fbpx
Wikipedia

Cohomology

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.

From its start in topology, this idea became a dominant method in the mathematics of the second half of the twentieth century. From the initial idea of homology as a method of constructing algebraic invariants of topological spaces, the range of applications of homology and cohomology theories has spread throughout geometry and algebra. The terminology tends to hide the fact that cohomology, a contravariant theory, is more natural than homology in many applications. At a basic level, this has to do with functions and pullbacks in geometric situations: given spaces X and Y, and some kind of function F on Y, for any mapping f : XY, composition with f gives rise to a function Ff on X. The most important cohomology theories have a product, the cup product, which gives them a ring structure. Because of this feature, cohomology is usually a stronger invariant than homology.

Singular cohomology edit

Singular cohomology is a powerful invariant in topology, associating a graded-commutative ring with any topological space. Every continuous map f: XY determines a homomorphism from the cohomology ring of Y to that of X; this puts strong restrictions on the possible maps from X to Y. Unlike more subtle invariants such as homotopy groups, the cohomology ring tends to be computable in practice for spaces of interest.

For a topological space X, the definition of singular cohomology starts with the singular chain complex:[1]

 
By definition, the singular homology of X is the homology of this chain complex (the kernel of one homomorphism modulo the image of the previous one). In more detail, Ci is the free abelian group on the set of continuous maps from the standard i-simplex to X (called "singular i-simplices in X"), and ∂i is the i-th boundary homomorphism. The groups Ci are zero for i negative.

Now fix an abelian group A, and replace each group Ci by its dual group   and   by its dual homomorphism

 

This has the effect of "reversing all the arrows" of the original complex, leaving a cochain complex

 

For an integer i, the ith cohomology group of X with coefficients in A is defined to be ker(di)/im(di−1) and denoted by Hi(X, A). The group Hi(X, A) is zero for i negative. The elements of   are called singular i-cochains with coefficients in A. (Equivalently, an i-cochain on X can be identified with a function from the set of singular i-simplices in X to A.) Elements of ker(d) and im(d) are called cocycles and coboundaries, respectively, while elements of ker(d)/im(d) = Hi(X, A) are called cohomology classes (because they are equivalence classes of cocycles).

In what follows, the coefficient group A is sometimes not written. It is common to take A to be a commutative ring R; then the cohomology groups are R-modules. A standard choice is the ring Z of integers.

Some of the formal properties of cohomology are only minor variants of the properties of homology:

  • A continuous map   determines a pushforward homomorphism   on homology and a pullback homomorphism   on cohomology. This makes cohomology into a contravariant functor from topological spaces to abelian groups (or R-modules).
  • Two homotopic maps from X to Y induce the same homomorphism on cohomology (just as on homology).
  • The Mayer–Vietoris sequence is an important computational tool in cohomology, as in homology. Note that the boundary homomorphism increases (rather than decreases) degree in cohomology. That is, if a space X is the union of open subsets U and V, then there is a long exact sequence:
     
  • There are relative cohomology groups   for any subspace Y of a space X. They are related to the usual cohomology groups by a long exact sequence:
     
  • The universal coefficient theorem describes cohomology in terms of homology, using Ext groups. Namely, there is a short exact sequence
     
    A related statement is that for a field F,   is precisely the dual space of the vector space  .
  • If X is a topological manifold or a CW complex, then the cohomology groups   are zero for i greater than the dimension of X.[2] If X is a compact manifold (possibly with boundary), or a CW complex with finitely many cells in each dimension, and R is a commutative Noetherian ring, then the R-module Hi(X,R) is finitely generated for each i.[3]

On the other hand, cohomology has a crucial structure that homology does not: for any topological space X and commutative ring R, there is a bilinear map, called the cup product:

 
defined by an explicit formula on singular cochains. The product of cohomology classes u and v is written as uv or simply as uv. This product makes the direct sum
 
into a graded ring, called the cohomology ring of X. It is graded-commutative in the sense that:[4]
 

For any continuous map   the pullback   is a homomorphism of graded R-algebras. It follows that if two spaces are homotopy equivalent, then their cohomology rings are isomorphic.

Here are some of the geometric interpretations of the cup product. In what follows, manifolds are understood to be without boundary, unless stated otherwise. A closed manifold means a compact manifold (without boundary), whereas a closed submanifold N of a manifold M means a submanifold that is a closed subset of M, not necessarily compact (although N is automatically compact if M is).

  • Let X be a closed oriented manifold of dimension n. Then Poincaré duality gives an isomorphism HiXHniX. As a result, a closed oriented submanifold S of codimension i in X determines a cohomology class in HiX, called [S]. In these terms, the cup product describes the intersection of submanifolds. Namely, if S and T are submanifolds of codimension i and j that intersect transversely, then
     
    where the intersection ST is a submanifold of codimension i + j, with an orientation determined by the orientations of S, T, and X. In the case of smooth manifolds, if S and T do not intersect transversely, this formula can still be used to compute the cup product [S][T], by perturbing S or T to make the intersection transverse.
    More generally, without assuming that X has an orientation, a closed submanifold of X with an orientation on its normal bundle determines a cohomology class on X. If X is a noncompact manifold, then a closed submanifold (not necessarily compact) determines a cohomology class on X. In both cases, the cup product can again be described in terms of intersections of submanifolds.
    Note that Thom constructed an integral cohomology class of degree 7 on a smooth 14-manifold that is not the class of any smooth submanifold.[5] On the other hand, he showed that every integral cohomology class of positive degree on a smooth manifold has a positive multiple that is the class of a smooth submanifold.[6] Also, every integral cohomology class on a manifold can be represented by a "pseudomanifold", that is, a simplicial complex that is a manifold outside a closed subset of codimension at least 2.
  • For a smooth manifold X, de Rham's theorem says that the singular cohomology of X with real coefficients is isomorphic to the de Rham cohomology of X, defined using differential forms. The cup product corresponds to the product of differential forms. This interpretation has the advantage that the product on differential forms is graded-commutative, whereas the product on singular cochains is only graded-commutative up to chain homotopy. In fact, it is impossible to modify the definition of singular cochains with coefficients in the integers   or in   for a prime number p to make the product graded-commutative on the nose. The failure of graded-commutativity at the cochain level leads to the Steenrod operations on mod p cohomology.

Very informally, for any topological space X, elements of   can be thought of as represented by codimension-i subspaces of X that can move freely on X. For example, one way to define an element of   is to give a continuous map f from X to a manifold M and a closed codimension-i submanifold N of M with an orientation on the normal bundle. Informally, one thinks of the resulting class   as lying on the subspace   of X; this is justified in that the class   restricts to zero in the cohomology of the open subset   The cohomology class   can move freely on X in the sense that N could be replaced by any continuous deformation of N inside M.

Examples edit

In what follows, cohomology is taken with coefficients in the integers Z, unless stated otherwise.

  • The cohomology ring of a point is the ring Z in degree 0. By homotopy invariance, this is also the cohomology ring of any contractible space, such as Euclidean space Rn.
  •  
    The first cohomology group of the 2-dimensional torus has a basis given by the classes of the two circles shown.
    For a positive integer n, the cohomology ring of the sphere   is Z[x]/(x2) (the quotient ring of a polynomial ring by the given ideal), with x in degree n. In terms of Poincaré duality as above, x is the class of a point on the sphere.
  • The cohomology ring of the torus   is the exterior algebra over Z on n generators in degree 1.[7] For example, let P denote a point in the circle  , and Q the point (P,P) in the 2-dimensional torus  . Then the cohomology of (S1)2 has a basis as a free Z-module of the form: the element 1 in degree 0, x := [P × S1] and y := [S1 × P] in degree 1, and xy = [Q] in degree 2. (Implicitly, orientations of the torus and of the two circles have been fixed here.) Note that yx = −xy = −[Q], by graded-commutativity.
  • More generally, let R be a commutative ring, and let X and Y be any topological spaces such that H*(X,R) is a finitely generated free R-module in each degree. (No assumption is needed on Y.) Then the Künneth formula gives that the cohomology ring of the product space X × Y is a tensor product of R-algebras:[8]
     
  • The cohomology ring of real projective space RPn with Z/2 coefficients is Z/2[x]/(xn+1), with x in degree 1.[9] Here x is the class of a hyperplane RPn−1 in RPn; this makes sense even though RPj is not orientable for j even and positive, because Poincaré duality with Z/2 coefficients works for arbitrary manifolds.
    With integer coefficients, the answer is a bit more complicated. The Z-cohomology of RP2a has an element y of degree 2 such that the whole cohomology is the direct sum of a copy of Z spanned by the element 1 in degree 0 together with copies of Z/2 spanned by the elements yi for i=1,...,a. The Z-cohomology of RP2a+1 is the same together with an extra copy of Z in degree 2a+1.[10]
  • The cohomology ring of complex projective space CPn is Z[x]/(xn+1), with x in degree 2.[9] Here x is the class of a hyperplane CPn−1 in CPn. More generally, xj is the class of a linear subspace CPnj in CPn.
  • The cohomology ring of the closed oriented surface X of genus g ≥ 0 has a basis as a free Z-module of the form: the element 1 in degree 0, A1,...,Ag and B1,...,Bg in degree 1, and the class P of a point in degree 2. The product is given by: AiAj = BiBj = 0 for all i and j, AiBj = 0 if ij, and AiBi = P for all i.[11] By graded-commutativity, it follows that BiAi = −P.
  • On any topological space, graded-commutativity of the cohomology ring implies that 2x2 = 0 for all odd-degree cohomology classes x. It follows that for a ring R containing 1/2, all odd-degree elements of H*(X,R) have square zero. On the other hand, odd-degree elements need not have square zero if R is Z/2 or Z, as one sees in the example of RP2 (with Z/2 coefficients) or RP4 × RP2 (with Z coefficients).

The diagonal edit

The cup product on cohomology can be viewed as coming from the diagonal map Δ: XX × X, x ↦ (x,x). Namely, for any spaces X and Y with cohomology classes uHi(X,R) and vHj(Y,R), there is an external product (or cross product) cohomology class u × vHi+j(X × Y,R). The cup product of classes uHi(X,R) and vHj(X,R) can be defined as the pullback of the external product by the diagonal:[12]

 

Alternatively, the external product can be defined in terms of the cup product. For spaces X and Y, write f: X × YX and g: X × YY for the two projections. Then the external product of classes uHi(X,R) and vHj(Y,R) is:

 

Poincaré duality edit

Another interpretation of Poincaré duality is that the cohomology ring of a closed oriented manifold is self-dual in a strong sense. Namely, let X be a closed connected oriented manifold of dimension n, and let F be a field. Then Hn(X,F) is isomorphic to F, and the product

 

is a perfect pairing for each integer i.[13] In particular, the vector spaces Hi(X,F) and Hni(X,F) have the same (finite) dimension. Likewise, the product on integral cohomology modulo torsion with values in Hn(X,Z) ≅ Z is a perfect pairing over Z.

Characteristic classes edit

An oriented real vector bundle E of rank r over a topological space X determines a cohomology class on X, the Euler class χ(E) ∈ Hr(X,Z). Informally, the Euler class is the class of the zero set of a general section of E. That interpretation can be made more explicit when E is a smooth vector bundle over a smooth manifold X, since then a general smooth section of X vanishes on a codimension-r submanifold of X.

There are several other types of characteristic classes for vector bundles that take values in cohomology, including Chern classes, Stiefel–Whitney classes, and Pontryagin classes.

Eilenberg–MacLane spaces edit

For each abelian group A and natural number j, there is a space   whose j-th homotopy group is isomorphic to A and whose other homotopy groups are zero. Such a space is called an Eilenberg–MacLane space. This space has the remarkable property that it is a classifying space for cohomology: there is a natural element u of  , and every cohomology class of degree j on every space X is the pullback of u by some continuous map  . More precisely, pulling back the class u gives a bijection

 

for every space X with the homotopy type of a CW complex.[14] Here   denotes the set of homotopy classes of continuous maps from X to Y.

For example, the space   (defined up to homotopy equivalence) can be taken to be the circle  . So the description above says that every element of   is pulled back from the class u of a point on   by some map  .

There is a related description of the first cohomology with coefficients in any abelian group A, say for a CW complex X. Namely,   is in one-to-one correspondence with the set of isomorphism classes of Galois covering spaces of X with group A, also called principal A-bundles over X. For X connected, it follows that   is isomorphic to  , where   is the fundamental group of X. For example,   classifies the double covering spaces of X, with the element   corresponding to the trivial double covering, the disjoint union of two copies of X.

Cap product edit

For any topological space X, the cap product is a bilinear map

 

for any integers i and j and any commutative ring R. The resulting map

 

makes the singular homology of X into a module over the singular cohomology ring of X.

For i = j, the cap product gives the natural homomorphism

 

which is an isomorphism for R a field.

For example, let X be an oriented manifold, not necessarily compact. Then a closed oriented codimension-i submanifold Y of X (not necessarily compact) determines an element of Hi(X,R), and a compact oriented j-dimensional submanifold Z of X determines an element of Hj(X,R). The cap product [Y] ∩ [Z] ∈ Hji(X,R) can be computed by perturbing Y and Z to make them intersect transversely and then taking the class of their intersection, which is a compact oriented submanifold of dimension ji.

A closed oriented manifold X of dimension n has a fundamental class [X] in Hn(X,R). The Poincaré duality isomorphism

 
is defined by cap product with the fundamental class of X.

Brief history of singular cohomology edit

Although cohomology is fundamental to modern algebraic topology, its importance was not seen for some 40 years after the development of homology. The concept of dual cell structure, which Henri Poincaré used in his proof of his Poincaré duality theorem, contained the beginning of the idea of cohomology, but this was not seen until later.

There were various precursors to cohomology.[15] In the mid-1920s, J. W. Alexander and Solomon Lefschetz founded intersection theory of cycles on manifolds. On a closed oriented n-dimensional manifold M an i-cycle and a j-cycle with nonempty intersection will, if in the general position, have as their intersection a (i + j − n)-cycle. This leads to a multiplication of homology classes

 

which (in retrospect) can be identified with the cup product on the cohomology of M.

Alexander had by 1930 defined a first notion of a cochain, by thinking of an i-cochain on a space X as a function on small neighborhoods of the diagonal in Xi+1.

In 1931, Georges de Rham related homology and differential forms, proving de Rham's theorem. This result can be stated more simply in terms of cohomology.

In 1934, Lev Pontryagin proved the Pontryagin duality theorem; a result on topological groups. This (in rather special cases) provided an interpretation of Poincaré duality and Alexander duality in terms of group characters.

At a 1935 conference in Moscow, Andrey Kolmogorov and Alexander both introduced cohomology and tried to construct a cohomology product structure.

In 1936, Norman Steenrod constructed Čech cohomology by dualizing Čech homology.

From 1936 to 1938, Hassler Whitney and Eduard Čech developed the cup product (making cohomology into a graded ring) and cap product, and realized that Poincaré duality can be stated in terms of the cap product. Their theory was still limited to finite cell complexes.

In 1944, Samuel Eilenberg overcame the technical limitations, and gave the modern definition of singular homology and cohomology.

In 1945, Eilenberg and Steenrod stated the axioms defining a homology or cohomology theory, discussed below. In their 1952 book, Foundations of Algebraic Topology, they proved that the existing homology and cohomology theories did indeed satisfy their axioms.

In 1946, Jean Leray defined sheaf cohomology.

In 1948 Edwin Spanier, building on work of Alexander and Kolmogorov, developed Alexander–Spanier cohomology.

Sheaf cohomology edit

Sheaf cohomology is a rich generalization of singular cohomology, allowing more general "coefficients" than simply an abelian group. For every sheaf of abelian groups E on a topological space X, one has cohomology groups Hi(X,E) for integers i. In particular, in the case of the constant sheaf on X associated with an abelian group A, the resulting groups Hi(X,A) coincide with singular cohomology for X a manifold or CW complex (though not for arbitrary spaces X). Starting in the 1950s, sheaf cohomology has become a central part of algebraic geometry and complex analysis, partly because of the importance of the sheaf of regular functions or the sheaf of holomorphic functions.

Grothendieck elegantly defined and characterized sheaf cohomology in the language of homological algebra. The essential point is to fix the space X and think of sheaf cohomology as a functor from the abelian category of sheaves on X to abelian groups. Start with the functor taking a sheaf E on X to its abelian group of global sections over X, E(X). This functor is left exact, but not necessarily right exact. Grothendieck defined sheaf cohomology groups to be the right derived functors of the left exact functor EE(X).[16]

That definition suggests various generalizations. For example, one can define the cohomology of a topological space X with coefficients in any complex of sheaves, earlier called hypercohomology (but usually now just "cohomology"). From that point of view, sheaf cohomology becomes a sequence of functors from the derived category of sheaves on X to abelian groups.

In a broad sense of the word, "cohomology" is often used for the right derived functors of a left exact functor on an abelian category, while "homology" is used for the left derived functors of a right exact functor. For example, for a ring R, the Tor groups ToriR(M,N) form a "homology theory" in each variable, the left derived functors of the tensor product MRN of R-modules. Likewise, the Ext groups ExtiR(M,N) can be viewed as a "cohomology theory" in each variable, the right derived functors of the Hom functor HomR(M,N).

Sheaf cohomology can be identified with a type of Ext group. Namely, for a sheaf E on a topological space X, Hi(X,E) is isomorphic to Exti(ZX, E), where ZX denotes the constant sheaf associated with the integers Z, and Ext is taken in the abelian category of sheaves on X.

Cohomology of varieties edit

There are numerous machines built for computing the cohomology of algebraic varieties. The simplest case being the determination of cohomology for smooth projective varieties over a field of characteristic  . Tools from Hodge theory, called Hodge structures help give computations of cohomology of these types of varieties (with the addition of more refined information). In the simplest case the cohomology of a smooth hypersurface in   can be determined from the degree of the polynomial alone.

When considering varieties over a finite field, or a field of characteristic  , more powerful tools are required because the classical definitions of homology/cohomology break down. This is because varieties over finite fields will only be a finite set of points. Grothendieck came up with the idea for a Grothendieck topology and used sheaf cohomology over the étale topology to define the cohomology theory for varieties over a finite field. Using the étale topology for a variety over a field of characteristic   one can construct  -adic cohomology for  . This is defined as

 

If we have a scheme of finite type

 

then there is an equality of dimensions for the Betti cohomology of   and the  -adic cohomology of   whenever the variety is smooth over both fields. In addition to these cohomology theories there are other cohomology theories called Weil cohomology theories which behave similarly to singular cohomology. There is a conjectured theory of motives which underlie all of the Weil cohomology theories.

Another useful computational tool is the blowup sequence. Given a codimension   subscheme   there is a Cartesian square

 

From this there is an associated long exact sequence

 

If the subvariety   is smooth, then the connecting morphisms are all trivial, hence

 

Axioms and generalized cohomology theories edit

There are various ways to define cohomology for topological spaces (such as singular cohomology, Čech cohomology, Alexander–Spanier cohomology or sheaf cohomology). (Here sheaf cohomology is considered only with coefficients in a constant sheaf.) These theories give different answers for some spaces, but there is a large class of spaces on which they all agree. This is most easily understood axiomatically: there is a list of properties known as the Eilenberg–Steenrod axioms, and any two constructions that share those properties will agree at least on all CW complexes.[17] There are versions of the axioms for a homology theory as well as for a cohomology theory. Some theories can be viewed as tools for computing singular cohomology for special topological spaces, such as simplicial cohomology for simplicial complexes, cellular cohomology for CW complexes, and de Rham cohomology for smooth manifolds.

One of the Eilenberg–Steenrod axioms for a cohomology theory is the dimension axiom: if P is a single point, then Hi(P) = 0 for all i ≠ 0. Around 1960, George W. Whitehead observed that it is fruitful to omit the dimension axiom completely: this gives the notion of a generalized homology theory or a generalized cohomology theory, defined below. There are generalized cohomology theories such as K-theory or complex cobordism that give rich information about a topological space, not directly accessible from singular cohomology. (In this context, singular cohomology is often called "ordinary cohomology".)

By definition, a generalized homology theory is a sequence of functors hi (for integers i) from the category of CW-pairs (XA) (so X is a CW complex and A is a subcomplex) to the category of abelian groups, together with a natural transformation i: hi(X, A) → hi−1(A) called the boundary homomorphism (here hi−1(A) is a shorthand for hi−1(A,∅)). The axioms are:

  1. Homotopy: If   is homotopic to  , then the induced homomorphisms on homology are the same.
  2. Exactness: Each pair (X,A) induces a long exact sequence in homology, via the inclusions f: AX and g: (X,∅) → (X,A):
     
  3. Excision: If X is the union of subcomplexes A and B, then the inclusion f: (A,AB) → (X,B) induces an isomorphism
     
    for every i.
  4. Additivity: If (X,A) is the disjoint union of a set of pairs (Xα,Aα), then the inclusions (Xα,Aα) → (X,A) induce an isomorphism from the direct sum:
     
    for every i.

The axioms for a generalized cohomology theory are obtained by reversing the arrows, roughly speaking. In more detail, a generalized cohomology theory is a sequence of contravariant functors hi (for integers i) from the category of CW-pairs to the category of abelian groups, together with a natural transformation d: hi(A) → hi+1(X,A) called the boundary homomorphism (writing hi(A) for hi(A,∅)). The axioms are:

  1. Homotopy: Homotopic maps induce the same homomorphism on cohomology.
  2. Exactness: Each pair (X,A) induces a long exact sequence in cohomology, via the inclusions f: AX and g: (X,∅) → (X,A):
     
  3. Excision: If X is the union of subcomplexes A and B, then the inclusion f: (A,AB) → (X,B) induces an isomorphism
     
    for every i.
  4. Additivity: If (X,A) is the disjoint union of a set of pairs (Xα,Aα), then the inclusions (Xα,Aα) → (X,A) induce an isomorphism to the product group:
     
    for every i.

A spectrum determines both a generalized homology theory and a generalized cohomology theory. A fundamental result by Brown, Whitehead, and Adams says that every generalized homology theory comes from a spectrum, and likewise every generalized cohomology theory comes from a spectrum.[18] This generalizes the representability of ordinary cohomology by Eilenberg–MacLane spaces.

A subtle point is that the functor from the stable homotopy category (the homotopy category of spectra) to generalized homology theories on CW-pairs is not an equivalence, although it gives a bijection on isomorphism classes; there are nonzero maps in the stable homotopy category (called phantom maps) that induce the zero map between homology theories on CW-pairs. Likewise, the functor from the stable homotopy category to generalized cohomology theories on CW-pairs is not an equivalence.[19] It is the stable homotopy category, not these other categories, that has good properties such as being triangulated.

If one prefers homology or cohomology theories to be defined on all topological spaces rather than on CW complexes, one standard approach is to include the axiom that every weak homotopy equivalence induces an isomorphism on homology or cohomology. (That is true for singular homology or singular cohomology, but not for sheaf cohomology, for example.) Since every space admits a weak homotopy equivalence from a CW complex, this axiom reduces homology or cohomology theories on all spaces to the corresponding theory on CW complexes.[20]

Some examples of generalized cohomology theories are:

  • Stable cohomotopy groups   The corresponding homology theory is used more often: stable homotopy groups  
  • Various different flavors of cobordism groups, based on studying a space by considering all maps from it to manifolds: unoriented cobordism   oriented cobordism   complex cobordism   and so on. Complex cobordism has turned out to be especially powerful in homotopy theory. It is closely related to formal groups, via a theorem of Daniel Quillen.
  • Various different flavors of topological K-theory, based on studying a space by considering all vector bundles over it:   (real periodic K-theory),   (real connective K-theory),   (complex periodic K-theory),   (complex connective K-theory), and so on.
  • Brown–Peterson cohomology, Morava K-theory, Morava E-theory, and other theories built from complex cobordism.
  • Various flavors of elliptic cohomology.

Many of these theories carry richer information than ordinary cohomology, but are harder to compute.

A cohomology theory E is said to be multiplicative if   has the structure of a graded ring for each space X. In the language of spectra, there are several more precise notions of a ring spectrum, such as an E ring spectrum, where the product is commutative and associative in a strong sense.

Other cohomology theories edit

Cohomology theories in a broader sense (invariants of other algebraic or geometric structures, rather than of topological spaces) include:

See also edit

Citations edit

  1. ^ Hatcher 2001, p. 108.
  2. ^ Hatcher (2001), Theorem 3.5; Dold (1972), Proposition VIII.3.3 and Corollary VIII.3.4.
  3. ^ Dold 1972, Propositions IV.8.12 and V.4.11.
  4. ^ Hatcher 2001, Theorem 3.11.
  5. ^ Thom 1954, pp. 62–63.
  6. ^ Thom 1954, Theorem II.29.
  7. ^ Hatcher 2001, Example 3.16.
  8. ^ Hatcher 2001, Theorem 3.15.
  9. ^ a b Hatcher 2001, Theorem 3.19.
  10. ^ Hatcher 2001, p. 222.
  11. ^ Hatcher 2001, Example 3.7.
  12. ^ Hatcher 2001, p. 186.
  13. ^ Hatcher 2001, Proposition 3.38.
  14. ^ May 1999, p. 177.
  15. ^ Dieudonné 1989, Section IV.3.
  16. ^ Hartshorne 1977, Section III.2.
  17. ^ May 1999, p. 95.
  18. ^ Switzer 1975, p. 117, 331, Theorem 9.27; Corollary 14.36; Remarks.
  19. ^ "Are spectra really the same as cohomology theories?". MathOverflow.
  20. ^ Switzer 1975, 7.68.

References edit

cohomology, mathematics, specifically, homology, theory, algebraic, topology, cohomology, general, term, sequence, abelian, groups, usually, associated, with, topological, space, often, defined, from, cochain, complex, viewed, method, assigning, richer, algebr. 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 From its start in topology this idea became a dominant method in the mathematics of the second half of the twentieth century From the initial idea of homology as a method of constructing algebraic invariants of topological spaces the range of applications of homology and cohomology theories has spread throughout geometry and algebra The terminology tends to hide the fact that cohomology a contravariant theory is more natural than homology in many applications At a basic level this has to do with functions and pullbacks in geometric situations given spaces X and Y and some kind of function F on Y for any mapping f X Y composition with f gives rise to a function F f on X The most important cohomology theories have a product the cup product which gives them a ring structure Because of this feature cohomology is usually a stronger invariant than homology Contents 1 Singular cohomology 2 Examples 3 The diagonal 4 Poincare duality 5 Characteristic classes 6 Eilenberg MacLane spaces 7 Cap product 8 Brief history of singular cohomology 9 Sheaf cohomology 10 Cohomology of varieties 11 Axioms and generalized cohomology theories 12 Other cohomology theories 13 See also 14 Citations 15 ReferencesSingular cohomology editSingular cohomology is a powerful invariant in topology associating a graded commutative ring with any topological space Every continuous map f X Y determines a homomorphism from the cohomology ring of Y to that of X this puts strong restrictions on the possible maps from X to Y Unlike more subtle invariants such as homotopy groups the cohomology ring tends to be computable in practice for spaces of interest For a topological space X the definition of singular cohomology starts with the singular chain complex 1 Ci 1 i 1Ci i Ci 1 displaystyle cdots to C i 1 stackrel partial i 1 to C i stackrel partial i to C i 1 to cdots nbsp By definition the singular homology of X is the homology of this chain complex the kernel of one homomorphism modulo the image of the previous one In more detail Ci is the free abelian group on the set of continuous maps from the standard i simplex to X called singular i simplices in X and i is the i th boundary homomorphism The groups Ci are zero for i negative Now fix an abelian group A and replace each group Ci by its dual group Ci Hom Ci A displaystyle C i mathrm Hom C i A nbsp and i displaystyle partial i nbsp by its dual homomorphismdi 1 Ci 1 Ci displaystyle d i 1 C i 1 to C i nbsp This has the effect of reversing all the arrows of the original complex leaving a cochain complex Ci 1 di Ci di 1Ci 1 displaystyle cdots leftarrow C i 1 stackrel d i leftarrow C i stackrel d i 1 leftarrow C i 1 leftarrow cdots nbsp For an integer i the ith cohomology group of X with coefficients in A is defined to be ker di im di 1 and denoted by Hi X A The group Hi X A is zero for i negative The elements of Ci displaystyle C i nbsp are called singular i cochains with coefficients in A Equivalently an i cochain on X can be identified with a function from the set of singular i simplices in X to A Elements of ker d and im d are called cocycles and coboundaries respectively while elements of ker d im d Hi X A are called cohomology classes because they are equivalence classes of cocycles In what follows the coefficient group A is sometimes not written It is common to take A to be a commutative ring R then the cohomology groups are R modules A standard choice is the ring Z of integers Some of the formal properties of cohomology are only minor variants of the properties of homology A continuous map f X Y displaystyle f X to Y nbsp determines a pushforward homomorphism f Hi X Hi Y displaystyle f H i X to H i Y nbsp on homology and a pullback homomorphism f Hi Y Hi X displaystyle f H i Y to H i X nbsp on cohomology This makes cohomology into a contravariant functor from topological spaces to abelian groups or R modules Two homotopic maps from X to Y induce the same homomorphism on cohomology just as on homology The Mayer Vietoris sequence is an important computational tool in cohomology as in homology Note that the boundary homomorphism increases rather than decreases degree in cohomology That is if a space X is the union of open subsets U and V then there is a long exact sequence Hi X Hi U Hi V Hi U V Hi 1 X displaystyle cdots to H i X to H i U oplus H i V to H i U cap V to H i 1 X to cdots nbsp There are relative cohomology groups Hi X Y A displaystyle H i X Y A nbsp for any subspace Y of a space X They are related to the usual cohomology groups by a long exact sequence Hi X Y Hi X Hi Y Hi 1 X Y displaystyle cdots to H i X Y to H i X to H i Y to H i 1 X Y to cdots nbsp The universal coefficient theorem describes cohomology in terms of homology using Ext groups Namely there is a short exact sequence 0 ExtZ1 Hi 1 X Z A Hi X A HomZ Hi X Z A 0 displaystyle 0 to operatorname Ext mathbb Z 1 operatorname H i 1 X mathbb Z A to H i X A to operatorname Hom mathbb Z H i X mathbb Z A to 0 nbsp A related statement is that for a field F Hi X F displaystyle H i X F nbsp is precisely the dual space of the vector space Hi X F displaystyle H i X F nbsp If X is a topological manifold or a CW complex then the cohomology groups Hi X A displaystyle H i X A nbsp are zero for i greater than the dimension of X 2 If X is a compact manifold possibly with boundary or a CW complex with finitely many cells in each dimension and R is a commutative Noetherian ring then the R module Hi X R is finitely generated for each i 3 On the other hand cohomology has a crucial structure that homology does not for any topological space X and commutative ring R there is a bilinear map called the cup product Hi X R Hj X R Hi j X R displaystyle H i X R times H j X R to H i j X R nbsp defined by an explicit formula on singular cochains The product of cohomology classes u and v is written as u v or simply as uv This product makes the direct sum H X R iHi X R displaystyle H X R bigoplus i H i X R nbsp into a graded ring called the cohomology ring of X It is graded commutative in the sense that 4 uv 1 ijvu u Hi X R v Hj X R displaystyle uv 1 ij vu qquad u in H i X R v in H j X R nbsp For any continuous map f X Y displaystyle f colon X to Y nbsp the pullback f H Y R H X R displaystyle f H Y R to H X R nbsp is a homomorphism of graded R algebras It follows that if two spaces are homotopy equivalent then their cohomology rings are isomorphic Here are some of the geometric interpretations of the cup product In what follows manifolds are understood to be without boundary unless stated otherwise A closed manifold means a compact manifold without boundary whereas a closed submanifold N of a manifold M means a submanifold that is a closed subset of M not necessarily compact although N is automatically compact if M is Let X be a closed oriented manifold of dimension n Then Poincare duality gives an isomorphism HiX Hn iX As a result a closed oriented submanifold S of codimension i in X determines a cohomology class in HiX called S In these terms the cup product describes the intersection of submanifolds Namely if S and T are submanifolds of codimension i and j that intersect transversely then S T S T Hi j X displaystyle S T S cap T in H i j X nbsp where the intersection S T is a submanifold of codimension i j with an orientation determined by the orientations of S T and X In the case of smooth manifolds if S and T do not intersect transversely this formula can still be used to compute the cup product S T by perturbing S or T to make the intersection transverse More generally without assuming that X has an orientation a closed submanifold of X with an orientation on its normal bundle determines a cohomology class on X If X is a noncompact manifold then a closed submanifold not necessarily compact determines a cohomology class on X In both cases the cup product can again be described in terms of intersections of submanifolds Note that Thom constructed an integral cohomology class of degree 7 on a smooth 14 manifold that is not the class of any smooth submanifold 5 On the other hand he showed that every integral cohomology class of positive degree on a smooth manifold has a positive multiple that is the class of a smooth submanifold 6 Also every integral cohomology class on a manifold can be represented by a pseudomanifold that is a simplicial complex that is a manifold outside a closed subset of codimension at least 2 For a smooth manifold X de Rham s theorem says that the singular cohomology of X with real coefficients is isomorphic to the de Rham cohomology of X defined using differential forms The cup product corresponds to the product of differential forms This interpretation has the advantage that the product on differential forms is graded commutative whereas the product on singular cochains is only graded commutative up to chain homotopy In fact it is impossible to modify the definition of singular cochains with coefficients in the integers Z displaystyle mathbb Z nbsp or in Z p displaystyle mathbb Z p nbsp for a prime number p to make the product graded commutative on the nose The failure of graded commutativity at the cochain level leads to the Steenrod operations on mod p cohomology Very informally for any topological space X elements of Hi X displaystyle H i X nbsp can be thought of as represented by codimension i subspaces of X that can move freely on X For example one way to define an element of Hi X displaystyle H i X nbsp is to give a continuous map f from X to a manifold M and a closed codimension i submanifold N of M with an orientation on the normal bundle Informally one thinks of the resulting class f N Hi X displaystyle f N in H i X nbsp as lying on the subspace f 1 N displaystyle f 1 N nbsp of X this is justified in that the class f N displaystyle f N nbsp restricts to zero in the cohomology of the open subset X f 1 N displaystyle X f 1 N nbsp The cohomology class f N displaystyle f N nbsp can move freely on X in the sense that N could be replaced by any continuous deformation of N inside M Examples editIn what follows cohomology is taken with coefficients in the integers Z unless stated otherwise The cohomology ring of a point is the ring Z in degree 0 By homotopy invariance this is also the cohomology ring of any contractible space such as Euclidean space Rn nbsp The first cohomology group of the 2 dimensional torus has a basis given by the classes of the two circles shown For a positive integer n the cohomology ring of the sphere Sn displaystyle S n nbsp is Z x x2 the quotient ring of a polynomial ring by the given ideal with x in degree n In terms of Poincare duality as above x is the class of a point on the sphere The cohomology ring of the torus S1 n displaystyle S 1 n nbsp is the exterior algebra over Z on n generators in degree 1 7 For example let P denote a point in the circle S1 displaystyle S 1 nbsp and Q the point P P in the 2 dimensional torus S1 2 displaystyle S 1 2 nbsp Then the cohomology of S1 2 has a basis as a free Z module of the form the element 1 in degree 0 x P S1 and y S1 P in degree 1 and xy Q in degree 2 Implicitly orientations of the torus and of the two circles have been fixed here Note that yx xy Q by graded commutativity More generally let R be a commutative ring and let X and Y be any topological spaces such that H X R is a finitely generated free R module in each degree No assumption is needed on Y Then the Kunneth formula gives that the cohomology ring of the product space X Y is a tensor product of R algebras 8 H X Y R H X R RH Y R displaystyle H X times Y R cong H X R otimes R H Y R nbsp The cohomology ring of real projective space RPn with Z 2 coefficients is Z 2 x xn 1 with x in degree 1 9 Here x is the class of a hyperplane RPn 1 in RPn this makes sense even though RPj is not orientable for j even and positive because Poincare duality with Z 2 coefficients works for arbitrary manifolds With integer coefficients the answer is a bit more complicated The Z cohomology of RP2a has an element y of degree 2 such that the whole cohomology is the direct sum of a copy of Z spanned by the element 1 in degree 0 together with copies of Z 2 spanned by the elements yi for i 1 a The Z cohomology of RP2a 1 is the same together with an extra copy of Z in degree 2a 1 10 The cohomology ring of complex projective space CPn is Z x xn 1 with x in degree 2 9 Here x is the class of a hyperplane CPn 1 in CPn More generally xj is the class of a linear subspace CPn j in CPn The cohomology ring of the closed oriented surface X of genus g 0 has a basis as a free Z module of the form the element 1 in degree 0 A1 Ag and B1 Bg in degree 1 and the class P of a point in degree 2 The product is given by AiAj BiBj 0 for all i and j AiBj 0 if i j and AiBi P for all i 11 By graded commutativity it follows that BiAi P On any topological space graded commutativity of the cohomology ring implies that 2x2 0 for all odd degree cohomology classes x It follows that for a ring R containing 1 2 all odd degree elements of H X R have square zero On the other hand odd degree elements need not have square zero if R is Z 2 or Z as one sees in the example of RP2 with Z 2 coefficients or RP4 RP2 with Z coefficients The diagonal editThe cup product on cohomology can be viewed as coming from the diagonal map D X X X x x x Namely for any spaces X and Y with cohomology classes u Hi X R and v Hj Y R there is an external product or cross product cohomology class u v Hi j X Y R The cup product of classes u Hi X R and v Hj X R can be defined as the pullback of the external product by the diagonal 12 uv D u v Hi j X R displaystyle uv Delta u times v in H i j X R nbsp Alternatively the external product can be defined in terms of the cup product For spaces X and Y write f X Y X and g X Y Y for the two projections Then the external product of classes u Hi X R and v Hj Y R is u v f u g v Hi j X Y R displaystyle u times v f u g v in H i j X times Y R nbsp Poincare duality editMain article Poincare duality Another interpretation of Poincare duality is that the cohomology ring of a closed oriented manifold is self dual in a strong sense Namely let X be a closed connected oriented manifold of dimension n and let F be a field Then Hn X F is isomorphic to F and the product Hi X F Hn i X F Hn X F F displaystyle H i X F times H n i X F to H n X F cong F nbsp is a perfect pairing for each integer i 13 In particular the vector spaces Hi X F and Hn i X F have the same finite dimension Likewise the product on integral cohomology modulo torsion with values in Hn X Z Z is a perfect pairing over Z Characteristic classes editMain article Characteristic class An oriented real vector bundle E of rank r over a topological space X determines a cohomology class on X the Euler class x E Hr X Z Informally the Euler class is the class of the zero set of a general section of E That interpretation can be made more explicit when E is a smooth vector bundle over a smooth manifold X since then a general smooth section of X vanishes on a codimension r submanifold of X There are several other types of characteristic classes for vector bundles that take values in cohomology including Chern classes Stiefel Whitney classes and Pontryagin classes Eilenberg MacLane spaces editMain article Eilenberg MacLane space For each abelian group A and natural number j there is a space K A j displaystyle K A j nbsp whose j th homotopy group is isomorphic to A and whose other homotopy groups are zero Such a space is called an Eilenberg MacLane space This space has the remarkable property that it is a classifying space for cohomology there is a natural element u of Hj K A j A displaystyle H j K A j A nbsp and every cohomology class of degree j on every space X is the pullback of u by some continuous map X K A j displaystyle X to K A j nbsp More precisely pulling back the class u gives a bijection X K A j Hj X A displaystyle X K A j stackrel cong to H j X A nbsp for every space X with the homotopy type of a CW complex 14 Here X Y displaystyle X Y nbsp denotes the set of homotopy classes of continuous maps from X to Y For example the space K Z 1 displaystyle K mathbb Z 1 nbsp defined up to homotopy equivalence can be taken to be the circle S1 displaystyle S 1 nbsp So the description above says that every element of H1 X Z displaystyle H 1 X mathbb Z nbsp is pulled back from the class u of a point on S1 displaystyle S 1 nbsp by some map X S1 displaystyle X to S 1 nbsp There is a related description of the first cohomology with coefficients in any abelian group A say for a CW complex X Namely H1 X A displaystyle H 1 X A nbsp is in one to one correspondence with the set of isomorphism classes of Galois covering spaces of X with group A also called principal A bundles over X For X connected it follows that H1 X A displaystyle H 1 X A nbsp is isomorphic to Hom p1 X A displaystyle operatorname Hom pi 1 X A nbsp where p1 X displaystyle pi 1 X nbsp is the fundamental group of X For example H1 X Z 2 displaystyle H 1 X mathbb Z 2 nbsp classifies the double covering spaces of X with the element 0 H1 X Z 2 displaystyle 0 in H 1 X mathbb Z 2 nbsp corresponding to the trivial double covering the disjoint union of two copies of X Cap product editMain article Cap product For any topological space X the cap product is a bilinear map Hi X R Hj X R Hj i X R displaystyle cap H i X R times H j X R to H j i X R nbsp for any integers i and j and any commutative ring R The resulting map H X R H X R H X R displaystyle H X R times H X R to H X R nbsp makes the singular homology of X into a module over the singular cohomology ring of X For i j the cap product gives the natural homomorphism Hi X R HomR Hi X R R displaystyle H i X R to operatorname Hom R H i X R R nbsp which is an isomorphism for R a field For example let X be an oriented manifold not necessarily compact Then a closed oriented codimension i submanifold Y of X not necessarily compact determines an element of Hi X R and a compact oriented j dimensional submanifold Z of X determines an element of Hj X R The cap product Y Z Hj i X R can be computed by perturbing Y and Z to make them intersect transversely and then taking the class of their intersection which is a compact oriented submanifold of dimension j i A closed oriented manifold X of dimension n has a fundamental class X in Hn X R The Poincare duality isomorphismHi X R Hn i X R displaystyle H i X R overset cong to H n i X R nbsp is defined by cap product with the fundamental class of X Brief history of singular cohomology editAlthough cohomology is fundamental to modern algebraic topology its importance was not seen for some 40 years after the development of homology The concept of dual cell structure which Henri Poincare used in his proof of his Poincare duality theorem contained the beginning of the idea of cohomology but this was not seen until later There were various precursors to cohomology 15 In the mid 1920s J W Alexander and Solomon Lefschetz founded intersection theory of cycles on manifolds On a closed oriented n dimensional manifold M an i cycle and a j cycle with nonempty intersection will if in the general position have as their intersection a i j n cycle This leads to a multiplication of homology classes Hi M Hj M Hi j n M displaystyle H i M times H j M to H i j n M nbsp which in retrospect can be identified with the cup product on the cohomology of M Alexander had by 1930 defined a first notion of a cochain by thinking of an i cochain on a space X as a function on small neighborhoods of the diagonal in Xi 1 In 1931 Georges de Rham related homology and differential forms proving de Rham s theorem This result can be stated more simply in terms of cohomology In 1934 Lev Pontryagin proved the Pontryagin duality theorem a result on topological groups This in rather special cases provided an interpretation of Poincare duality and Alexander duality in terms of group characters At a 1935 conference in Moscow Andrey Kolmogorov and Alexander both introduced cohomology and tried to construct a cohomology product structure In 1936 Norman Steenrod constructed Cech cohomology by dualizing Cech homology From 1936 to 1938 Hassler Whitney and Eduard Cech developed the cup product making cohomology into a graded ring and cap product and realized that Poincare duality can be stated in terms of the cap product Their theory was still limited to finite cell complexes In 1944 Samuel Eilenberg overcame the technical limitations and gave the modern definition of singular homology and cohomology In 1945 Eilenberg and Steenrod stated the axioms defining a homology or cohomology theory discussed below In their 1952 book Foundations of Algebraic Topology they proved that the existing homology and cohomology theories did indeed satisfy their axioms In 1946 Jean Leray defined sheaf cohomology In 1948 Edwin Spanier building on work of Alexander and Kolmogorov developed Alexander Spanier cohomology Sheaf cohomology editMain article Sheaf cohomology Sheaf cohomology is a rich generalization of singular cohomology allowing more general coefficients than simply an abelian group For every sheaf of abelian groups E on a topological space X one has cohomology groups Hi X E for integers i In particular in the case of the constant sheaf on X associated with an abelian group A the resulting groups Hi X A coincide with singular cohomology for X a manifold or CW complex though not for arbitrary spaces X Starting in the 1950s sheaf cohomology has become a central part of algebraic geometry and complex analysis partly because of the importance of the sheaf of regular functions or the sheaf of holomorphic functions Grothendieck elegantly defined and characterized sheaf cohomology in the language of homological algebra The essential point is to fix the space X and think of sheaf cohomology as a functor from the abelian category of sheaves on X to abelian groups Start with the functor taking a sheaf E on X to its abelian group of global sections over X E X This functor is left exact but not necessarily right exact Grothendieck defined sheaf cohomology groups to be the right derived functors of the left exact functor E E X 16 That definition suggests various generalizations For example one can define the cohomology of a topological space X with coefficients in any complex of sheaves earlier called hypercohomology but usually now just cohomology From that point of view sheaf cohomology becomes a sequence of functors from the derived category of sheaves on X to abelian groups In a broad sense of the word cohomology is often used for the right derived functors of a left exact functor on an abelian category while homology is used for the left derived functors of a right exact functor For example for a ring R the Tor groups ToriR M N form a homology theory in each variable the left derived functors of the tensor product M RN of R modules Likewise the Ext groups ExtiR M N can be viewed as a cohomology theory in each variable the right derived functors of the Hom functor HomR M N Sheaf cohomology can be identified with a type of Ext group Namely for a sheaf E on a topological space X Hi X E is isomorphic to Exti ZX E where ZX denotes the constant sheaf associated with the integers Z and Ext is taken in the abelian category of sheaves on X Cohomology of varieties editThere are numerous machines built for computing the cohomology of algebraic varieties The simplest case being the determination of cohomology for smooth projective varieties over a field of characteristic 0 displaystyle 0 nbsp Tools from Hodge theory called Hodge structures help give computations of cohomology of these types of varieties with the addition of more refined information In the simplest case the cohomology of a smooth hypersurface in Pn displaystyle mathbb P n nbsp can be determined from the degree of the polynomial alone When considering varieties over a finite field or a field of characteristic p displaystyle p nbsp more powerful tools are required because the classical definitions of homology cohomology break down This is because varieties over finite fields will only be a finite set of points Grothendieck came up with the idea for a Grothendieck topology and used sheaf cohomology over the etale topology to define the cohomology theory for varieties over a finite field Using the etale topology for a variety over a field of characteristic p displaystyle p nbsp one can construct ℓ displaystyle ell nbsp adic cohomology for ℓ p displaystyle ell neq p nbsp This is defined as Hk X Qℓ lim Hetk X Z ℓn ZℓQℓ displaystyle H k X mathbb Q ell varprojlim H et k X mathbb Z ell n otimes mathbb Z ell mathbb Q ell nbsp If we have a scheme of finite type X Proj Z x0 xn f1 fk displaystyle X text Proj left frac mathbb Z left x 0 ldots x n right left f 1 ldots f k right right nbsp then there is an equality of dimensions for the Betti cohomology of X C displaystyle X mathbb C nbsp and the ℓ displaystyle ell nbsp adic cohomology of X Fq displaystyle X mathbb F q nbsp whenever the variety is smooth over both fields In addition to these cohomology theories there are other cohomology theories called Weil cohomology theories which behave similarly to singular cohomology There is a conjectured theory of motives which underlie all of the Weil cohomology theories Another useful computational tool is the blowup sequence Given a codimension 2 displaystyle geq 2 nbsp subscheme Z X displaystyle Z subset X nbsp there is a Cartesian square E BlZ X Z X displaystyle begin matrix E amp longrightarrow amp Bl Z X downarrow amp amp downarrow Z amp longrightarrow amp X end matrix nbsp From this there is an associated long exact sequence Hn X Hn Z Hn BlZ X Hn E Hn 1 X displaystyle cdots to H n X to H n Z oplus H n Bl Z X to H n E to H n 1 X to cdots nbsp If the subvariety Z displaystyle Z nbsp is smooth then the connecting morphisms are all trivial hence Hn BlZ X Hn Z Hn X Hn E displaystyle H n Bl Z X oplus H n Z cong H n X oplus H n E nbsp Axioms and generalized cohomology theories editSee also List of cohomology theories There are various ways to define cohomology for topological spaces such as singular cohomology Cech cohomology Alexander Spanier cohomology or sheaf cohomology Here sheaf cohomology is considered only with coefficients in a constant sheaf These theories give different answers for some spaces but there is a large class of spaces on which they all agree This is most easily understood axiomatically there is a list of properties known as the Eilenberg Steenrod axioms and any two constructions that share those properties will agree at least on all CW complexes 17 There are versions of the axioms for a homology theory as well as for a cohomology theory Some theories can be viewed as tools for computing singular cohomology for special topological spaces such as simplicial cohomology for simplicial complexes cellular cohomology for CW complexes and de Rham cohomology for smooth manifolds One of the Eilenberg Steenrod axioms for a cohomology theory is the dimension axiom if P is a single point then Hi P 0 for all i 0 Around 1960 George W Whitehead observed that it is fruitful to omit the dimension axiom completely this gives the notion of a generalized homology theory or a generalized cohomology theory defined below There are generalized cohomology theories such as K theory or complex cobordism that give rich information about a topological space not directly accessible from singular cohomology In this context singular cohomology is often called ordinary cohomology By definition a generalized homology theory is a sequence of functors hi for integers i from the category of CW pairs X A so X is a CW complex and A is a subcomplex to the category of abelian groups together with a natural transformation i hi X A hi 1 A called the boundary homomorphism here hi 1 A is a shorthand for hi 1 A The axioms are Homotopy If f X A Y B displaystyle f X A to Y B nbsp is homotopic to g X A Y B displaystyle g X A to Y B nbsp then the induced homomorphisms on homology are the same Exactness Each pair X A induces a long exact sequence in homology via the inclusions f A X and g X X A hi A f hi X g hi X A hi 1 A displaystyle cdots to h i A overset f to h i X overset g to h i X A overset partial to h i 1 A to cdots nbsp Excision If X is the union of subcomplexes A and B then the inclusion f A A B X B induces an isomorphism hi A A B f hi X B displaystyle h i A A cap B overset f to h i X B nbsp for every i Additivity If X A is the disjoint union of a set of pairs Xa Aa then the inclusions Xa Aa X A induce an isomorphism from the direct sum ahi Xa Aa hi X A displaystyle bigoplus alpha h i X alpha A alpha to h i X A nbsp for every i The axioms for a generalized cohomology theory are obtained by reversing the arrows roughly speaking In more detail a generalized cohomology theory is a sequence of contravariant functors hi for integers i from the category of CW pairs to the category of abelian groups together with a natural transformation d hi A hi 1 X A called the boundary homomorphism writing hi A for hi A The axioms are Homotopy Homotopic maps induce the same homomorphism on cohomology Exactness Each pair X A induces a long exact sequence in cohomology via the inclusions f A X and g X X A hi X A g hi X f hi A dhi 1 X A displaystyle cdots to h i X A overset g to h i X overset f to h i A overset d to h i 1 X A to cdots nbsp Excision If X is the union of subcomplexes A and B then the inclusion f A A B X B induces an isomorphism hi X B f hi A A B displaystyle h i X B overset f to h i A A cap B nbsp for every i Additivity If X A is the disjoint union of a set of pairs Xa Aa then the inclusions Xa Aa X A induce an isomorphism to the product group hi X A ahi Xa Aa displaystyle h i X A to prod alpha h i X alpha A alpha nbsp for every i A spectrum determines both a generalized homology theory and a generalized cohomology theory A fundamental result by Brown Whitehead and Adams says that every generalized homology theory comes from a spectrum and likewise every generalized cohomology theory comes from a spectrum 18 This generalizes the representability of ordinary cohomology by Eilenberg MacLane spaces A subtle point is that the functor from the stable homotopy category the homotopy category of spectra to generalized homology theories on CW pairs is not an equivalence although it gives a bijection on isomorphism classes there are nonzero maps in the stable homotopy category called phantom maps that induce the zero map between homology theories on CW pairs Likewise the functor from the stable homotopy category to generalized cohomology theories on CW pairs is not an equivalence 19 It is the stable homotopy category not these other categories that has good properties such as being triangulated If one prefers homology or cohomology theories to be defined on all topological spaces rather than on CW complexes one standard approach is to include the axiom that every weak homotopy equivalence induces an isomorphism on homology or cohomology That is true for singular homology or singular cohomology but not for sheaf cohomology for example Since every space admits a weak homotopy equivalence from a CW complex this axiom reduces homology or cohomology theories on all spaces to the corresponding theory on CW complexes 20 Some examples of generalized cohomology theories are Stable cohomotopy groups pS X displaystyle pi S X nbsp The corresponding homology theory is used more often stable homotopy groups p S X displaystyle pi S X nbsp Various different flavors of cobordism groups based on studying a space by considering all maps from it to manifolds unoriented cobordism MO X displaystyle MO X nbsp oriented cobordism MSO X displaystyle MSO X nbsp complex cobordism MU X displaystyle MU X nbsp and so on Complex cobordism has turned out to be especially powerful in homotopy theory It is closely related to formal groups via a theorem of Daniel Quillen Various different flavors of topological K theory based on studying a space by considering all vector bundles over it KO X displaystyle KO X nbsp real periodic K theory ko X displaystyle ko X nbsp real connective K theory K X displaystyle K X nbsp complex periodic K theory ku X displaystyle ku X nbsp complex connective K theory and so on Brown Peterson cohomology Morava K theory Morava E theory and other theories built from complex cobordism Various flavors of elliptic cohomology Many of these theories carry richer information than ordinary cohomology but are harder to compute A cohomology theory E is said to be multiplicative if E X displaystyle E X nbsp has the structure of a graded ring for each space X In the language of spectra there are several more precise notions of a ring spectrum such as an E ring spectrum where the product is commutative and associative in a strong sense Other cohomology theories editCohomology theories in a broader sense invariants of other algebraic or geometric structures rather than of topological spaces include Algebraic K theory Andre Quillen cohomology Bounded cohomology BRST cohomology Cech cohomology Coherent sheaf cohomology Crystalline cohomology Cyclic cohomology Deligne cohomology Equivariant cohomology Etale cohomology Ext groups Flat cohomology Floer homology Galois cohomology Group cohomology Hochschild cohomology Intersection cohomology Khovanov homology Lie algebra cohomology Local cohomology Motivic cohomology Non abelian cohomology Quantum cohomologySee also editcomplex oriented cohomology theoryCitations edit Hatcher 2001 p 108 Hatcher 2001 Theorem 3 5 Dold 1972 Proposition VIII 3 3 and Corollary VIII 3 4 Dold 1972 Propositions IV 8 12 and V 4 11 Hatcher 2001 Theorem 3 11 Thom 1954 pp 62 63 Thom 1954 Theorem II 29 Hatcher 2001 Example 3 16 Hatcher 2001 Theorem 3 15 a b Hatcher 2001 Theorem 3 19 Hatcher 2001 p 222 Hatcher 2001 Example 3 7 Hatcher 2001 p 186 Hatcher 2001 Proposition 3 38 May 1999 p 177 Dieudonne 1989 Section IV 3 Hartshorne 1977 Section III 2 May 1999 p 95 Switzer 1975 p 117 331 Theorem 9 27 Corollary 14 36 Remarks Are spectra really the same as cohomology theories MathOverflow Switzer 1975 7 68 References editDieudonne Jean 1989 History of Algebraic and Differential Topology Birkhauser ISBN 0 8176 3388 X MR 0995842 Dold Albrecht 1972 Lectures on Algebraic Topology Springer Verlag ISBN 978 3 540 58660 9 MR 0415602 Eilenberg Samuel Steenrod Norman 1952 Foundations of Algebraic Topology Princeton University Press ISBN 9780691627236 MR 0050886 Hartshorne Robin 1977 Algebraic Geometry Graduate Texts in Mathematics vol 52 New York Heidelberg Springer Verlag ISBN 0 387 90244 9 MR 0463157 Hatcher Allen 2001 Algebraic Topology Cambridge University Press ISBN 0 521 79540 0 MR 1867354 Cohomology Encyclopedia of Mathematics EMS Press 2001 1994 May J Peter 1999 A Concise Course in Algebraic Topology PDF University of Chicago Press ISBN 0 226 51182 0 MR 1702278 Switzer Robert 1975 Algebraic Topology Homology and Homotopy Springer Verlag ISBN 3 540 42750 3 MR 0385836 Thom Rene 1954 Quelques proprietes globales des varietes differentiables Commentarii Mathematici Helvetici 28 17 86 doi 10 1007 BF02566923 MR 0061823 S2CID 120243638 Retrieved from https en wikipedia org w index php title Cohomology amp oldid 1215220665 Axioms and generalized cohomology theories, wikipedia, wiki, book, books, library,

article

, read, download, free, free download, mp3, video, mp4, 3gp, jpg, jpeg, gif, png, picture, music, song, movie, book, game, games.