Author: Eiko

Tags: D-modules, differential equation, flat connections, derivations, differential operator, differential form

Time: 2024-12-12 01:59:49 - 2024-12-12 02:00:55 (UTC)

D-modules is the study of differential equations using algebraic methods. The main example to consider is the followinf differential equation (i=1mai(z)i)f=0. Here ai are polynomials and =z is a differential operator. It is evident that the solutions of the equation forms a linear space, but on different open sets, the solution sets might be very different. An observation of Cauchy is that, on a small ball B(z,ε) which does not have any zeros of am(z), the equation admits m solutions.

When talking about differential equations, there will be both O-modules and C-sheaves. For algebraic variety X, let OX be the sheaf of regular functions in Zariski topology and

    • there are natural multiplication actions Γ(OX)EndC(OX)OHomC(O,O).

    • A derivation of OX is a C-linear morphism δ:OXOX satisfying δU(sUtU)=δU(sU)tU+sUδU(tU).

    • So derivations have to be k-linear rather than OX-linear, so they eventually live in the ring Endk(OX). But this ring also lives Γ(OX), which are multiplications by scalars.

  1. One defines

    1. The sheaf of derivations ΘX or SDerXSEndk(OX) given by ΘX(U)=Der(OU).

    2. The ring of differential operators D(X)Endk(OX) generated by elements Γ(OX) and Der(OX).

    3. The sheaf of differential operators DXSEndk(OX) given by DX(U)=D(U)Endk(OU).

    4. We can talk about the category of DX-modules, which equipps with both OX action and derivation, fixing the problem of not being able to talk about OX-linearity and OX-modules when dealing with derivations.

  2. The simplest example of sheaf of differential operators is DA1. In this case we have Der(OA1)=k[x]x,Γ(OA1)=k[x], and therefore D(A1)=k[x,x]/([x,x]=1).

  3. Let M be a coherent OX-module, an integrable connection on M is a morphism of k-sheaves :ΘXSEndk(M) such that fUθUsU=fUθUsU, θU(fs)=(θf)s+fθs, [S,T]=[S,T]. Here the third condition is telling you that curvature is zero, thus being an ‘integrable connection’. Without the third condition it is a connection in general sense.

  4. Let M be coherent OX-module, an integrable connection on M is equivalent to given a structure of DX-module on M. Moreover, if admits an integral connection, it is locally free (vector bundles).

Introduction to D-modules

Let X be a smooth variety, OX,ΘX be the sheaf of regular functions and sheaf of vector fields (derivations). For any point of X we can take an affine neighborhood U and a local coordinate system {xi,i}1in, this is because OX,x is regular, ΩX,x1 is locally free and we can take local coordinates xi and its dual basis, satisfying ΘU=i=1nOUi,[i,j]=0,[i,xj]=δij. DU=DX|U=αNnOUα

Algebraic Structure of DX

The following example is a good way to understand the abstract structure of DX.

  • Example.

    Let U be affine, then D(U) is abstractly the algebra generated by abstract symbols {f~,θ~:fO(U),θΘ(U)} subject to relations

    1. f~g~=fg~,

    2. f~+g~=f+g~,

    3. θ1~+θ2~=θ1+θ2~,

    4. [θ1~,θ2~]=[θ1,θ2]~,

    5. f~θ~=fθ~,

    6. [θ~,f~]=θ(f)~.

    i.e. ()~ preserves the algebraic structure of OU, the Lie algebra structure of ΘU, the linear action of O on Θ and the Leibniz rule.

  • In summary, we can say in terms of algebraic structures, DX=Ring(OX)+Lie(ΘX)+OXΘX+Leibniz.

  • Example.

    In terms of local coordinates of any element P=αaααD(U), the multiplication of polynomials is obvioius. By the Leibniz rule θffθ=θ(f), the multiplication of any derivation (on the left) is given by the following linear maps =M+C where M:aααaαα multiplies on the multiple derivations and C:aαα(aα)α applies the derivation on the coefficients. Clearly M and C are C-linear maps and they commute with each other. This easily give us the following Leibniz rule α=i(Mi+Ci)αi=βα(αβ)MβCαβ. As a result α=βα(αβ)MβCαβ=β1β!α!(αβ)!MαβCβ Extending linearly we have by total symbol σ:D(U)O(U)[ξ], σ(PQ)=β1β!ξβσ(P)(x,ξ)xβσ(Q)(x,ξ).

  • Example.

    If we write everything reversely as P=()αaα and consider the action of any derivation multiplied on the right, we have similarly =M()+C since (βb)()=()β(b+(b)). Therefore ()α=β(αβ)M()βCαβ

Order Filtration

  • There is a ascending filtration called the order filtration on DU given by local coordinates on affine U FlDU=|α|lOUα. This filtration is however not dependent on the choice of local coordinates.

  • Proposition.

    (X smooth)

    1. FlDX is locally free.

    2. F0DX=OX, (FlDX)(FmDX)Fl+mDX.

    3. [FlDX,FmDX]Fl+m1DX.

  • Recall that DX is non-commutative, but since we have a filtration we can derive associated graded ring (commutative) given by (with the convention that F1DX=0) grDX=l0FlDX/Fl1DX=l0grlDX

  • Proposition.

    Let U be an affine chart with coordinates {xi,i}1in, then If we take ξi to be the image of igr1DU, then

    1. grlDU=|α|=lOUξα,

    2. grDU=OU[ξ1,,ξn].

    The associated polynomial σl:FlO[ξ] of P is called the principal symbol σl(P).

D-modules and Connections

Giving a DX-module structure is the same as giving a OX module M and a (flat) connection :ΘXSEndk(M)Mor(Abk) such that

  1. fθs=fθs,

  2. θ(fs)=θ(f)s+fθs,

  3. (flatness/integrability) [θ1,θ2]=[θ1,θ2].

We have

{DX-modules}{Flat connections on OX-modules}

M(M,θs:=θs),(M,θm=θm)(M,) Note that connection only requires the first two conditions, the third condition is called the flatness or integrability condition.

  • This can be seen as the DX module relations are defined by the algebraic structure of OX, Lie algebra structure of ΘX, linear action of OΘ and the Leibniz rule. To be precise,

    1. The first condition reflects OΘ

    2. The second condition reflects the Leibniz rule [θ,f]=θ(f).

    3. The flatness is derived from the Lie algebra structure of Θ.

    4. As the algebraic structure of O, it is automatically encoded in the OX-module structure of M.

  • Remarks on multiplication

    As a little warning, the fθ in fθ is the OΘ structure and not the multiplication in DX. In DX the multiplications fθ is different from θf, on ly the former equals the action of OΘ, the latter is the action of multiplying f followed by θ.

  • Remarks on terminologies

    Although the term integrable connections and flat connections are the same, in the theory of DX-modules people usually use integrable connections to mean something stronger, requiring an extra condition of being locally free and finitely generated. The category of such integrable connections are denoted by Conn(X). They are the most important examples of left DX-modules. You can also say they are vector bundles with flat connections.

Left and Right D-modules

Recall that given a left G-module M, we can use the anti-homomorphism ()1:GopG to obtain a right G-module M by acting mg=g1m. On a general non-commutative ring R however, there is no obvious anti-homomorphism and thus no obvious way to get a right R-module from a left R-module. However, for DX-modules, this is possible due to the following adjoint anti-involution.

  • Adjoint Anti-involution.

    Let U be affine with coordinates {xi,i}1in, then the anti-involution ():DUDU given by (aαα)=()αaα defines a anti-involution on DX, i.e. (PQ)=QP.

    Note that this involution map depends on the choice of coordinate.

  • Proof. It suffices to verify (aαbβ)=(bβ)(aα)=()βb()αa. This is a direct computation. (aαbβ)=(aγ(αγ)γ(b)αγ+β)=γ(αγ)()αγ+βaγ(b) and by the formulas in previous examples we have ()βb()αa=γ(αγ)(M()αγCγ()βb)a.=γ(αγ)()αγ+βγ(b)a. ◻

  • Corollary.

    Any left DX-module M can be viewed as a DXop-module or equivalently a right DX-module by sP:=Ps.

Recall that left DX-modules are OX-modules with a flat-connection. However right DX-modules are [not]{.underline} connections. To understand what it does, we can examine the relations

  1. s(fθ)=(sf)θ=(fs)θfθs=θ(fs),

  2. s(θf)=s(fθ+θ(f))=(sθ)ffθs+θ(f)s=fθs.

  3. s[θ1,θ2]=s(θ1θ2θ2θ1)[θ1,θ2]s=[θ2,θ1]s.

These relations seem a bit weird, in general we use = to simplify them as

  1. fθs=θ(fs),

  2. fθs=fθs+θ(f)sθ(fs)=θ(f)s+fθs,

  3. [θ1,θ2]s=[θ1,θ2]s.

So most relations are similar to that of a connection except the first one. This is infact a Lie derivative. On the top differential ΩX=ΩX1 there is a natural action of ΘX by Lie derivative, (Lθω)(θ1n)=θ(ω(θ1n))iω(,[θ,θi],), which satisfy

  1. Lfθω=Lθ(fω),

  2. Lθ(fω)=θ(f)ω+fLθω,

  3. L[θ1,θ2]ω=[Lθ1,Lθ2]ω.

Given such or L, the right DX-module structure is given by sθ:=θs. i.e. {DXop-modules}{Lie derivatives on OX-modules} M(M,θs:=sθ),M(M,)

Right Module Structure on Top Differential ΩX

In terms of local coordinates, the right module structure on ΩX is given by fdx1dxnP=(Pf)dx1dxn.

The right DX module structure on ΩX gives a map DXopSEndC(ΩX) of CX algebras, by the locally-freeness of ΩX we have an isomorphism of sheaves of rings SEndC(ΩX)ΩXOXSEndC(OX)OXΩX. which gives a canonical isomorphism of CX-algebras DXopΩXOXDXOXΩX. This is seen by computing the image of DXop by the previous results on the right actions of DXop on local coordinates of the top differentials.

  • Remarks On Why Right Module Structure

    One may wonder why ΩX is a right DX-module and not a left DX-module. One explanation is that the right module structure is naturally given by the Lie derivative. To see why using fdx1dxnPfdx1dxn will not work, we can see how coordinate transformation laws determine the right module structure and not the left in the example of ΩX1 in the 1-dimensional case. Consider two different coordinates x,y, and assume ω=adx=bdy,a=by and we compute the action of P=x=yy.

    1. If ΩX1 is a left DX-module, we would expect Pω=axdx=ybydy,ax=(y)2by which is in general not true.

    2. For the right action, we should expect ωP=axdx=y(yb)dy,ax=yy(yb), this is exactly ax=yay.

    Another curious question is, why isn’t Ω1 a D or Dop module? Why is only top differential ΩX a right module? The above example might provide an intuitive explanation, only in the top differential, we can divide two differentials and get a function. When dimX>1, the quotient dy/dx no longer make sense since m/m2 is not 1-dimensional. But nm/m2 is.

Tensor and Hom Over O

Given left DX modules M,N, right DX-modules M,N, we have the following module structures

rendering math failed o.o

  • These results might be a bit surprising, but remember that we are tensoring over O and so there are some subtleties. It is automatically O-module, to become D-module three extra conditions need to satisfy. To check the module structure quickly we can test the relation of fθ acting on left or right to see if (fθ)m=f(θm) or m(fθ)=(mf)θ, this works a lot of times.

    For example we can see why MON with the left D-structure given by θ(mn)=θmnmnθ does not work by observing that (fθ)(mn)=fθmnmnfθ while f(θ(mn))=fθmnmnθf. they differ by a term θ(f)mn.

    There is another interesting fact, if X is a smooth curve with genus g, we have a DX-module degO=0 and a right DX-module degΩX=2g2. In fact it is proved that a line bundle is equippable with DX-module structure iff degL=0, and with right DX-module structure iff degL=2g2.

  • Remarks on dual

    Note that for a left DX-module M, the O-dual M=HomO(M,O) is still a left DX-module!

    This might sound weird, but remember that we are taking dual over O. If you take the DX-dual M=HomDX(M,DX), then this is a right DX-module.

  • 2+1 Mixed Associativity

    Let M,N and M be two left and one right DX-modules. There is a canonical isomorphism of CX-modules (MOXN)DXMMDX(MOXN)(MDXM)OXN. Note that all three of them are of the form VDDDW.

  • Category of left and right D-modules are equivalent

    There is an equivalence of categories {DX-modules}{DXop-modules}. given by the two functors ΩXOX():DXModModDX,HomOX(ΩX,):ModDXDXMod. In fact here ΩXOX()=HomOX(ΩX,).

Inverse Images

Given a morphism f:XY of smooth varieties, we can pushforward tangent vectors and pullback cotangent vectors naturally. Therefore we have the tangent map T(f):ΘXfΘY=OXf1OYf1ΘY and the cotangent map T(f):OXf1OYf1ΩY1=fΩY1ΩX1 given by a(x,y)dya(x,f(x))df(x) naturally. In fact the tangent map is the dual of the cotangent map, HomOX(ΩX1,OX)HomOX(fΩY1,OX)=fHomOY(ΩY1,OY). The latter equality comes from the canonical map fHomOY(F,G)HomOX(fF,fG) which is an isomorphism when F is locally free and finitely generated, satisfied by ΩY1.

These natural operations allow us to define a DX-module structure on a pullback fM of a DY-module M by f:DYModDXMod,MfM, θ(ψsf)=θψ+(Tfθs)f. To be precise, let ψsOXf1OYf1M, then for θΘX we define θ(ψs)=θψs+ψTfθs. Given local good coordinate system {yi,yi} on Y, the tangent map can be computed as Tf(θ)=iθ(yif)yi, so θ(ψs)=θψs+iψθ(yif)yis.

As such, any MDYMod gives a DX-module fM=OXf1OYf1M. The pullback fDY=OXf1OYf1DY is therefore a left DX module. But at the same time it is a right f1DY module, and the two module structures are compatible (note that the left module structure acts on both components, but the right module structure only acts on the right component). We denote this (DX,f1DY)-module by DXY.

  • Proposition.

    Given a morphism f:XY of smooth varieties, the pullback functor f:DYModDXMod,MfM also given by MDXYf1DYf1M by the associativity of tensor product.

  • Proof.

    DXYf1DYf1M=(OXf1OYf1DY)f1DYf1M=OXf1OYf1M.

     ◻

Direct Images

Left tensoring DXY=OXf1OYf1DY allows us to pullback DY modules to DX-modules. What about tensoring DXY on the right? Intuitively this could pushforward a right DX module M into a right f1DY-module MDXDXY. Pushing forward again we reach a DY module ModDXModDY:Mf(MDXDXY). But this is not very homologically friendly since it involves both a left exact functor f and a right exact functor . We will give the right answer in terms of derived categories later, here we explain the above process can be used to pushforward left DX-modules as well.

It suffices to use the categorical equivalence of left and right D-modules given by ΩX and ΩY rendering math failed o.o i.e. we map MΩYOYf((ΩXOXM)DXDXY). From the isomorphism (ΩXOXM)DXDXY(ΩXOXDXY)DXM note that the right DX-module structure and right f1DY-module structure on ΩXOXDXY are compatible, the above isomorphism is actually an isomorphism of f1DY-modules.

  • Definition.

    Define the following (f1DY;DX)-bimodule DYX=(ΩXODXY)f1OYf1ΩY. Intuitively, this is just changing the direction of two D-module structures by tensoring with corresponding functors.

This gives a clean definition of the pushforward functor for left DX-modules f:DXModDYMod,Mf(DYXDXM).

Some Categories of D-Modules

Our module DX is locally free and thus quasi-coherent over OX. These quasi-coherent sheaves are fundamental in algebraic geometry, and we shall mainly deal with DX-modules that are quasi-coherent.

  • The category of quasi-coherent OX-modules is denoted by Modqc(OX). For the special case of affine varieties,

    1. the global sections functor Γ(X,):Modqc(OX)Mod(Γ(X,OX)) is exact.

    2. If Γ(X,M)=0 for MModqc(OX), then M=0.

  • The category of quasi-coherent DX-modules is denoted by Modqc(DX). This is an abelian category. We say X is D-affine if

    1. The global sections functor Γ(X,):Modqc(DX)Mod(Γ(X,DX)) is exact.

    2. If Γ(X,M)=0 for MModqc(DX), then M=0.

  • Let X be D-affine, then

    1. MModqc(DX) is generated by global sections.

    2. Γ(X,):Modqc(DX)Mod(Γ(X,DX)) is an equivalence of categories.

Inverse Images and Direct Images Using Derived Categories

We shall define several functors on derived categories of D-modules and study its fundamental properties.

Recall the following fundamental lemma about the category of modules over a sheaves of rings.

  • Lemma. Let O be a sheaf of rings on a topological space X, for the category Mod(O), we have

    1. Any object M can be embedded into an injective object MI.

    2. Any object can be is a quotient of a flat object FM.

This means any object MD+(O) is quasi-isomorphic to a complex of injective modules in D+(O), and any object in D(O) is quasi-isomorphic to a complex of flat modules in D(O).

The usual push-forward (direct image) functor f and its derived functor Rf in sheaf theory can be displayed and understood in the following diagram

rendering math failed o.o

For a morphism of algebraic varieties f:XY and a sheaf of rings RY on Y, Rf sends Db(f1RY)Db(RY) and commutes with direct sums. These come from familiar properties about abelian sheaves (in the case R=ZY which is the second line in the above diagram).

Let X be smooth algebraic variety. As a note for notations, D(DX) is the derived category of sheaves of DX-modules, while Dqc and Dc means the full subcategory of D consisting of complexes with quasi-coherent or coherent cohomology. So that

  • any object of Db(DX) is represented by a bounded complex of flat DX-module,

  • any object of Dqcb(DX) is represented by a bounded complex of locally projective quasi-coherent DX-modules.

Theorem (Equivalence Of Quasi-coherent Cohomology)

Let X be smooth. The inclusion functors Db(Modqc(DX))Dqcb(DX) Db(Modc(DX))Dcb(DX) from the complexes of quasi-coherent objects to the complexes having quasi-coherent cohomologies, are categorical equivalences.

Let f:XY be a morphism of smooth algebraic varieties, we can define the left derived functor Lf using flat resolution, Lf:Db(DY)Db(DX)MDXYf1DYf1M, and it preserves Dqcb(DY)Dqcb(DX). Note that it does not necessarily preserve Dcb to Dcb, since when f is a closed embedding with dimX<dimY, DXY can be of infinite rank.

The functor Lf defined above is called the inverse image functor. In practice the shifted inverse image functor is more useful f:=Lf[dimXdimY]:Db(DY)Db(DX). (you can think of the shift as matching their top cohomology).