nLab classifying space

Contents

Context

Yoneda lemma

Geometry

Contents

Idea

A classifying space for some sort of data refers to a space (or a more general object), usually written (data)\mathcal{B}(data), such that maps X(data)X\to \mathcal{B}(data) correspond to data over XX.

The classical example is the classifying space G\mathcal{B}G of a group GG, which has the property that there is a bijection between homotopy classes of maps XGX\to \mathcal{B}G and isomorphism classes of GG-bundles over XX. (In fact, one can jack this up to an equivalence of groupoids or \infty-groupoids.) Various improvements of this are possible which classify bundles with extra structure or fibrations.

Categorically, the corresponding statement is that Grothendieck fibrations over a category XX correspond to pseudofunctors X opCatX^{op}\to Cat. Thus CatCat is the “classifying space for categories.” Similarly, discrete fibrations over XX correspond to functors X opSetX^{op}\to Set.

To see the connection between the two, consider the case when XX is a groupoid and we restrict the fibers of the fibration to be isomorphic to a given set FF. Then the functor X opSetX^{op}\to Set must land in the subcategory of SetSet consisting of just the automorphisms of FF, which is the one-object groupoid corresponding to the automorphism group Aut(F)Aut(F). If we further restrict the automorphisms appearing to preserve some given structure on FF, so that they lie in some smaller group GG, then the “classifying space” will be the one-object groupoid corresponding to GG. Under the homotopy hypothesis, groupoids correspond to homotopy 1-types, and the one-object groupoid of a group GG corresponds precisely to the usual topological classifying space G\mathcal{B}G (in fact, this is one construction of G\mathcal{B}G). For this reason, BG\mathbf{B}G is often used to denote that one-object groupoid; see the delooping hypothesis and the discussion at category algebra.

The phrase “classifying space” is also sometimes used for the realization of the nerve of any category, although it is more complicated to say what exactly this space “classifies.” (One answer is “torsors modulo concordance.”)

Examples

For principal bundles

A good account is in Rudolf & Schmidt 17, Sec. 3, esp. Thm. 3.5.1 & Prop. 3.6.2.

For GG a topological group there is a classifying space BGB G \in Top for topological GG-principal bundles, hence a space such that for XX any sufficiently nice topological space there is a natural isomorphism

GBund(X) 0π 0Top(X,BG) G Bund(X)_0 \simeq \pi_0 Top(X, B G)

between the set of isomorphism classes of GG-principal bundles on XX and the set of homotopy-classes of continuous functions XBGX \to B G.

This space may be constructed as follows:

write BGTop Δ op\mathbf{B}G \in Top^{\Delta^{op}} for the simplicial topological space obtained as the nerve of the one-object topological groupoid associated to GG, the simplicial space given by

(BG) n=G ×n (\mathbf{B}G)_n = G^{\times n}

whose face maps are induced by the product operation on GG and whose degeneracy maps are induced from the unit map.

If GG is well-pointed, then the geometric realization of simplicial topological spaces of GG is a model for the homotopy type of the classifying space

BG|BG|. B G \simeq \vert \mathbf{B}G\vert \,.

For more details on this construction see the section classifying spaces at geometric realization of simplicial topological spaces.

As discussed there, too, this construction generalizes to more general simplicial topological groups and classifying spaces for their principal ∞-bundles.


Theorem

If GGrp(HausSp)G \,\in\, Grp(HausSp) is a Hausdorff topological group, then the Milnor join construction classifies topological GG-principal bundles over all paracompact Hausdorff spaces.

(due to Milnor, laid out in Rudolf & Schmidt 17, Thm. 3.5.1)

For orthogonal and unitary principal bundles

Idea

For G=O(n)G = O(n) the orthogonal group and G=U(n)G = U(n) the unitary group, there are standard realizations of the corresponding classifying spaces as direct limits of Grassmannian spaces.

Let V n( q)V_n(\mathbb{R}^q) be the Stiefel manifold of orthonormal nn-frames in the Cartesian space q\mathbb{R}^q. Its points are nn-tuples of orthonormal vectors in q\mathbb{R}^q, and it is topologized as a subspace of ( q) n(\mathbb{R}^q)^n, or, equivalently, as a subspace of (S q1) n(S^{q-1})^n. It is a compact manifold.

Let G n( q)G_n(\mathbb{R}^q) be the Grassmannian of nn-planes in q\mathbb{R}^q. Its points are the n-dimensional subspaces of q\mathbb{R}^q. Sending an nn-tuple of orthonormal vectors to the nn-plane they span gives a surjective function V n( q)G n( q)V_n(\mathbb{R}^q) \to G_n(\mathbb{R}^q), and we topologize G n( q)G_n(\mathbb{R}^q) as a quotient space of V n( q)V_n(\mathbb{R}^q). It too is a compact manifold.

The standard inclusion of q\mathbb{R}^q in q+1\mathbb{R}^{q+1} induces inclusions V n( q)V n( q+1)V_n(\mathbb{R}^q) \hookrightarrow V_n(\mathbb{R}^{q+1}) and G n( q)G n( q+1)G_n(\mathbb{R}^q) \hookrightarrow G_n(\mathbb{R}^{q+1}). We define V n( )V_n(\mathbb{R}^\infty) and G n( )G_n(\mathbb{R}^\infty) to be the unions of the V n( q)V_n(\mathbb{R}^q) and G n( q)G_n(\mathbb{R}^q), with the topology of the union.

Then G n( )G_n(\mathbb{R}^\infty) is a model for the classifying space B O ( n ) B O(n) .

Definitions

In the following we take Top to denote compactly generated topological spaces. For these the Cartesian product X×()X \times (-) is a left adjoint and hence preserves colimits.

Definition

For n,kn, k \in \mathbb{N} and nkn \leq k, then the nnth real Stiefel manifold of k\mathbb{R}^k is the coset topological space.

V n( k)O(k)/O(kn), V_n(\mathbb{R}^k) \coloneqq O(k)/O(k-n) \,,

where the action of O(kn)O(k-n) is via its canonical embedding O(kn)O(k)O(k-n)\hookrightarrow O(k).

Similarly the nnth complex Stiefel manifold of k\mathbb{C}^k is

V n( k)U(k)/U(kn), V_n(\mathbb{C}^k) \coloneqq U(k)/U(k-n) \,,

here the action of U(kn)U(k-n) is via its canonical embedding U(kn)U(k)U(k-n)\hookrightarrow U(k).

Definition

For n,kn, k \in \mathbb{N} and nkn \leq k, then the nnth real Grassmannian of k\mathbb{R}^k is the coset topological space.

Gr n( k)O(k)/(O(n)×O(kn)), Gr_n(\mathbb{R}^k) \coloneqq O(k)/(O(n) \times O(k-n)) \,,

where the action of the product group is via its canonical embedding O(n)×O(kn)O(n)O(n)\times O(k-n) \hookrightarrow O(n) into the orthogonal group.

Similarly the nnth complex Grassmannian of k\mathbb{C}^k is the coset topological space.

Gr n( k)U(k)/(U(n)×U(kn)), Gr_n(\mathbb{C}^k) \coloneqq U(k)/(U(n) \times U(k-n)) \,,

where the action of the product group is via its canonical embedding U(n)×U(kn)U(n)U(n)\times U(k-n) \hookrightarrow U(n) into the unitary group.

Example
Proposition

For all nkn \leq k \in \mathbb{N}, the canonical projection from the real Stiefel manifold (def. ) to the Grassmannian is a O(n)O(n)-principal bundle

O(n) V n( k) Gr n( k) \array{ O(n) &\hookrightarrow& V_n(\mathbb{R}^k) \\ && \downarrow \\ && Gr_n(\mathbb{R}^k) }

and the projection from the complex Stiefel manifold to the Grassmannian us a U(n)U(n)-principal bundle:

U(n) V n( k) Gr n( k). \array{ U(n) &\hookrightarrow& V_n(\mathbb{C}^k) \\ && \downarrow \\ && Gr_n(\mathbb{C}^k) } \,.
Proof

By (this cor. and this prop.).

Definition

By def. there are canonical inclusions

Gr n( k)Gr n( k+1) Gr_n(\mathbb{R}^k) \hookrightarrow Gr_n(\mathbb{R}^{k+1})

and

Gr n( k)Gr n( k+1) Gr_n(\mathbb{C}^k) \hookrightarrow Gr_n(\mathbb{C}^{k+1})

for all kk \in \mathbb{N}. The colimit (in Top, see there) over these inclusions is denoted

BO(n)lim kGr n( k) B O(n) \coloneqq \underset{\longrightarrow}{\lim}_k Gr_n(\mathbb{R}^k)

and

BU(n)lim kGr n( k), B U(n) \coloneqq \underset{\longrightarrow}{\lim}_k Gr_n(\mathbb{C}^k) \,,

respectively.

Moreover, by def. there are canonical inclusions

V n( k)V n( k+1) V_n(\mathbb{R}^k) \hookrightarrow V_n(\mathbb{R}^{k+1})

and

V n( k)V n( k+1), V_n(\mathbb{C}^k) \hookrightarrow V_n(\mathbb{C}^{k+1}) \,,

respectively, that are compatible with the O(n)O(n)-action and the U(n)U(n)-action, respectively. The colimit (in Top, see there) over these inclusions, regarded as equipped with the induced action, is denoted

EO(n)lim kV n( k) E O(n) \coloneqq \underset{\longrightarrow}{\lim}_k V_n(\mathbb{R}^k)

and

EU(n)lim kV n( k), E U(n) \coloneqq \underset{\longrightarrow}{\lim}_k V_n(\mathbb{C}^k) \,,

respectively. The inclusions are in fact compatible with the bundle structure from prop. , so that there are induced projections

(EO(n) BO(n))lim k(V n( k) Gr n( k)) \left( \array{ E O(n) \\ \downarrow \\ B O(n) } \right) \;\; \simeq \;\; \underset{\longrightarrow}{\lim}_k \left( \array{ V_n(\mathbb{R}^k) \\ \downarrow \\ Gr_n(\mathbb{R}^k) } \right)

and

(EU(n) BU(n))lim k(V n( k) Gr n( k)), \left( \array{ E U(n) \\ \downarrow \\ B U(n) } \right) \;\; \simeq \;\; \underset{\longrightarrow}{\lim}_k \left( \array{ V_n(\mathbb{C}^k) \\ \downarrow \\ Gr_n(\mathbb{C}^k) } \right) \,,

respectively. These are the standard models for the universal principal bundles for OO and UU, respectively. The corresponding associated vector bundles

EO(n)×O(n) n E O(n) \underset{O(n)}{\times} \mathbb{R}^n

and

EU(n)×U(n) n E U(n) \underset{U(n)}{\times} \mathbb{C}^n

are the corresponding universal vector bundles.

Since the Cartesian product O(n)×()O(n)\times (-) in compactly generated topological spaces preserves colimits, it follows that the colimiting bundle is still an O(n)O(n)-principal bundle

(EO(n))/O(n) (lim kV n( k))/O(n) lim k(V n( k)/O(n)) lim kGr n( k) BO(n) \begin{aligned} (E O(n))/O(n) & \simeq (\underset{\longrightarrow}{\lim}_k V_{n}(\mathbb{R}^k))/O(n) \\ & \simeq \underset{\longrightarrow}{\lim}_k (V_n(\mathbb{R}^k)/O(n)) \\ & \simeq \underset{\longrightarrow}{\lim}_k Gr_n(\mathbb{R}^k) \\ & \simeq B O(n) \end{aligned}

and anlogously for EU(n)E U(n).

As such this is the standard presentation for the O(n)O(n)-universal principal bundle. Its base space B O ( n ) B O(n) is the corresponding classifying space.

Definition

There are canonical inclusions

Gr n( k)Gr n+1( k+1) Gr_n(\mathbb{R}^k) \hookrightarrow Gr_{n+1}(\mathbb{R}^{k+1})

and

Gr n( k)Gr n+1( k+1) Gr_n(\mathbb{C}^k) \hookrightarrow Gr_{n+1}(\mathbb{C}^{k+1})

given by adjoining one coordinate to the ambient space and to any subspace. Under the colimit of def. these induce maps of classifying spaces

BO(n)BO(n+1) B O(n) \longrightarrow B O(n+1)

and

BU(n)BU(n+1). B U(n) \longrightarrow B U(n+1) \,.
Definition

There are canonical maps

Gr n 1( k 1)×Gr n 2( k 2)Gr n 1+n 2( k 1+k 2) Gr_{n_1}(\mathbb{R}^{k_1}) \times Gr_{n_2}(\mathbb{R}^{k_2}) \longrightarrow Gr_{n_1 + n_2}(\mathbb{R}^{k_1 + k_2})

and

Gr n 1( k 1)×Gr n 2( k 2)Gr n 1+n 2( k 1+k 2) Gr_{n_1}(\mathbb{C}^{k_1}) \times Gr_{n_2}(\mathbb{C}^{k_2}) \longrightarrow Gr_{n_1 + n_2}(\mathbb{C}^{k_1 + k_2})

given by sending ambient spaces and subspaces to their direct sum.

Under the colimit of def. these induce maps of classifying spaces

BO(n 1)×BO(n 2)BO(n 1+n 2) B O(n_1) \times B O(n_2) \longrightarrow B O(n_1 + n_2)

and

BU(n 1)×BU(n 2)BU(n 1+n 2) B U(n_1) \times B U(n_2) \longrightarrow B U(n_1 + n_2)

Properties

Proposition

The real Grassmannians Gr n( k)Gr_n(\mathbb{R}^k) and the complex Grassmannians Gr n( k)Gr_n(\mathbb{C}^k) of def. admit the structure of CW-complexes. Moreover the canonical inclusions

Gr n( k)Gr n( k+1) Gr_n(\mathbb{R}^k) \hookrightarrow Gr_n(\mathbb{R}^{k+1})

and

Gr n( k)Gr n( k+1) Gr_n(\mathbb{C}^k) \hookrightarrow Gr_n(\mathbb{C}^{k+1})

are subcomplex incusions (hence relative cell complex inclusions).

Accordingly there is an induced CW-complex structure on the classifying spaces B O ( n ) B O(n) and B U ( n ) B U(n) (def. ).

A general proof is spelled out in Hatcher, section 1.2 (pages 31-34). For the case of real- , complex- and quaternionic vector bundles see at cell structure on projective space.

Proposition

The Stiefel manifold V n( k)V_n(\mathbb{R}^k) from def. admits the structure of a CW-complex.

e.g. (James 59, p. 3, James 76, p. 5 with p. 21, Blaszczyk 07)

(And I suppose with that cell structure the inclusions V n( k)V n( k+1)V_n(\mathbb{R}^k) \hookrightarrow V_n(\mathbb{R}^{k+1}) are subcomplex inclusions.)

Proposition

The Stiefel manifold V n( k)V_n(\mathbb{R}^k) (def. ) is (k-n-1)-connected.

Proof

Consider the coset quotient projection

O(kn)O(k)O(k)/O(kn)=V n( k). O(k-n) \longrightarrow O(k) \longrightarrow O(k)/O(k-n) = V_n(\mathbb{R}^k) \,.

Since the orthogonal groups is compact (prop.) and by this corollary the projection O(k)O(k)/O(kn)O(k)\to O(k)/O(k-n) is a Serre fibration. Therefore there is induced the long exact sequence of homotopy groups of this fiber sequence, and by this prop. it has the following form in degrees bounded by nn:

π kn1(O(kn))epiπ kn1(O(k))0π kn1(V n( k))0π 1<kn1(O(k))π 1<kn1(O(kn)). \cdots \to \pi_{\bullet \leq k-n-1}(O(k-n)) \overset{epi}{\longrightarrow} \pi_{\bullet \leq k-n-1}(O(k)) \overset{0}{\longrightarrow} \pi_{\bullet \leq k-n-1}(V_n(\mathbb{R}^k)) \overset{0}{\longrightarrow} \pi_{\bullet-1 \lt k-n-1}(O(k)) \overset{\simeq}{\longrightarrow} \pi_{\bullet-1 \lt k-n-1}(O(k-n)) \to \cdots \,.

This implies the claim. (Exactness of the sequence says that every element in π n1(V n( k))\pi_{\bullet \leq n-1}(V_n(\mathbb{R}^k)) is in the kernel of zero, hence in the image of 0, hence is 0 itself.)

Similarly:

Proposition

The complex Stiefel manifold V n( k)V_n(\mathbb{C}^k) (def. ) is 2(k-n)-connected.

Proof

Consider the coset quotient projection

U(kn)U(k)U(k)/U(kn)=V n( k). U(k-n) \longrightarrow U(k) \longrightarrow U(k)/U(k-n) = V_n(\mathbb{C}^k) \,.

By prop. and by this corollary the projection U(k)U(k)/U(kn)U(k)\to U(k)/U(k-n) is a Serre fibration. Therefore there is induced the long exact sequence of homotopy groups of this fiber sequence, and by prop. it has the following form in degrees bounded by nn:

π 2(kn)(U(kn))epiπ 2(kn)(U(k))0π 2(kn)(V n( k))0π 1<2(kn)(U(k))π 1<2(kn)(U(kn)). \cdots \to \pi_{\bullet \leq 2(k-n)}(U(k-n)) \overset{epi}{\longrightarrow} \pi_{\bullet \leq 2(k-n)}(U(k)) \overset{0}{\longrightarrow} \pi_{\bullet \leq 2(k-n)}(V_n(\mathbb{C}^k)) \overset{0}{\longrightarrow} \pi_{\bullet-1 \lt 2(k-n)}(U(k)) \overset{\simeq}{\longrightarrow} \pi_{\bullet-1 \lt 2(k-n)}(U(k-n)) \to \cdots \,.

This implies the claim.

Corollary

The colimiting space EO(n)=lim kV n( k)E O(n) = \underset{\longrightarrow}{\lim}_k V_n(\mathbb{R}^k) from def. is weakly contractible.

The colimiting space EU(n)=lim kV n( k)E U(n) = \underset{\longrightarrow}{\lim}_k V_n(\mathbb{C}^k) from def. is weakly contractible.

Proposition

The homotopy groups of the classifying spaces B O ( n ) B O(n) and B U ( n ) B U(n) (def. ) are those of the orthogonal group O(n)O(n) and of the unitary group U(n)U(n), respectively, shifted up in degree: there are isomorphisms

π +1(BO(n))π O(n) \pi_{\bullet+1}(B O(n)) \simeq \pi_\bullet O(n)

and

π +1(BU(n))π U(n) \pi_{\bullet+1}(B U(n)) \simeq \pi_\bullet U(n)

(for homotopy groups based at the canonical basepoint).

Proof

Consider the sequence

O(n)EO(n)BO(n) O(n) \longrightarrow E O(n) \longrightarrow B O(n)

from def. , with O(n)O(n) the fiber. Since (by this prop.) the second map is a Serre fibration, this is a fiber sequence and so it induces a long exact sequence of homotopy groups of the form

π (O(n))π (EO(n))π (BO(n))π 1(O(n))π 1(EO(n)). \cdots \to \pi_\bullet(O(n)) \longrightarrow \pi_\bullet(E O(n)) \longrightarrow \pi_\bullet(B O(n)) \longrightarrow \pi_{\bullet-1}(O (n)) \longrightarrow \pi_{\bullet-1}(E O(n)) \to \cdots \,.

Since by cor. π (EO(n))=0\pi_\bullet(E O(n))= 0, exactness of the sequence implies that

π (BO(n))π 1(O(n)) \pi_\bullet(B O(n)) \overset{\simeq}{\longrightarrow} \pi_{\bullet-1}(O (n))

is an isomorphism.

The same kind of argument applies to the complex case.

Proposition

For nn \in \mathbb{N} there are homotopy fiber sequences

S nBO(n)BO(n+1) S^n \longrightarrow B O(n) \longrightarrow B O(n+1)

and

S 2n+1BU(n)BU(n+1), S^{2n+1} \longrightarrow B U(n) \longrightarrow B U(n+1) \,,

exhibiting the n-sphere ((2n+1)(2n+1)-sphere) as the homotopy fiber of the canonical maps from def. .

This means that there is a replacement of the canonical inclusion BO(n)BO(n+1)B O(n) \hookrightarrow B O(n+1) (induced via def. ) by a Serre fibration

BO(n) BO(n+1) weakhomotopyequivalence Serrefib. B˜O(n) \array{ B O(n) &\hookrightarrow& B O(n+1) \\ {}^{\mathllap{{weak \, homotopy} \atop equivalence}}\downarrow & \nearrow_{\mathrlap{Serre \, fib.}} \\ \tilde B O(n) }

such that S nS^n is the ordinary fiber of BO(n)B˜O(n+1)B O(n)\to \tilde B O(n+1), and analogously for the complex case.

Proof

Take B˜O(n)(EO(n+1))/O(n)\tilde B O(n) \coloneqq (E O(n+1))/O(n).

To see that the canonical map BO(n)(EO(n+1))/O(n)B O(n)\longrightarrow (E O(n+1))/O(n) is a weak homotopy equivalence consider the commuting diagram

O(n) id O(n) EO(n) EO(n+1) BO(n) (EO(n+1))/O(n). \array{ O(n) &\overset{id}{\longrightarrow}& O(n) \\ \downarrow && \downarrow \\ E O(n) &\longrightarrow& E O(n+1) \\ \downarrow && \downarrow \\ B O(n) &\longrightarrow& (E O(n+1))/O(n) } \,.

By this prop. both bottom vertical maps are Serre fibrations and so both vertical sequences are fiber sequences. By prop. part of the induced morphisms of long exact sequences of homotopy groups looks like this

π (BO(n)) π ((EO(n+1))/O(n)) π 1(O(n)) = π 1(O(n)), \array{ \pi_\bullet(B O(n)) &\overset{}{\longrightarrow}& \pi_\bullet( (E O(n+1))/O(n) ) \\ {}^{\mathllap{\simeq}}\downarrow && \downarrow^{\mathrlap{\simeq}} \\ \pi_{\bullet-1}(O(n)) &\overset{=}{\longrightarrow}& \pi_{\bullet-1}(O(n)) } \,,

where the vertical and the bottom morphism are isomorphisms. Hence also the to morphisms is an isomorphism.

That BO(n)B˜O(n+1)B O(n)\to \tilde B O(n+1) is indeed a Serre fibration follows again with this prop., which gives the fiber sequence

O(n+1)/O(n)(EO(n+1))/O(n)(EO(n+1))/O(n+1). O(n+1)/O(n) \longrightarrow (E O(n+1))/O(n) \longrightarrow (E O(n+1))/O(n+1) \,.

The claim in then follows since (this exmpl.)

O(n+1)/O(n)S n. O(n+1)/O(n) \simeq S^n \,.

The argument for the complex case is of the same form, concluding now with the identification (this exmpl.)

U(n+1)/U(n)S 2n+1. U(n+1)/U(n) \simeq S^{2n+1} \,.

Classification of bundles

Proposition

For XX a paracompact topological space, the operation of pullback of the universal principal bundle EO(n)BO(n)E O(n) \to B O(n) from def. along continuous functions f:XBO(n)f \colon X \to B O(n) establishes a bijection

[X,BO(n)]isoff *EO(n)O(n)Bund/ [X, B O(n)] \underoverset{iso}{f \mapsto f^\ast E O(n)}{\longrightarrow} O(n) Bund/_\sim

between homotopy classes of functions from XX to B O ( n ) B O(n) and isomorphism classes of O(n)O(n)-principal bundles on XX.

A full proof is spelled out in Hatcher, section 1.2, theorem 1.16.

For symmetric principal bundles

For crossed complexes

We discuss here classifying spaces of crossed complexes.

The notion of classifying space should be regarded in general terms as giving a functor

:(algebraicdata)(topologicaldata). \mathcal{B} :(algebraic data) \to (topological data).

Composition with a forgetful functor U:(topologicaldata)(topologicalspaces)U: (topological data) \to (topological spaces) gives a classifying space. In such cases one would also like a homotopically defined functor

Ξ:(topologicaldata)(algebraicdata) \Xi: (topological data) \to (algebraic data)

such that

  1. Ξ\Xi \circ \mathcal{B} is equivalent to the identity;

  2. Ξ\Xi preserves certain colimits (Generalised van Kampen theorem) allowing some calculation;

  3. there are notions of homotopy for both types of data leading to a bijection of homotopy classes for some XX

[X,UC][ΞX *,C].[X,U\mathcal{B}C] \cong [\Xi X_*, C].

This happens for the algebraic data of crossed complexes and the topological data of filtered spaces, when XX is a CW-complex, and Ξ\Xi is the fundamental crossed complex of a filtered space. Thus in this case the classifying space does classify homotopy classes of maps, and more work is needed to sort out the data over XX which this classifies (gerbes?).

However C\mathcal{B}C is in this case defined by a nerve construction which generalises that for groupoids, and can also be applied to topological crossed crossed complexes, giving a simplicial space.

Mike: I don’t really get any intuition from that. There might be lots of functors from “algebraic data” to “topological data” but it seems to me that only particular sorts of them deserve the name “classifying space.” Can you say more specifically what sorts of functors you have in mind, and relate it to the more basic ideas that I am familiar with? What do these classifying spaces classify?

Ronnie What I am trying to characterise is that higher categories carry structure such as a filtration by lower dimensional higher categories, or, for multiple structures, a multiple filtration. Thus one expects a classifying space to inherit this extra structure. Conversely, the construction of an infinity-groupoid from a space might depend on this extra structure.

So I spent 9 years trying to construct a strict homotopy double groupoid of a space, yet Philip Higgins and I did this overnight in 1974 when we tried the simplest relative example we could think of: take homotopy classes of maps from a square to XX which take the edges to a subspace X 1X_1 and the vertices to a base point x 0x_0. Then the filtered case took another 4 years or so to complete.

Then Loday constructed a cat-n-group from an n-cube of spaces, published in 1982. Its multi-nerve is an (n+1)(n+1)-simplicial set, whose realisation is (n+1)(n+1)-filtered.

A strict homotopy double groupoid of a Hausdorff space has been constructed but this needs a subtle notion of thin homotopy.

Of course the filtration for a group is not so apparent, but it is more clear that groupoids carry structure in dimension 0 and 1, and hence are useful for representing non connected homotopy 1-types, and their identifications in dimension 0, as explained in the first edition (1968) of my Topology book.

The intuition for the higher homotopy van Kampen theorem is that you need structure in all dimensions from 0 to nto get colimit theorems in dimension n, because in homotopy, low dimensional identifications, even in dimension 0, usually effect high dimensional homotopy information. In effect, the higher homotopy van Kampen theorem is about gluing homotopy n-types.

Mike: Thanks, that is helpful.

Some such constructions arise from generalisations of the Dold-Kan correspondence, with values in simplicial sets. For example, from a crossed complex CC one obtains a simplicial set Nerve(C)Nerve(C) which in dimension nn is Crs(Π(Δ * n),C)Crs(\Pi(\Delta^n_*),C). The geometric realisation C\mathcal{B}C of this is canonically filtered by the skeleta of CC, so \mathcal{B} is really a functor to filtered spaces. This ties in with the functor Π\Pi which goes in the opposite direction. But note that there is a different filtration of the space C\mathcal{B}C since it is a CW-complex, and so Π\Pi of this filtration gives a free crossed complex.

Special cases of crossed complexes are groupoids, and so we get the classifying space of a groupoid; and similarly of a crossed module.

A crossed module is equivalent to a category object in groups, and so a nerve of this can be constructed as a bisimplicial set. The geometric realisation of this is naturally bifiltered, in several ways!

In considering what is desirable for a fundamental infinity-groupoid one should bring the notion of classifying space, and its inherited structure, into account.

For simplicial groups

The simplicial classifying space W¯()\overline{W}(-)-construction (see simplicial group and groupoid object in an (∞,1)-category) which gives the classifying space functor for simplicial groups and simplicially enriched groupoids is given in the entry on simplicial groups. It provides a good example of the above as the W-bar functor is right adjoint to the Dwyer-Kan loop groupoid functor and induces an equivalence of homotopy categories between that of simplicial sets and that of simplicially enriched groupoids. The simplicial sets here are playing the role of ‘topological data’.

Properties

For classical Lie groups

Let O(n)O(n) be the orthogonal group and U(n)U(n) the unitary group in real/complex dimension nn, respectively.

Proposition

The real Grassmannians Gr n( k)Gr_n(\mathbb{R}^k) and the complex Grassmannians Gr n( k)Gr_n(\mathbb{C}^k) admit the structure of CW-complexes. Moreover the canonical inclusions

Gr n( k)Gr n( k+1) Gr_n(\mathbb{R}^k) \hookrightarrow Gr_n(\mathbb{R}^{k+1})

are subcomplex incusion (hence relative cell complex inclusions).

Accordingly there is an induced CW-complex structure on the classifying space

BO(n)lim kGr n( k). B O(n) \simeq \underset{\longrightarrow}{\lim}_k Gr_n(\mathbb{R}^k) \,.

A proof is spelled out in (Hatcher, section 1.2 (pages 31-34)).

Proposition

The classifying spaces B O ( n ) B O(n) are paracompact spaces.

An early source of this statement is (Cartan-Schwartz 63, exposé 5). It follows for instance by prop. the fact that every CW-complex is paracompact.

Cohomology

(equivariant) cohomologyrepresenting
spectrum
equivariant cohomology
of the point *\ast
cohomology
of classifying space BGB G
(equivariant)
ordinary cohomology
HZBorel equivariance
H G (*)H (BG,)H^\bullet_G(\ast) \simeq H^\bullet(B G, \mathbb{Z})
(equivariant)
complex K-theory
KUrepresentation ring
KU G(*)R (G)KU_G(\ast) \simeq R_{\mathbb{C}}(G)
Atiyah-Segal completion theorem
R(G)KU G(*)compl.KU G(*)^KU(BG)R(G) \simeq KU_G(\ast) \overset{ \text{compl.} }{\longrightarrow} \widehat {KU_G(\ast)} \simeq KU(B G)
(equivariant)
complex cobordism cohomology
MUMU G(*)MU_G(\ast)completion theorem for complex cobordism cohomology
MU G(*)compl.MU G(*)^MU(BG)MU_G(\ast) \overset{ \text{compl.} }{\longrightarrow} \widehat {MU_G(\ast)} \simeq MU(B G)
(equivariant)
algebraic K-theory
K𝔽 pK \mathbb{F}_prepresentation ring
(K𝔽 p) G(*)R p(G)(K \mathbb{F}_p)_G(\ast) \simeq R_p(G)
Rector completion theorem
R 𝔽 p(G)K(𝔽 p) G(*)compl.(K𝔽 p) G(*)^Rector 73K𝔽 p(BG)R_{\mathbb{F}_p}(G) \simeq K (\mathbb{F}_p)_G(\ast) \overset{ \text{compl.} }{\longrightarrow} \widehat {(K \mathbb{F}_p)_G(\ast)} \!\! \overset{\text{<a href="https://ncatlab.org/nlab/show/Rector+completion+theorem">Rector 73</a>}}{\simeq} \!\!\!\!\!\! K \mathbb{F}_p(B G)
(equivariant)
stable cohomotopy
K𝔽 1Segal 74K \mathbb{F}_1 \overset{\text{<a href="stable cohomotopy#StableCohomotopyIsAlgebraicKTheoryOverFieldWithOneElement">Segal 74</a>}}{\simeq} SBurnside ring
𝕊 G(*)A(G)\mathbb{S}_G(\ast) \simeq A(G)
Segal-Carlsson completion theorem
A(G)Segal 71𝕊 G(*)compl.𝕊 G(*)^Carlsson 84𝕊(BG)A(G) \overset{\text{<a href="https://ncatlab.org/nlab/show/Burnside+ring+is+equivariant+stable+cohomotopy+of+the+point">Segal 71</a>}}{\simeq} \mathbb{S}_G(\ast) \overset{ \text{compl.} }{\longrightarrow} \widehat {\mathbb{S}_G(\ast)} \!\! \overset{\text{<a href="https://ncatlab.org/nlab/show/Segal-Carlsson+completion+theorem">Carlsson 84</a>}}{\simeq} \!\!\!\!\!\! \mathbb{S}(B G)

Moduli spaces

The notion of moduli space is closely related to that of classifying space, but has some subtle differences. See there for more on this.

References

General

Original discussion of classifying spaces in topological homotopy theory:

Review:

  • William Dwyer, Homotopy theory of classifying spaces, Lecture notes, Copenhagen 2008, (pdf, pdf)

  • Stephen Mitchell, Notes on principal bundles and classifying spaces, Lecture Notes. University of Washington, 2011 (pdf, pdf)

Textbook accounts:

Discussion in simplicial homotopy theory:

Discussion in topos theory relating to classifying toposes:

  • Ieke Moerdijk, Classifying spaces and classifying topoi, Lecture Notes in Math. 1616, Springer-Verlag, New York, 1995.

Discussion of characterization of principal bundles by rational universal characteristic classes and torsion information is in the appendices of

  • Igor Belegradek, Vitali Kapovitch, Obstructions to nonnegative curvature and rational homotopy theory (arXiv:math/0007007)

  • Igor Belegradek, Pinching, Pontrjagin classes, and negatively curved vector bundles (arXiv:math/0001132)

Discussion of classifying spaces in the context of measure theory is in

For equivariant bundles

Discussion of classifying G-spaces for GG-equivariant principal bundles:

Last revised on March 4, 2024 at 20:25:28. See the history of this page for a list of all contributions to it.