The intersection cohomology Hodge module of toric varieties
Abstract.
We study the Hodge filtration of the intersection cohomology Hodge module for toric varieties. More precisely, we study the cohomology sheaves of the graded de Rham complex of the intersection cohomology Hodge module and give a precise formula relating it with the stalks of the intersection cohomology as a constructible complex. The main idea is to use the Ishida complex in order to compute the higher direct images of the sheaf of reflexive differentials.
2020 Mathematics Subject Classification:
14B05, 14C30, 14F10, 14M25, 14Q99, 32S35, 52B221. Introduction
A toric variety is a normal complex algebraic variety with an open subset isomorphic to the algebraic torus , along with an extension of the natural action of the torus to an action on . Toric varieties provide an interesting interplay between algebraic geometry and convex geometry since they admit an alternate description in terms of convex geometric objects. As a consequence, algebro-geometric concepts on toric varieties correspond to much more elementary and tractable notions in convex geometry. One particular example where this relation is exploited is in studying intersection cohomology on toric varieties. There has been a long and fruitful study of the intersection cohomology complex and the intersection cohomology groups on toric varieties starting with the works of Stanley ([Stanley:Intersection-cohomology-toric-varieties]) and Fieseler ([Fieseler-ICprojtoric]), and more recently, the works of de Cataldo-Migliorini-Mustaţă ([dCMM-toricmaps]) and Saito ([saito2020intersection]).
However, the intersection cohomology complex has a richer structure as a (pure) Hodge module in the sense of Saito’s theory (see [saito1988modulesdeHodge], [saito1990mixedHodgemodules]). A (pure) Hodge module is a tuple satisfying certain conditions, where is a holonomic -module, is a good filtration on (called the Hodge filtration), is a perverse sheaf on defined over , and is an isomorphism between and the analytic de Rham complex of
For more details, see Section 2.1. Moreover, the Hodge filtration on induces a natural filtration on , and the graded pieces lie inside the derived category of coherent sheaves on .
Given a variety , it follows from Saito’s theory that there exists a (pure) Hodge module , whose underlying perverse sheaf is the intersection complex . The main goal of this paper is to study the graded de Rham complex , which can only be captured after enhancing to a Hodge module . We now elaborate on how we study this object.
Observe that any cohomology sheaf of the graded de Rham complex on an affine toric variety is an -module, with a natural grading by , where is the group of characters of the torus. We consider the generating function of for . In order to compute this, we consider the following five generating functions for an dimensional affine toric variety defined by a cone , a proper birational toric morphism with simplicial, and two faces of . These generating functions encode the following corresponding data:
(Cohomology of fibers) | ||||
(Intersection cohomology stalks) | ||||
(Decomposition theorem) | ||||
(Kähler differentials) | ||||
(Graded de Rham complex). |
For a detailed explanation of the notation, we refer to Section 2.2 for basic notation for toric varieties, Section 2.4 for , and , Section 3 for , and Section 4 for . The relation between , , and is purely topological and well understood. The two extra pieces of data and are related to the Hodge filtration on the intersection cohomology Hodge module . However, we show that is completely determined by the topological data of the toric variety by the following formula.
Theorem 1.1 (Main Theorem = Theorem 4.1).
With the above notation, we have
Moreover, and hence , can be computed explicitly in an algorithmic way.
A rough sketch of the proof of the equality goes as follows. We know that , , and are related by the Decomposition theorem. Similarly, the decomposition theorem for Hodge modules gives a similar relation between the , , and . The main new ingredient is to compute using the Ishida complex, which is presented in Section 3. Then we show that can be expressed explicitly in terms of when is given by a barycentric subdivision of the fan (Proposition 3.4). Along the way, we also show that barycentric subdivisions are shellable in Proposition 2.16, which is an interesting combinatorial result in its own right. Finally, the rest of the proof follows using the two Decomposition theorems and induction.
For the explicit calculation of (and hence ), we again use the relation between , , and given by the Decomposition theorem. We have an explicit description of from [dCMM-toricmaps]. We then follow the strategy of [CFS-Effectivedecompositiontheorem] to prove that once we have explicitly, we can calculate both and (Remark 2.18). We demonstrate this strategy in the appendix by explicitly computing in dimensions .
2. Preliminaries
2.1. Hodge modules
We give a brief summary on Hodge modules and state some results relevant to our situation. We will mostly follow the notation in [saito1988modulesdeHodge] and [saito1990mixedHodgemodules]. In these papers, Saito defines two abelian categories and which are the categories of polarizable Hodge modules of weight , and the category of polarizable mixed Hodge modules. The objects in are holonomic -modules with a filtration by coherent sheaves with some extra structure satisfying suitable conditions. The objects in are also holonomic -modules with a filtration and an additional filtration by holonomic -modules satisfying some suitable conditions. The most important condition is that the graded piece should be an object in . For general terminology associated to -modules, we refer to [HTT-Dmodulesbook].
The most important piece of data in our case is the filtration , known as the Hodge filtration. For a mixed Hodge module on a smooth variety , we can consider the de Rham complex
which sits in cohomological degrees . Moreover, induces a natural filtration on and the graded pieces are given by
which also sits in cohomological degrees . The maps in this complex are -linear and we thus view as an object in , the derived category of coherent sheaves of . We mention that even if is singular, one can define the categories and by embedding into a smooth variety and considering the objects in and , respectively, which are supported on . The categories and do not depend on the choice of the embedding. If cannot be embedded in a smooth variety, we locally embed each open set and impose suitable compatibility conditions on the intersections. The de Rham complex and the graded pieces are defined as above by locally embedding into a smooth variety. The graded pieces of the de Rham complex also do not depend on the choice of the embedding as objects in .
We say a pure Hodge module has strict support if is supported on and has no nonzero subobjects or quotients supported on a strictly smaller subset of . The category admits a decomposition by strict support, that means, for any , there is a decomposition
such that has strict support , where the sum runs over all irreducible subvarieties of . Also, a pure Hodge module with strict support is a variation of Hodge structures on an open subset of weight . Conversely, any variation of Hodge structures on an open subset of the smooth locus of can be uniquely extended to a pure Hodge module on with strict support . In this case, the underlying -module is the intermediate extension of the -module corresponding to the variation of Hodge structures. In this sense the intersection cohomology -module on underlies a pure Hodge module of weight since it is associated to the trivial variation of Hodge structures on the smooth locus of . We denote by the intersection cohomology Hodge module of in order to distinguish this from the perverse sheaf .
The derived category of mixed Hodge modules has a six functor formalism and moreover, these functors are compatible with the functors at the level of perverse sheaves. The most important functor that we use is the pushforward , where is a proper morphism. We recall Saito’s decomposition theorem for Hodge modules. We say a complex is pure of weight if is a pure Hodge module of weight for all . In this case, we always have the following
Proposition 2.1 ([saito1990mixedHodgemodules]*4.5.4).
If is pure of weight , then
Similarly, we say that is of weight (resp. ), if the following condition is satisfied:
By [saito1990mixedHodgemodules]*4.5.2, if is of weight (resp. ), then (resp. ) is also of weight (resp. ). Since for proper morphisms, in this case takes pure complexes to pure complexes. Therefore, we have the following decomposition theorem.
Theorem 2.2 (Saito’s decomposition theorem).
Let be a proper morphism and be a polarizable pure Hodge module. Then we have a decomposition
and for all .
We also recall that taking the graded de Rham complex commutes with the pushforward by a proper morphism [saito1988modulesdeHodge]*2.3.7:
(1) |
We end this section by discussing the relation between the Du Bois complex and mixed Hodge modules. In [DuBois:complexe-de-deRham], Du Bois introduced a filtered complex which can be thought of as a replacement of the de Rham complex when is singular. By taking the graded quotients, the -th Du Bois complex is defined as
We have a natural comparison map which is an isomorphism if is smooth. Note that is an object in .
In [Saito-MixedHodgecomplexes], Saito gives a description of the Du Bois complex using the trivial mixed Hodge module . The category of mixed Hodge modules over a point can be identified with the category of mixed Hodge structures. Hence, we have the Hodge module with weight zero given by the following mixed Hodge structure :
For an arbitrary variety , is defined as
where is the structure morphism. It is a consequence of [Saito-MixedHodgecomplexes]*Theorem 4.2 that the graded de Rham complex of this Hodge modules is related to the Du Bois complex in the following way:
One can also get this easily using [Mustata-Popa:localcohomologyHodge]*Proposition 5.5 and duality.
The two objects and have nice descriptions when the variety has quotient singularities. In general, we have a natural morphism in the derived category of mixed Hodge modules (see [saito1990mixedHodgemodules, 4.5.11]). If has quotient singularities, this is an isomorphism at the level of perverse sheaves by [borho-macpherson]*Section 1.4, which implies that is also an isomorphism of mixed Hodge modules. Moreover, coincides with the reflexive Kähler differentials if has quotient singularities [DuBois:complexe-de-deRham]*Théorème 5.3. Wrapping all up, we have the following lemma:
Lemma 2.3.
If has quotient singularities, then
2.2. Toric varieties
Fix a free abelian group of rank and let . Denote and . To a strongly convex rational polyhedral cone , we associate an -dimensional affine toric variety . More generally, to a fan , we associate an -dimensional toric variety by gluing the affine toric varieties corresponding to the cones of . For general notions regarding toric varieties, we refer to [Fulton-ToricVar] and [CoxLittleSchenck-ToricVar].
Notation and terminology. We collect some notation for convex cones that we will use. Here, denotes a strongly convex rational polyhedral cone, and denote faces of .
-
(1)
-
(2)
-
(3)
-
(4)
.
-
(5)
is the subspace spanned by .
-
(6)
.
-
(7)
We denote by the image of under the projection .
-
(8)
is full-dimensional if .
-
(9)
is simplicial if the 1-dimensional faces (i.e. rays) are linearly independent over in .
We collect several facts on toric varieties that we will need.
Remark 2.4.
-
(1)
[CoxLittleSchenck-ToricVar]*Theorem 9.2.5 For a proper toric morphism , we have for all .
-
(2)
[Fulton-ToricVar]*Section 3.1 Let be the affine toric variety corresponding to a strongly convex rational polyhedral cone . For an -dimensional face , we get a torus invariant subvariety of codimension . This is the affine toric variety corresponding to the cone , where the lattice and the dual lattice are given by
We denote by the torus orbit corresponding to . Also, is an open subset of and we have the diagram of torus equivariant morphisms
After fixing a non-canonical splitting and the corresponding splitting , we can identify as the projection , where is the full-dimensional toric variety .
For two faces , we denote by the full-dimensional affine toric variety corresponding to the cone . We have an analogous diagram
Note that
as a fan in , , and , respectively.
-
(3)
We say that a toric variety is simplicial if all the cones in the fan are simplicial. If is simplicial, then has quotient singularities [CoxLittleSchenck-ToricVar, Theorem 11.4.8]. By Lemma 2.3, in this case we have a canonical isomorphism
-
(4)
[oda-ConvexBodies, Lemma 3.5] (see also [Ishida]) There is a natural resolution of the sheaf of reflexive differentials of a simplicial toric variety , called the Ishida complex, that we now recall. Let be an -dimensional cone. Since is simplicial, is generated by rays . We set
where . For instance, . Then we have an exact complex
(2)
The relevant situation for us will be when is obtained by a simplicial subdivision of the fan . In this case, for , we denote by the minimal cone of containing . For , we let
We point out that our notation for is slightly different from the one in [dCMM-toricmaps]. The following proposition describes how the fibers look like for arbitrary proper toric morphisms.
Proposition 2.5 ([dCMM-toricmaps]*Lemma 2.6, Proposition 2.7).
Let be a proper toric morphism.
-
(i)
Then every irreducible component of the fiber is a toric variety. Moreover, this is smooth (resp. simplicial) if is smooth (resp. simplicial).
-
(ii)
For any , we have an isomorphism such that the restriction of to corresponds to the projection onto the second component. In particular, for every .
The following proposition gives a combinatorial formula for the cohomology of the fibers.
Proposition 2.6 ([dCMM-toricmaps]*Theorem C).
Let be the proper birational toric morphism obtained by a simplicial subdivision of . For every and every , we have the following formula
Remark 2.7.
At last, we describe the construction of what we call a barycentric resolution of an affine toric variety of dimension . Let be the affine toric variety corresponding to a cone . Since is strictly convex, there exists a linear functional such that on and . Then consider the polytope . Note that the cone generated by is . For each face , choose lying inside the relative interior of . We construct the fan by describing its maximal cones. Each maximal cone in is of the form
where each is an -dimensional face of satisfying . Note that is simplicial by construction. We have a proper birational toric morphism corresponding to the map induced by the identity on . We call a barycentric subdivision of and the corresponding toric morphism a barycentric resolution of . Observe that we also get a simplicial polytopal complex (see Definition 2.9), the cone over which is . Each maximal simplex of corresponds to a sequence of faces
where is a facet of for . Observe that is a vertex of . Given this sequence, the set of vertices of the corresponding simplex consists of
Even though the toric variety constructed in this way depends on the choice of the generators in the relative interior, this will not affect the arguments throughout this article.
Remark 2.8.
Though it will not be important for what follows, we mention that every barycentric resolution of an affine toric variety is a projective morphism.
2.3. Polytopes and Shellability
In this section we review the concept of shellability, as we will later use it to prove results about the pushforward of the Ishida complex. We follow [ziegler-polytopes, Chapter 8], where the reader can find additional details.
Definition 2.9.
[ziegler-polytopes] A polytopal complex is a finite, non-empty collection of polytopes (called faces of ) in that contains the faces of all its polytopes and such that the intersection of any two of its polytopes is a face of each of them. The inclusion-maximal faces of are called the facets of .
A polytopal complex is pure if all its facets have the same dimension and is simplicial if all its faces are simplices.
Example 2.10.
If is a polytope, then the boundary complex , which is defined to be the set of all proper faces of , is a pure polytopal complex of dimension .
Definition 2.11 ([ziegler-polytopes]*Definition 8.1).
Let be a pure -dimensional polytopal complex. A shelling of is a linear ordering of the facets of such that either is a set of points, or it satisfies the following condition:
-
(1)
The boundary complex of the first facet has a shelling.
-
(2)
For ,
for some shelling of
A pure polytopal complex is shellable if it has a shelling.
We will use the following theorem of Bruggesser and Mani:
Theorem 2.12 ([Bruggesser-Mani:Shellable]).
For a polytope , the polytopal complex is shellable.
Definition 2.13 (Type).
Let be a shelling of a simplicial polytopal complex. Define to be of type . For , we have that is a pure -dimensional complex. Define to be of type , where is the number of facets in the pure complex .
Notation 2.14.
Let be a simplex whose vertices are . In this case, we sometimes use the notation
in order to denote .
Remark 2.15.
With notation as in Definition 2.13, let be of type , and let for be the facets in the pure complex , so that
Observe that we have:
We will be interested in the shellability of the pure simplicial polytopal complex as defined in Remark 2.7.
Proposition 2.16.
With notation as in Remark 2.7, the polytopal complex is shellable.
Proof.
We describe a shelling of by defining a lexicographic order on the set of maximal simplices of .
Let us recall the maximal simplices of the simplicial polytopal complex . Each maximal simplex corresponds to a sequence of faces
where is a facet of for . Observe that is a vertex of . From now on, for a maximal simplex in corresponding to the chain of faces , we use the notation
to represent the chain of faces.
First, let be an ordering of the facets of the polytope such that it gives a shelling of (such an ordering exists by Theorem 2.12). This defines an ordering, call it , on the set of facets of , given by:
Similarly, given an dimensional face of , let us now define an order on the set of facets of . Observe that
where is a shelling of . Define:
In the same vein, given a chain , we now inductively define an order on the set of facets of . Observe that, by induction, the chain defines an order on the set of facets of (by defining a shelling on ). We will use this information to get an order on the set of facets of . Observe:
where is a shelling of . Define:
We can now finally define the lexicographic order on the set of all simplices of : Given two simplices and , let (observe that as ), then define if
Now, we finally prove that if we arrange the simplices in the lexicographical order, this is a shelling of . Consider a simplex
and write the vertices of as where , with notation as in Remark 2.7. One observation is the following: if we have a chain of faces , then there is a unique face such that . Denote as the unique vertex of which is not .
We describe the simplices which share a facet with . There are exactly of them:
Let be the integers such that exactly for these indices.
Note that the facets are the facets of which are contained in . We claim that any face of containing is not a face of with . Provided that, it is straightforward to see that
following Remark 2.15. This also shows that is of type and that is a shelling order.
We are left with proving that any face of containing is not a face of with , that is, we want to prove that the set of simplices satisfying the following properties:
-
(1)
-
(2)
for
is empty.
Assume for the sake of contradiction that . For each consider
Pick such that is the minimum among these quantities. We note that cannot be one of the by the definition of .
First, we suppose that . We have . Also, we have which implies that contains . This implies that there exists a facet of such that is a -dimensional face and is a face of . We replace with
Then we see that . If , then we replace with
Then we see that and this violates the minimality of . On the contrary, if , this means that
since is contained on the right hand side. However, this implies that , which is a contradiction since are exactly those satisfying and is not one of them.
The second case is when . This implies that there exists such that is a facet of . Then we follow the same lines as above and get a contradiction on the minimality of . ∎
2.4. The Decomposition Theorem
In this subsection, we study the decomposition theorem for a proper birational toric morphism. We note that the singular cohomology of the fibers determine both the intersection cohomology stalks and the coefficients in the decomposition theorem, and we describe how to compute both of these explicitly (see Remark 2.18 below). We follow along the lines of [CFS-Effectivedecompositiontheorem], which discusses the decomposition theorem for Schubert varieties. However, the same argument works in the setting of toric varieties.
Let be the affine toric variety of dimension associated to a full-dimensional rational polyhedral cone and let be the associated fan. Consider a toric variety obtained by a subdivision of , and let be the corresponding toric morphism. The decomposition theorem in this setting (see [dCMM-toricmaps, Theorem D]) tells us that
(3) |
This is a more precise version of the decomposition theorem in [BBD-Faisceauxpervers], which is achieved by exploiting the action of the torus.
For , we define
where and
Observe that is independent of the choice of by Proposition 2.5, and is independent of the choice of by Remark 2.18 below.
We fix the notation for , and from now on. If there is an ambiguity regarding the ambient toric variety, we will sometimes denote by . We also point out that is nonzero only for . If we additionally assume that is a simplicial toric variety, we have . By taking the stalk of both sides of Equation 3 at a point , we get
(4) |
Now we list some basic properties of these polynomials.
-
(1)
by Poincaré duality [dCM-Decomposition]*§1.6. (10).
-
(2)
is strictly supported in negative degrees if by [dCM-Decomposition]*§2.1.(12).
-
(3)
.
-
(4)
.
For convenience, we put
Lemma 2.17.
For every pair of faces of with , we have the equality
where in the second term is considered as a face of the cone and the in the third term is considered as the cone .
Proof.
The first equality is straightforward. For the second equality, we use the description of . Note that . Since , we get
due to the shift of cohomological degrees in the intersection complex. This proves the second equality. ∎
Remark 2.18.
We describe how to explicitly compute and in terms of the combinatorial data of the map . Proposition 2.6 tells us how to compute explicitly in terms of the combinatorial data of . By induction on the dimension of , we can assume that we have computed for all , since . Moreover, we can also assume that we have computed for using Lemma 2.17 and by induction on the dimension of . We proceed to compute and . By Equation (4) and using and , we get
After multiplying by , we get
(5) |
Note that and can be completely determined provided that the left-hand side of Equation (5) is known since is supported in strictly negative degrees and . Hence, we can compute and inductively. Additionally, we observe from the computation above that does not depend on the choice of .
In the appendix, we use this idea to explicitly compute in dimensions .
We finally upgrade the decomposition in (3) at the level of Hodge modules by accounting for the Tate twists. Since is a pure Hodge module of weight , the direct image is a pure Hodge module of weight . Hence, the appropriate formula for Hodge modules is
since is pure of weight . One can notice that is automatically zero if is odd due to weight reasons, although this can already be seen at the level of constructible sheaves in [dCMM-toricmaps]*Theorem D.
3. Higher direct images of Kähler differentials for toric morphisms
Let be a barycentric resolution of an affine toric variety as in Remark 2.7. The main goal of this section is to describe a systematic way to compute the higher direct images of the sheaves of reflexive differentials.
The Ishida complex (2) provides a resolution of . The terms appearing in the Ishida complex are structure sheaves of various torus invariant subvarieties. Combining this fact with Remark 2.4 (1), we see that the higher direct images of can be computed by calculating the cohomology of the pushforward of the Ishida complex along . The pushforward of the Ishida complex is given by:
(6) |
We first describe the morphisms in the complex. Let such that and . Then we see that if and only if . Let be the rays of . If , there is a unique ray such that is the span of . Recall from Remark 2.4 (4) that
Note that can be naturally identified with . From the short exact sequence , we get a surjection
The map between in (6) is given by the map above (up to a well-defined sign after choosing the order of the rays) tensored with and the restriction map if . Otherwise, the map is zero.
Using the action of the torus, the complex in (6) decomposes into various eigenspaces, hence it carries a natural -grading. For example, for viewed as a character of the torus, the degree -part of can be described as
after identifying with its space of global sections. First, let us consider this complex in degree . Since every contains , the pushforward of the Ishida complex in degree 0 is
(7) |
with in cohomological degree .
Let us write this in terms of the polytopal complex instead of . For , we denote by the vector space where is the cone in corresponding to . We point out that there is a difference of dimension by 1 between the dimension as a cone and the dimension as a simplex, hence (7) becomes
(8) |
We choose the lexicographic shelling of as described in Proposition 2.16. Let denote the shelling.
Remark 3.1.
Let us describe the complex associated to a simplex of type for . Let be a simplex of type with . Let denote the face of that is not a face of one of . We associate a Koszul-type complex to :
(9) |
with in cohomological degree . We note that the morphisms in the complex come from the push-forward of the Ishida complex in Equation 6. When , this is a complex concentrated in a single degree, namely . For , this is exactly [toric-SVV]*Equation (3) tensored by , and is hence exact. In either case, observe that the complex is exact in all cohomological degrees other than .
Proposition 3.2.
The complex (8) is exact in all cohomological degrees other than .
Proof.
Let be a shelling of . We compute the cohomology of the complex (8)
with in cohomological degree , by putting a filtration on from the shelling and computing the associated spectral sequence. We define by
for , and . First we observe that are indeed subcomplexes of . For this, it is enough to show that if with , and if with , then the morphism in is zero. But this is true because cannot be a face of in this situation, and hence is indeed a subcomplex of . We describe the graded pieces of this filtration. First, notice that
For , let be the face of that is not a face of any of . Notice that
where sits in cohomological degree . Observe that this is exactly the complex described in Remark 3.1. Consider the spectral sequence
We point out that unless by Remark 3.1. Therefore, the spectral sequence degenerates at , and the only non-trivial cohomology of is in cohomological degree . ∎
Remark 3.3.
Observe that Proposition 3.2 is actually a statement about a complex of vector spaces associated to the barycentric subdivision of the cone . In particular, this statement also holds for any face of and its barycentric subdivision as well.
Finally, let us see what happens in a degree . Observe that the terms that are non-zero in degree are precisely those for which . Therefore the pushforward of the Ishida complex in degree is
(10) |
Let and let (note that and ). Fix a non-canonical splitting . Observe that we have a non-canonical isomorphism
with the convention that for .
More generally, given , observe that . Define . The non-canonical splitting of we fixed earlier induces a splitting . We can now write the pushforward of the Ishida complex in degree as a direct sum
Now, observe that the complex
is the complex of vector spaces associated to the barycentric subdivision of the cone . Therefore by Remark 3.3, it follows that this complex is exact at all cohomological degrees other than .
Now, we define the generating function for the pushforward of the sheaves of reflexive differentials. For , we define
where . This is independent of the choice of , as shown by the following combinatorial formula for .
Proposition 3.4.
If is a barycentric resolution of , then
Moreover, is related to by the following formula:
Proof.
We compute the dimension of . Let and . Note that by Proposition 3.2, the cohomology of the complex
is concentrated in degree and is equal to . Furthermore, the number of summands of the -th term is equal to and the dimension of each summand of the -th term is equal to . Hence we have
Therefore, we have
This concludes the proof of the first part of the proposition. Now, we prove the second statement. By construction of the barycentric subdivision, we have
If , then we get
Observe that the fiber is an irreducible simplicial toric variety of dimension . First, we show that is irreducible. Let denote the unique ray in contained in the relative interior of , then we have
Since is irreducible, and since is an open subset (it is the inverse image of the open set under the natural map , we have that is irreducible. By Proposition 2.5, we have . Hence is irreducible and hence, irreducible simplicial by Proposition 2.5.
Now by Proposition 2.6, we have:
Since a proper simplicial toric variety satisfies Poincaré duality, we know that , and therefore
Hence, we get
Therefore,
∎
Remark 3.5.
Even though Proposition 3.4 is stated for barycentric resolutions, this method provides a general framework of computing the higher direct images of reflexive Kähler differentials. More precisely, if we have a proper toric morphism from a simplicial toric variety , the computation of essentially boils down to a linear algebra computation of finite dimensional vector spaces. Moreover, if is a birational toric morphism such that the polytopal complex associated to has every face shellable, then the first assertion of Proposition 3.4 holds in that setting as well. It would be interesting to investigate what happens for the non-shellable subdivisions.
4. Graded de Rham complex of the Intersection Cohomology Hodge module
In this section, we define the generating function associated to the graded de Rham complex of the intersection cohomology Hodge module and prove the main result of the paper.
For such that , we define as
for . This definition is independent of the choice of , as the next result shows.
Theorem 4.1.
Let be the affine toric variety associated to a full dimensional cone of dimension and let be the associated fan. Then is related to in the following way:
In particular, depends only on the graded poset structure of . Moreover, can be explicitly computed in terms of the combinatorics of .
Before giving the proof, we state a lemma relating the graded de Rham complex of the intersection cohomology and the pushforward of Kähler differentials.
Lemma 4.2.
Let be a birational toric morphism given by a simplicial subdivision of the fan . Then for each , we have
Proof.
This is a simple consequence of the decomposition theorem:
Note from Remark 2.4. (3) that
By taking and using (Equation 1), we get
By Proposition 3.4 and induction on dimension, we see that the dimension of the degree piece of does not depend on the choice of . By taking the degree piece for , we get
The assertion of the lemma follows because we have by Poincaré duality. ∎
Now, we give the proof of Theorem 4.1.
Proof of Theorem 4.1.
We prove this by induction on the dimension of . If is of dimension zero, then there is nothing to prove. For , we have the equality
by definition and the description of the fan of . Note that we also have
Hence, the equality follows by Lemma 2.17 and the induction hypothesis. Therefore, it is enough to show the equality when . Consider the proper toric morphism induced by the barycentric subdivision of . By Lemma 4.2 and the inductive hypothesis, we have
By Proposition 3.4, we have
Equation 5 in Section 2.4 gives
By multiplying on both sides, we get
Finally, observe that Remark 2.18 applied to the barycentric subdivision tells us that can be explicitly computed in terms of the combinatorics of . Therefore, can be explicitly computed in terms of the combinatorics of as well. ∎
Remark 4.3.
We end the section by relating Theorem 4.1 to a recent -theoretic result of Maxim and Schürmann. Roughly speaking, the graded de Rham complex gives a homomorphism from to , and one can consider the image of by this map. [maxim2024weighted]*Corollary 5.3 says that the image can be written as the sum
Theorem 4.1 applied to gives:
Since we are taking the image in , we specialize to and set and to get
We observe that to get
The comes from the fact that is defined for while [maxim2024weighted] work with .
Appendix A Explicit Formulas
In the appendix, we demonstrate that the polynomials and can be calculated rather explicitly by computing them for full dimensional affine toric varieties up to dimension 4.
A.1. Dimension 0, 1, and 2
Note that up to dimension 2, every toric variety is simplicial. Hence the intersection cohomology Hodge module agrees with the trivial one. For dimension zero, a zero dimensional toric variety is just a point. Hence
For dimension 1, we denote the nonzero ray by . It is easy to check that
Hence, we get
For dimension 2, let be the two extremal rays of the two dimensional cone . Then, it is clear that
The case when the first index is non-zero is redundant since it comes from a lower dimensional toric variety. Therefore,
A.2. Dimension 3
Let be a 3-dimensional affine toric variety corresponding to a cone and suppose there are extremal rays . Note that the number of two dimensional faces is also . Let be the two dimensional faces. Adding a ray in the interior of the cone gives a proper birational toric morphism where is simplicial. Note that is an isomorphism outside of the torus fixed point . Hence, the decomposition theorem tells us
It remains to calculate . Using (5) and Proposition 2.6, we get
Then we can conclude that we have
Note that all the other information for comes from lower-dimensional toric varieties because of Lemma 2.17. Therefore,
A.3. Dimension 4.
Let be a 4-dimensional affine toric variety corresponding to a cone . We denote the 1-dimensional faces by , the 2-dimensional faces by , and the 3-dimensional faces by . For each , we let the number of 1-dimensional faces contained in . We add rays and in the interior of and the interior of ’s. This gives a proper birational toric morphism where is simplicial. The morphism is an isomorphism outside a dimension 1 subset of and this implies
Hence, it remains to calculate ’s and . By considering the fiber , we can see that
This is exactly the same computation as in the previous section. Now, we compute the cohomology of the fiber . Notice that coming from . Also,
since coming from and coming from itself. The last equality follows from Euler’s identity . Also,
We notice that equals the number of 2-dimensional faces of , and for each 2-dimensional face , there are exactly two 3-dimensional faces containing . Therefore
Similarly,
Using (4) and Proposition 2.6, we get
Therefore, and
Therefore,
Acknowledgements. We would like to thank Mircea Mustaţă for numerous helpful discussions, and Claudiu Raicu for pointing out the reference [CFS-Effectivedecompositiontheorem] which helped us compute the polynomial explicitly, and Jörg Schürmann for kindly explaining the relation between [maxim2024weighted] and Theorem 4.1.