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 learned as an independent discipline.

Applications to other fields of mathematics

edit

Besides algebraic topology, the theory has also been used in other areas of mathematics such as:

Concepts

edit

Spaces and maps

edit

In homotopy theory and algebraic topology, the word "space" denotes a topological space. In order to avoid pathologies, one rarely works with arbitrary spaces; instead, one requires spaces to meet extra constraints, such as being compactly generated weak Hausdorff or a CW complex.

In the same vein as above, a "map" is a continuous function, possibly with some extra constraints.

Often, one works with a pointed space—that is, a space with a "distinguished point", called a basepoint. A pointed map is then a map which preserves basepoints; that is, it sends the basepoint of the domain to that of the codomain. In contrast, a free map is one which needn't preserve basepoints.

The Cartesian product of two pointed spaces   are not naturally pointed. A substitute is the smash product   which is characterized by the adjoint relation

 ,

that is, a smash product is an analog of a tensor product in abstract algebra (see tensor-hom adjunction). Explicitly,   is the quotient of   by the wedge sum  .

Homotopy

edit

Let I denote the unit interval  . A map

 

is called a homotopy from the map   to the map  , where  . Intuitively, we may think of   as a path from the map   to the map  . Indeed, a homotopy can be shown to be an equivalence relation. When X, Y are pointed spaces, the maps   are required to preserve the basepoint.

Given a pointed space X and an integer  , let   be the homotopy classes of based maps   from a (pointed) n-sphere   to X. As it turns out, for  ,   are groups called homotopy groups; in particular,   is called the fundamental group of X, while   can be identified with the set of path-connected components in  . Every group is the fundamental group of some space.[1]

If one prefers to work with a space instead of a pointed space, there is the notion of a fundamental groupoid (and higher variants): by definition, the fundamental groupoid of a space X is the category where the objects are the points of X and the morphisms are paths.

A map   is called a homotopy equivalence if there is another map   such that   and   are both homotopic to the identities. Two spaces are said to be homotopy equivalent if there is a homotopy equivalence between them. A homotopy equivalence class of spaces is then called a homotopy type. There is a weaker notion: a map   is said to be a weak homotopy equivalence if   is an isomorphism for each   and each choice of a base point. (Note the existence of such a map is more strict than saying that   have isomorphic homotopy set/groups.) In general, a homotopy equivalence is a weak homotopy equivalence but the converse need not be true.

CW complex

edit

A CW complex is a space that has a filtration   whose union is   and such that

  1.   is a discrete space, called the set of 0-cells (vertices) in  .
  2. Each   is obtained by attaching several n-disks, n-cells, to   via maps  ; i.e., the boundary of an n-disk is identified with the image of   in  .
  3. A subset   is open if and only if   is open for each  .

For example, a sphere   has two cells: one 0-cell and one  -cell, since   can be obtained by collapsing the boundary   of the n-disk to a point. In general, every manifold has the homotopy type of a CW complex;[2] in fact, Morse theory implies that a compact manifold has the homotopy type of a finite CW complex.[citation needed]

Remarkably, Whitehead's theorem says that for CW complexes, a weak homotopy equivalence and a homotopy equivalence are the same thing.

Another important result is the approximation theorem. First, the homotopy category of spaces is the category where an object is a space but a morphism is the homotopy class of a map. Then

CW approximation — [3] There exist a functor (called the CW approximation functor)

 

from the homotopy category of spaces to the homotopy category of CW complexes as well as a natural transformation

 

where  , such that each   is a weak homotopy equivalence.

Similar statements also hold for pairs and excisive triads.[4][5]

Explicitly, the above approximation functor can be defined as the composition of the singular chain functor   followed by the geometric realization functor; see § Simplicial set.

The above theorem justifies a common habit of working only with CW complexes. For example, given a space  , one can just define the homology of   to the homology of the CW approximation of   (the cell structure of a CW complex determines the natural homology, the cellular homology and that can be taken to be the homology of the complex.)

Cofibration and fibration

edit

A map   is called a cofibration if given:

  1. A map  , and
  2. A homotopy  

such that  , there exists a homotopy   that extends   and such that  . An example is a neighborhood deformation retract; that is,   contains a mapping cylinder neighborhood of a closed subspace   and   the inclusion (e.g., a tubular neighborhood of a closed submanifold).[6] In fact, a cofibration can be characterized as a neighborhood deformation retract pair.[7] Another basic example is a CW pair  ; i.e.,   is a subcomplex of  . Many often work only with CW complexes and the notion of a cofibration there is then often implicit.

A fibration in the sense of Hurewicz is the dual notion of a cofibration: that is, a map   is a fibration if given (1) a map   and (2) a homotopy   such that  , there exists a homotopy   that extends   and such that  . A basic example is a covering map (in fact, a fibration is a generalization of a covering map). If   is a principal G-bundle, that is, a space with a free and transitive (topological) group action of a (topological) group, then the projection map   is an example of a fibration.

A Hurewicz fibration is a Serre fibration; the converse holds for CW complexes.[8]

There are also based versions of a cofibration and a fibration (namely, the maps are required to be based).[9]

Loop and suspension

edit

On the category of pointed spaces, there are two important functors: the loop functor   and the (reduced) suspension functor  , which are in the adjoint relation. Precisely, they are defined as[10]

  •  , and
  •  .

Because of the adjoint relation between a smash product and a mapping space, we have:

 

These functors are used to construct fiber sequences and cofiber sequences. Namely, if   is a map, the fiber sequence generated by   is the exact sequence[11]

 

where   is the homotopy fiber of  ; i.e., a fiber obtained after replacing   by a (based) fibration. The cofibration sequence generated by   is   where   is the homotooy cofiber of   constructed like a homotopy fiber (use a quotient instead of a fiber.)

The functors   restrict to the category of CW complexes in the following weak sense: a theorem of Milnor says that if   has the homotopy type of a CW complex, then so does its loop space  .[12]

Classifying spaces and homotopy operations

edit

Given a topological group G, the classifying space for principal G-bundles ("the" up to equivalence) is a space   such that, for each space X,

  {principal G-bundle on X} / ~  

where

  • the left-hand side is the set of homotopy classes of maps  ,
  • ~ refers isomorphism of bundles, and
  • = is given by pulling-back the distinguished bundle   on   (called universal bundle) along a map  .

Brown's representability theorem guarantees the existence of classifying spaces.

Spectrum and generalized cohomology

edit

The idea that a classifying space classifies principal bundles can be pushed further. For example, one might try to classify cohomology classes: given an abelian group A (such as  ),

 

where   is the Eilenberg–MacLane space. The above equation leads to the notion of a generalized cohomology theory; i.e., a contravariant functor from the category of spaces to the category of abelian groups that satisfies the axioms generalizing ordinary cohomology theory. As it turns out, such a functor may not be representable by a space but it can always be represented by a sequence of (pointed) spaces with structure maps called a spectrum. In other words, to give a generalized cohomology theory is to give a spectrum. A K-theory is an example of a generalized cohomology theory.

A basic example of a spectrum is a sphere spectrum:  

Key theorems

edit

Obstruction theory and characteristic class

edit

See also: Characteristic class, Postnikov tower, Whitehead torsion

Localization and completion of a space

edit

Specific theories

edit

There are several specific theories

Homotopy hypothesis

edit

One of the basic questions in the foundations of homotopy theory is the nature of a space. The homotopy hypothesis asks whether a space is something fundamentally algebraic.

Abstract homotopy theory

edit

Abstract homotopy theory is an axiomatic approach to homotopy theory. Such axiomatization is useful for non-traditional applications of homotopy theory. One approach to axiomatization is by Quillen's model categories. A model category is a category with a choice of three classes of maps called weak equivalences, cofibrations and fibrations, subject to the axioms that are reminiscent of facts in algebraic topology. For example, the category of (reasonable) topological spaces has a structure of a model category where a weak equivalence is a weak homotopy equivalence, a cofibration a certain retract and a fibration a Serre fibration.[13] Another example is the category of non-negatively graded chain complexes over a fixed base ring.[14]

See also: Algebraic homotopy

Simplicial set

edit

A simplicial set is an abstract generalization of a simplicial complex and can play a role of a "space" in some sense. Despite the name, it is not a set but is a sequence of sets together with the certain maps (face and degeneracy) between those sets.

For example, given a space  , for each integer  , let   be the set of all maps from the n-simplex to  . Then the sequence   of sets is a simplicial set.[15] Each simplicial set   has a naturally associated chain complex and the homology of that chain complex is the homology of  . The singular homology of   is precisely the homology of the simplicial set  . Also, the geometric realization   of a simplicial set is a CW complex and the composition   is precisely the CW approximation functor.

Another important example is a category or more precisely the nerve of a category, which is a simplicial set. In fact, a simplicial set is the nerve of some category if and only if it satisfies the Segal conditions (a theorem of Grothendieck). Each category is completely determined by its nerve. In this way, a category can be viewed as a special kind of a simplicial set, and this observation is used to generalize a category. Namely, an  -category or an  -groupoid is defined as particular kinds of simplicial sets.

Since simplicial sets are sort of abstract spaces (if not topological spaces), it is possible to develop the homotopy theory on them, which is called the simplicial homotopy theory.[15]

See also

edit

References

edit
  1. ^ May, Ch 4. § 5.
  2. ^ Milnor 1959, Corollary 1. NB: "second countable" implies "separable".
  3. ^ May, Ch. 10., § 5
  4. ^ May, Ch. 10., § 6
  5. ^ May, Ch. 10., § 7
  6. ^ Hatcher, Example 0.15.
  7. ^ May, Ch 6. § 4.
  8. ^ https://ncatlab.org/nlab/show/a+Serre+fibration+between+CW-complexes+is+a+Hurewicz+fibration
  9. ^ May, Ch 8. § 3. and § 5.
  10. ^ May, Ch. 8, § 2.
  11. ^ May, Ch. 8, § 6.
  12. ^ Milnor 1959, Theorem 3.
  13. ^ Dwyer & Spalinski, Example 3.5.
  14. ^ Dwyer & Spalinski, Example 3.7.
  15. ^ a b May, Ch. 16, § 4.
  • May, J. A Concise Course in Algebraic Topology
  • J. P. May and K. Ponto, More Concise Algebraic Topology: Localization, completion, and model categories
  • George William Whitehead (1978). Elements of homotopy theory. Graduate Texts in Mathematics. Vol. 61 (3rd ed.). New York-Berlin: Springer-Verlag. pp. xxi+744. ISBN 978-0-387-90336-1. MR 0516508. Retrieved September 6, 2011.
  • Ronald Brown, Topology and groupoids (2006) Booksurge LLC ISBN 1-4196-2722-8.
  • https://ncatlab.org/nlab/show/homotopical+algebra
  • Homotopy Theories and Model Categories by W.G. Dwyer and J. Spalinski in Handbook of Algebraic Topology edited by I.M. James
  • Hatcher, Allen. "Algebraic topology".
  • Milnor, John (1959). "On spaces having the homotopy type of 𝐶𝑊-complex". Transactions of the American Mathematical Society. 90 (2): 272–280. doi:10.1090/S0002-9947-1959-0100267-4. ISSN 0002-9947.
  • Edwin Spanier, Algebraic topology
  • Dennis Sullivan. Genetics of homotopy theory and the Adams conjecture. Ann. of Math. (2), 100:1–79, 1974.

Further reading

edit
edit

"homotopy theory". ncatlab.org.