Local system

From HandWiki
Short description: Locally constant sheaf of abelian groups on topological space

In mathematics, a local system (or a system of local coefficients) on a topological space X is a tool from algebraic topology which interpolates between cohomology with coefficients in a fixed abelian group A, and general sheaf cohomology in which coefficients vary from point to point. Local coefficient systems were introduced by Norman Steenrod in 1943.[1]

Local systems are the building blocks of more general tools, such as constructible and perverse sheaves.

Definition

Let X be a topological space. A local system (of abelian groups/modules/...) on X is a locally constant sheaf (of abelian groups/modules...) on X. In other words, a sheaf [math]\displaystyle{ \mathcal{L} }[/math] is a local system if every point has an open neighborhood [math]\displaystyle{ U }[/math] such that the restricted sheaf [math]\displaystyle{ \mathcal{L}|_U }[/math] is isomorphic to the sheafification of some constant presheaf.[clarification needed]

Equivalent definitions

Path-connected spaces

If X is path-connected,[clarification needed] a local system [math]\displaystyle{ \mathcal{L} }[/math] of abelian groups has the same stalk L at every point. There is a bijective correspondence between local systems on X and group homomorphisms

[math]\displaystyle{ \rho: \pi_1(X,x) \to \text{Aut}(L) }[/math]

and similarly for local systems of modules. The map [math]\displaystyle{ \pi_1(X,x) \to \text{End}(L) }[/math] giving the local system [math]\displaystyle{ \mathcal{L} }[/math] is called the monodromy representation of [math]\displaystyle{ \mathcal{L} }[/math].

This shows that (for X path-connected) a local system is precisely a sheaf whose pullback to the universal cover of X is a constant sheaf.

This correspondence can be upgraded to an equivalence of categories between the category of local systems of abelian groups on X and the category of abelian groups endowed with an action of [math]\displaystyle{ \pi_1(X,x) }[/math] (equivalently, [math]\displaystyle{ \mathbb{Z}[\pi_1(X,x)] }[/math]-modules).[2]

Stronger definition on non-connected spaces

A stronger nonequivalent definition that works for non-connected X is: the following: a local system is a covariant functor

[math]\displaystyle{ \mathcal{L}\colon \Pi_1(X) \to \textbf{Mod}(R) }[/math]

from the fundamental groupoid of [math]\displaystyle{ X }[/math] to the category of modules over a commutative ring [math]\displaystyle{ R }[/math], where typically [math]\displaystyle{ R = \Q,\R,\Complex }[/math]. This is equivalently the data of an assignment to every point [math]\displaystyle{ x\in X }[/math] a module [math]\displaystyle{ M }[/math] along with a group representation [math]\displaystyle{ \rho_x: \pi_1(X,x) \to \text{Aut}_R(M) }[/math] such that the various [math]\displaystyle{ \rho_x }[/math] are compatible with change of basepoint [math]\displaystyle{ x \to y }[/math] and the induced map [math]\displaystyle{ \pi_1(X, x) \to \pi_1(X, y) }[/math] on fundamental groups.

Examples

  • Constant sheaves such as [math]\displaystyle{ \underline{\Q}_X }[/math]. This is a useful tool for computing cohomology since in good situations, there is an isomorphism between sheaf cohomology and singular cohomology:

[math]\displaystyle{ H^k(X,\underline{\Q}_X) \cong H^k_\text{sing}(X,\Q) }[/math]

  • Let [math]\displaystyle{ X=\R^2 \setminus \{(0,0)\} }[/math]. Since [math]\displaystyle{ \pi_1(\R^2 \setminus \{(0,0)\})=\mathbb{Z} }[/math], there is an [math]\displaystyle{ S^1 }[/math] family of local systems on X corresponding to the maps [math]\displaystyle{ n \mapsto e^{in\theta} }[/math]:

[math]\displaystyle{ \rho_\theta: \pi_1(X; x_0) \cong \Z \to \text{Aut}_\Complex(\Complex) }[/math]

  • Horizontal sections of vector bundles with a flat connection. If [math]\displaystyle{ E\to X }[/math] is a vector bundle with flat connection [math]\displaystyle{ \nabla }[/math], then there is a local system given by [math]\displaystyle{ E^\nabla_U=\left\{\text{sections }s\in \Gamma(U,E) \text{ which are horizontal: }\nabla s=0\right\} }[/math] For instance, take [math]\displaystyle{ X=\Complex \setminus 0 }[/math] and [math]\displaystyle{ E = X \times \Complex.^n }[/math] the trivial bundle. Sections of E are n-tuples of functions on X, so [math]\displaystyle{ \nabla_0(f_1,\dots,f_n)= (df_1,\dots,df_n) }[/math] defines a flat connection on E, as does [math]\displaystyle{ \nabla(f_1,\dots,f_n)= (df_1,\dots,df_n)-\Theta(x)(f_1,\dots,f_n)^t }[/math] for any matrix of one-forms [math]\displaystyle{ \Theta }[/math] on X. The horizontal sections are then

    [math]\displaystyle{ E^\nabla_U= \left\{(f_1,\dots,f_n)\in E_U: (df_1,\dots,df_n)=\Theta (f_1,\dots,f_n)^t\right\} }[/math] i.e., the solutions to the linear differential equation [math]\displaystyle{ df_i = \sum \Theta_{ij} f_j }[/math].

    If [math]\displaystyle{ \Theta }[/math] extends to a one-form on [math]\displaystyle{ \Complex }[/math] the above will also define a local system on [math]\displaystyle{ \Complex }[/math], so will be trivial since [math]\displaystyle{ \pi_1(\Complex) = 0 }[/math]. So to give an interesting example, choose one with a pole at 0:

    [math]\displaystyle{ \Theta= \begin{pmatrix} 0 & dx/x \\ dx & e^x dx \end{pmatrix} }[/math] in which case for [math]\displaystyle{ \nabla= d+ \Theta }[/math], [math]\displaystyle{ E^\nabla_U =\left\{ f_1,f_2: U \to \mathbb{C} \ \ \text{ with } f'_1= f_2/x \ \ f_2'=f_1+ e^x f_2\right\} }[/math]
  • An n-sheeted covering map [math]\displaystyle{ X\to Y }[/math] is a local system with fibers given by the set [math]\displaystyle{ \{1,\dots,n\} }[/math]. Similarly, a fibre bundle with discrete fibre is a local system, because each path lifts uniquely to a given lift of its basepoint. (The definition adjusts to include set-valued local systems in the obvious way).
  • A local system of k-vector spaces on X is equivalent to a k-linear representation of [math]\displaystyle{ \pi_1(X,x) }[/math].
  • If X is a variety, local systems are the same thing as D-modules which are additionally coherent O_X-modules (see O modules).
  • If the connection is not flat (i.e. its curvature is nonzero), then parallel transport of a fibre F_x over x around a contractible loop based at x_0 may give a nontrivial automorphism of F_x, so locally constant sheaves can not necessarily be defined for non-flat connections.

Cohomology

There are several ways to define the cohomology of a local system, called cohomology with local coefficients, which become equivalent under mild assumptions on X.

  • Given a locally constant sheaf [math]\displaystyle{ \mathcal{L} }[/math] of abelian groups on X, we have the sheaf cohomology groups [math]\displaystyle{ H^j(X,\mathcal{L}) }[/math] with coefficients in [math]\displaystyle{ \mathcal{L} }[/math].
  • Given a locally constant sheaf [math]\displaystyle{ \mathcal{L} }[/math] of abelian groups on X, let [math]\displaystyle{ C^n(X;\mathcal{L}) }[/math] be the group of all functions f which map each singular n-simplex [math]\displaystyle{ \sigma\colon\Delta^n\to X }[/math] to a global section [math]\displaystyle{ f(\sigma) }[/math] of the inverse-image sheaf [math]\displaystyle{ \sigma^{-1}\mathcal{L} }[/math]. These groups can be made into a cochain complex with differentials constructed as in usual singular cohomology. Define [math]\displaystyle{ H^j_\mathrm{sing}(X;\mathcal{L}) }[/math] to be the cohomology of this complex.
  • The group [math]\displaystyle{ C_n(\widetilde{X}) }[/math] of singular n-chains on the universal cover of X has an action of [math]\displaystyle{ \pi_1(X,x) }[/math] by deck transformations. Explicitly, a deck transformation [math]\displaystyle{ \gamma\colon\widetilde{X}\to\widetilde{X} }[/math] takes a singular n-simplex [math]\displaystyle{ \sigma\colon\Delta^n\to\widetilde{X} }[/math] to [math]\displaystyle{ \gamma\circ\sigma }[/math]. Then, given an abelian group L equipped with an action of [math]\displaystyle{ \pi_1(X,x) }[/math], one can form a cochain complex from the groups [math]\displaystyle{ \operatorname{Hom}_{\pi_1(X,x)}(C_n(\widetilde{X}),L) }[/math] of [math]\displaystyle{ \pi_1(X,x) }[/math]-equivariant homomorphisms as above. Define [math]\displaystyle{ H^j_\mathrm{sing}(X;L) }[/math] to be the cohomology of this complex.

If X is paracompact and locally contractible, then [math]\displaystyle{ H^j(X,\mathcal{L})\cong H^j_\mathrm{sing}(X;\mathcal{L}) }[/math].[3] If [math]\displaystyle{ \mathcal{L} }[/math] is the local system corresponding to L, then there is an identification [math]\displaystyle{ C^n(X;\mathcal{L})\cong\operatorname{Hom}_{\pi_1(X,x)}(C_n(\widetilde{X}),L) }[/math] compatible with the differentials,[4] so [math]\displaystyle{ H^j_\mathrm{sing}(X;\mathcal{L})\cong H^j_\mathrm{sing}(X;L) }[/math].

Generalization

Local systems have a mild generalization to constructible sheaves -- a constructible sheaf on a locally path connected topological space [math]\displaystyle{ X }[/math] is a sheaf [math]\displaystyle{ \mathcal{L} }[/math] such that there exists a stratification of

[math]\displaystyle{ X = \coprod X_\lambda }[/math]

where [math]\displaystyle{ \mathcal{L}|_{X_\lambda} }[/math] is a local system. These are typically found by taking the cohomology of the derived pushforward for some continuous map [math]\displaystyle{ f:X \to Y }[/math]. For example, if we look at the complex points of the morphism

[math]\displaystyle{ f:X = \text{Proj}\left(\frac{\Complex[s,t][x,y,z]}{(stf(x,y,z))}\right) \to \text{Spec}(\Complex[s,t]) }[/math]

then the fibers over

[math]\displaystyle{ \mathbb{A}^2_{s,t} - \mathbb{V}(st) }[/math]

are the smooth plane curve given by [math]\displaystyle{ f }[/math], but the fibers over [math]\displaystyle{ \mathbb{V} }[/math] are [math]\displaystyle{ \mathbb{P}^2 }[/math]. If we take the derived pushforward [math]\displaystyle{ \mathbf{R}f_!(\underline{\Q}_X) }[/math] then we get a constructible sheaf. Over [math]\displaystyle{ \mathbb{V} }[/math] we have the local systems

[math]\displaystyle{ \begin{align} \mathbf{R}^0f_!(\underline{\mathbb{Q}}_X)|_{\mathbb{V}(st)} &= \underline{\Q}_{\mathbb{V}(st)} \\ \mathbf{R}^2f_!(\underline{\mathbb{Q}}_X)|_{\mathbb{V}(st)} &= \underline{\Q}_{\mathbb{V}(st)} \\ \mathbf{R}^4f_!(\underline{\mathbb{Q}}_X)|_{\mathbb{V}(st)} &= \underline{\Q}_{\mathbb{V}(st)} \\ \mathbf{R}^kf_!(\underline{\mathbb{Q}}_X)|_{\mathbb{V}(st)} &= \underline{0}_{\mathbb{V}(st)} \text{ otherwise} \end{align} }[/math]

while over [math]\displaystyle{ \mathbb{A}^2_{s,t} - \mathbb{V}(st) }[/math] we have the local systems

[math]\displaystyle{ \begin{align} \mathbf{R}^0f_!(\underline{\Q}_X)|_{\mathbb{A}^2_{s,t} - \mathbb{V}(st)} &= \underline{\Q}_{\mathbb{A}^2_{s,t} - \mathbb{V}(st)} \\ \mathbf{R}^1f_!(\underline{\Q}_X)|_{\mathbb{A}^2_{s,t} - \mathbb{V}(st)} &= \underline{\Q}_{\mathbb{A}^2_{s,t} - \mathbb{V}(st)}^{\oplus 2g} \\ \mathbf{R}^2f_!(\underline{\Q}_X)|_{\mathbb{A}^2_{s,t} - \mathbb{V}(st)} &= \underline{\Q}_{\mathbb{A}^2_{s,t} - \mathbb{V}(st)} \\ \mathbf{R}^kf_!(\underline{\Q}_X)|_{\mathbb{A}^2_{s,t} - \mathbb{V}(st)} &= \underline{0}_{\mathbb{A}^2_{s,t} - \mathbb{V}(st)} \text{ otherwise} \end{align} }[/math]

where [math]\displaystyle{ g }[/math] is the genus of the plane curve (which is [math]\displaystyle{ g = (\deg(f) - 1)(\deg(f) - 2)/2 }[/math]).

Applications

The cohomology with local coefficients in the module corresponding to the orientation covering can be used to formulate Poincaré duality for non-orientable manifolds: see Twisted Poincaré duality.

See also

References

  1. Steenrod, Norman E. (1943). "Homology with local coefficients". Annals of Mathematics 44 (4): 610–627. doi:10.2307/1969099. 
  2. Milne, James S. (2017). Introduction to Shimura Varieties. Proposition 14.7.
  3. Bredon, Glen E. (1997). Sheaf Theory, Second Edition, Graduate Texts in Mathematics, vol. 25, Springer-Verlag. Chapter III, Theorem 1.1.
  4. Hatcher, Allen (2001). Algebraic Topology, Cambridge University Press . Section 3.H.

External links