The Cech Expansion: nerves of open covers

The Whitehead theorem in homotopy theory basically says that to fully understand the homotopy type of a CW-complex one only needs to know about the homotopy groups (really, the weak homotopy type). It is very easy to produce spaces to which Whitehead’s theorem doesn’t apply. For instance, consider an infinite wedge X=\bigvee_{n\in\mathbb{N}}S^1 of circles and the one-dimensional space Y pictured below:

The space Y: a wedge of converging circles

There is clearly a continuous bijection X\to Y, which induces an isomorphism on all homotopy groups. However, X and Y are not homotopy equivalent because the topological fundamental group of X, which is a homotopy invariant, is discrete while the topological fundamental group of Y is not discrete.

Heavier machinery is required to differentiate spaces which are more complicated on a local level than CW-complexes. One of the most traditional approaches to this problem is Shape Theory, initially developed by Carl Borsuk and refined and expanded in the 1970’s by many, including the authors of [1].

A common technique in mathematics is to approximate complicated objects by simpler ones. For instance, approximating functions in Calculus by Taylor polynomials of increasing degree. This is basically the approach of shape theory: approximate complicated spaces by simpler ones, in particular polyhedra (spaces built out of line segments, triangles, tetrahedra, etc.). Borsuk, the inventor of Shape Theory, first used ANR’s to study the topology of compact metric spaces. The more modern approach pioneered by Segal and Mardesic [1] is categorical in nature and makes heavy use of inverse systems of polyhedra.

Though Shape Theory helps a great deal in our understanding of complicated spaces; however, it has its limits (pun intended). For instance Shape Theory tells you absolutely nothing about the Griffiths Twin Cone because it is “shape equivalent” to a point!

So where do we start?

Let’s look at the Cech Expansion of a space X. This is supposed to let us approximate spaces by simpler ones. We’ll start with an open cover \mathscr{U} of our space X. Even when you forget about the points in each set U\in\mathscr{U}, the open cover still gives a vague picture of X. For instance, take the following cover of the unit circle S^1 as a subspace of the plane.

Now forget about the points in the space.

The shape left still closely resembles that of a circle. We can even recover the circle from this data: replace each open set with a point. If two open sets intersect, draw a line segment between the two corresponding points.

We get a polyhedron homeomorphic to the circle. It isn’t usually true that we will get back the original space; this only happens in very special cases and for very fine open covers.

The nerve of an open cover

Definition: An abstract simplicial complex is a set S and a set K consisting of finite subsets of S such that if A\in K and B\subset A, then B\in K. A vertex or 0-simplex is a singleton \{s\}\in K and an n-simplex is a set \{s_1,...,s_{n+1}\}\in K containing n+1 elements. The n-skeleton of K is the set K_n=\{A\in K||A|=n+1\}

Now if \mathscr{U} is an open cover of X, we construct an abstract simplicial complex N(\mathscr{U}) called the nerve of \mathscr{U}. An element of \mathscr{U} is a finite set \{U_1,U_2,...,U_n\}\subset \mathscr{U} such that \bigcap_{i=1}^{n}U_i\neq \emptyset . The geometric realization |N(\mathscr{U})| is a geometric complex obtained by pasting simplices together using N(\mathscr{U}) as instructions.

Definition: The geometric realization of an abstract simplicial complex K with vertex set K_0 is the topological space |K| defined as a subset of the product P=[0,1]^{K_0} of functions f:K_0\to [0,1]. In particular, |K| is the set of functions f\in P such that

  1. \{v\in K_0|f(v)>0\}\in K  (in particular all but finitely many f(v) are zero),
  2. \sum_{v} f(v)=1.

Give  P the weak (or induced) topology so that U is open in P if and only if U\cap [0,1]^{F} is open in [0,1]^{F} for all finite sets F\subset K_0. |K| is topologized with the subspace topology of P.

It is common to write the simplex in |K| spanned by vertices s_1,...,s_n in |K| as [s_1,...,s_n]=\{f\in |K||f(s_i)>0\text{ for some }1\leq i\leq n\}.

While |N(\mathscr{U})| is defined as the geometric realization of the nerve, it is a bit more intuitive to think of it in the following way.

Here is the cover:

An open cover

0 skeleton – A vertex of N(\mathscr{U}) is a set U\in \mathscr{U}

0-skeleton

1 skeleton – If U_1\cap U_2 \neq \emptyset , then place an edge (1-simplex) between U_1 and U_2.

1-skeleton

2 skeleton – If U_1\cap U_2\cap U_3\neq \emptyset, then there are three edges joining each pair of the vertices U_1,U_2,U_3. Place a triangle (or 2-simplex) so that the edges of the triangle match up with these three edges. In the picture, fill in the each empty triangle with a triangle.

3 skeleton – If U_1\cap U_2\cap U_3\cap U_4\neq \emptyset, attach a tetrahedron to fill in the boundary that exists from the four triangles.

tetrahedron

Input the tetrahedron into the place where it obviously goes (on the right).

In our example, we stop here and leave it embedded in \mathbb{R}^3. In general, you would continue to add higher dimensional simplices and give the resulting geometric simplicial complex the weak topology. In addition, the space might not be compact and open covers would typically contain infinitely many sets.

Here’s an observation we can go ahead and make. The order of an open cover \mathscr{U} of X (if it exists) is the smallest number n\in\mathbb{N} such that every point x\in X lies in n elements of \mathscr{U}. Hence, if the order of \mathscr{U} is n, then at most n distinct elements can intersect and N(\mathscr{U}) can only have simplices of dimension \leq n.

Proposition: If the order of \mathscr{U} is n, then dim(N(\mathscr{U}))=n.

Refinements:

The nerve |N(\mathscr{U})| is supposed to be an “approximation” of the original space X. What if it is a bad approximation? You should take a “closer look” at X by using a finer open cover of X, i.e. one consisting of smaller open sets.

Definition: An open cover \mathscr{V} is a refinement of another cover \mathscr{U} if for each V\in \mathscr{V} there is a U\in \mathscr{U} such that V\subseteq U.

If \mathscr{V} refines \mathscr{U}, then |N(\mathscr{V})| is “larger” than |N(\mathscr{U})| since there are more sets in \mathscr{V}. It makes sense to think of |N(\mathscr{V})| as being a better approximation to X since if we collapse the appropriate simplices of |N(\mathscr{V})|, we get back something homotopy equivalent to |N(\mathscr{U})|. This is captured in the next proposition.

Proposition: If  \mathscr{V} is a refinement of another cover \mathscr{U}, there is a there is an onto simplicial map p_{\mathscr{V}\mathscr{U}}:N(\mathscr{V})\to N(\mathscr{U}). The map |p_{\mathscr{V}\mathscr{U}}|:|N(\mathscr{V})|\to |N(\mathscr{U})| induced on geometric realizations is unique up to homotopy.

Proof. First define p_{\mathscr{V}\mathscr{U}} on vertices (i.e. elements of \mathscr{V}). If V\in \mathscr{V} and V\subseteq U_{V} for U_{V}\in\mathscr{U}, define p_{\mathscr{V}\mathscr{U}}(V)=U_{V}. If V\cap V'\neq \emptyset, then clearly U_{V}\cap U_{V'}\neq \emptyset so we define p_{\mathscr{V}\mathscr{U}} on the 1-simplex [V,V'] spanned by V and V' to the 1-simplex \left[U_{V},U_{V'}\right] spanned by U_{V} and U_{V'}. Any map defined in this way is called a projection.

The same goes for higher simplices; if \bigcap_{i=1}^{n}V_i\neq \emptyset, then \bigcap_{i=1}^{n}U_{V_{i}}\neq \emptyset and we send the simplex [V_1,\dots,V_n] to [U_{V_{1}},...,U_{V_{n}}]. This gives a well-defined simplicial map on the nerves.

Though it seems like there is a lot of freedom in defining a projection, the choice is independent of homotopy. By construction, any two projections N(\mathscr{V})\to N(\mathscr{U}) induce contiguous maps on geometric realizations. But contiguous maps of simplicial complexes are homotopic, proving the proposition. \square

Note that if \mathscr{W} is a refinement of \mathscr{V} and \mathscr{V} is a refinement of \mathscr{U}, the composition p_{\mathscr{V}\mathscr{U}} \circ p_{\mathscr{W}\mathscr{V}} is a canonical map. Thus if p_{\mathscr{W}\mathscr{U}} is another projection, then there is a homotopy \left|p_{\mathscr{W}\mathscr{U}}\right| \simeq |p_{\mathscr{V}\mathscr{U}}| \circ |p_{\mathscr{W}\mathscr{V}}|. Now if we let [p_{\mathscr{V}\mathscr{U}}] denote the homotopy class of |p_{\mathscr{V}\mathscr{U}}|, we have strict equality [p_{\mathscr{W}\mathscr{U}}] = [p_{\mathscr{V}\mathscr{U}}] \circ [p_{\mathscr{W}\mathscr{V}}]. Therefore, since open covers of X form a directed set \mathcal{O}(X), we have an inverse system \left(|N(\mathscr{U})|,[p_{\mathscr{V}\mathscr{U}}],\mathcal{O}(X)\right) of homotopy classes of nerves of covers.

Definition: For a paracompact Hausdorff space X, the Cech Expansion of X is the inverse system \left(|N(\mathscr{U})|,[p_{\mathscr{V}\mathscr{U}}],\mathcal{O}(X)\right) in the homotopy category of polyhedra.

Of course, even if X is not paracompact Hausdorff, you still get an inverse system; the problem with non-paracompact spaces is that it is much harder to relate X to the inverse system without “enough” partitions of unity to build maps X\to |N(\mathscr{U})|.

References:

[1] S. Mardsic and J. Segal, Shape theory, North-Holland Publishing Company, 1982.

This entry was posted in Cech expansion, Shape theory, Simplicial complexes and tagged , , , , , . Bookmark the permalink.

4 Responses to The Cech Expansion: nerves of open covers

  1. tghyde says:

    In your definition of the geometric realization, you list two criterion defining the space P. For the second criterion, do you mean the sum is > 1 or did you mean = 1?

    Like

  2. Mathieu says:

    Is there a theorem relating the Cech expansion of X to X?

    Like

    • Yes, when X is a paracompact Hausdorff space. Given a cover U of X, you choose a partition of unity subordinate to U and construct a map X to |N(U)|, that is uniquely determined by the partition of unity. This map is called a “canonical map” in most shape theory books. What’s really interesting and useful is that canonical maps are unique up to homotopy, that is, given any two partitions of unity subordinate to U, the two resulting canonical maps will be homotopic to each other. A nice detailed description of this stuff is in the Appendix to Mardesic & Segal’s Shape Theory book.

      Like

Leave a Reply

Fill in your details below or click an icon to log in:

WordPress.com Logo

You are commenting using your WordPress.com account. Log Out /  Change )

Facebook photo

You are commenting using your Facebook account. Log Out /  Change )

Connecting to %s