empty if . The converse can be seen as follows. Let be a family consisting of integral points on , ordered according one of the orientation of this boundary. Two consecutive elements of this family form a basis of , and
It follows that the set of irreducible elements in the semigroup are presicely the elements on the boundary . In particular, if , then for some . It follows that . ∎
\thecor Corollary.
Consider the notation from the proof of section 4(ii), that is, irreducible and with . Then the multiplicity of is
\therem Remark.
Let be any two linearly independent integral vectors in any free module, and let be the free module generated by them. Then the definitions from 4 and 4 can be repeated in . 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 as rows, see [27]. Note that . Moreover, , 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 in this section. This is a (non-direct) generalization of a result of Kouchnirenko valid in the classical case [15].
5.1 Theorem.
Let be a Newton nondegenerate singularity and assume . The following are equivalent
-
(i)
has an isolated singularity.
-
(ii)
If satisfies for all , then for all .
-
(iii)
Let and and assume that with . If so that and , then .
-
(iv)
Let and . Then there is at most one with and . If such a exists, then is one of the following
(5.1) and, furthermore, there exists an so that
(5.2) (see section 4 for ) and
(5.3)
Proof.
By section 3, the singular locus of the punctured germ is a union of orbits for some . For such a , we have if and only if the projection of in is nontrivial, by section 3. By the same lemma, if , then the generic transverse type to in is a Newton nondegenerate curve with Newton polyhedron the projection of to . Therefore item (i)item (iii) follows from section 4. The equivalence of item (ii) and item (iii) is an exercise.
The generic transverse type to in is smooth if and only if its diagram has a single face corresponding to a as in eq. 5.1, and this face has length one. item (i)item (iv) follows, once we show that given such a , an 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 of the Newton diagram corresponding to .
Take a smooth subdivision of containing as a ray, and let be the ray adjacent to between and . Then there exists a so that
(5.4) |
Furthermore, for , we may assume that
where is the continued fraction associated with and . As a result, eq. 5.4 can be rewritten as (5.2) with . Let be the endpoints of so that and . Since is a primitive function on the affine hull of the face of , the length of . We find
6. Resolution of Newton nondegenerate surface singularities
In this section, we retain the notation introduced in section 3, with the assumption that . We describe Oka’s algorithm which describes explicitly the graph of a resolution of a Newton nondegenerate Weil divisor of dimension . This algorithm was originally described by Oka [27] for Newton nondegenerate hypersurface singularities in . 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 is a subdivision satisfying the following.
-
(i)
is a regular subdivision of .
-
(ii)
If , then is the canonical subdivision of given in section 4.
\theblock.
The existence of a canonical subdivision is proved in [27, §3]. We fix such a subdivision . We will denote by the toric variety associated with . The map is denoted by , and the strict transform of under this map is denoted by . We denote by the restriction . By section 2, the map is proper, hence is proper as well.
\thedefinition Definition.
For , define
\thedefinition Definition.
We start by defining a graph as follows. Index the set by a set , i.e. write in such a way that the map , is bijective. Similarly, index the set by . Hence . The elements of are referred to as extended nodes, while as nodes.
Denote by the face of corresponding to and by the primitive integral generator of . Note that if and only if is bounded. For , let be the length of the segment if this is a bounded segment of dimension 1. If is unbounded, or has dimension 0, then we set . Now, for every pair and , we join by bamboos of type , as in fig. 3. This finishes the construction of the graph . Denote its set of vertices .
Define the graph as the induced full subgraph of on the set of vertices .
In order to have a plumbing graph structure on , 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 associated with is defined as the number of integral interior points in the polygon .
To every extended node we have associated the cone and its primitive integral generator . If are the vertices appearing on a bamboo, in this order, from to , let be the canonical primitive sequence associated with . We then set for , , and . This induces a map with the property that for , and generate an element of if are adjacent in .
For any , let and be the set of neighbours of in and , respectively. Then there exists a unique satisfying
The number is the Euler number associated with . We note that if lies on a bamboo, with the notation of the previous paragraph, , then and unless .
\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 lie on its boundary .
Indeed, if is a vertex which is not on the boundary, then the nodes corresponding to faces of containing lie on an embedded cycle. Similarly, if is a face of dimension which is not a subset of the boundary, and contains integral interior points, then the nodes corresponding to the two faces containing are joined by more than one bamboo, inducing an embedded cycle in . Finally, if 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 is discussed in details in [7].
\theexample Example.
Let , and consider standard coordinates on , and the function
The Newton diagram 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.
\theprop Proposition.
Let be a Newton nondegenerate surface singularity. Then the map is a resolution of whose resolution graph is .
More precisely, is smooth and the exceptional set is a normal crossing divisor. For each , we can enumerate the irreducible components of by so that , where is a smooth curve.
If , then is compact, has genus , and its normal bundle in has Euler number . If , then is a smooth germ, transverse to a smooth point of the exceptional divisor.
Furthermore, if , then the number of intersection points equals the number of edges between and in .
Proof.
The proof goes exactly as in [27] ∎
\thelemma Lemma.
For , we have
Proof.
\therem Remark.
-
(i)
The exceptional divisor is the union of for which is a cone which is not contained in , or, equivalently, is compact.
-
(ii)
If , then is a compact smooth irreducible curve. If , then is the union of disjoint smooth compact rational curves. For , the intersection is the disjoint union of smooth curve germs, where is the length of the segment . If , then (the global divisor does not intersect ).
\thedefinition Definition.
We denote by the lattice of integral cycles in supported on the exceptional divisor .
\thedefinition Definition.
Let and denote its restriction by . For any , we define
For , we also write instead of , as this is independent of .
Similarly, for any , let be the valuation on associated with the divisor , that is, for , denote by the order of vanishing of the function along . Set also
\therem Remark.
-
(i)
If and , then is not independent of the choice of .
-
(ii)
For , the function is the valuation on associated with the irreducible divisor , cf. eq. 3.2.
-
(iii)
In general, the functions and do not coincide on . However, for any and . Furthermore, if and , then . In particular, this defines a group homomorphism , .
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 of is under no restriction. Recall that we say that (or ) is pointed at , if for any we have , 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 in an analytic variety , denote by the normalization of its local ring . The delta invariant associated with is defined as
Let be a resolution of the singularity and assume that has dimension . Assume, furthermore, that , and that the higher direct image sheaves , , are concentrated at . The geometric genus is defined as
We say that is rational if and for .
7.1 Theorem.
Let be a Newton nondegenerate Weil divisor of dimension .
-
(i)
We have the following canonical identifications
In particular, if these vector spaces have finite dimension, then
where denotes the reduced Euler characteristic, that is, the alternating sum of ranks of reduced singular cohomology groups.
- (ii)
-
(iii)
Assume that is -pointed, that , and that has only rational singularities outside the origin. Then is normal and for .
\thecor Corollary.
Assume that and is normal. Then
This generalizes a result of Merle and Teissier [19] valid for the classical case .
\thecor Corollary.
Proof.
\therem Remark.
Assume that , and that is rational outside . Then, for , we have
Here, the first isomorphism comes from the long exact sequence for cohomology with support in , and the vanishing , for [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 , and the fact that for , if we choose a Stein representative of the germ . This last long exact sequence furthermore gives
Therefore, in this case, the groups described in theorem 7.1 are closely related with the depth of . In particular, the conclusion of theorem 7.1item (iii) is that is a Cohen–Macaulay ring.
Proof of theorem 7.1.
To prove item (i), we use results and notation from [10, §7], see also [13, 3.5]. Define
Then is the divisor of the pullback of to , and we have a short exact sequence
By [10, Corollary 7.4], we have for all . Furthermore, , and the image of in is . Therefore,
Denote by the order function defined in [10, §6] (using the natural trivialization of on the open torus)
and define the sets
We note that is a convex cone and that for all . By [10, Theorem 7.2], we have isomorphisms
Since is a convex set, the long exact sequence associated with cohomology with supports provides, for any
To finish the proof of (i), we will show that for any , the spaces and are in fact homotopically equivalent. We start by noting that the the condition (including the case when ) is equivalent to . If this happens then we can choose a small so that as well, and so is star-shaped with center . In particular, in this case,
where denotes the homotopy equivalence. Thus, in what follows, we assume that contains an interior point in , equivalently, .
Choose and satisfying and define the hyperplanes
Then, seeing and as linear spaces by choosing origins , the pairing is nondegenerate and the polyhedrons and are each others polar sets as in [13, 1.5].
Since , we have
By the assumptions made above, there is an . Both and are compact convex polyhedrons in . Projection away from onto then induces a homotopy equivalence
By projection, we mean that any element in a ray maps to the unique element in . By section 7item (i), this has the subset
as a strong deformation retract. All this yields
(7.1) |
Using a projection, this time onto in , having as center any element in , we get a homotopy equivalence
By section 7item (ii), we have a homotopy equivalence
Since, by assumption made above, , and so the right hand side above has a free action by which has a section given by intersection with . Furthermore, one checks that if , then
Here, the condition on the left is equivalent to , so
(7.2) |
Now, consider the CW structure given by the cells in and given by cells in . Using barycentric subdivision, one obtains a homeomorphism , sending the center of a cell to the center of the dual cell , thus identifying with the dual of . By this identification, the left hand side of eq. 7.2 is a regular neighbourhood around the image under 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 or . Assuming that this is not the case, the complex, say, , on the right hand side of eq. 7.1 is a closed subset of . Then , unless , in which case . But this is equivalent to for all , that is, .
For (iii), we will show that has trivial homology in degrees for all . By assumption, there is a so that for we have . We can again assume that . We must show that for , where is the right hand side of eq. 7.1. We note that by definition, consists of cells for satisfying . Define similarly
Define
This space can be either , an dimensional ball, or empty. In each case, for . We will show that , 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 the -skeleton of the complex , and define similarly
We will prove by induction on that and that this is a homotopy equivalence. The case follows from the pointed condition: assuming is a ray, there is a so that . By assumption, we have , so that if and only if . Since consists only of such zero-cells, we get .
Next, assume that for some we have an inclusion which is a homotopy equivalence. Let provide an -cell in . In this case, we want to show that . In fact, we have , hence , by induction. But by the rationality assumption on the transverse type, it follows from (ii) and section 3 that we must have , thus .
To show that this inclusion is a homotopy equivalence, let provide an -cell which is not in . By definition, we see that as well. In fact, similarly as in the proof of (i), the inclusion is a strong deformation retract. Since these cells, along with provide a finite closed covering, these glue together to form a strong deformation retract . ∎
\thelemma Lemma.
Let . Assume that is given as a finite disjoint union of relatively open convex polyhedrons , i.e. each is given by a finite number of affine equations and strict inequalities. Furthermore, assume the following two conditions:
-
❀
If is the face of for some , then for some .
-
❀
For any , the intersection is a face of both and .
Note that the polyhedrons may be unbounded. In this case
-
(i)
Assume that is compact and is convex. Then the inclusion
(7.3) is a strong deformation retract.
-
(ii)
Assume that is convex. Then the inclusion
is a strong deformation retract.
Proof.
We prove item (i), similar arguments work for (ii). We use induction on the number of with . Indeed, if this number is zero, then the inclusion in eq. 7.3 is an equality.
Otherwise, there is an with . Define
Then the left hand side of eq. 7.3 does not change if we replace by . Therefore, using the induction hypothesis, it is enough to show that the inclusion is a homotopy equivalence. We do this by constructing a deformation retract . For this, we use the finite closed covering , of . It is then enough to define the restriction of to for in such a way that these definitions coincide on intersections.
For any , we define . Let . If , then , and we define by projecting away from , that is, for any there is a unique in the intersection of and they ray starting at passing through . We define . 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 and . Furthermore, in the case , we give a formula for the canonical cycle.
\thedefinition Definition.
Let be a resolution of singularities of an -dimensional singularity. A canonical divisor on is any divisor satisfying .
If then let be the exceptional divisor of a resolution , where are the irreducible components of . Recall that we denoted by the lattice of integral cycles in supported on the exceptional divisor : that is, . We also set and
where denotes the intersection form, extended linearly to . Moreover, set for the unique rational cycle satisfying and for .
In this surface singularity case the canonical cycle is the unique rational cycle on supported on the exceptional divisor, satisfying the adjunction formula
for any irreducible component of the exceptional divisor, where is the Euler number of the normal bundle of , and is the genus of (we assume here that the components of the exceptional divisor are smooth).
\therem Remark.
The cycles and are well defined, since the intersection matrix, with entries , associated with any resolution is negative definite. Notice also that any two canonical divisors are linearly equivalent, and that any canonical divisor is numerically equivalent to . However, it can happen that has infinite order in the Picard group.
\theprop Proposition.
Fix any . Let be a Newton nondegenerate Weil divisor, and a subdivision of the normal fan so that is an embedded resolution. Then the divisors
(8.1) |
are possible canonical divisors for and , respectively.
Furthermore, in the surface case (), the canonical cycle on is given by the formula
(8.2) |
where the sum to the right runs through edges in the graph so that and (and the identity is in ).
Proof.
For , see e.g. 4.3 of [13]. Since the divisor is principal in (and ), the adjunction formula gives
which proves eq. 8.1. To prove eq. 8.2, it is enough to show that in for all ,
(8.3) |
where the sum is as in eq. 8.2. Recall that . We note that the adjunction formula gives for all , where is the valency of the vertex in , and is the genus of . Furthermore, it follows from section 6 that if , then
-
❀
if and only is on the end of a bamboo joining a node and an extended node . In this case, has exactly one neighbour in in the graph .
-
❀
if and only is on a bamboo joining two extended nodes, and is not of the form described in the previous item.
-
❀
if and only if is a node.
Consider first the case , and let be the unique neighbour of in . It follows from section 6 that , since is a segment, and so has area zero. As a result, the right hand side of eq. 8.3 is .
Assume finally that . Then, has no neighbours in . Furthermore, coincides with the number of integral points on the boundary of , since each edge adjacent to can be seen to correspond to a primitive segment of the boundary. By using Pick’s theorem and section 6, we therefore get
which finishes the proof. ∎
\therem Remark.
As depends on the choice of up to a multiplication, the right hand side of the second formula from eq. 8.1 depends on this choice too. In fact, the monomial rational function realizes the linear equivalence between the two divisors 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 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 and be as above. We say that , or , is (-)Gorenstein-pointed if there exists a () such that for all .
\theexample Example.
Recall that is Gorenstein if and only if there is a satisfying for all , see e.g. [8], Theorem 6.32. Therefore, if is Cartier, and , then is Gorenstein pointed (since for ). Furthermore, is Gorenstein since is Gorenstein and forms a regular sequence.
Similarly, is -Gorenstein if there is a satisfying for all , see e.g. [1]. Therefore, if is Cartier, and , then is -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 is Cartier, see section 3. However, being ‘Gorenstein pointed’ codifies an abstract property of the germ , namely its Gorenstein property, see theorem 9.1 below.
\theblock.
Recall also that is Gorenstein if it admits a Gorenstein form. A Gorenstein form is a nowhere vanishing section in . A Gorenstein pluri-form is a nowhere vanishing section in for some .
In this section and are canonical divisors with a choice as in eq. 8.1.
\thedefinition Definition.
Let be some meromorphic 2-form on whose divisor is .
9.1 Theorem.
Assume that is a normal Newton nondegenerate surface singularity (i.e. ). The following conditions are equivalent:
-
(i)
is Gorenstein-pointed at some .
-
(ii)
There exists a so that for all we have .
-
(iii)
There exists a so that for all we have .
-
(iv)
There exists a so that is a Gorenstein form.
-
(v)
is Gorenstein.
When these conditions hold, (i), (ii), (iii) and (iv) uniquely identify the same point .
\theprop Proposition.
Under the assumption of theorem 9.1, the following conditions are equivalent, and imply that is -Gorenstein:
-
(i)
is -Gorenstein-pointed at some .
-
(ii)
There exists a so that for all we have .
-
(iii)
There exists a so that for all we have .
-
(iv)
There exists a so that is a Gorenstein pluri-form for some .
Furthermore, all these these conditions identify the very same uniquely.
Proof.
(ii)(iii) For any consider the cycles
where the sum runs over edges in so that and (as in eq. 8.2), and (where all these ’s are the noncompact curves in ).
If is an edge as above, then . Moreover, . Therefore, by assumption (ii), for any . On the other hand, by section 6, for any as well. Hence . But by eq. 8.2 .
\therem Remark.
Similarly as in theorem 9.1, one may ask whether the equivalent cases in 9 are equivalent with the property that is -Gorenstein. If is -Gorenstein-pointed at , then (iv) implies that is -Gorenstein. The converse does not hold, as seen by the following example.
Let and
Write , for the rays generated by the vector specified above and denote by the corresponding multiplicities. We find and . As a result, since the linear equation
has no solution, hence is not -Gorenstein pointed.

On the other hand, one verifies that the Weil divisor defined by 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 . Such a singularity is a cyclic quotient singularity. In particular, it is -Gorenstein.
\theblock.
\thelemma Lemma.
Let , that is, is a meromorphic function on the complement of the exceptional divisor in satisfying
(9.1) |
Then, there exists a Laurent series satisfying and
(9.2) |
Proof.
Let and let be the set of meromorphic functions obtained as a restriction of Laurent series satisfying eq. 9.2. We want to show that .
We immediately see . In fact, this inclusion fits into an exact sequence as follows. Recall the notation from the proof of theorem 7.1, and . Also, define as the union of compact divisors in , that is, for . Since , we have a short exact sequence of sheaves
yielding a long exact sequence of cohomology groups. We have
since . Furthermore, since is normal, is the set of Laurent series satisfying eq. 9.2. Thus, its image in is . Therefore, the quotient injects into . On the other hand,
(9.3) |
where, following Fulton [13], is the unique function restricting to linear function on all , and satisfying for , and for we set
Firstly, since is contractible, we find
Secondly, define as the union of those cones satisfying (i.e. for all ), and let be the set of satisfying . By section 7, the inclusions
are strong deformation retracts. But the right hand side above is either a contractible set, or it has the homotopy of . In particular, it is connected, by our assumption , and so eq. 9.3 vanishes. ∎
\thelemma Lemma.
Assume that is a Gorenstein normal surface singularity, i.e. , and that we have a Gorenstein form on . Thus, is linearly trivial, and there exists
Then there is a satisfying
(9.4) |
Proof.
By the previous section 9, we can find a satisfying and eq. 9.2. Let and . We want to prove that . Both and are supported in the exceptional divisor and the noncompact curves for , and by our assumptions, they have the same multiplicity along this noncompact part. Thus, is supported on the exceptional divisor. Furthermore, we have for , thus .
For the reverse inequality, note first that for all since is principal. For any , let be an element of the support of the principal part of with respect to , i.e. and . By definition, we also have for all . Therefore,
As a result, is in the Lipman cone, and so, , proving eq. 9.4. ∎
Proof of theorem 9.1.
Assuming that is Gorenstein, let be a Gorenstein form. Then there is meromorphic so that on . By section 9, is the restriction of a Laurent series satisfying eq. 9.4.
For any , denote by the principal part of with respect to the weight . We make the
Claims:
-
(a)
For any , is a monomial, that is, there is a so that for some .
-
(b)
If is a vertex on a bamboo connecting and some other node in , then .
By (b), the exponent does not depend on , finishing the proof since hence is a Gorenstein form.
(a) is proved as follows. Set arbitrarily. We then have , and also , for any other , since . In particular,
The right hand side is sero since for . On the other hand, by the section 9, we have for all , thus, the left hand side above equals . Furthermore, since , does not have any zeroes or poles outside the exceptional divisor, in a neighbourhood around , hence . Therefore, the inequality above is an equality, and we have for .
This fact is true for any choice of , therefore, for any and for any other choice . But the vectors form a generator set, hence necessarily .
For (b), assume that and are joined by a bamboo, consisting of vertices , with and , and neighbours for . For convenience, we set and . We start by showing that using induction (we replace the subscript by just for legibility). Indeed, for this is clear, and we showed in the proof of (a) that this holds for . For the induction step we use the recursive formulas
The first one holds by section 6, and the second one follows from similarly as above, although for the case , we may have to use a component of the noncompact curve .
We now see that for any , the support of consists of points for which and . But these equations are equivalent to and . Therefore, for these . ∎
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, -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, denotes a resolution of a normal surface singularity with exceptional divisor , whose irreducible decomposition is .
We assume that has a rational homology sphere link; thus for all .
We use the notations , and as in section 8. For with we write . denotes the canonical cycle. Note that if and only if is an germ. Otherwise, it is known that in the minimal resolution, or, even in the minimal good resolution, all the coefficients of are strictly positive. However, usually this is not the case in non-minimal resolutions, i.e. in our it is not automatically guaranteed.
\thelemma Lemma.
In any resolution of a normal surface singularity with we have
(10.1) |
Proof.
By the generalized version of Grauert–Riemenschneider vanishing we have the two vanishings
(10.2) |
Hence, if then too. Otherwise, from the long exact sequence of cohomology groups associated with
we obtain that the right hand side of eq. 10.1 equals . By Serre duality, this equals . Now, the short exact sequence
with the above vanishing eq. 10.2 gives . ∎
\thedefinition Definition.
A computation sequence is a sequence of cycles from ,
such that
-
(i)
for all there is a so that , and
-
(ii)
and is the union of some reduced and non-intersecting rational -curves
Given such a sequence , we define
Then is a line bundle on . Denote by its degree. Since is numerically equivalent to zero, we have . In particular, since , we get and
\theblock.
Given a computation sequence , the inclusion induces an isomorphism
Indeed, let be such that . Then we have a short exact sequence
which induces an exact sequence
and the right hand side vanishes, since .
\thecor Corollary.
Let be a computation sequence. Then
(10.3) |
with equality if and only if the map is surjective for all . ∎
\therem Remark.
-
(i)
We note in particular that if there exists a computation sequence so that for all , then , that is, is rational. In general, if for some , then the inequality between the terms in the sums eq. 10.3 is an equality.
-
(ii)
Let be the sum from the right hand side of eq. 10.3 associated with . Then we have
(10.4) where the minimum is taken over all computation sequences. Note that 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 . 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 , 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 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 . We claim that for any cycle , there is a smallest cycle satisfying
(10.5) |
The existence and uniqueness of such an element is explained in [21] in the case when and in general in [16, 25, 31]. The name comes from a construction of Laufer in [17, Proposition 4.1]. Note that only depends on the multiplicities of for and the class .
The following properties hold for the operator , assuming :
Monotonicity: If then .
Idempotency: We have for any .
Lower bound by intersection numbers: If and so that for and for all , then .
Generalized Laufer sequence: Assume that . First note that if for some , then we have as well, similarly as in the proof of Proposition 4.1 [17]. We claim that there exists a generalized Laufer sequence which connects with . It is determined recursively as follows. Start by setting . Assume that we have constructed . By induction, we then have . If for all then by the minimality of we get ; hence the construction is finished and we stop. Otherwise, there is a so that . We then define (for some choice of such ).
\therem Remark.
The computation sequence (as in corollary 10), what we will construct, will have several intermediate parts formed by generalized Laufer sequences as above. Note that if and are two consecutive elements in a Laufer sequence, then , hence , 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 of Newton nondegenerate Weil divisor as in section 6. Let denote a canonical divisor as in section 8. In this section we will assume that in the dual resolution graph we have for any node . This assumption will be justified in section 11.
From the assumption , valid for any node , an immediate application of the construction of from section 6 gives that . Thus .
\thelemma Lemma.
-
(i)
.
-
(ii)
Let be the set of -vertices appearing on bamboos joining with in section 6. Then . In particular, the sequence constructed in section 10 satisfies ii in section 10.
Proof.
i Since for any , it is enough to show that . We can analyse each component of independently, let be such a bamboo formed from , with dual vectors in denoted by . If and are the multiplicities of along the neighboring nodes of in (with convention that if there is only one such node), we search for a cycle with for all . Thus, is in the Lipman cone of , hence .
ii Using the lower bound by intersection numbers, we find that . Since , there exists a Laufer sequence from to . Now, one verifies that the construction/algorithm of this sequence chooses each vertex once, and each vertex in twice. ∎
\thedefinition Definition.
A (coarse) diagonal computation sequence with respect to is defined as follows. Start with , and define . Assuming () has been defined, and that , choose a minimizing the ratio
(10.6) |
Then set . If , then we record . If , then we stop, and set .
We refine the above choice as follows. Choose some node and define a partial order on the set : for , define if lies on the geodesic joining and (here we make use of the assumption that the link is a rational homology sphere, in particular, is a tree). When choosing , if given a choice of several nodes minimizing , and , then, we choose minimal of those with respect to . If , let be the set of nodes for which . If has one element we have to chose that one. Otherwise, let be the minimal connected subgraph of containing , and we choose as a leaf of .
Note that by section 10i, , hence there exists a Laufer sequence connecting with . Furthermore, using idempotency and monotonicity of the Laufer operator section 10, we find
As a result, we can join and by a Laufer sequence. This way, we obtain a computation sequence , connecting with . Finally, by section 10ii, 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 to , we have , we find, using sections 10 and 10
(10.7) |
10.1 Theorem.
Let be a normal Newton nondegenerate Weil divisor given by a function , with a rational homology sphere link, and assume that the polyhedron is -Gorenstein pointed at . Then, a diagonal computation sequence 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 , by the results of the next section this assumption can be removed.
\thedefinition Definition.
Let , corresponding to the face . Denote by the convex hull of the union of and . Set also
where we use the partial ordering on defined in section 10. For , let be the hyperplane in defined as the set of points satisfying . For , we set
\therem Remark.
The affine plane contains an affine lattice , that is there is an affine isomorphism , inducing a bijection . The polyhedron is then the image of a lattice polyhedron with no integral integer points under a homothety with ratio in if . These properties allow us to apply section 10 in the proof of theorem 10.1. Furthermore, the polygon is always nonempty, even if may be empty.
\theblock.
The sets form a partitioning of the union of segments starting at and ending in points on , that is, . This follows from the construction as follows. The partially ordered set is an lower semilattice, i.e. any subset has a largest lower bound. If , and is the set of nodes for which , then , and for , where is the largerst lower bound of .
The integral points in the union of the sets are presicely the integral points satisfying for all and for some . Indeed, by the rational homology sphere assumption, any integral point on the Newton diagram must lie on the boundary , see section 6. These are the points “under the Newton diagram”; by theorem 7.1, the number of these points is . It follows from construction that the family forms a partition of these points. We conclude:
(10.8) |
\thedefinition Definition.
For , denote by the smallest real number larger or equal to and congruent to modulo . That is,
\therem Remark.
The number depends on only up to an integer. For all , we have . In particular, given an , we have .
\thelemma Lemma.
Proof.
We prove eq. 10.9, eq. 10.10 follows similarly. Let be the vertices on the bamboo with Euler numbers as in fig. 3. Set and . There exists a unique set of numbers so that the equations
(10.11) |
are satisfied. This follows from the fact that the intersection matrix of the bamboo is invertible over . In fact, it follows from [12, Lemma 20.2] that in fact,
This, and the lower bound by intersection numbers from section 10, implies that , and therefore , since .
For the inverse inequality, we must show that there exist numbers satisfying
(10.12) |
for , and so that is the right hand side of eq. 10.9. Let be the canoncial primitive sequence as in section 6, and note that . Set recursively
Note that, by definition, . The assumption therefore implies that the left hand side of eq. 10.12 is an integer. It is then enough to prove eq. 10.12 for . This equation is clear if , so we assume that . Setting , we find
where . 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
since . Since the left hand side is an integer, eq. 10.12 follows. ∎
\thelemma Lemma.
If , then . As a result, the corresponding terms in eq. 10.7 vanish.
Proof.
Let be a neighbour of . Assume first that lies on a bamboo connecting and . We then have . Furthermore, , where equals or . By the previous lemma, we find
with as in the lemma.
Next, assume that lies on a bamboo connecting and . Name the vertices on the bamboo as in the proof of the previous lemma. We then have for , and . By the lower bound on intersection numbers, we find . A Laufer sequence which computes from may start with . This shows that in this case.
As a result, for every , we have , with an equality for at most one neighbour. As a result, since we find
The final statement of the lemma is now clear. ∎
\thelemma Lemma.
Let be an integral polygon with no internal integral points. Let be the faces of and let be the integral lenght of . Let , . Then let be the unique integral affine function whose minimal set on is and this minimal value is if and if . Set . Then there exists an satifying
Proof.
This is [31, Theroem 4.2.2]. ∎
Proof of theorem 10.1.
Recall the order on the set defined in section 10. We extend this order in the obvious way to all of . Also, by assumption, is -pointed at the point . Fix an and set . For , define
(recall that is nonempty, see section 10) and
Define the affine functions , . By construction, these are primitive integral functions on with respect to the affine lattice . It now follows from section 10 that there is an so that and .
On the other hand, we claim that for . Using section 6, and the definition of , i.e. for , it follows that
where is any element of . As a result, using eq. 10.7 and section 10, as well as eq. 10.8, we have
and so these inequalities are in fact equalities.
We are left with proving the claim for . Fix , and let so that lies on a bamboo connecting and . Let . Then is the minimal set of on , i.e., . Let be the affine hull of . Since the two affine functions
both take value on and on , by theorem 9.1item (iii), they conincide on . Let
Using the minimality of eq. 10.6, we get for any
(10.13) |
and so . In the case when , or equivalently, , this inequality is strict. It follows, using section 10, that
(10.14) |
Therefore, since , we find
This proves the claim, unless and . In that case, the numbers and are both integers. Since form a part of an integral basis of , we can assume that , hence,
As a result, since we have a strict inequality we get a strict inequality in eq. 10.14 as well. Therefore, we have
and so , which finishes the proof of the claim. ∎
11. Removing -facets
In this section we consider only surface singularities, i.e. we assume that . We consider removable -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 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 with rational homology sphere link.
The concept of a -facet appears in [11] in the case of hypersurfaces in , where is a -adic field, and is further studied in [18, 6].
\thedefinition Definition.
Let be a compact facet, i.e. of dimension . Then is a -facet if has exactly vertices so that there is a so that . A -facet is removable if furthermore, the segment is contained in the boundary of .

\thedefinition Definition.
Let be closure in of the union of cones in which correspond to compact facets of which have dimension . This is the tropicalization of . We say that is generated by the tropicalization of , if is generated as a cone by the set .
Let be the cone generated by . This is a finitely generated rational strictly convex cone, and if is not rational, then has dimension . This cone induces an affine toric variety , and the function defines a Weil divisor . Furthermore, the inclusion induces a morphism , which restricts to a morphism .
\therem Remark.
The closure of in a certain partial compactification of is called the local tropicalization of [30].

\thelemma Lemma.
Let . Then the orbit intersects if and only if .
Proof.
The orbit is an affine variety (recall , and if is an element of the affine hull of , then and
Therefore, is empty if and only if is a unit in , which is equivalent to being a monomial, i.e. . ∎
\thelemma Lemma.
Let and be as in section 11. If is normal, then the morphism is an isomorphism.
Proof.
We can assume that the smooth subdivision subdivides the cone , so that we get a subdivision of the cone . Let be the corresponding toric variety. Let be the fan consisting of cones which are contained in . We then get open inclusions
where is the toric variety associated with the fan .
It follows from section 11 that the strict transforms and of and , respectively, are contained in , and so . As a result, . Since is normal, the morphism is an isomorphism. ∎
\theblock.
Assume that is a removable -face, and let and be as in section 11. If is the only facet of , then we leave as an exercise to show that the graph is equivalent to a string of rational curves, and so is rational. We will always assume that is not the only facet of . There exists an element of which is constant on the segment (e.g. the normal vector to ). As a result, the boundary intersects the hyperplane of elements which are constant on in two rays, and , where satisfies , and satisfies .
Let be a primitive generator of , set and define
where are the coefficients of as in eq. 3.1. Let be the Weil divisor defined by . We get a Newton polyhedron , from which we calculate invariants of as described in previous sections. It follows from this construction that , and that is Newton nondegenerate.
Now, assume that is generated by the tropicalization of . Let and be the rays corresponding to the noncompact faces of containing the segments and , respectively. Let be primitive generators of . By construction, and the above assumption that is generated by , we have , and so .
In fact, we have where . Indeed, is the unique positive linear combination of and which vanishes on , and is primitive. Since , by definition of , and since , we have
Furthermore, we have and , and so by section 4, form a part of an integral basis, which implies that is primitive.
Now, define as the combinatorial length of the segment . We have and via Oka’s algorithm (section 6), this segment corresponds to bamboos in , each consisting of a single -curve, whereas corresponds to one bamboo with determinant .
\theprop Proposition.
Let , and be as above, and assume that is Newton nondegenerate. Assume also that is generated by the tropicalization as described in section 11. Then
-
(i)
is Newton nondegenerate.
-
(ii)
.
-
(iii)
The singularities and have diffeomorphic links.
-
(iv)
The singularities and have equal geometric genera and -invariants.
-
(v)
If is normal, then is normal.
-
(vi)
If is -Gorenstein-pointed at , then so is . In particular, if is Gorenstein, then is also Gorenstein.
Proof.
We now prove item (iii). We have , the output of Oka’s algorithm for the Newton polyhedron , and , the output of Oka’s algorithm for . Let be the ray dual to and let be the unique face of adjacent to , i.e. . Then , and we can subdivide the canonical subdivision of so that we can assume that . We can therefore identify vertices of and corresponding to the same ray and . It is then clear from construction that the components of and in the direction of are isomorphic. After blowing down the -curves corresponding to the segment , we must show
-
❀
The two bamboos joining with on one hand, and with on the other, are isomorphic.
-
❀
The vertex has the same Euler number in and in .

For the first of these, we prove that
We calculate as the greatest common divisor of maximal minors of the matrix having coordinate vectors for and as rows. But , and so adding a multiple of to does not modify the greatest common divisor of these determinants, hence .
The invariant can be calculated as the unique number so that is a multiple of . On the other hand, we find, setting and ,
Finally, we show that has the same Euler number in the graphs and . Denote these by and . After blowing down the curves associated with the segment , the vertex has two neighbors in either graph or . Denote by and the neighbor of contained in the same component of and as . It is then clear that .
Denote by the neighbours of in the direction of , respectively, and the other neighbours, as in fig. 8. Then we have and
where are as above. The two numbers and are identified by section 6
which leads to their equality.
Next, we prove item (iv) and item (v). By theorem 7.1, it suffices to show that
have the same cohomology for all . By shifting , we simplify the following proof by assuming . The inclusion
is a homotopy equivalence. Indeed, one can construct a suitable vectorfied on pointing in the direction of , whose trajectories end up in , thus giving a homotopy inverse to the above inclusion.
Now, let be the union of faces of which do not intersect . By section 7, the inclusion is a homotopy equivalence. Define similarly, using . Thus it suffices to prove that vanish for all . By excision, this is equivalent to showing
(11.1) |
If does not intersect the face , then . Also, if , then . In either case, eq. 11.1 holds. We can therefore assume that and . With these assumptions at hand, it is then enough to prove that excactly one of the segements and is contained in , i.e. it cannot happen that either both or neither is contained in .
Let be the affine hull of , i.e. the hyperplane in defined by , and let . Define a point by
This is well defined, since the functions are linearly independent. Then is a convex polygon in , and is a vertex of . Furthermore, is the unique point in where both functions and take their minimal values.
If neither of the segments , are contained in , i.e. both intersect , then we can choose and . Furthermore, we have , and . Therefore, is in the convex hull of , and so , contrary to the assumption .

Next, assume that both segments , are contained in . We start by showing that in this case, we have . By assumption, we can choose .
We have . One verifies (see fig. 9) that if , then we would have , but is a minimum for . Therefore, we can assume that , similarly, . It follows, since , that , so we can assume that . But, since , we find
and so . But this is a contradiction, since .
Next we prove item (vi). Assume that is -Gorenstein pointed at . It suffices to show that , where is the minimal value of on . We immediately find
\theexample Example.
Consider the cone and the function
which defines a nonrational singularity . In this case, has a -facet
corresponding to a node . The normal vector to is and eq. 8.2 gives . By the above computations, removing the monomial from gives another singularity with the same link and geometric genus, but is nonnegative on the other node. After removing we find
Note that is generated by the tropicalization of , but the tropicalization of generates the cone .

\theblock.
In what follows, we connect the above construction with the coefficients of . We introduce a simplified graph, whose vertices are the nodes of . whose vertices are the nodes of , and a bamboo of connecting two nodes of is replaced in by an edge. Then is a tree, with an edge connecting if and only if and intersect in a segment (of length ). Recall that a leaf of a tree is a vertex with exactly one neighbour. If we assume that , then we see that the following are equivalent, since is a tree:
-
❀
is a leaf in ,
-
❀
is connected,
-
❀
all edges of , except for one, lie on the boundary of the Newton diagram.
If , then there is a unique , and , in particular, . Finally, if , and if we assume that is normal, then is rational.

The following lemma is elementary:
\thelemma Lemma.
Let be an integral polyhedron in , having no integral interior points. Then, up to an integral affine automorphism of , is one one the following:
-
❀
Big triangle The convex hull of , , .
-
❀
Small triangle of type The convex hull of , , .
-
❀
Trapezoid of type The convex hull of , , , , where , and . ∎
\thelemma Lemma.
Assume that is normal, Gorenstein-pointed at , and not rational. If is a leaf in and , then is a removable -facet of (See section 11 for the definition of ).
Proof.
By assumption, has two adjacent edges contained in , say and . Let be the noncompact faces of containing the segments and , respectively, and let be the primitive functions having as minimal sets on , denote these minimal values by .
Let and and and . Then, the bamboos corresponding to the segements and have determininats , see section 4.
Assume first that is a small triangle of type , that the segment has length , and that . This implies that is a removable -facet.
Otherwise, let be the affine hull of . If is a big triangle, a trapezoid, or a small triangle as above, but with , then the square
is contained in . In particular, its vertex , the unique point in satisfying for , is contained in . The set
is a one dimensional face of (here we use the condition that is generated by the tropicalization of ). By our assumption we have , contradicting the assumption that is not rational. ∎
\theprop Proposition.
Assume that is normal, Gorenstein-pointed at , and not rational. If there is an so that , then has a removable -facet.
Proof.
If is a leaf in (see section 11), then is removable by section 11. So let us assume that is not a leaf in , i.e. that is disconnected. The inclusion
is a strong homotopy retract (here we set ). In particular, the right hand side is disconnected as well. But it follows from our assumptions that the point is in the right hand side above. Let be a component of contained in a component of the right hand side which does not contain . Then, for any so that we have , i.e. . Let be the induced subgraph of having vertices for . 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 of is a leaf of . In the second case, has at least two leaves, so we can choose a leaf of which is not adjacent to in . In either case, is a removable -facet by section 11. ∎
\theprop Proposition.
Assume that defines a normal Newton nondegenerate Weil divisor , which is not rational. Then there exists a normal Newton nondegenerate Weil divisor , defined by a function and a cone (possibly different than ) satisfying the following conditions:
-
❀
and have diffeomorphic links.
-
❀
.
-
❀
If is Gorenstein or pointed at , then so is .
-
❀
If is a compact facet, then .
In fact, is the union of those facets of for which .
Proof.
By section 11, we can assume that is generated by , since is normal (see section 11). The result therefore follows, using induction on the number of facets of , and sections 11 and 11 below. ∎
12. Examples
\theexample Example.
Let and let be natural numbers with no common factor, and let be coprime with . Take
The cone is then generated by
Corresponding to these, we have irreducible invariant divisors and multiplicities
The Newton diagram consists of a single face with normal vector and . Fulton shows in 3.4 of [13] that the group of Weil divisors modulo linear equivalence on is generated by , and that is Cartier if and only if there is a so that for .
In our case, is equivalent to . Therefore, if is Cartier, then there is a so that . Therefore, we find , and
Therefore, is Cartier if and only if , i.e. .

\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.

They show that for this graph, the path lattice cohomology is , but that the maximal geometric genus among analytic structures with this topological type is . 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
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 -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 -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 -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.