Dold–Thom theorem

Last updated

In algebraic topology, the Dold-Thom theorem states that the homotopy groups of the infinite symmetric product of a connected CW complex are the same as its reduced homology groups. The most common version of its proof consists of showing that the composition of the homotopy group functors with the infinite symmetric product defines a reduced homology theory. One of the main tools used in doing so are quasifibrations. The theorem has been generalised in various ways, for example by the Almgren isomorphism theorem.

Contents

There are several other theorems constituting relations between homotopy and homology, for example the Hurewicz theorem. Another approach is given by stable homotopy theory. Thanks to the Freudenthal suspension theorem, one can see that the latter actually defines a homology theory. Nevertheless, none of these allow one to directly reduce homology to homotopy. This advantage of the Dold-Thom theorem makes it particularly interesting for algebraic geometry.

The theorem

Dold-Thom theorem. For a connected CW complex X one has πnSP(X) ≅ n(X), where n denotes reduced homology and SP stands for the infinite symmetric product.

It is also very useful that there exists an isomorphism φ : πnSP(X) → n(X) which is compatible with the Hurewicz homomorphism h: πn(X) → n(X), meaning that one has a commutative diagram

Hurewicz (corrected).svg

where i* is the map induced by the inclusion i: X = SP1(X) → SP(X).

The following example illustrates that the requirement of X being a CW complex cannot be dropped offhand: Let X = CHCH be the wedge sum of two copies of the cone over the Hawaiian earring. The common point of the two copies is supposed to be the point 0 ∈ H meeting every circle. On the one hand, H1(X) is an infinite group [1] while H1(CH) is trivial. On the other hand, π1(SP(X)) ≅ π1(SP(CH)) × π1(SP(CH)) holds since φ : SP(X) × SP(Y) → SP(XY) defined by φ([x1, ..., xn], [y1, ..., yn]) = ([x1, ..., xn, y1, ..., yn]) is a homeomorphism for compact X and Y.

But this implies that either π1(SP(CH)) ≅ H1(CH) or π1(SP(X)) ≅ H1(X) does not hold.

Sketch of the proof

One wants to show that the family of functors hn = πn ∘ SP defines a homology theory. Dold and Thom chose in their initial proof a slight modification of the Eilenberg-Steenrod axioms, namely calling a family of functors (n)nN0 from the category of basepointed, connected CW complexes to the category of abelian groups a reduced homology theory if they satisfy

  1. If fg: XY, then f* = g*: n(X) → n(Y), where ≃ denotes pointed homotopy equivalence.
  2. There are natural boundary homomorphisms ∂ : n(X/A) → n−1(A) for each pair (X, A) with X and A being connected, yielding an exact sequence
    where i: AX is the inclusion and q: XX/A is the quotient map.
  3. n(S1) = 0 for n ≠ 1, where S1 denotes the circle.
  4. Let (Xλ) be the system of compact subspaces of a pointed space X containing the basepoint. Then (Xλ) is a direct system together with the inclusions. Denote by respectively the inclusion if XλXμ. n(Xλ) is a direct system as well with the morphisms . Then the homomorphism
    induced by the is required to be an isomorphism.

One can show that for a reduced homology theory (n)nN0 there is a natural isomorphism n(X) ≅ n(X; G) with G = 1(S1). [2]

Clearly, hn is a functor fulfilling property 1 as SP is a homotopy functor. Moreover, the third property is clear since one has SP(S1) ≃ S1. So it only remains to verify the axioms 2 and 4. The crux of this undertaking will be the first point. This is where quasifibrations come into play:

The goal is to prove that the map p*: SP(X) → SP(X/A) induced by the quotient map p: XX/A is a quasifibration for a CW pair (X, A) consisting of connected complexes. First of all, as every CW complex is homotopy equivalent to a simplicial complex, [3] X and A can be assumed to be simplicial complexes. Furthermore, X will be replaced by the mapping cylinder of the inclusion AX. This will not change anything as SP is a homotopy functor. It suffices to prove by induction that p* : EnBn is a quasifibration with Bn = SPn(X/A) and En = p*−1(Bn). For n = 0 this is trivially fulfilled. In the induction step, one decomposes Bn into an open neighbourhood of Bn−1 and BnBn−1 and shows that these two sets are, together with their intersection, distinguished, i.e. that p restricted to each of the preimages of these three sets is a quasifibration. It can be shown that Bn is then already distinguished itself. Therefore, p* is indeed a quasifibration on the whole SP(X) and the long exact sequence of such a one implies that axiom 2 is satisfied as p*−1([e]) ≅ SP(A) holds.

One may wonder whether p* is not even a fibration. However, that turns out not to be the case: Take an arbitrary path xt for t ∈ [0, 1) in XA approaching some aA and interpret it as a path in X/A ⊂ SP(X/A). Then any lift of this path to SP(X) is of the form xtαt with αtA for every t. But this means that its endpoint aα1 is a multiple of a, hence different from the basepoint, so the Homotopy lifting property fails to be fulfilled.

Verifying the fourth axiom can be done quite elementary, in contrast to the previous one.

One should bear in mind that there is a variety of different proofs although this one is seemingly the most popular. For example, proofs have been established via factorisation homology or simplicial sets. One can also proof the theorem using other notions of a homology theory (the Eilenberg-Steenrod axioms e.g.).

Compatibility with the Hurewicz homomorphism

In order to verify the compatibility with the Hurewicz homomorphism, it suffices to show that the statement holds for X = Sn. This is because one then gets a prism

Hurewicz2 (corrected).svg

for each Element [f] ∈ πn(X) represented by a map f: SnX. All sides except possibly the one at the bottom commute in this diagram. Therefore, one sees that the whole diagram commutes when considering where 1 ∈ πn(Sn) ≅ Z gets mapped to. However, by using the suspension isomorphisms for homotopy respectively homology groups, the task reduces to showing the assertion for S1. But in this case the inclusion SP1(S1) → SP(S1) is a homotopy equivalence.

Applications

Mayer-Vietoris sequence

One direct consequence of the Dold-Thom theorem is a new way to derive the Mayer-Vietoris sequence. One gets the result by first forming the homotopy pushout square of the inclusions of the intersection AB of two subspaces A, BX into A and B themselves. Then one applies SP to that square and finally π* to the resulting pullback square. [4]

A theorem of Moore

Another application is a new proof of a theorem first stated by Moore. It basically predicates the following:

Theorem. A path-connected, commutative and associative H-space X with a strict identity element has the weak homotopy type of a generalised Eilenberg-MacLane space.

Note that SP(Y) has this property for every connected CW complex Y and that it therefore has the weak homotopy type of a generalised Eilenberg-MacLane space. The theorem amounts to saying that all k-invariants of a path-connected, commutative and associative H-space with strict unit vanish.

Proof

Let Gn = πn(X). Then there exist maps M(Gn, n) → X inducing an isomorphism on πn if n ≥ 2 and an isomorphism on H1 if n = 1 for a Moore space M(Gn, n). [5] These give a map

if one takes the maps to be basepoint-preserving. Then the special H-space structure of X yields a map

given by summing up the images of the coordinates. But as there are natural homeomorphisms

with Π denoting the weak product, f induces isomorphisms on πn for n ≥ 2. But as π1(X) → π1SP(X) = H1(X) induced by the inclusion X → SP(X) is the Hurewicz homomorphism and as H-spaces have abelian fundamental groups, f also induces isomorphisms on π1. Thanks to the Dold-Thom theorem, each SP(M(Gn, n)) is now an Eilenberg-MacLane space K(Gn, n). This also implies that the natural inclusion of the weak product Πn SP(M(Gn, n)) into the cartesian product is a weak homotopy equivalence. Therefore, X has the weak homotopy type of a generalised Eilenberg-MacLane space.

Algebraic geometry

What distinguishes the Dold-Thom theorem from other alternative foundations of homology like Cech or Alexander-Spanier cohomology is that it is of particular interest for algebraic geometry since it allows one to reformulate homology only using homotopy. Since applying methods from algebraic topology can be quite insightful in this field, one tries to transfer these to algebraic geometry. This could be achieved for homotopy theory, but for homology theory only in a rather limited way using a formulation via sheaves. So the Dold-Thom theorem yields a foundation of homology having an algebraic analogue. [6]

Notes

  1. Dold and Thom (1958), Example 6.11
  2. Dold and Thom (1958), Satz 6.8
  3. Hatcher (2002), Theorem 2C.5
  4. The Dold-Thom theorem on nLab
  5. Hatcher (2002), Lemma 4.31
  6. The Dold-Thom theorem An essay by Thomas Barnet-Lamb

Related Research Articles

<span class="mw-page-title-main">Algebraic topology</span> Branch of mathematics

Algebraic topology is a branch of mathematics that uses tools from abstract algebra to study topological spaces. The basic goal is to find algebraic invariants that classify topological spaces up to homeomorphism, though usually most classify up to homotopy equivalence.

In mathematics, specifically in homology theory and algebraic topology, cohomology is a general term for a sequence of abelian groups, usually one associated with a topological space, often defined from a cochain complex. Cohomology can be viewed as a method of assigning richer algebraic invariants to a space than homology. Some versions of cohomology arise by dualizing the construction of homology. In other words, cochains are functions on the group of chains in homology theory.

<span class="mw-page-title-main">Homotopy</span> Continuous deformation between two continuous functions

In topology, a branch of mathematics, two continuous functions from one topological space to another are called homotopic if one can be "continuously deformed" into the other, such a deformation being called a homotopy between the two functions. A notable use of homotopy is the definition of homotopy groups and cohomotopy groups, important invariants in algebraic topology.

The notion of a fibration generalizes the notion of a fiber bundle and plays an important role in algebraic topology, a branch of mathematics.

In mathematics, particularly algebraic topology and homology theory, the Mayer–Vietoris sequence is an algebraic tool to help compute algebraic invariants of topological spaces, known as their homology and cohomology groups. The result is due to two Austrian mathematicians, Walther Mayer and Leopold Vietoris. The method consists of splitting a space into subspaces, for which the homology or cohomology groups may be easier to compute. The sequence relates the (co)homology groups of the space to the (co)homology groups of the subspaces. It is a natural long exact sequence, whose entries are the (co)homology groups of the whole space, the direct sum of the (co)homology groups of the subspaces, and the (co)homology groups of the intersection of the subspaces.

In algebraic topology, singular homology refers to the study of a certain set of algebraic invariants of a topological space X, the so-called homology groups Intuitively, singular homology counts, for each dimension n, the n-dimensional holes of a space. Singular homology is a particular example of a homology theory, which has now grown to be a rather broad collection of theories. Of the various theories, it is perhaps one of the simpler ones to understand, being built on fairly concrete constructions.

In homotopy theory, the Whitehead theorem states that if a continuous mapping f between CW complexes X and Y induces isomorphisms on all homotopy groups, then f is a homotopy equivalence. This result was proved by J. H. C. Whitehead in two landmark papers from 1949, and provides a justification for working with the concept of a CW complex that he introduced there. It is a model result of algebraic topology, in which the behavior of certain algebraic invariants determines a topological property of a mapping.

In mathematics, specifically algebraic topology, an Eilenberg–MacLane space is a topological space with a single nontrivial homotopy group.

In mathematics, the Thom space,Thom complex, or Pontryagin–Thom construction of algebraic topology and differential topology is a topological space associated to a vector bundle, over any paracompact space.

In algebraic topology, a branch of mathematics, the excision theorem is a theorem about relative homology and one of the Eilenberg–Steenrod axioms. Given a topological space and subspaces and such that is also a subspace of , the theorem says that under certain circumstances, we can cut out (excise) from both spaces such that the relative homologies of the pairs into are isomorphic.

In mathematics, specifically in algebraic topology, the Eilenberg–Steenrod axioms are properties that homology theories of topological spaces have in common. The quintessential example of a homology theory satisfying the axioms is singular homology, developed by Samuel Eilenberg and Norman Steenrod.

In mathematics, Kan complexes and Kan fibrations are part of the theory of simplicial sets. Kan fibrations are the fibrations of the standard model category structure on simplicial sets and are therefore of fundamental importance. Kan complexes are the fibrant objects in this model category. The name is in honor of Daniel Kan.

In mathematics, in the field of algebraic topology, the Eilenberg–Moore spectral sequence addresses the calculation of the homology groups of a pullback over a fibration. The spectral sequence formulates the calculation from knowledge of the homology of the remaining spaces. Samuel Eilenberg and John C. Moore's original paper addresses this for singular homology.

In algebraic topology, a discipline within mathematics, the acyclic models theorem can be used to show that two homology theories are isomorphic. The theorem was developed by topologists Samuel Eilenberg and Saunders MacLane. They discovered that, when topologists were writing proofs to establish equivalence of various homology theories, there were numerous similarities in the processes. Eilenberg and MacLane then discovered the theorem to generalize this process.

In algebraic topology, the nth symmetric product of a topological space consists of the unordered n-tuples of its elements. If one fixes a basepoint, there is a canonical way of embedding the lower-dimensional symmetric products into the higher-dimensional ones. That way, one can consider the colimit over the symmetric products, the infinite symmetric product. This construction can easily be extended to give a homotopy functor.

In mathematics, more precisely, in the theory of simplicial sets, the Dold–Kan correspondence states that there is an equivalence between the category of chain complexes and the category of simplicial abelian groups. Moreover, under the equivalence, the th homology group of a chain complex is the th homotopy group of the corresponding simplicial abelian group, and a chain homotopy corresponds to a simplicial homotopy.

In algebraic topology, a quasifibration is a generalisation of fibre bundles and fibrations introduced by Albrecht Dold and René Thom. Roughly speaking, it is a continuous map p: EB having the same behaviour as a fibration regarding the (relative) homotopy groups of E, B and p−1(x). Equivalently, one can define a quasifibration to be a continuous map such that the inclusion of each fibre into its homotopy fibre is a weak equivalence. One of the main applications of quasifibrations lies in proving the Dold-Thom theorem.

This is a glossary of properties and concepts in algebraic topology in mathematics.

In mathematics, homotopy theory is a systematic study of situations in which maps can come with homotopies between them. It originated as a topic in algebraic topology but nowadays is studied as an independent discipline. Besides algebraic topology, the theory has also been used in other areas of mathematics such as algebraic geometry (e.g., A1 homotopy theory) and category theory (specifically the study of higher categories).

References