Spectrum (topology)

Summary

In algebraic topology, a branch of mathematics, a spectrum is an object representing a generalized cohomology theory. Every such cohomology theory is representable, as follows from Brown's representability theorem. This means that, given a cohomology theory

,

there exist spaces such that evaluating the cohomology theory in degree on a space is equivalent to computing the homotopy classes of maps to the space , that is

.

Note there are several different categories of spectra leading to many technical difficulties,[1] but they all determine the same homotopy category, known as the stable homotopy category. This is one of the key points for introducing spectra because they form a natural home for stable homotopy theory.

The definition of a spectrum edit

There are many variations of the definition: in general, a spectrum is any sequence   of pointed topological spaces or pointed simplicial sets together with the structure maps  , where   is the smash product. The smash product of a pointed space   with a circle is homeomorphic to the reduced suspension of  , denoted  .

The following is due to Frank Adams (1974): a spectrum (or CW-spectrum) is a sequence   of CW complexes together with inclusions   of the suspension   as a subcomplex of  .

For other definitions, see symmetric spectrum and simplicial spectrum.

Homotopy groups of a spectrum edit

One of the most important invariants of spectra are the homotopy groups of the spectrum. These groups mirror the definition of the stable homotopy groups of spaces since the structure of the suspension maps is integral in its definition. Given a spectrum   define the homotopy group   as the colimit

 

where the maps are induced from the composition of the map   (that is,   given by functoriality of  ) and the structure map  . A spectrum is said to be connective if its   are zero for negative k.

Examples edit

Eilenberg–Maclane spectrum edit

Consider singular cohomology   with coefficients in an abelian group  . For a CW complex  , the group   can be identified with the set of homotopy classes of maps from   to  , the Eilenberg–MacLane space with homotopy concentrated in degree  . We write this as

 

Then the corresponding spectrum   has  -th space  ; it is called the Eilenberg–MacLane spectrum of  . Note this construction can be used to embed any ring   into the category of spectra. This embedding forms the basis of spectral geometry, a model for derived algebraic geometry. One of the important properties of this embedding are the isomorphisms

 

showing the category of spectra keeps track of the derived information of commutative rings, where the smash product acts as the derived tensor product. Moreover, Eilenberg–Maclane spectra can be used to define theories such as topological Hochschild homology for commutative rings, a more refined theory than classical Hochschild homology.

Topological complex K-theory edit

As a second important example, consider topological K-theory. At least for X compact,   is defined to be the Grothendieck group of the monoid of complex vector bundles on X. Also,   is the group corresponding to vector bundles on the suspension of X. Topological K-theory is a generalized cohomology theory, so it gives a spectrum. The zeroth space is   while the first space is  . Here   is the infinite unitary group and   is its classifying space. By Bott periodicity we get   and   for all n, so all the spaces in the topological K-theory spectrum are given by either   or  . There is a corresponding construction using real vector bundles instead of complex vector bundles, which gives an 8-periodic spectrum.

Sphere spectrum edit

One of the quintessential examples of a spectrum is the sphere spectrum  . This is a spectrum whose homotopy groups are given by the stable homotopy groups of spheres, so

 

We can write down this spectrum explicitly as   where  . Note the smash product gives a product structure on this spectrum

 

induces a ring structure on  . Moreover, if considering the category of symmetric spectra, this forms the initial object, analogous to   in the category of commutative rings.

Thom spectra edit

Another canonical example of spectra come from the Thom spectra representing various cobordism theories. This includes real cobordism  , complex cobordism  , framed cobordism, spin cobordism  , string cobordism  , and so on. In fact, for any topological group   there is a Thom spectrum  .

Suspension spectrum edit

A spectrum may be constructed out of a space. The suspension spectrum of a space  , denoted   is a spectrum   (the structure maps are the identity.) For example, the suspension spectrum of the 0-sphere is the sphere spectrum discussed above. The homotopy groups of this spectrum are then the stable homotopy groups of  , so

 

The construction of the suspension spectrum implies every space can be considered as a cohomology theory. In fact, it defines a functor

 

from the homotopy category of CW complexes to the homotopy category of spectra. The morphisms are given by

 

which by the Freudenthal suspension theorem eventually stabilizes. By this we mean

  and  

for some finite integer  . For a CW complex   there is an inverse construction   which takes a spectrum   and forms a space

 

called the infinite loop space of the spectrum. For a CW complex  

 

and this construction comes with an inclusion   for every  , hence gives a map

 

which is injective. Unfortunately, these two structures, with the addition of the smash product, lead to significant complexity in the theory of spectra because there cannot exist a single category of spectra which satisfies a list of five axioms relating these structures.[1] The above adjunction is valid only in the homotopy categories of spaces and spectra, but not always with a specific category of spectra (not the homotopy category).

Ω-spectrum edit

An Ω-spectrum is a spectrum such that the adjoint of the structure map (i.e., the map ) is a weak equivalence. The K-theory spectrum of a ring is an example of an Ω-spectrum.

Ring spectrum edit

A ring spectrum is a spectrum X such that the diagrams that describe ring axioms in terms of smash products commute "up to homotopy" (  corresponds to the identity.) For example, the spectrum of topological K-theory is a ring spectrum. A module spectrum may be defined analogously.

For many more examples, see the list of cohomology theories.

Functions, maps, and homotopies of spectra edit

There are three natural categories whose objects are spectra, whose morphisms are the functions, or maps, or homotopy classes defined below.

A function between two spectra E and F is a sequence of maps from En to Fn that commute with the maps ΣEn → En+1 and ΣFn → Fn+1.

Given a spectrum  , a subspectrum   is a sequence of subcomplexes that is also a spectrum. As each i-cell in   suspends to an (i + 1)-cell in  , a cofinal subspectrum is a subspectrum for which each cell of the parent spectrum is eventually contained in the subspectrum after a finite number of suspensions. Spectra can then be turned into a category by defining a map of spectra   to be a function from a cofinal subspectrum   of   to  , where two such functions represent the same map if they coincide on some cofinal subspectrum. Intuitively such a map of spectra does not need to be everywhere defined, just eventually become defined, and two maps that coincide on a cofinal subspectrum are said to be equivalent. This gives the category of spectra (and maps), which is a major tool. There is a natural embedding of the category of pointed CW complexes into this category: it takes   to the suspension spectrum in which the nth complex is  .

The smash product of a spectrum   and a pointed complex   is a spectrum given by   (associativity of the smash product yields immediately that this is indeed a spectrum). A homotopy of maps between spectra corresponds to a map  , where   is the disjoint union   with   taken to be the basepoint.

The stable homotopy category, or homotopy category of (CW) spectra is defined to be the category whose objects are spectra and whose morphisms are homotopy classes of maps between spectra. Many other definitions of spectrum, some appearing very different, lead to equivalent stable homotopy categories.

Finally, we can define the suspension of a spectrum by  . This translation suspension is invertible, as we can desuspend too, by setting  .

The triangulated homotopy category of spectra edit

The stable homotopy category is additive: maps can be added by using a variant of the track addition used to define homotopy groups. Thus homotopy classes from one spectrum to another form an abelian group. Furthermore the stable homotopy category is triangulated (Vogt (1970)), the shift being given by suspension and the distinguished triangles by the mapping cone sequences of spectra

 .

Smash products of spectra edit

The smash product of spectra extends the smash product of CW complexes. It makes the stable homotopy category into a monoidal category; in other words it behaves like the (derived) tensor product of abelian groups. A major problem with the smash product is that obvious ways of defining it make it associative and commutative only up to homotopy. Some more recent definitions of spectra, such as symmetric spectra, eliminate this problem, and give a symmetric monoidal structure at the level of maps, before passing to homotopy classes.

The smash product is compatible with the triangulated category structure. In particular the smash product of a distinguished triangle with a spectrum is a distinguished triangle.

Generalized homology and cohomology of spectra edit

We can define the (stable) homotopy groups of a spectrum to be those given by

 ,

where   is the sphere spectrum and   is the set of homotopy classes of maps from   to  . We define the generalized homology theory of a spectrum E by

 

and define its generalized cohomology theory by

 

Here   can be a spectrum or (by using its suspension spectrum) a space.

Technical complexities with spectra edit

One of the canonical complexities while working with spectra and defining a category of spectra comes from the fact each of these categories cannot satisfy five seemingly obvious axioms concerning the infinite loop space of a spectrum  

 

sending

 

a pair of adjoint functors  , the and the smash product   in both the category of spaces and the category of spectra. If we let   denote the category of based, compactly generated, weak Hausdorff spaces, and   denote a category of spectra, the following five axioms can never be satisfied by the specific model of spectra:[1]

  1.   is a symmetric monoidal category with respect to the smash product  
  2. The functor   is left-adjoint to  
  3. The unit for the smash product   is the sphere spectrum  
  4. Either there is a natural transformation   or a natural transformation   which commutes with the unit object in both categories, and the commutative and associative isomorphisms in both categories.
  5. There is a natural weak equivalence   for   which that there is a commuting diagram:

     

    where   is the unit map in the adjunction.

Because of this, the study of spectra is fractured based upon the model being used. For an overview, check out the article cited above.

History edit

A version of the concept of a spectrum was introduced in the 1958 doctoral dissertation of Elon Lages Lima. His advisor Edwin Spanier wrote further on the subject in 1959. Spectra were adopted by Michael Atiyah and George W. Whitehead in their work on generalized homology theories in the early 1960s. The 1964 doctoral thesis of J. Michael Boardman gave a workable definition of a category of spectra and of maps (not just homotopy classes) between them, as useful in stable homotopy theory as the category of CW complexes is in the unstable case. (This is essentially the category described above, and it is still used for many purposes: for other accounts, see Adams (1974) or Rainer Vogt (1970).) Important further theoretical advances have however been made since 1990, improving vastly the formal properties of spectra. Consequently, much recent literature uses modified definitions of spectrum: see Michael Mandell et al. (2001) for a unified treatment of these new approaches.

See also edit

References edit

  1. ^ a b c Lewis, L. Gaunce (1991-08-30). "Is there a convenient category of spectra?". Journal of Pure and Applied Algebra. 73 (3): 233–246. doi:10.1016/0022-4049(91)90030-6. ISSN 0022-4049.

Introductory edit

Modern articles developing the theory edit

Historically relevant articles edit

  • Lima, Elon Lages (1959), "The Spanier–Whitehead duality in new homotopy categories", Summa Brasil. Math., 4: 91–148, MR 0116332
  • Lima, Elon Lages (1960), "Stable Postnikov invariants and their duals", Summa Brasil. Math., 4: 193–251
  • Vogt, Rainer (1970), Boardman's stable homotopy category, Lecture Notes Series, No. 21, Matematisk Institut, Aarhus Universitet, Aarhus, MR 0275431
  • Whitehead, George W. (1962), "Generalized homology theories", Transactions of the American Mathematical Society, 102 (2): 227–283, doi:10.1090/S0002-9947-1962-0137117-6

External links edit

  • Spectral Sequences - Allen Hatcher - contains excellent introduction to spectra and applications for constructing Adams spectral sequence
  • An untitled book project about symmetric spectra
  • "Are spectra really the same as cohomology theories?".