Home

Introduction to Verdier Duality

Xinyu Zhou
Abstract.

We give a brief introduction to the Verdier duality. The main references are [1], [2], and [3].

Notations

In this note, X and Y are topological manifolds or more generally locally compact Hausdorff topological spaces of finite cohomological dimensions. We will fix a base field, which will be denoted as k or A. The corresponding constant sheaf on X is denoted as AΒ―X. A sheaf β„± on X is understood as a sheaf of AΒ―X-modules. Therefore, one can also regard X as a ringed space (X,π’ͺX), where π’ͺX:=AΒ―X; so β„± is an π’ͺX-module. We will work with the derived category Db⁒(X):=Db⁒(Modπ’ͺX) of bounded complexes of π’ͺX-modules, although many constructions would work for the derived category D+⁒(X) of bounded-below complexes. If β„± and 𝒒 are two sheaves on X, the sheaf-hom will be denoted as Hom¯⁒(β„±,𝒒).

Let β„± be a sheaf on X. For a subset Z of X (not necessarily open or closed), write i:Zβ†’X for the inclusion map and consider the subspace topology on Z. We endow Z with a structure sheaf π’ͺZ:=iβˆ’1⁒(π’ͺX). In the case π’ͺX=AΒ―X, the structure sheaf π’ͺZ is the constant sheaf AΒ―Z. The pullback of β„± is defined as iβˆ—β’β„±:=iβˆ’1β’β„±βŠ—iβˆ’1⁒π’ͺXπ’ͺZ. The group Γ⁒(Z,β„±) is understood as Γ⁒(Z,iβˆ—β’β„±). Also notice iβˆ’1=iβˆ—.

1. Motivation

Let X be an n-dimensional connected orientable topological manifold, and let k be a field. For any i, there is a natural pairing of singular cohomology (with compact support)

Hci⁒(X;k)Γ—Hnβˆ’i⁒(X;k)β†’Hcn⁒(X;k)β†’βˆΌk

induced by the cup product and an orientation map Hcn⁒(M;k)β†’βˆΌk. The PoincarΓ© duality states the following.

Theorem 1.1 (PoincaΓ© duality).

The pairing above is non-degenerate and thus induces an isomorphism

Hci⁒(X;k)βˆ¨β‰…Hnβˆ’i⁒(X;k)

where Hci⁒(X;k)∨ is the dual as a k-vector space.

Example 1.2.

Let X=ℂ⁑𝐏n. Recall

Hi⁒(ℂ⁑𝐏n,k)={k0≀i≀2⁒nΒ Β even0otherwise
Example 1.3.

Let X=Ξ£g be the compact orientable surface of genus g. Then

Hi⁒(Σg,k)={ki=0,2k2⁒gi=10otherwise
Example 1.4.

Let X=ℝ⁑𝐏n be the n-dimensional real projective space. ℝ⁑𝐏n is orientable if and only if n is odd. If n is odd, its cohomology is

Hi⁒(ℝ⁑𝐏n,β„€)={β„€i=0,nβ„€/2⁒℀0<i<nΒ even0otherwise

If n is even, its cohomology is

Hi⁒(ℝ⁑𝐏n,β„€)={β„€i=0β„€/2⁒℀0<i≀nΒ Β even0otherwise

Observe that in the case that ℝ⁑𝐏n is not orientable, the symmetry of the cohomology groups is broken.

It is natural to envision a generalization of the PoincarΓ© duality to more general spaces. But a naΓ―ve extension would fail as shown by the next example.

Example 1.5.

Let X be the figure-eight. The expected dimension of X should be 1. One has H0⁒(X,ℝ)=ℝ and H1⁒(X,ℝ)=ℝ2. So the cohomology groups don’t satisfy the PoincarΓ© duality.

The problem is solved by the Verdier duality. But before going into that, let us first reinterpret the PoincarΓ© duality using sheaf cohomology.

Let X be an n-dimensional topological manifold (not necessarily orientable). We also fix a coefficient field A111Many constructions would work for A being a (commutative) ring. But for simplicity, we only assume A is a field.. Write AΒ―X for the constant sheaf on X with values in A. The orientation sheaf OrX of X is the sheaf associated to the presheaf

U↦Hcn⁒(U,A)∨

The orientation sheaf OrX can be equivalently defined as the sheaf associated to the presheaf

U↦Hn⁒(X,Xβˆ’U;A)
Proposition 1.6.

The two definitions are equivalent.

Proof.

Denote by Or1 and Or2 the two sheaves, and denote by OrXpre and OrXpre the corresponding presheaves. For any open subset UβŠ†X, there is a natural map

(1) Hn⁒(X,Xβˆ’U;A) β†’Hcn⁒(U;A)∨
(2) [ΞΎ] ↦(βˆ«ΞΎβ‹…:Hcn(U,A)β†’A)

If U is small enough, it is homeomorphic to an open subset of ℝn. In this case, Hn⁒(X,Xβˆ’U;A) is naturally identified with Hn⁒(Sn;A)β‰…A, and Hcn⁒(U;A) is identified with Hn⁒(Sn;A)β‰…A. The natural map 1 becomes an isomorphism. This should the natural map Or2β†’Or1 is an isomorphism. Furthermore, this makes clear that the orientation sheaf is locally constant. ∎

In the case that X is orientable, there are non-canonical isomorphisms OrX≅A¯X, and each such isomorphism corresponds to a choice of orientation on X. The Poincaré duality can be restated as: there is a non-degenerate pairing

Hci⁒(X,AΒ―X)Γ—Hnβˆ’i⁒(X,OrX)β†’Hcn⁒(X,OrX)β†’βˆΌA

for any i.

2. f! and f!

Let f:Xβ†’Y be a continuous map between locally compact topological spaces. Let β„± be a sheaf on X. Recall the support of a section sβˆˆΞ“β’(X,β„±) is the closed subset

Supp⁑s:={x∈X∣s⁒(x)β‰ 0}Β―

where s⁒(x) is the image of s in the stalk β„±x at the point x. Using this notion, we can define:

Definition 2.1.

Define a subgroup Ξ“c⁒(X,β„±) of Γ⁒(X,β„±) by

Ξ“c⁒(X,β„±):={sβˆˆΞ“β’(X,β„±)∣Supp⁑s⁒ is compact}.

The functor Ξ“c⁒(X,β‹…) is left-exact and has right derived functor R⁒Γc⁒(X,β‹…). The i-th derived functor is called the (hyper)cohomology in degree i with compact support and is often denoted as

ℍi⁒(X,β„±βˆ™):=hi⁒(R⁒Γc⁒(X,β„±βˆ™))

or as Hci⁒(X,β„±) if β„±βˆ™=ℱ⁑[0] is concentrated in degree 0.

Recall that a map between locally compact topological spaces is called proper if the inverse image of a compact set is compact.

Definition 2.2.

For any sheaf β„±βˆˆSh⁒(X), the pushforward with proper support f!β’β„±βˆˆSh⁒(Y) is defined by assigning to every open subset UβŠ†Y an A-vector space

Γ⁒(U,f!⁒ℱ):={sβˆˆΞ“β’(fβˆ’1⁒(U),β„±)∣f|Supp⁑s:Supp⁑sβ†’U⁒ is proper}

The assignment ℱ↦f!⁒ℱ defines a left-exact functor

f!:Sh⁒(X)β†’Sh⁒(Y)

and therefore also gives a derived functor

R⁒f!:Db⁒(X)β†’Db⁒(Y)

The functor f! can be thought as the relative version of Ξ“c⁒(X,β‹…) as illustrated by the following example.

Example 2.3.

If Y is a point, then f!⁒(β„±)=Ξ“c⁒(X,β„±) and Ri⁒f!⁒(β„±)=Hci⁒(X,β„±).

Example 2.4.

If f is proper (e.g. a closed inclusion), then f!=fβˆ—.

We want to define a functor f!:Sh⁒(Y)β†’Sh⁒(X) right adjoint to f!. That is, f! should satisfy

Hom¯⁒(f!⁒ℱ,𝒒)β‰…fβˆ—β’Hom¯⁒(β„±,f!⁒𝒒)

for any β„±βˆˆSh⁒(X) and π’’βˆˆSh⁒(Y). However, the following example shows such f! does not always exist.

Example 2.5.

Let j:Uβ†’X be the inclusion of an open subset UβŠ†X. For β„±=AΒ―U and π’’βˆˆSh⁒(Y), we have

Γ⁒(U,j!⁒𝒒) =Hom⁑(AΒ―U,j!⁒𝒒)
=Γ⁒(U,Hom¯⁒(AΒ―U,j!⁒𝒒))
=Γ⁒(U,jβˆ—β’Hom¯⁒(AΒ―U,j!⁒𝒒))
=Γ⁒(U,Hom¯⁒(j!⁒AΒ―U,𝒒))
=Hom⁑(jβˆ—β’j!⁒AΒ―U,jβˆ—β’π’’)
=Hom⁑(AΒ―U,jβˆ—β’π’’)
=Γ⁒(U,jβˆ—β’π’’)

Performing the same calculation for every open VβŠ‚U, one obtains

j!⁒𝒒≅jβˆ—β’π’’

for open inclusions j.
Now for a continuous map f:X→Y, by a similar calculation, one gets

Γ⁒(U,f!⁒𝒒)=Hom⁑(f!⁒(j!⁒AΒ―U),𝒒)

This implies that, if f! existed, then f!⁒𝒒 should be given by the assignment

U↦Hom⁑(f!⁒(j!⁒AΒ―U),𝒒)

But this is in general not a sheaf.

However, we claim that f! can be defined on the derived level. That is, we can define a functor

R⁒f!=f!:Db⁒(Y)β†’Db⁒(X)

which satisfies the derived version of adjointness condition, which we will state later.

To define f! on the derived level, we need the notions of c-soft and flat sheaves.

Definition 2.6.

A sheaf β„±βˆˆSh⁒(X) is c-soft if the map Γ⁒(X,β„±)→Γ⁒(K,β„±) is surjective for any compact subset KβŠ‚X.

The following characterization of c-softness is useful.

Lemma 2.7.

Let β„± be a sheaf on X. The following are equivalent:

  1. (1)

    β„± is c-soft;

  2. (2)

    Hc1⁒(U,β„±|U)=0 for all open subsets UβŠ‚X;

  3. (3)

    Hci⁒(U,β„±|U)=0 for all iβ‰₯1 and all open subsets UβŠ‚X.

Corollary 2.8.

text

  1. (1)

    Any injective or flasuqe sheaf is c-soft .

  2. (2)

    If a sheaf on the open subset UβŠ‚X is c-soft, then its extension by zero is c-soft on X.

  3. (3)

    The restriction of a c-soft sheaf to a closed subspace is c-soft.

  4. (4)

    The restriction of a c-soft sheaf to an open subspace is c-soft.

Proof.

Easy. Omitted. ∎

Proposition 2.9.

Assume X has cohomological dimension n. Consider an exact sequence of sheaves on X

0β†’Aβ†’F0β†’β‹―β†’Fkβˆ’1β†’Bβ†’0

where Fi are c-soft.

  1. (1)

    If kβ‰₯n, then B is c-soft.

  2. (2)

    If kβ‰₯0 and A is also c-soft, then B is c-soft.

Proof.

Since each Fi is Ξ“c-acyclic, it is easy to see

Hci+k⁒(U,A)β‰…Hci⁒(U,B)

for any open subset UβŠ†X and iβ‰₯1. If kβ‰₯n, then Hci+k⁒(U,A)=0. Hence Hci⁒(U,B)=0, and consequently, B is c-soft. If A is c-soft, then already Hci⁒(U,A)=0 for any iβ‰₯1. Therefore, B is c-soft. ∎

Definition 2.10.

A sheaf π’¦βˆˆSh⁒(X) is flat if for any injective map β„±β†ͺ𝒒 of sheaves on X, the map β„±βŠ—π’¦β†’π’’βŠ—π’¦ is again injective. A complex π’¦βˆ™ of sheaves on X is flat if each term 𝒦i is flat.

Proposition 2.11.

Assume X has cohomological dimension n. Consider an exact sequence of sheaves on X

0β†’Aβ†’F0β†’β‹―β†’Fkβˆ’1β†’Bβ†’0

where Fi are flat.

  1. (1)

    If kβ‰₯n, then B is flat.

  2. (2)

    If kβ‰₯0 and A is also flat, then B is flat.

Definition 2.12 (Cohomological dimension).

The cohomological dimension dimX of a locally compact Hausdorff space X is the smallest integer n for which

Hci⁒(X,β„±)=0

for all i>n and all sheaves β„± on X.

Example 2.13.

A stratified pseudomanifold as introduced in [1] Section 4.1.2 has a finite cohomological dimension. In particular, any topological manifold has a finite cohomological dimension.

For now on, assume X and Y have finite cohomological dimension. Assume dimX=n. Then any sheaf β„± on X admits a resolution by c-soft flat sheaves

0→ℱ→K0→⋯→Kn→0

This is obtained by first considering the Godement resolution

0β†’β„±β†’G0β†’G1β†’β‹―

Let us first review how the Godement resolution is defined. For a sheaf β„± on X, we set Gode⁒(β„±) to be the sheaf

Uβ†¦βˆx∈Uβ„±x.

There is a natural map β„±β†’Gode⁒(β„±) given by on any open subset U the maps ℱ⁑(U)⁒t⁒o⁒ℱx to stalks at x∈U. It is easy to see the natural map β„±β†’Gode⁒(β„±) is injective. Its cokernel is denoted as 𝒒.

0β†’β„±β†’Gode⁒(β„±)→𝒒→0

We can now apply Gode⁒(β‹…) to 𝒒 and obtain a sheaf morphism Gode⁒(β„±)β†’Gode⁒(𝒒) by composing Gode⁒(β„±)→𝒒 and 𝒒→Gode⁒(𝒒). Now write G0:=Gode⁒(β„±) and G1:=Gode⁒(𝒒). Inductively, we define Gi:=Gode⁒(coker⁑(Giβˆ’2β†’Giβˆ’1)). We therefore get a complex of sheaves

0→ℱ→𝒒0→𝒒1β†’β‹―

Furthermore, one verifies easily that each term Gi is flasque.

We now can obtain a desired c-soft flat resolution of ℱ by truncating the Godement resolution. Denote by di:Gi→Gi+1 the differential map. Set

Ki:={Gi0≀i<nim⁑dnβˆ’1i=n
Proposition 2.14.

All Ki are c-soft and flat.

Proof.

If 0≀i<n, Ki is c-soft since it is flasque. It is also soft by construction. If i=n, the statement follows from PropositionΒ 2.9 and PropositionΒ 2.11. ∎

If β„±βˆ™ is a bounded complex of sheaves on X, then similar to the Cartan-Eilenberg resolution, we can construct a quasi-isomorphism

β„±βˆ™β†’qisKβˆ™

where Kβˆ™ is a bounded complex of c-soft flat sheaves on X.

Fix a c-soft sheaf K on X. For any sheaf β„± on U and for any open inclusion i:Uβ†ͺX, we will write β„±X:=i!⁒ℱ to avoid the need of a symbol for the open inclusion. One can verify that the assignment

U↦Hom⁑(f!⁒(K|U)X,𝒒)

is a sheaf on X; we denote this sheaf by fK!⁒𝒒. For a fixed bounded c-soft flat resolution π’¦βˆ™ of AΒ―X and a bounded complex π’’βˆ™ of sheaves on X, define fπ’¦βˆ™!β’π’’βˆ™ to be the simple complex associated to the double complex (fKi!⁒𝒒j)i,jβˆˆβ„€. In other words, fπ’¦βˆ™!β’π’’βˆ™ has associated presheaf

U↦Homβˆ™β‘(f!⁒(π’¦βˆ™|U)X,π’’βˆ™)
Proposition 2.15.

If π’¦βˆ™ and β„’βˆ™ are two bounded c-soft flat resolutions of AΒ―X with a quasi-isomorphism Ο„:π’¦βˆ™β†’qisβ„’βˆ™ extending the identity map on AΒ―X, then Ο„ induces a quasi-isomorphism fβ„’βˆ™!⁒𝒒→fπ’¦βˆ™!⁒𝒒. In particular, we obtain a well-defined functor

f!:Db⁒(Y)β†’Db⁒(X)

independent of the choice of c-soft flat resolutions of AΒ―X.

3. Dualizing sheaf and Verdier duality

Definition 3.1.

Assume Y=pt is a point. The dualizing complex of X is defined to be

𝔻Xβˆ™=f!⁒AΒ―pt

The Verdier dual of a bounded complex β„±βˆ™βˆˆDb⁒(X) is defined to be

π’ŸXβ‘β„±βˆ™:=R⁒Homβˆ™β‘(β„±βˆ™,𝔻Xβˆ™)

The next important theorem shows the connection between the dualizing complex and the orientation sheaf on a manifold.

Theorem 3.2.

Let X be an n-dimensional topological manifold with orientation sheaf OrX. There is a canonical isomorphism in Db⁒(X):

𝔻Xβˆ™β‰…OrX⁑[n].

Now we have the statement of the (local) Verdier duality.

Theorem 3.3 (Verdier duality).

For β„±βˆ™βˆˆDb⁒(X) and π’’βˆ™βˆˆDb⁒(Y), there is a canonical isomorphism

R⁒HomΒ―βˆ™β’(R⁒f!⁒ℱ,π’’βˆ™)β‰…R⁒fβˆ—β’R⁒HomΒ―βˆ™β’(β„±βˆ™,R⁒f!β’π’’βˆ™)

Notice this is a kind of adjointness property, which we promised to state in the previous section. Applying the functor h0⁒Γ⁒(Y,β‹…), one obtains the global version of the Verdier duality.

Corollary 3.4 (Global Verdier duality).

With notations as above, there is an isomorphism

HomDb⁒(Y)⁑(R⁒f!⁒ℱ,π’’βˆ™)β‰…HomDb⁒(X)⁑(β„±βˆ™,R⁒f!β’π’’βˆ™).

4. Examples

We now consider some example of the Verdier duality. The primary one is of course the PoincarΓ© duality.

Proposition 4.1 (PoincarΓ© duality).

Let X be an n-dimensional orientable topological manifold. Then there is a canonical isomorphism of A-vector spaces

Hnβˆ’i⁒(X;A)β‰…Hci⁒(X;A)∨

for any i.

Proof.

Since Y=pt is a point, the Verdier duality for f:X→pt simplifies to

(3) Hom⁑(R⁒Γ⁒(pt,R⁒f!⁒AΒ―X),AΒ―pt)β‰…R⁒fβˆ—β’R⁒HomΒ―βˆ™β’(AΒ―X,AΒ―X⁒[n])

in Db⁒(pt)β‰…Db⁒(A). Choose an injective resolution AΒ―Xβ†’Iβˆ™. For Hom¯⁒(AΒ―X,Iβˆ™β’[n])=Hom¯⁒(AΒ―X,Iβˆ™)⁒[n], we choose an injective resolution

Hom¯⁒(AΒ―X,Iβˆ™)⁒[n]β†’qisJβˆ™.

Now we apply the functor hβˆ’i⁒Γ⁒(pt,β‹…) on Jβˆ™; this gives

hβˆ’i⁒Γ⁒(pt,Jβˆ™)=β„βˆ’i⁒(X,Hom¯⁒(AΒ―X,Iβˆ™)⁒[n])=Hnβˆ’i⁒(X,A)

Similarly, applying hβˆ’i⁒Γ⁒(pt,β‹…) to the left-hand side gives Hci⁒(X,A)∨. Therefore, we conclude that there is an isomorphism

Hci⁒(X,A)βˆ¨β‰…Hnβˆ’i⁒(X,A)

∎

References

  • [1] Banagl, M. (2007). Topological invariants of stratified spaces. Springer Berlin, Heidelberg.
  • [2] Borel, A. (1984). Intersection Cohomology (Vol. 50). BirkhΓ€user Boston.
  • [3] Maxim, L. G. (2019). Intersection Homology & Perverse Sheaves: With Applications to Singularities (Graduate Texts in Mathematics vol. 281). Springer Cham.