Local system

{{Short description|Locally constant sheaf of abelian groups on topological space}}

{{More citations needed|date=January 2021}}

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.{{cite journal | last=Steenrod | first=Norman E. | author-link=Norman Steenrod| title=Homology with local coefficients | journal=Annals of Mathematics | volume=44 | issue=4 | year=1943 | doi=10.2307/1969099 | pages=610–627 | jstor=1969099 | mr=9114}}

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/of modules...) on X. In other words, a sheaf \mathcal{L} is a local system if every point has an open neighborhood U such that the restricted sheaf \mathcal{L}|_U is isomorphic to the sheafification of some constant presheaf. {{clarify| there is some ambiguity of constant sheaf, some uses the definition that restriction is identity; others uses the convention that isomorphic to such a sheaf suffices.|date=September 2022}}

=Equivalent definitions=

==Path-connected spaces==

If X is path-connected,{{clarify|Don't you also need "locally simply connected?"|date=August 2022}} a local system \mathcal{L} of abelian groups has the same stalk L at every point. There is a bijective correspondence between local systems on X and group homomorphisms

: \rho: \pi_1(X,x) \to \text{Aut}(L)

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

{{math proof|title=Proof of equivalence|proof=Take local system \mathcal{L} and a loop \gamma at x. It's easy to show that any local system on [0,1] is constant. For instance, \gamma^* \mathcal{L} is constant. This gives an isomorphism (\gamma^*\mathcal{L})_0\simeq \Gamma([0,1], \mathcal{L}) \simeq (\gamma^*\mathcal{L})_1 , i.e. between L and itself.

Conversely, given a homomorphism \rho: \pi_1(X,x)\to \text{Aut}(L), consider the constant sheaf \underline{L} on the universal cover \widetilde{X} of X. The deck-transform-invariant sections of \underline{L} gives a local system on X. Similarly, the deck-transform-ρ-equivariant sections give another local system on X: for a small enough open set U, it is defined as

: \mathcal{L}(\rho)_U\ = \ \left\{ \text{sections }s \in \underline{L}_{\pi^{-1}(U)} \text{ with }\theta\circ s=\rho(\theta) s \text{ for all }\theta \in\text{ Deck}(\widetilde{X}/X)=\pi_1(X,x) \right\}

where \pi:\widetilde{X}\to X is the universal covering.

}}

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 \pi_1(X,x) (equivalently, \mathbb{Z}[\pi_1(X,x)]-modules). Milne, James S. (2017). [https://www.jmilne.org/math/xnotes/svi.pdf Introduction to Shimura Varieties]. Proposition 14.7.

==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

: \mathcal{L}\colon \Pi_1(X) \to \textbf{Mod}(R)

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

Examples

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

H^k(X,\underline{\Q}_X) \cong H^k_\text{sing}(X,\Q)

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

\rho_\theta: \pi_1(X; x_0) \cong \Z \to \text{Aut}_\Complex(\Complex)

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

    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\} i.e., the solutions to the linear differential equation df_i = \sum \Theta_{ij} f_j.

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

    \Theta= \begin{pmatrix} 0 & dx/x \\ dx & e^x dx \end{pmatrix} in which case for \nabla= d+ \Theta , 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\}
  • An n-sheeted covering map X\to Y is a local system with fibers given by the set \{1,\dots,n\} . 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 \pi_1(X,x).
  • 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 \mathcal{L} of abelian groups on X, we have the sheaf cohomology groups H^j(X,\mathcal{L}) with coefficients in \mathcal{L}.
  • Given a locally constant sheaf \mathcal{L} of abelian groups on X, let C^n(X;\mathcal{L}) be the group of all functions f which map each singular n-simplex \sigma\colon\Delta^n\to X to a global section f(\sigma) of the inverse-image sheaf \sigma^{-1}\mathcal{L}. These groups can be made into a cochain complex with differentials constructed as in usual singular cohomology. Define H^j_\mathrm{sing}(X;\mathcal{L}) to be the cohomology of this complex.
  • The group C_n(\widetilde{X}) of singular n-chains on the universal cover of X has an action of \pi_1(X,x) by deck transformations. Explicitly, a deck transformation \gamma\colon\widetilde{X}\to\widetilde{X} takes a singular n-simplex \sigma\colon\Delta^n\to\widetilde{X} to \gamma\circ\sigma. Then, given an abelian group L equipped with an action of \pi_1(X,x), one can form a cochain complex from the groups \operatorname{Hom}_{\pi_1(X,x)}(C_n(\widetilde{X}),L) of \pi_1(X,x)-equivariant homomorphisms as above. Define H^j_\mathrm{sing}(X;L) to be the cohomology of this complex.

If X is paracompact and locally contractible, then H^j(X,\mathcal{L})\cong H^j_\mathrm{sing}(X;\mathcal{L}). Bredon, Glen E. (1997). Sheaf Theory, Second Edition, Graduate Texts in Mathematics, vol. 25, Springer-Verlag. Chapter III, Theorem 1.1. If \mathcal{L} is the local system corresponding to L, then there is an identification C^n(X;\mathcal{L})\cong\operatorname{Hom}_{\pi_1(X,x)}(C_n(\widetilde{X}),L) compatible with the differentials, Hatcher, Allen (2001). Algebraic Topology, Cambridge University Press. Section 3.H. so H^j_\mathrm{sing}(X;\mathcal{L})\cong H^j_\mathrm{sing}(X;L).

Generalization

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

:X = \coprod X_\lambda

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

:f:X = \text{Proj}\left(\frac{\Complex[s,t][x,y,z]}{(st\cdot h(x,y,z))}\right) \to \text{Spec}(\Complex[s,t])

then the fibers over

:\mathbb{A}^2_{s,t} - \mathbb{V}(st)

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

:

\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}

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

:\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}

where g is the genus of the plane curve (which is g = (\deg(f) - 1)(\deg(f) - 2)/2).

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

{{reflist}}