Halloween 4: The Return of Adams Spectral Sequence π
Enrique De Alba ; October 31, 2025
Happy Halloween! We're first going to define the Steenrod algebra A which by the end of this post will be the best friend we never knew we always wanted.
The Steenrod algebra, denoted A, is the graded, associative, non-commutative Z2β-algebra generated by the Steenrod squares {Sqi}iβ₯0β subject to the Adem relations:
where a,b are integers such that 0<a<2b, and where Sq0=1 is the identity [1]. The grading of A is given by deg(Sqi)=i, and deg(SqaβSqb)=a+b. While {Sqi}iβ₯0β generates A, it does not form a basis. For that, we'll need the concept of admissible monomials.
A monomial Sqinβββ―βSqi1β is called admissible iff ikββ₯2ikβ1β for all k. For example, Sq2βSq2 is not an admissible monomial since 2β±4. However, we can use the Adem relations to get that
Sq2βSq2=Sq3βSq1
where Sq3βSq1is admissible since 3β₯2β 1.
Theorem: The set of all admissible monomials forms a vector space basis for the A, i.e. the Steenrod algebra [2]. This means that any composition of Steenrod squares in A can be written uniquely as a linear combination of admissible monomials.
The main goal this post is to introduce the Adams spectral sequence and construct the E2β page for the sphere spectrum S, which is the bigraded algebra:
ExtAs,tβ(Z2β,Z2β)
While the Serre spectral sequence was concerned about the (co)homology of fibrations, the Adams spectral sequence is a machine for computing stable homotopy groups of spheres, i.e. ΟβSβ. These guys are way spookier than homology or even cohomology π». Note that the E2β page for S is:
E2s,tβ=ExtAs,tβ(Hβ(S;Z2β),Z2β),
and while we won't prove this here, it turns out that Hβ(S;Z2β)β Z2β, hence why we'll be doing a deep-dive into ExtAs,tβ(Z2β,Z2β).
Ext: What is Ext? Ext is short for extension and is a regular in the stage of homological algebra. Consider the following short exact sequence (SES) of modules:
0βAβBβCβ0
Let's apply the (contravariant) functor HomAβ(β,Z2β) to get:
Suppose our modules are left A-modules, then this new sequence will be left-exact which just means that the induced map HomAβ(C,Z2β)βHomAβ(B,Z2β) is injective. Now, the right map HomAβ(B,Z2β)βHomAβ(A,Z2β) is not generally surjective, so the sequence as a whole is not (in general) exact. The Ext groups are the algebraic objects that measure this "failure of exactness."
ExtAsβ(M,N): Let's see how we construct these extensions. We'll start by looking at a general ExtAsβ(M,N) case which is equivalent to our case study, just with M=Z2β and N=Z2β. We construct ExtAsβ(M,N) in three simple steps:
Resolve: We first look at the projective resolution of M. This is an exact sequence of projective modules (i.e. projective resolution). These are modules that are "nice" and behave like free modules. This LES "approximates" M:
β―β2ββP1ββ1ββP0βΟ΅βMβ0,
where each Psβ is a projective A-module and Ο΅ is a surjective map.
Apply Hom: We apply the HomAβ(β,N) functor to the sequence after first removing the M object from the sequence to get:
This new sequence is not guaranteed to be exact, but we do have the property βs+1ββββsββ=0, like that of cochain complexes in singular cohomology.
Take the Cohomology: There's not much else to it. The Ext groups are defined as the cohomology of this complex:
where ker(βs+1ββ)/im(βsββ) is the s-th cohomology group Hs(HomAβ(Pββ,N)).
Unfortunately, for the Steenrod algebra A, finding a simple projective resolution of Z2β is extremely difficult. Instead, we'll be using a different resolution that works for any algebra: the bar resolution. For this, we'll be using the normalized bar resolution which is defined using the augmentation ideal of A.
Let AΛ:=ker(Ο΅:AβZ2β) denote the augmentation ideal. This is the vector space spanned by all the Steenrod squares of positive degree, i.e. {Sqi}i>0β. Importantly, note that Sq0=1β/AΛ. The bar resolution for Z2β is given by
β―ββββPβββZ2ββ0
where the s-th term is:
Psβ=Aβ(AΛ)βs
For example, we have: P0β=A,P1β=AβAΛ,P2β=AβAΛβAΛ, and so on [4]. The differential
βsβ:PsββPsβ1β
makes use of the following notation [a1ββ£a2ββ£β―β£asβ] for elements in (AΛ)βs. Hence, we denote elements in Psβ=Aβ(AΛ)βs as aβ[a1ββ£β―β£asβ]. The differential maps these elements to βsβ1β the following way:
Note that the (β1)i term is just 1 in our case since A is over Z2β. Additionally, note that the first term has the A-action with aβ a1β. In the sum, we just multiply all adjacent aiβai+1β terms inside the (AΛ)β bracket. Example:
Note that here we use + to show that β1β‘1 (mod 2). Side note: the summation for βsβ usually includes a term
(β1)saβ[a1ββ£β―β£asβ1β]Ο΅(asβ),
but since asββAΛ then Ο΅(asβ)=0, so this term is just zero in our case.
Before we proceed, let's briefly look at the following map called the augmentation map:
Ο΅:AβZ2β
Ο΅ is a homomorphism that strips away all the complexity in A. Note that for any element aβ²βAΛβA, we get Ο΅(aβ²)=0 (since we defined AΛ to be the kernel of this Ο΅ map). Additionally, note that Ο΅(1)=1 since it's a homomorphism. We define the trivial action of A on Z2β using this Ο΅ map , where we have the trivial action aβ k for any aβA,kβZ2β such that:
that satisfies the trivial action we defined above, i.e.
Ο(aβ x)=aβ Ο(x)=Ο΅(a)Ο(x),
since Ο(x)βZ2β, hence why aβA acts trivially on it via the augmentation map Ο΅. Note that Ο΅(a)=0, as well as Ο(aβ x)=0, for anything not starting with 1. This means that Ο is completely determined by what it does to elements of the form
1β[a1ββ£β―β£asβ]
The next thing we do is apply HomAβ(β,Z2β) to our bar complex, so we'll have the groups HomAβ(Psβ,Z2β). However, note that these map homomorphisms are the same as our Ο, so we get the following isomorphism of vector spaces:
where we get this HomZ2ββ(β) simplification since we're now dealing with just AΛ and we don't have to worry about the A-module structure from A since we've basically factored it out. Furthermore, since AΛ is just a Z2β vector space, then note that this HomZ2ββ((AΛ)βs,Z2β) is precisely the definition of the dual of (AΛ)βs. We'll denote this dual as (AΛβs)β and will denote most other duals with Vβ-notation, however, we'll be making a special exception for the Steenrod algebra itself which is denoted Aββ and we'll explain why later (spoiler alert: it's because A is a Hopf algebra). Altogether, we have the following isomorphism:
HomAβ(Psβ,Z2β)β (AΛβs)β
We also get the following induced differentials βsββ which map from HomAβ(Psβ1β,Z2β) to HomAβ(Psβ,Z2β) (reversed due to contravariant functor), i.e.
βsββ:(AΛβsβ1)ββ(AΛβs)β
We denote this entire dual bar complex object as ((AΛβs)β,βsββ). Before in our general HomAβ(β,N) case we took the cohomology of the complex (HomAβ(Pββ,N),βsββ), but now we'll be taking the cohomology of ((AΛβs)β,βsββ) to get ExtAβ(Z2β,Z2β)[5].
Let's compute ExtA1β(Z2β,Z2β). Looking at the cohomology of ((AΛβs)β,βsββ), we get:
Let's first look at β1ββ:(AΛβ0)ββ(AΛβ1)β. Note that this dual (AΛβ0)ββ HomZ2ββ(AΛβ0,Z2β), and since AΛ is a Z2β vector space then AΛβ0β Z2β, so we get:
HomZ2ββ(Z2β,Z2β)β Z2β
so we just get the following differential:
β1ββ:Z2ββAΛβ
This is dual to β1β:AβAΛβA such that β1β(aβ[a1β])=aβ a1β. Let ΟβHomAβ(P0β,Z2β)β Z2β, where this space is spanned by the augmentation map Ο΅:AβZ2β. Consider the pull-back β1ββ(Ο)=Οββ1β where we use inputs from AβAΛ. Let aβ[a1β]βAβAΛ, then consider:
Note that since ΟβZ2β then Ο(a1β)=0 if Ο is the zero map, so suppose Ο=Ο΅. Note that Ο(a1β)=Ο΅(a1β)=0 for all a1β in AΛ. This means that Ο΅(a)Ο΅(a1β)=0, hence
β1ββ(Ο)=Οββ1β(aβ[a1β])=0
for any aβ[a1β]βAβAΛ. This means that β1ββ is the zero map, and hence:
im(β1ββ)=0
Therefore, we get a major simplification for Ext1 where we have:
ExtA1β(Z2β,Z2β)β ker(β2ββ)
Let's look at this differential β2ββ:AΛββ(AΛβAΛ)β. Note that AΛββ HomZ2ββ(AΛ,Z2β) and ker(β2ββ)βAΛβ. This means that for ker(β2ββ), we're looking for all the AΛβZ2β map homomorphisms, i.e. let fβHomZ2ββ(AΛ,Z2β), such that β2ββ(f)=0 in (AΛβAΛ)β. Consider the pull-back:
β2ββ(f)=fββ2β=0,
where β2ββ is the dual of β2β such that β2β:AβAΛβAΛβAβAΛ.
Let aβ[a1ββ£a2β]βAβAΛβAΛ, and note that
Note that this term f(aβ a1ββ[a2β])=0 since we have the term Ο΅(aβ a1β)=Ο΅(a)Ο΅(a1β)=0. For the second term we get:
f(aβ[a1βa2β])=Ο΅(a)f(1β[a1βa2β])
Which means that fββ2β=0, hence we want f(1β[a1βa2β])=0. This is equivalent to looking for maps f:AΛβZ2β such that f([a1βa2β])=0 for all a1β,a2ββAΛ since we can just factor out the (1ββ) term. Additionally, another way to see this is that we're looking for maps f such that for any elements [a]βAΛ that can be written as a=a1βa2β for a1β,a2ββAΛ we get f([a])=0.
Let's take a tiny step back and then tie this back to f([a])=0 for a=a1βa2β. Elements in the Steenrod algebra A are either decomposable or indecomposable. A simple example is Sq3 which can be rewritten as Sq3=Sq1βSq2 via the Adem relations. Since we can write Sq3=Sq1βSq2, this means that Sq3 is decomposable [6]. Examples of indecomposable elements are Sq1, Sq2, and Sq4. Spoiler alert: it turns out that indecomposable elements are of the form Sq2i which becomes relevant very soon.
Recall that AΛ is spanned by {Sqi}i>0β, so Sq3βAΛ such that we can write Sq3=a1βa2β where a1β=Sq1 and a2β=Sq2. Note that this means that Sq3 is such that f([Sq3])=0 for fβker(β2ββ).
So we want our maps f to vanish for any decomposable element in AΛ. Note that when we write these decomposable elements like a1βa2β, these are elements in AΛβ AΛβAΛ. This means that ker(β2ββ) is the set of all AΛβZ2β maps that are zero at AΛβ AΛ. Note the following equivalence:
Importantly, note that since we're modding out by all the decomposable elements in AΛ then we're left with the indecomposable elements.
Theorem: The vector space of indecomposable elements of A, i.e. AΛ/(AΛβ AΛ)βA, denoted as Q(A), has a Z2β-basis given by the elements {Sq2i}iβ₯0β. Note that ExtA1β(Z2β,Z2β) is the dual of Q(A) which means that Ext1 must have a dual basis, and this is dual to the indecomposable elements in AΛ/(AΛβ AΛ).
The dual basis of Ext1 is the set of generators we call hiβ:
The class dual to the first indecomposable element, which is Sq1=Sq20, is h0β. Its bidegree (s,t) is (1,deg(Sq1))=(1,1).
The class dual to Sq2=Sq21 is h1β with bidegree (1,deg(Sq2))=(1,2).
In general, the class dual to Sq2i is hiβ with bidegree (1,2i).
With this, we've constructed the s=1 column of the Adams E2β page. It's just an infinite column of generators hiβ, each corresponding to the indecomposable elements Sq2i.
Next, we'll be building the s=2 column of ExtA2β(Z2β,Z2β) using the hiβ generators from the s=1 column. These are of the form hiβhjβ such that:
h02β=h0βh0β has bidegree (s,t) such that (1,1)+(1,1)=(2,2).
h0βh1β has bidegree (1,1)+(1,2)=(2,3).
h12β=h1βh1β has bidgree (1,2)+(1,2)=(2,4).
Warning: For h02β=h0βh0β, one may be tempted to look at its corresponding (dual) Sq1βSq1 which is just zero due to the Adem relations in A, and then incorrectly conclude that Sq1βSq1=0 implies that h02β=0 in Ext2. This turns out to not be the case. Since we're looking at these hiβhjβ elements in
then we'll have to deal with lots of messy Adem relations. This would be a nightmare. Is there something else we can do? We'll be switching to a different machine that will at first look more spooky π», but turns out to make life in Ext2 a lot easier. Since A is a Hopf algbra, then we can use the following theorem:
Hβ(Bar(A))β Hβ(Cobar(Aββ)),
where the cohomology of the bar complex of the Steenrod algebra A is isomorphic to the cohomology of the cobar complex of the dual for A which we denote as Aββ[7, 8]. This is the bar-cobar isomorphism. So far we've been using the dual of the reduced bar complex:
((AΛβs)β,βsββ)
where we have the differentials βsββ:(AΛβsβ1)ββ(AΛβs)β. When we switch to this new cobar complex, we get the following:
((AΛβ)βs,ds)
where we have these new differentials ds:(AΛβ)βsβ(AΛβ)βs+1. Importantly, note that these are not isomorphic when s>1:
(AΛβ)βsβ(AΛβs)β
So we build this new cobar complex from the dual algebra Aββ which is a graded-commutative polynomial algebra which we denote:
Aββ=Z2β[ΞΎ1β,ΞΎ2β,ΞΎ3β,β¦]
where these generators ΞΎkβ have internal degree deg(ΞΎkβ)=2kβ1. Similar to the A case, we look at the augmentation ideal AΛβ of Aββ to get Cs=(AΛβ)βs where our differentials are maps ds:CsβCs+1. Furthermore, note that s in this case is the same s-degree in Exts,β, and we also denote the degree t with a subscript Ctsβ. Let's use this new cobar framework to view our previous results for Ext1 again:
h0ββExt1,1 is represented by the cocycle ΞΎ1ββC11β since deg(ΞΎ1β)=21β1=1.
h1ββExt1,2 is represented by the cocycle ΞΎ12ββC21β since deg(ΞΎ12β)=deg(ΞΎ1β)+deg(ΞΎ1β)=2.
h2ββExt1,4 is represented by the cocycle ΞΎ14ββC41β since deg(ΞΎ14β)=4.
In general, hiββExt1,2i is represented by the cocycle ΞΎ12iββC2i1β with such that deg(ΞΎ12iβ)=2i. Note that this degree is different from the degree of ΞΎkβ as a generator which is 2kβ1.
We take the cohomology of this cobar complex ((AΛβ)βs,ds)=(Cs,ds) to get Ext:
h02β: We're finally ready to compute h02β. Recall that h02β=h0βh0β has bidegree (s,t)=(2,2), so we're computing ExtA2,2ββ ker(d2:C22ββC23β)/im(d1:C21ββC22β). Furthermore, h0β is represented by the 1-cocycle ΞΎ1ββC11β, so h02β=h0βh0β is represented by the 2-cochain ΞΎ1ββΞΎ1ββC22β. Side note: we call general elements xβCscochains, and if xβker(ds) then it's a cocycle, and if xβim(dsβ1) then it's a coboundary. Additionally, as we do a deep dive into ker(ds) and im(dsβ1), note that the differential ds follows the Leibniz rule (see blog post #3) since it's induced by the cup product. Let's see this in action to see if ΞΎ1ββΞΎ1ββker(d2), i.e. to check if ΞΎ1ββΞΎ1β is a cocycle:
Note that (β1)deg(ΞΎkβ)=1 for any k since we're working in Z2β (mod 2). Note that our d1 operator is the same as the reduced coproduct ΞΛ:AΛββAΛββAΛβ. This map is derived from the full coproduct Ξ on the dual Steenrod algebra Aββ by the relation:
ΞΛ(x)=Ξ(x)β(xβ1)β(1βx),
where ΞΛ captures the non-trivial part of the coproduct by subtracting the trivial parts xβ1 and 1βx, etc. With this, we get:
where ΞΛ(ΞΎ1β)=0 since Ξ(ΞΎ1β)=ΞΎ1ββ1+1βΞΎ1β. Therefore, ΞΎ1ββΞΎ1ββker(d2), so it's a cocycle. Next, let's see if it's a coboundary, i.e. if ΞΎ1ββΞΎ1ββim(d1:C21ββC22β). Looking at the source space C21β, note that this has degree 2, and we have the following available generators from AΛβ: ΞΎ1β,ΞΎ2β,β¦ such that deg(ΞΎ1β)=1,deg(ΞΎ2β)=3, etc. There's no single degree 2 generator, so the only way to get a generator for C21β is to consider ΞΎ12β with degree deg(ΞΎ12β)=2. We don't have anything else (since even considering ΞΎ1βΞΎ2β has degree 4), so our source space C21β is spanned by the single element ΞΎ12β, i.e.
C21β=span{ΞΎ12β}
Let's see how d1=ΞΛ maps ΞΎ12β to C22β. First, let's look at Ξ(ΞΎ12β):
h0βh1β: As an additional example, let's look at h0βh1β. Similar to before, recall that h0βh1β has bidegree (s,t)=(2,3), so we're looking at
Let's look at ker(d2). We have the source space C32β with elements from AΛββAΛβ. Note that since 1β/AΛβ then we can't have elements like 1βΞΎ2β or ΞΎ2ββ1. Since any variation of ΞΎ2ββΞΎ1β will have degree β₯4 then we can only span C32β using ΞΎ1β. We get:
C32β=span{ΞΎ12ββΞΎ1β,ΞΎ1ββΞΎ12β},
where the degree of both ΞΎ12ββΞΎ1β and ΞΎ1ββΞΎ12β is 3. Let's see how d2 maps these. Let's start with ΞΎ12ββΞΎ1β:
where d1(ΞΎ12β)=0 from our previous h02β calculation, and d1(ΞΎ1β)=0 since ΞΎ1β is primitive. Note that d2(ΞΎ1ββΞΎ12β) is a similar story where we get d2(ΞΎ1ββΞΎ12β)=0. Since d2 maps everything from the source space to zero, then we get:
Next, let's look at the coboundaries im(d1). In this case, since we're getting elements of C31β from AΛβ then we get the following:
C31β=span{ΞΎ13β,ΞΎ2β}
Let's apply d1=ΞΛ to this basis. First, let's look at Ξ for ΞΎ13β and then just remove the trivial terms ΞΎ13ββ1 and 1βΞΎ13β. Note that we'll be using the (mod 2) expansion of (a+b)3=a3+a2b+ab2+b3:
where we get ΞΎ12ββΞΎ1β from (ΞΎ12ββ1)β(1βΞΎ1β), and ΞΎ1ββΞΎ12β from (ΞΎ1ββ1)β(1βΞΎ12β). We can remove the trivial terms (ΞΎ13ββ1 and 1βΞΎ13β) from this expression to get:
Note that this space is equivalent to span{ΞΎ12ββΞΎ1β,ΞΎ1ββΞΎ12β} since the generating sets span the same space. We get the following for Ext2,3:
This implies that h0βh1β is a coboundary, thus h0βh1β=0. Overall, we could continue analyzing ExtA2,ββ to fully characterize the s=2 column (but we won't do that here). So far we've only scratched the surface of the Adams E2β page, but this feels like a good spot to wrap things up. π»
[1] N. E. Steenrod, Products of cocycles and extensions of mappings,
Ann. of Math. (2) 48 (1947), 290β320.
[2] N. E. Steenrod and D. B. A. Epstein,
Cohomology Operations,
Princeton University Press, Princeton, N.J., 1962.
[3] H. R. Margolis, Spectra and the Steenrod Algebra,
North-Holland Mathematical Library, vol. 29, North-Holland, Amsterdam, 1983.
[4] G. C. Rota, Review of "Spectra and the Steenrod Algebra" by H. R. Margolis,
Adv. Math.59 (1986), 1β2.
[5] R.E. Mosher and M.C. Tangora, Cohomology Operations and Applications in Homotopy Theory, Dover Publications, 2008.
[6] J. Milnor, The Steenrod algebra and its dual, Ann. of Math.67 (1958), 150β171.
[7] J.F. Adams, On the cobar construction, Proc. Nat. Acad. Sci. U.S.A.42 (1956), 409β412.
[8] J.F. Adams, On the structure and applications of the Steenrod algebra, Comment. Math. Helv.32 (1958), 180β214.