We will begin this talk with the axiomatic approach towards (ordinary)Cohomology theory. We will spend most of the time on 'Cohomology Ring/algebra', It's application in Hopf division algebra, HH-spaces, how Cohomology operation motivates James construction which gives us EHP Sequence. Finally, we conclude our talk with Serre finiteness conditions, which we will try to look in the aspect of Cohomology operations.
0.1. Axiomatic ordinary Cohomology
We will quickly define cohomology over a topological space XX. We will define Delta_(q)(X)\Delta_{q}(X) be the free Abelian group generated by the continuous functions sigma:Delta^(q)rarr X\sigma: \Delta^{q} \rightarrow X. Let, sigma^((i))\sigma^{(i)} be the map sigma:Delta^(q)rarr X\sigma: \Delta^{q} \rightarrow X restricted on the face opposite to the ii-th vertex. We can define a homomorphism del_(q):Delta_(q)(X)rarrDelta_(q-1)(X)\partial_{q}: \Delta_{q}(X) \rightarrow \Delta_{q-1}(X) as,
It can be verified that, del_(q)del_(q+1)=0\partial_{q} \partial_{q+1}=0. We will have the following sequence of free Abelian groups with del_(q+1)del_(q)=0\partial_{q+1} \partial_{q}=0.
This is called Singular Chain Complex. From this chain complex we can construct a co chain complex Delta^(q)(X):=Hom(Delta_(q),Z)\Delta^{q}(X):=\operatorname{Hom}\left(\Delta_{q}, \mathbb{Z}\right). The corresponding cohomology H^(q)(X)H^{q}(X) is called Singular cohomology of XX. If we have a subspace A sub XA \subset X then Delta^(q)(X,A):=Hom(Delta^(q)(X)//Delta^(q)(A),Z)\Delta^{q}(X, A):=\operatorname{Hom}\left(\Delta^{q}(X) / \Delta^{q}(A), \mathbb{Z}\right) forms a co-chain complex, the corresponding cohomology H^(q)(X;A)H^{q}(X ; A) is called Cohomology of pairs (X,A)(X, A). Singular Cohomology satisfy the following properties, [AX][\mathrm{AX}]
(AX0) H^(q)(X;A)H^{q}(X ; A) is a contravariant functor with, a natural map delta:H^(q)(A)rarrH^(q+1)(X,A)\delta: H^{q}(A) \rightarrow H^{q+1}(X, A).
(DIMENSION) If XX is a point then H^(0)(X)=ZH^{0}(X)=\mathbb{Z} and trivial for other dimension.
(EXACTNESS) There is a Long exact sequence,
cdotsH^(q)(X,A)rarrH^(q)(X)rarrH^(q)(A)rarr"delta"H^(q+1)(X,A)cdots\cdots H^{q}(X, A) \rightarrow H^{q}(X) \rightarrow H^{q}(A) \xrightarrow{\delta} H^{q+1}(X, A) \cdots
(EXCISION) U sube A sube XU \subseteq A \subseteq X be a closed set in AA then the inclusion (X-U,A-U)↪(X,A)(X-U, A-U) \hookrightarrow(X, A) will give us isomorphism in cohomology.
H^(q)(X,A)≃H^(q)(X-U,A-U)H^{q}(X, A) \simeq H^{q}(X-U, A-U)
(ADDITIVE) If (X_(i),A_(i))\left(X_{i}, A_{i}\right) are disjoint pair then
(HOMOTOPY) If f:X rarr Yf: X \rightarrow Y is a homotopy equivalence it will induce isomorphism in cohomology groups.
Any contravariant functors E^(q)E^{q} from homotopy category of pairs to modules over Z\mathbb{Z} satisfy the above properties [AX][\mathrm{AX}], will give us a cohomology theory which is called ordinary cohomology theory. Any cohomology theory over a fixed coefficient ring RR are equivalent (this is an application of Acyclic model theorem). (refer report)
0.2. Cohomology Ring
One of the very important feature of Cohomology is that, it naturally gives us a product which helps us to give the o+H^(**)(X)\oplus H^{*}(X) a graded ring structure. (Every co-chain complex has coefficient in Z\mathbb{Z} unless anything mentioned) For a topological space XX consider the diagonal map, Delta:X rarr X xx X\Delta: X \rightarrow X \times X. For each kk It will give us a map in co-chain complex C^(k)(X xx X)rarr"Delta^(**)"C^(k)(X)C^{k}(X \times X) \xrightarrow{\Delta^{*}} C^{k}(X). Using Künneth formula we have,
Consider R=o+C^(k)(X)R=\oplus C^{k}(X). Note that if a inC^(p)(X),b inC^(q)(X)a \in C^{p}(X), b \in C^{q}(X) then Delta^(**)@k(a ox b)inC^(p+q)(X)\Delta^{*} \circ k(a \otimes b) \in C^{p+q}(X). Thus we can give RR a graded ring structure with the product a⌣b=Delta^(**)@k(a ox b)a \smile b=\Delta^{*} \circ k(a \otimes b). It can be shown that
delta(a⌣b)=delta a⌣b+(-1)^(p)a⌣delta b\delta(a \smile b)=\delta a \smile b+(-1)^{p} a \smile \delta b
thus this product respects both co-boundary and co-cycles, so it will induce a product (hence a graded ring structure) in Cohomology-groups. We call this ring Cohomology Ring and denote it by H^(**)(X)=o+H^(k)(X)H^{*}(X)=\oplus H^{k}(X). We will now look at some examples. For computational purpose we will use Poincaré duality.
Theorem 5.1. (Poincaré Duality) If MM is a n-dimensional, closed, orientable manifold with fundamental class [M][M], then the following bilinear pairing is non-degenerate,
H^(k)(M;R)xxH^(n-k)(M;R)rarr RH^{k}(M ; R) \times H^{n-k}(M ; R) \rightarrow R
given by (alpha,beta)|->(alpha⌣beta)([M])(\alpha, \beta) \mapsto(\alpha \smile \beta)([M]).
The above theorem tells us for every alpha inH^(k)(M;k)\alpha \in H^{k}(M ; k) there is a beta inH^(n-k)(M;k)\beta \in H^{n-k}(M ; k), such that alpha⌣beta\alpha \smile \beta is generator of H^(n)(M;k)H^{n}(M ; k). Where kk is a field. Now we will work out some examples.
Example 1. The first example is RP^(n)\mathbb{R} P^{n}. Orientation depends on the units of the underlying coefficient ring. If we take underlying ring tobe Z//2Z\mathbb{Z} / 2 \mathbb{Z} (which is also a field), then there is only one choice of generator and hence RP^(n)\mathbb{R} P^{n} is an orientable manifold with coefficients in the above ring. It can be computed via universal coefficient theorem that H^(k)(RP^(n);Z//2Z)=Z//2ZH^{k}\left(\mathbb{R} P^{n} ; \mathbb{Z} / 2 \mathbb{Z}\right)=\mathbb{Z} / 2 \mathbb{Z} for k <= nk \leq n It is also known that RP^(n)\mathbb{R} P^{n} admits a filtration (this gives CW\mathrm{CW}-structure),
The inclusion RP^(k)↪RP^(k+1)\mathbb{R} P^{k} \hookrightarrow \mathbb{R} P^{k+1} gives isomorphism in cohomology for i <= ki \leq k. Thus if alpha\alpha is a generator of H^(1)(RP^(n);Z//2Z),alpha^(k)H^{1}\left(\mathbb{R} P^{n} ; \mathbb{Z} / 2 \mathbb{Z}\right), \alpha^{k} will be generator of H^(k)(RP^(n);Z//2Z)H^{k}\left(\mathbb{R} P^{n} ; \mathbb{Z} / 2 \mathbb{Z}\right) and alpha^(n+1)=0\alpha^{n+1}=0 as there is no non-trivial cohomology above dimension nn. Thus the cohomology ring is F_(2)[alpha]//(alpha^(n+1))\mathbb{F}_{2}[\alpha] /\left(\alpha^{n+1}\right).
Example 2. Following the same way as above we can compute cohomology ring of CP^(n)\mathbb{C} P^{n} and HP^(n)\mathbb{H} P^{n}.
Example 3. (Torus) The above examples were sort of abstractly done. But we will understand a case more geometrically. Consider, T=S^(1)xxS^(1)T=\mathbb{S}^{1} \times \mathbb{S}^{1}
With this simplicial decomposition of TT, we can compute the cohomology. It will be
Here let, alpha\alpha and beta\beta be the generator of H^(1)(T)H^{1}(T), then alpha\alpha and beta\beta are such that they takes value 1 on sides aa and value 0 on sides bb. Some computation will give us
which means H^(**)(T)H^{*}(T) is a Z\mathbb{Z}-module with generator alpha,beta\alpha, \beta and alpha beta=-beta alpha,alpha^(2),beta^(2)=0\alpha \beta=-\beta \alpha, \alpha^{2}, \beta^{2}=0. This is called exterior algebra denoted as Lambda_(Z)(alpha,beta)\Lambda_{\mathbb{Z}}(\alpha, \beta).
Example 4. In general for X=prod_(k)S^(1)X=\prod_{k} \mathbb{S}^{1} will be the exterior algebra Lambda_(Z)[x_(1),cdots,x_(k)]\Lambda_{\mathbb{Z}}\left[x_{1}, \cdots, x_{k}\right] which is a Z\mathbb{Z} module generated by {x_(i)}\left\{x_{i}\right\} and with the relation x_(i)x_(j)x_{i} x_{j} is -x_(j)x_(i)-x_{j} x_{i} if i!=ji \neq j and 0 if i=ji=j.
Some important properties of cohomology ring
If X,YX, Y are two topological spaces then H^(**)(X xx Y)=H^(**)(X)oxH^(**)(Y)H^{*}(X \times Y)=H^{*}(X) \otimes H^{*}(Y).
If X,YX, Y are two topological spaces, H^(**)(X vv Y)=H^(**)(X)xxH^(**)(Y)H^{*}(X \vee Y)=H^{*}(X) \times H^{*}(Y)
Example 5. As an example of above properties consider X,Y=RP^(oo)X, Y=\mathbb{R} P^{\infty}, then with coefficients in Z_(2)\mathbb{Z}_{2} we can say H^(**)(X xx Y)=Z_(2)[x]oxZ_(2)[y]=Z_(2)[x,y]H^{*}(X \times Y)=\mathbb{Z}_{2}[x] \otimes \mathbb{Z}_{2}[y]=\mathbb{Z}_{2}[x, y].
Now we will explore some crucial applications of this product structure.
0.3. Application 1: Hopf Invariant And HH-Spaces
Now we will talk about the case when R^(n)\mathbb{R}^{n} can be a division algebra. The following theorem will tell us R^(n)\mathbb{R}^{n} could be a division algebra
Theorem 5.2. If R^(n)\mathbb{R}^{n} has a division algebra structure over field R\mathbb{R}, then R\mathbb{R} must be a power of 2
Proof. For a division algebra structure we must have x|->axx \mapsto a x and x|->xax \mapsto x a are linear isomorphism. Now the multiplication in R^(n)\mathbb{R}^{n} will give rise to a map t:RP^(n-1)xxRP^(n-1)rarrRP^(n-1)t: \mathbb{R} P^{n-1} \times \mathbb{R} P^{n-1} \rightarrow \mathbb{R} P^{n-1}. It is a homeomorphism when restricted to {x}xxRP^(n-1)\{x\} \times \mathbb{R} P^{n-1} and RP^(n-1)xx{y}\mathbb{R} P^{n-1} \times\{y\}. Now,
tells us t^(**)(alpha)=k_(1)alpha_(1)+k_(2)alpha_(2)t^{*}(\alpha)=k_{1} \alpha_{1}+k_{2} \alpha_{2}. Now, RP^(n-1)↪RP^(n-1)xxRP^(n-1)\mathbb{R} P^{n-1} \hookrightarrow \mathbb{R} P^{n-1} \times \mathbb{R} P^{n-1} in two ways. First, a|->(x,a)a \mapsto(x, a) and second one a|->(a,y)a \mapsto(a, y). Thus we can say the first map, maps alpha_(1)rarr0\alpha_{1} \rightarrow 0 and alpha _(2){ }_{2} to zero. Thus taking composition with t^(**)t^{*} will tell us alpha|->k_(2)alpha\alpha \mapsto k_{2} \alpha since this composition is isomorphism it must be identity, and hence k_(2)=1k_{2}=1 similarly k_(1)=1k_{1}=1 and thus t^(**)(alpha)=alpha_(1)+alpha_(2)t^{*}(\alpha)=\alpha_{1}+\alpha_{2}. Since alpha^(n)\alpha^{n} is zero we must have (alpha_(1)+alpha_(2))^(n)\left(\alpha_{1}+\alpha_{2}\right)^{n} is zero in other words ([n],[k])=0\left(\begin{array}{l}n \\ k\end{array}\right)=0 for each kk modulo 2 . It can be easily proved that n=2^(m)n=2^{m} for some mm using Number theory argument.
In-fact we can say more about nn. From the following discussion we will conclude R^(n)\mathbb{R}^{n} is division algebra over R\mathbb{R} only for n=1,2,4,8n=1,2,4,8.
Let, f:S^(2n-1)rarrS^(n)f: \mathbb{S}^{2 n-1} \rightarrow \mathbb{S}^{n} be a continous map and consider the adjunction space X=S^(n)uu_(f)D^(2n)X=\mathbb{S}^{n} \cup_{f} D^{2 n}. If we compute the cohomology of XX it will be Z\mathbb{Z} for index nn and 2n2 n and it will be zero for other indices. In the cohomology ring H^(**)(X)H^{*}(X) if we take alpha inH^(n)(X)\alpha \in H^{n}(X) to be the generator of this group and beta\beta to be the generator of H^(2n)(X)H^{2 n}(X). Then we must have alpha^(2)=k beta\alpha^{2}=k \beta the constant kk will depends only on the homotopy class of ff we call it h(f)h(f). This h(f)h(f) is known as Hopf invariant. Note that this gives a map(homomorphism) h:pi_(2n-1)(S^(n))rarrZh: \pi_{2 n-1}\left(\mathbb{S}^{n}\right) \rightarrow \mathbb{Z}.
Question 1 ? For which nn there exist a map f:S^(2n-1)rarrS^(n)f: \mathbb{S}^{2 n-1} \rightarrow \mathbb{S}^{n} such that h(f)h(f) is 1 ?
For example treat S^(3)subeC^(2)\mathbb{S}^{3} \subseteq \mathbb{C}^{2}, such that S^(3)={(z_(1),z_(2)):|z_(1)|^(2)+|z_(2)|^(2)=1}\mathbb{S}^{3}=\left\{\left(z_{1}, z_{2}\right):\left|z_{1}\right|^{2}+\left|z_{2}\right|^{2}=1\right\} and there is a natural projection from S^(3)rarrCP^(1)\mathbb{S}^{3} \rightarrow \mathbb{C} P^{1}. Call this projection map pi:S^(3)rarrS^(2)\pi: \mathbb{S}^{3} \rightarrow \mathbb{S}^{2}. It can be shown this hh has fibre S^(1)\mathbb{S}^{1} at each point. In other words S^(1)↪S^(3)rarr"pi"S^(2)\mathbb{S}^{1} \hookrightarrow \mathbb{S}^{3} \xrightarrow{\pi} \mathbb{S}^{2} is a Fibre-Bundle. It can be shown, pi\pi has hopf invariant 1 (the easiest way to do is to use the equivalence of De-Rahm cohomology to ordinary cohomology theories). For which nn such map exist with hopf invarient 1 ? It was a very open problem untill J.F.Admas (1960) proved, it can exist only for n=1,2,4,8n=1,2,4,8. We will now look into two more problems. He proved it using KK-theory. There is one more proof using steenrod operations and existence of non-trivial class in Ext_(A)^(1)[F_(2),F_(2)[n]]\operatorname{Ext}_{A}^{1}\left[\mathbb{F}_{2}, \mathbb{F}_{2}[n]\right], Adams spectral sequence.
Definition 5.3. (H-Space) An H-space consists of a topological space XX, together with an element ee of XX and a continuous map mu:X xx X rarr X\mu: X \times X \rightarrow X, such that mu(e,e)=e\mu(e, e)=e and the maps x|->mu(x,e)x \mapsto \mu(x, e) and x|->mu(e,x)x \mapsto \mu(e, x) are both homotopic to the identity map through maps sending ee to ee.
This is generalization of the definition of topological groups. We know, S^(0),S^(1),S^(3)\mathbb{S}^{0}, \mathbb{S}^{1}, \mathbb{S}^{3} admits a topological group structure from reals, complex, quaternions and S^(7)\mathbb{S}^{7} admits a HH-space structure coming from octonions, it's not a group because the multiplication of octonions are not associative.
Question 2 ? for which n,S^(n-1)n, \mathbb{S}^{n-1} admits a HH-space structure?
Surprisingly, for n=1,2,4,8n=1,2,4,8 the above Question can be true. In-fact it can be shown Question 1 and Question 2 are equivalent. Thre is one more fact,
Lemma 5.4. If R^(n)\mathbb{R}^{n} had a real division algebra structure then, for that nn there exist a Hopf invariant 1 map ( S^(n-1)\mathbb{S}^{n-1} is HH-space).
Proof of the lemma is not very hard to prove. Let R^(n)\mathbb{R}^{n} be a real division algebra and ee be a vector of 11 nit norm. We can compose the multiplication R^(n)xxR^(n)rarrR^(n)\mathbb{R}^{n} \times \mathbb{R}^{n} \rightarrow \mathbb{R}^{n} with an invertible map R^(n)rarrR^(n)\mathbb{R}^{n} \rightarrow \mathbb{R}^{n} so that e^(2)=ee^{2}=e. With this modified multiplication, let alpha(x)=cx\alpha(x)=c x and beta(x)=xe\beta(x)=x e We can then define a new division algebra product x o.y=alpha^(-1)(x)beta^(-1)(y)x \odot y=\alpha^{-1}(x) \beta^{-1}(y). Then o.o.x=x\odot \odot x=x and x o.e=xx \odot e=x. This product has no zero divisors, so the map (x,y)|->(x o.y)//|x o.y|(x, y) \mapsto(x \odot y) /|x \odot y| defires an H-space structure on S^(n-1)subR^(n)S^{n-1} \subset \mathbb{R}^{n} with identity ee.
Thus by Hopf invariant one problem we can say n=1,2,4,8n=1,2,4,8 are the only case when R^(n)\mathbb{R}^{n} can-be division algebra and we know reals, complex, quaternions, octonions are the ones.
0.4. Application 2: motivation towards EHP Sequence
We have seen the cohomology rings of RP^(oo),CP^(oo)\mathbb{R} P^{\infty}, \mathbb{C} P^{\infty} are polynomial ring. If we look at the CW\mathrm{CW} structure of CP^(oo)\mathbb{C} P^{\infty} it consist of one cell in each even dimension and thus the CW\mathrm{CW} cohomology is Z\mathbb{Z}
at every even dimension. The cohomology ring which is Z[x]\mathbb{Z}[x] is basically direct sum of these copies of Z\mathbb{Z}. James tried to generalize this construction to S^(2n)\mathbb{S}^{2 n} so that we get a new space JS^(2n)J \mathbb{S}^{2 n}, which has a cohomology ring Z[x]\mathbb{Z}[x]. It turns out that the space he constructed is almost a polynomial ring on Z,H^(**)(JS^(2n))\mathbb{Z}, H^{*}\left(J \mathbb{S}^{2 n}\right) is a Z\mathbb{Z} module with {x^(i)//i!}\left\{x^{i} / i !\right\} are the generator. We denote this module as Gamma_(Z)[x]\Gamma_{\mathbb{Z}}[x]. If the coefficients were in Q\mathbb{Q} it must have been the polynomial ring Q[x]\mathbb{Q}[x].
James's construction : Let, XX is a CW complex with a base point ee. Let, J_(i)XJ_{i} X be the quotient space of X^(i)(i:}X^{i}\left(i\right.-many products of XX ) under the equivalence relation, (x_(1),cdots,x_(k),e,cdots,x_(i))∼\left(x_{1}, \cdots, x_{k}, e, \cdots, x_{i}\right) \sim(x_(1),cdots,e,x_(k),cdots,x_(i))\left(x_{1}, \cdots, e, x_{k}, \cdots, x_{i}\right). There is a natural inclusion J_(i)X↪J_(i+1)XJ_{i} X \hookrightarrow J_{i+1} X, which is given by (x_(1),cdots,x_(i))rarr\left(x_{1}, \cdots, x_{i}\right) \rightarrow(x_(1),cdots,e,cdots,x_(i))\left(x_{1}, \cdots, e, \cdots, x_{i}\right). Now if we take colimit under this inclusion we will get a space,
JX:=colim(cdotsJ_(i)X↪J_(i+1)X cdots)J X:=\operatorname{colim}\left(\cdots J_{i} X \hookrightarrow J_{i+1} X \cdots\right)
This gives us a monoid structure on JXJ X.
We will focus on the case X=S^(2n)X=\mathbb{S}^{2 n}. In this case JXJ X has a CW structure with only one cell at the 2kn2 k n dimension for 0 <= k0 \leq k. Thus, if we compute the cellular cohomology of JXJ X it will be isomorphic to Z\mathbb{Z} for the indices 2kn2 k n, and J_(k)S^(2n)J_{k} \mathbb{S}^{2 n} is the 2nk2 n k-dimension skeleta in CW decomposition of JXJ X. Thus the cohomology ring is a Z\mathbb{Z} module with countably many generators. Let, x_(k)x_{k} be the generator of H^(2kn)(JX;Z)H^{2 k n}(J X ; \mathbb{Z}). Now we will establish relations between these x_(k)x_{k} 's. Let, q_(k):X^(k)rarrJ_(k)Xq_{k}: X^{k} \rightarrow J_{k} X be the quotient map. Every cell in J_(k)XJ_{k} X is homeomorphic to some cells in X^(k)X^{k}. It is not hard to note qq is a cellular map. In cohomology ring it will induce a map preserving degree
q_(k)^(**):H^(**)(J_(k)X;Z)rarrH^(**)(X^(k);Z)q_{k}^{*}: H^{*}\left(J_{k} X ; \mathbb{Z}\right) \rightarrow H^{*}\left(X^{k} ; \mathbb{Z}\right)
Note that H^(**)(X^(k);Z)≃Z[alpha_(1),cdots,alpha_(k)]//(alpha_(1)^(2),cdots,alpha_(k)^(2))H^{*}\left(X^{k} ; \mathbb{Z}\right) \simeq \mathbb{Z}\left[\alpha_{1}, \cdots, \alpha_{k}\right] /\left(\alpha_{1}^{2}, \cdots, \alpha_{k}^{2}\right), so q_(k)^(**)(x_(1))q_{k}^{*}\left(x_{1}\right) must be sumt_(i)alpha_(i).q_(k)^(**)\sum t_{i} \alpha_{i} . q_{k}^{*} identifies all the 2n2 n-cell of X^(k)X^{k} to get one 2n2 n-cell. Thus t_(i)=1t_{i}=1. We will must have q_(k)^(**)(x_(k))=alpha_(1)cdotsalpha_(k)q_{k}^{*}\left(x_{k}\right)=\alpha_{1} \cdots \alpha_{k}. Now note that,
Now, the quotient map induces an isomorphism in H^(2nk)H^{2 n k} (as both have only one 2kn2 k n-cell). We can say, x_(k)=x_(1)^(k)//kx_{k}=x_{1}^{k} / k !. We also know J_(k)X↪JXJ_{k} X \hookrightarrow J X is homotopy equivalence and hence induce isomorphism in Cohomology ring. Thus we can say H^(**)(JS^(2n);Z)H^{*}\left(J \mathbb{S}^{2 n} ; \mathbb{Z}\right) is a Z\mathbb{Z} module with {x^(i)//i!}\left\{x^{i} / i !\right\} are generators.
Even more can be said about H^(**)(JS^(n);Z)H^{*}\left(J \mathbb{S}^{n} ; \mathbb{Z}\right), which is given by the following theorem:
Theorem 5.5. (James) If nn is even, H^(**)(JS^(n);Z)H^{*}\left(J \mathbb{S}^{n} ; \mathbb{Z}\right) is isomorphic to Gamma_(Z)[x]\Gamma_{\mathbb{Z}}[x] (defined above). If nn is odd then H^(**)(JS^(n);Z)H^{*}\left(J \mathbb{S}^{n} ; \mathbb{Z}\right) isomorphic to H^(**)(S^(n);Z)oxH^(**)(JS^(2n);Z)H^{*}\left(\mathbb{S}^{n} ; \mathbb{Z}\right) \otimes H^{*}\left(J \mathbb{S}^{2 n} ; \mathbb{Z}\right) as a graded ring.
This James construction on S^(n)\mathbb{S}^{n} will lead us to a beautiful proof of EHP sequence. Whitehead introduced this sequence of related to homotopy group of spheres. He proved that for n >= 1n \geq 1 there is an exact sequence of homotopy groups
For any fibration, F↪E rarr BF \hookrightarrow E \rightarrow B there is a fibre sequence
cdots rarr Omega F rarr Omega E rarr Omega B rarr F rarr E rarr B\cdots \rightarrow \Omega F \rightarrow \Omega E \rightarrow \Omega B \rightarrow F \rightarrow E \rightarrow B
Any triple in the above sequence will give us a long exact sequence on homotopy groups.
We would like to get a fibration/fibre sequence so that the EHP sequence will turn out to be the corresponding LES on those triple. Recall that we have the adjoint relation [A,Omega X]=[Sigma A,X][A, \Omega X]=[\Sigma A, X] it will give us,
since the homotopy group above is non-trivial, we must have a map f:S^(n)rarr Omega SigmaS^(n)=OmegaS^(n+1)f: \mathbb{S}^{n} \rightarrow \Omega \Sigma \mathbb{S}^{n}=\Omega \mathbb{S}^{n+1}. If we consider the homotopy fibre of the above map FF.
Remark : Given any map f:X rarr Yf: X \rightarrow Y we can get a homotopy fibre of the map which is pull back of the following diagram
basically F=Xxx_(Y)PYF=X \times_{Y} P Y. It is the space of (gamma,x)(\gamma, x) where gamma\gamma is a path in YY starts at the base point and ends at f(x)f(x).
Thus F rarrS^(n)rarr OmegaS^(n+1)F \rightarrow \mathbb{S}^{n} \rightarrow \Omega \mathbb{S}^{n+1} is a fibre sequence and the corresponding LES will be
comparing with the EHP sequence we guess pi_(k-1)(F)≃pi_(k+1)(S^(2n+1))\pi_{k-1}(F) \simeq \pi_{k+1}\left(\mathbb{S}^{2 n+1}\right), we can guess F=Omega^(2)S^(2n+1)F=\Omega^{2} \mathbb{S}^{2 n+1}. In-fact this is true. James construction gives us a map OmegaS^(n+1)rarr OmegaS^(2n+1)\Omega \mathbb{S}^{n+1} \rightarrow \Omega \mathbb{S}^{2 n+1}, whose fibre is S^(n)\mathbb{S}^{n}. We can note Omega^(2)S^(2n+1)rarrS^(n)rarr OmegaS^(n+1)\Omega^{2} \mathbb{S}^{2 n+1} \rightarrow \mathbb{S}^{n} \rightarrow \Omega \mathbb{S}^{n+1} is part of the fibre sequence of S^(n)↪OmegaS^(n+1)rarr OmegaS^(2n+1)\mathbb{S}^{n} \hookrightarrow \Omega \mathbb{S}^{n+1} \rightarrow \Omega \mathbb{S}^{2 n+1}.
James construction of a space XX, call it JXJ X has the following properties:
(1) There is a map JXJ X to Omega Sigma X\Omega \Sigma X, which is map of monoids. It takes ee to the constant loop and to the point xx to the loop t|->(x,t)t \mapsto(x, t).
(2) If XX is a CW-complex it gives us a homotopy equivalence JX≃_("htop ")Omega Sigma XJ X \simeq_{\text {htop }} \Omega \Sigma X (otherwise it is a weak equivalence).
(3) It can be shown that, Sigma JX=vvv_(i)(Sigma(X^(^^i)))\Sigma J X=\bigvee_{i}\left(\Sigma\left(X^{\wedge i}\right)\right)
Using the above properties we will get a map Sigma OmegaS^(n+1)rarrS^(2n+1)\Sigma \Omega \mathbb{S}^{n+1} \rightarrow \mathbb{S}^{2 n+1} as follows:
here HH is composition of all consequent maps. By the adjoint property of loop-suspension we can say there must exist a map h:OmegaS^(n+1)rarr OmegaS^(2n+1)h: \Omega \mathbb{S}^{n+1} \rightarrow \Omega \mathbb{S}^{2 n+1}. It can be shown that h_(**)h_{*} induce isomorphism in the homology groups (It will follow from the cohomology ring computation) for the indices ii multiple of 2n+12 n+1. If FF is the homotopy fibre of the map ff, it can be shown that FF is simply connected (LES of homotopy groups) H_(i)(F)≃H_(i)(S^(n))H_{i}(F) \simeq H_{i}\left(\mathbb{S}^{n}\right) and hence there will exist a map F rarrS^(n)F \rightarrow \mathbb{S}^{n} which is weak equivalence. LES associated to this fibre sequence will give us EHP sequence. Which is remarkably used in computation of homotopy groups of sphere.
Remark : The proof the theorem 5.4 can be done using product structure of Serre spectral sequence made out of the fibration OmegaS^(n+1)↪PS^(n+1)rarrS^(n+1)\Omega \mathbb{S}^{n+1} \hookrightarrow P \mathbb{S}^{n+1} \rightarrow \mathbb{S}^{n+1}. This can be found in (refer Kirk).
0.5. Application 3: Serre finiteness Condition
There is a deep connection between cohomology and spectral sequence. For those hearing the term 'spectral sequence', let me introduce it.
Definition 5.6. A cohomological spectral sequence is a sequence {E_(**,**)^(r),d_(r)}_(r > 0)\left\{E_{*, *}^{r}, d_{r}\right\}_{r>0} of co-chain complexes such that E_(**,**)^(r+1)=H^(**)(E_(**,**)^(r))E_{*, *}^{r+1}=H^{*}\left(E_{*, *}^{r}\right). In more details we have Abelian groups d_(p,q)^(r):E_(p,q)^(r)rarrd_{p, q}^{r}: E_{p, q}^{r} \rightarrowE_(p+r,q-r+1)^(r)E_{p+r, q-r+1}^{r}, such that d^(r)@d^(r)=0d^{r} \circ d^{r}=0 and
such that E_(p,q)^(oo)=F_(p)^(p+q)//F_(p+1)^(p+q)E_{p, q}^{\infty}=F_{p}^{p+q} / F_{p+1}^{p+q}. For this moment take this as the definition. Although spectral sequence comes very naturally from filtered complex, double complex, due to shortage of time we would not be able to go through that. Rather one can look at [?]. Now we will state a theorem by Larey and Serre.
Theorem 5.7. (Larey and Serre) Given a fibration F↪E rarr BF \hookrightarrow E \rightarrow B, with FF being the simply connected space there is a spectral sequence {E_(p,q)^(r),d^(r)}\left\{E_{p, q}^{r}, d^{r}\right\} with E^(2)E^{2}-page
and E^(oo)E^{\infty} converges to {H^(n)(E)}\left\{H^{n}(E)\right\}.
The above theorem helps us to define a product on the spectral sequence. Given a spectral sequence of cohomology we get a natural bilinear multiplication E_(p,q)^(r)xxE_(s,t)^(r)rarrE_(p+s,q+t)^(r)E_{p, q}^{r} \times E_{s, t}^{r} \rightarrow E_{p+s, q+t}^{r} satisfy the following properties,
The derivation d(xy)=d(x)y+(-1)^(p+q)xd(y)d(x y)=d(x) y+(-1)^{p+q} x d(y).
The product E_(p,q)^(2)xxE_(s,t)^(2)rarrE_(p+s,q+t)^(2)E_{p, q}^{2} \times E_{s, t}^{2} \rightarrow E_{p+s, q+t}^{2} is (-1)^(qs)(-1)^{q s} times the standard cup product
where the coefficient get multiplied by the cup product H^(q)(F;Z)xxH^(s)(F;Z)rarrH^(q+s)(F;Z)H^{q}(F ; \mathbb{Z}) \times H^{s}(F ; \mathbb{Z}) \rightarrow H^{q+s}(F ; \mathbb{Z}).
The cup product H^(**)(X;Z)H^{*}(X ; \mathbb{Z}) restricts to maps F_(p)^(p+q)xxF_(r)^(r+s)rarrF_(p+r)^(p+q+r+s)F_{p}^{p+q} \times F_{r}^{r+s} \rightarrow F_{p+r}^{p+q+r+s} it induces a map in quotients and hence there is a product structure in E^(oo)E^{\infty}.
Remark : The above theorems, product structure is true if we take the coefficient ring to be any arbitrary ring RR. There is also similar kind of theorem for the homology groups, but since the product structure in cohomology ring is more clear we will only talk about Serre cohomology spectral sequence.
Recall k(Z,n)k(\mathbb{Z}, n) denotes the Eilenberg-Maclane space whose, nn-th homotopy group is Z\mathbb{Z} and rest homotopy groups are trivial. It can be shown (very easily) H^(**)(K(Z,n);Q)H^{*}(K(\mathbb{Z}, n) ; \mathbb{Q}) is isomorphic to the polynomial ring Q[x]\mathbb{Q}[x]. In-fact we could do th calculation quickly. (Calculation)
Note that [S^(n),k(Z,n)]≃Z\left[\mathbb{S}^{n}, k(\mathbb{Z}, n)\right] \simeq \mathbb{Z} thus there must exist a map f:S^(n)rarr k(Z,n)f: \mathbb{S}^{n} \rightarrow k(\mathbb{Z}, n) which is not nullhomotopic. We know any continous map can be decomposed as a homotopy equivalence and a fibration.
Thus in the above picture we have a fibre sequence F↪X rarr k(Z,n)F \hookrightarrow X \rightarrow k(\mathbb{Z}, n) and if we can calulate the homotopy group (pi_(i)(F))\left(\pi_{i}(F)\right) of FF it will be trivial for i <= ni \leq n and it is isomorphic to the homotopy group of XX (i.e S^(n)\mathbb{S}^{n} ) for i > ni>n. corresponding the inclusion F↪XF \hookrightarrow X there is again a fibre sequence,
Again from the long exact sequence it can be shown hat(F)\hat{F} is k(Z,n-1)k(\mathbb{Z}, n-1). Upto homotopy the above fibre sequence in equivalent to the sequence in the right side. If nn is odd and n > 1n>1 we can see k(Z,n-1)k(\mathbb{Z}, n-1) is simply connected and by Larey Serre theorem there is a spectral sequence coefficient in Q\mathbb{Q} with
now note that, E_(n,0)^(2)≃QE_{n, 0}^{2} \simeq \mathbb{Q} and E_(k(n-1),0)^(2)≃QE_{k(n-1), 0}^{2} \simeq \mathbb{Q} and rest are zero. So till nn-th page these will remain unchanged. Now E^(n)E^{n} ( nn-th page looks like the following),
where d^(n):Qa rarrQxd^{n}: \mathbb{Q} a \rightarrow \mathbb{Q} x will be isomorphism. Otherwise, this map will be a zero map then Qa\mathbb{Q} a survives till E^(oo)E^{\infty} but since F\mathscr{F} is (n-1)(n-1) connected it's (n-1)(n-1) homology groups must be trivial (by Hurewicz theorem) and hence ( n-1n-1 ) cohomology is trivial. Thus d^(n)d^{n} will be an isomorphism. By the product structure we can say Qa^(2)rarrQax\mathbb{Q} a^{2} \rightarrow \mathbb{Q} a x is an isomorphism and hence at (n+1)(n+1)-page every thing is trivial. So, for higher i > ni>n we must have, H^(**)(F)H^{*}(F) are trivial and hence homology groups are also trivial. Thus, pi_(i)(F)oxZ\pi_{i}(F) \otimes \mathbb{Z} are trivial. SO, pi_(i)(S^(n))\pi_{i}\left(\mathbb{S}^{n}\right) must contain the torsion part only. Since pi_(i)(S^(n))\pi_{i}\left(\mathbb{S}^{n}\right) finitely generated we can say, pi_(i)(S^(n))\pi_{i}\left(\mathbb{S}^{n}\right) are finite. Using EHP sequence we can say except for pi_(4k-1)(S^(2k))\pi_{4 k-1}\left(\mathbb{S}^{2 k}\right) every pi_(i)(S^(n))\pi_{i}\left(\mathbb{S}^{n}\right) is finite for i > ni>n. This is Serre's finiteness condition.
Theorem 5.8. (Serre) Except for pi_(4k-1)(S^(2k))\pi_{4 k-1}\left(\mathbb{S}^{2 k}\right) every pi_(i)(S^(n))\pi_{i}\left(\mathbb{S}^{n}\right) is finite for i > ni>n.