Halloween 4: The Return of Adams Spectral Sequence πŸŽƒ

Enrique De Alba ; October 31, 2025


Happy Halloween! We're first going to define the Steenrod algebra A\mathcal{A} which by the end of this post will be the best friend we never knew we always wanted.

The Steenrod algebra, denoted A,\mathcal{A}, is the graded, associative, non-commutative Z2\mathbb{Z}_2-algebra generated by the Steenrod squares {Sqi}iβ‰₯0\{Sq^i\}_{i\geq 0} subject to the Adem relations:

Sqa∘Sqb=βˆ‘k=0⌊a/2βŒ‹(bβˆ’kβˆ’1aβˆ’2k)Sqa+bβˆ’k∘SqkSq^a \circ Sq^b = \sum_{k=0}^{\lfloor a / 2 \rfloor} {{b-k-1} \choose {a-2k}} Sq^{a+b-k} \circ Sq^{k}

where a,ba,b are integers such that 0<a<2b,0 < a < 2b, and where Sq0=1Sq^0 = 1 is the identity [1]. The grading of A\mathcal{A} is given by deg⁑(Sqi)=i,\deg(Sq^i) = i, and deg⁑(Sqa∘Sqb)=a+b.\deg(Sq^a \circ Sq^b) = a+b. While {Sqi}iβ‰₯0\{Sq^i\}_{i\geq 0} generates A,\mathcal{A}, it does not form a basis. For that, we'll need the concept of admissible monomials.

A monomial Sqinβˆ˜β‹―βˆ˜Sqi1Sq^{i_n} \circ \cdots \circ Sq^{i_1} is called admissible iff ikβ‰₯2ikβˆ’1i_k \geq 2 i_{k-1} for all k.k. For example, Sq2∘Sq2Sq^2 \circ Sq^2 is not an admissible monomial since 2≱4.2 \ngeq 4. However, we can use the Adem relations to get that

Sq2∘Sq2=Sq3∘Sq1Sq^2 \circ Sq^2 = Sq^3 \circ Sq^1

where Sq3∘Sq1Sq^3 \circ Sq^1 is admissible since 3β‰₯2β‹…1.3 \geq 2\cdot 1.

Theorem: The set of all admissible monomials forms a vector space basis for the A,\mathcal{A}, i.e. the Steenrod algebra [2]. This means that any composition of Steenrod squares in A\mathcal{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 E2E_2 page for the sphere spectrum S,\mathbb{S}, which is the bigraded algebra:

ExtAs,t(Z2,Z2)\text{Ext}_{\mathcal{A}}^{s, t}(\mathbb{Z}_2, \mathbb{Z}_2)

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. Ο€βˆ—S.\pi_*^S. These guys are way spookier than homology or even cohomology πŸ‘». Note that the E2E_2 page for S\mathbb{S} is:

E2s,t=ExtAs,t(Hβˆ—(S;Z2),Z2),E_2^{s,t} = \text{Ext}_{\mathcal{A}}^{s,t}(H^*(\mathbb{S};\mathbb{Z}_2), \mathbb{Z}_2),

and while we won't prove this here, it turns out that Hβˆ—(S;Z2)β‰…Z2,H^*(\mathbb{S};\mathbb{Z}_2) \cong \mathbb{Z}_2, hence why we'll be doing a deep-dive into ExtAs,t(Z2,Z2).\text{Ext}_{\mathcal{A}}^{s, t}(\mathbb{Z}_2, \mathbb{Z}_2).

Ext\text{Ext}: What is Ext\text{Ext}? Ext\text{Ext} is short for extension and is a regular in the stage of homological algebra. Consider the following short exact sequence (SES) of modules:

0→    A→    B→    C→    00 \xrightarrow{\,\, \,\,} A \xrightarrow{\,\, \,\,} B \xrightarrow{\,\, \,\,} C \xrightarrow{\,\, \,\,} 0

Let's apply the (contravariant) functor HomA(βˆ’,Z2)\text{Hom}_{\mathcal{A}}(-, \mathbb{Z}_2) to get:

0→    HomA(C,Z2)→    HomA(B,Z2)→    HomA(A,Z2)→    00 \xrightarrow{\,\, \,\,} \text{Hom}_{\mathcal{A}}(C, \mathbb{Z}_2) \xrightarrow{\,\, \,\,} \text{Hom}_{\mathcal{A}}(B, \mathbb{Z}_2) \xrightarrow{\,\, \,\,} \text{Hom}_{\mathcal{A}}(A, \mathbb{Z}_2) \xrightarrow{\,\, \,\,} 0

Suppose our modules are left A\mathcal{A}-modules, then this new sequence will be left-exact which just means that the induced map HomA(C,Z2)→    HomA(B,Z2)\text{Hom}_{\mathcal{A}}(C, \mathbb{Z}_2) \xrightarrow{\,\, \,\,} \text{Hom}_{\mathcal{A}}(B, \mathbb{Z}_2) is injective. Now, the right map HomA(B,Z2)→    HomA(A,Z2)\text{Hom}_{\mathcal{A}}(B, \mathbb{Z}_2) \xrightarrow{\,\, \,\,} \text{Hom}_{\mathcal{A}}(A, \mathbb{Z}_2) is not generally surjective, so the sequence as a whole is not (in general) exact. The Ext\text{Ext} groups are the algebraic objects that measure this "failure of exactness."

ExtAs(M,N)\text{Ext}_{\mathcal{A}}^{s}(M, N): Let's see how we construct these extensions. We'll start by looking at a general ExtAs(M,N)\text{Ext}_{\mathcal{A}}^s(M,N) case which is equivalent to our case study, just with M=Z2M=\mathbb{Z}_2 and N=Z2.N=\mathbb{Z}_2. We construct ExtAs(M,N)\text{Ext}_{\mathcal{A}}^s(M,N) in three simple steps:

  1. Resolve: We first look at the projective resolution of M.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" MM:

    β‹―β†’β€‰β€‰βˆ‚2  P1β†’β€‰β€‰βˆ‚1  P0→  ϡ  M→    0,\cdots \xrightarrow{\,\, \partial_2 \,\,} P_1 \xrightarrow{\,\, \partial_1 \,\,} P_0 \xrightarrow{\,\, \epsilon \,\,} M \xrightarrow{\,\, \,\,} 0,

    where each PsP_s is a projective A\mathcal{A}-module and Ο΅\epsilon is a surjective map.

  2. Apply Hom: We apply the HomA(βˆ’,N)\text{Hom}_{\mathcal{A}}(-,N) functor to the sequence after first removing the MM object from the sequence to get:

    0→    HomA(P0,N)β†’β€‰β€‰βˆ‚1βˆ—β€‰β€‰HomA(P1,N)β†’β€‰β€‰βˆ‚2βˆ—β€‰β€‰β‹―0 \xrightarrow{\,\, \,\,} \text{Hom}_\mathcal{A}(P_0, N) \xrightarrow{\,\, \partial_1^* \,\,} \text{Hom}_\mathcal{A}(P_1, N) \xrightarrow{\,\, \partial_2^* \,\,} \cdots

    This new sequence is not guaranteed to be exact, but we do have the property βˆ‚s+1βˆ—βˆ˜βˆ‚sβˆ—=0,\partial_{s+1}^* \circ \partial_{s}^* = 0, like that of cochain complexes in singular cohomology.

  3. Take the Cohomology: There's not much else to it. The Ext\text{Ext} groups are defined as the cohomology of this complex:

    ExtAs(M,N):=ker⁑(βˆ‚s+1βˆ—)im(βˆ‚sβˆ—),\text{Ext}_\mathcal{A}^s(M,N) := \frac{\ker(\partial_{s+1}^*)}{\text{im}(\partial_{s}^*)},

    where ker⁑(βˆ‚s+1βˆ—)/im(βˆ‚sβˆ—)\ker(\partial_{s+1}^*) / \text{im}(\partial_{s}^*) is the ss-th cohomology group Hs(HomA(Pβˆ—,N)).H^s(\text{Hom}_\mathcal{A}(P_*, N)).

Unfortunately, for the Steenrod algebra A,\mathcal{A}, finding a simple projective resolution of Z2\mathbb{Z}_2 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.\mathcal{A}.

Let AΛ‰:=ker⁑(Ο΅:Aβ†’Z2)\bar{\mathcal{A}}:= \ker(\epsilon: \mathcal{A}\to\mathbb{Z}_2) denote the augmentation ideal. This is the vector space spanned by all the Steenrod squares of positive degree, i.e. {Sqi}i>0.\{Sq^i\}_{i>0}. Importantly, note that Sq0=1βˆ‰AΛ‰.Sq^0 = 1 \notin \bar{\mathcal{A}}. The bar resolution for Z2\mathbb{Z}_2 is given by

β‹―β†’β€‰β€‰βˆ‚βˆ—β€‰β€‰Pβˆ—β†’β€‰β€‰β€‰β€‰Z2→    0\cdots \xrightarrow{\,\, \partial_* \,\,} P_* \xrightarrow{\,\, \,\,} \mathbb{Z}_2 \xrightarrow{\,\, \,\,} 0

where the ss-th term is:

Ps=AβŠ—(AΛ‰)βŠ—sP_s = \mathcal{A} \otimes (\bar{\mathcal{A}})^{\otimes s}

For example, we have: P0=A,P_0 = \mathcal{A}, P1=AβŠ—AΛ‰,P_1 = \mathcal{A} \otimes \bar{\mathcal{A}}, P2=AβŠ—AΛ‰βŠ—AΛ‰,P_2 = \mathcal{A} \otimes \bar{\mathcal{A}} \otimes \bar{\mathcal{A}}, and so on [4]. The differential

βˆ‚s:Psβ†’Psβˆ’1\partial_s: P_s \to P_{s-1}

makes use of the following notation [a1∣a2βˆ£β‹―βˆ£as][a_1 \mid a_2 \mid \cdots \mid a_s] for elements in (AΛ‰)βŠ—s.(\bar{\mathcal{A}})^{\otimes s}. Hence, we denote elements in Ps=AβŠ—(AΛ‰)βŠ—sP_s = \mathcal{A} \otimes (\bar{\mathcal{A}})^{\otimes s} as aβŠ—[a1βˆ£β‹―βˆ£as].a \otimes [a_1 \mid \cdots \mid a_s]. The differential maps these elements to βˆ‚sβˆ’1\partial_{s-1} the following way:

βˆ‚s ⁣(aβŠ—[a1βˆ£β‹―βˆ£as])= aβ‹…a1βŠ—[a2∣a3βˆ£β‹―βˆ£as]+βˆ‘i=1sβˆ’1(βˆ’1)iaβŠ—[a1βˆ£β‹―βˆ£aiai+1βˆ£β‹―βˆ£as]\begin{aligned} \partial_s\!\left(a \otimes [a_1\mid \cdots \mid a_s]\right) = \,&a\cdot a_1 \otimes [a_2\mid a_3 \mid \cdots \mid a_s] \\ + &\sum_{i=1}^{s-1} (-1)^i a \otimes [a_1 \mid \cdots \mid a_ia_{i+1} \mid \cdots \mid a_s] \end{aligned}

Note that the (βˆ’1)i(-1)^i term is just 11 in our case since A\mathcal{A} is over Z2.\mathbb{Z}_2. Additionally, note that the first term has the A\mathcal{A}-action with aβ‹…a1.a \cdot a_1. In the sum, we just multiply all adjacent aiai+1a_{i}a_{i+1} terms inside the (AΛ‰)βŠ—(\bar{\mathcal{A}})^{\otimes} bracket. Example:

βˆ‚2(aβŠ—[a1∣a2])=aβ‹…a1βŠ—[a2]+aβŠ—[a1a2]\begin{aligned} \partial_2(a \otimes [a_1 \mid a_2]) = &a \cdot a_1 \otimes [a_2] \\ + &a \otimes [a_1a_2] \end{aligned}

Note that here we use + to show that βˆ’1≑1-1 \equiv 1 (mod 2). Side note: the summation for βˆ‚s\partial_s usually includes a term

(βˆ’1)saβŠ—[a1βˆ£β‹―βˆ£asβˆ’1]Ο΅(as),(-1)^sa\otimes[a_1 \mid \cdots \mid a_{s-1}]\epsilon(a_s),

but since as∈AΛ‰a_s \in \bar{\mathcal{A}} then Ο΅(as)=0\epsilon(a_s) = 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:

ϡ:A→Z2\epsilon: \mathcal{A} \to \mathbb{Z}_2

Ο΅\epsilon is a homomorphism that strips away all the complexity in A.\mathcal{A}. Note that for any element aβ€²βˆˆAΛ‰βŠ‚A,a' \in \bar{\mathcal{A}} \subset \mathcal{A}, we get Ο΅(aβ€²)=0\epsilon(a') = 0 (since we defined AΛ‰\bar{\mathcal{A}} to be the kernel of this Ο΅\epsilon map). Additionally, note that Ο΅(1)=1\epsilon(1) = 1 since it's a homomorphism. We define the trivial action of A\mathcal{A} on Z2\mathbb{Z}_2 using this Ο΅\epsilon map , where we have the trivial action aβ‹…ka \cdot k for any a∈A,a \in \mathcal{A}, k∈Z2k \in \mathbb{Z}_2 such that:

aβ‹…k=Ο΅(a)ka \cdot k = \epsilon(a)k

Example 1: Sq1β‹…1=Ο΅(Sq1)1=0,Sq^1 \cdot 1 = \epsilon(Sq^1)1 = 0, since Ο΅(Sq1)=0.\epsilon(Sq^1) = 0.

Example 2:

(1+Sq2)β‹…1=Ο΅(1+Sq2)1=Ο΅(1)1+Ο΅(Sq2)1=1+0=1\begin{aligned} (1+Sq^2) \cdot 1 &= \epsilon(1 + Sq^2)1 \\ &= \epsilon(1)1 + \epsilon(Sq^2)1 \\ &= 1 + 0 = 1 \end{aligned}

An A\mathcal{A}-module map ϕ:Ps→Z2\phi: P_s \to \mathbb{Z}_2 is a map

Ο•:AβŠ—(AΛ‰)βŠ—sβ†’Z2\phi: \mathcal{A} \otimes (\bar{\mathcal{A}})^{\otimes s} \to \mathbb{Z}_2

that satisfies the trivial action we defined above, i.e.

Ο•(aβ‹…x)=aβ‹…Ο•(x)=Ο΅(a)Ο•(x),\phi(a \cdot x) = a \cdot \phi(x) = \epsilon(a)\phi(x),

since Ο•(x)∈Z2,\phi(x) \in \mathbb{Z}_2, hence why a∈Aa \in \mathcal{A} acts trivially on it via the augmentation map Ο΅.\epsilon. Note that Ο΅(a)=0,\epsilon(a) = 0, as well as Ο•(aβ‹…x)=0,\phi(a \cdot x) = 0, for anything not starting with 1. This means that Ο•\phi is completely determined by what it does to elements of the form

1βŠ—[a1βˆ£β‹―βˆ£as]1 \otimes [a_1 \mid \cdots \mid a_s]

The next thing we do is apply HomA(βˆ’,Z2)\text{Hom}_{\mathcal{A}}(-, \mathbb{Z}_2) to our bar complex, so we'll have the groups HomA(Ps,Z2).\text{Hom}_\mathcal{A}(P_s, \mathbb{Z}_2). However, note that these map homomorphisms are the same as our Ο•,\phi, so we get the following isomorphism of vector spaces:

HomA(Ps,Z2)β‰…HomZ2((AΛ‰)βŠ—s,Z2),\text{Hom}_\mathcal{A}(P_s, \mathbb{Z}_2) \cong \text{Hom}_{\mathbb{Z}_2}( (\bar{\mathcal{A}})^{\otimes s}, \mathbb{Z}_2 ),

where we get this HomZ2(βˆ’)\text{Hom}_{\mathbb{Z}_2}(-) simplification since we're now dealing with just AΛ‰\bar{\mathcal{A}} and we don't have to worry about the A\mathcal{A}-module structure from A\mathcal{A} since we've basically factored it out. Furthermore, since AΛ‰\bar{\mathcal{A}} is just a Z2\mathbb{Z}_2 vector space, then note that this HomZ2((AΛ‰)βŠ—s,Z2)\text{Hom}_{\mathbb{Z}_2}( (\bar{\mathcal{A}})^{\otimes s}, \mathbb{Z}_2 ) is precisely the definition of the dual of (AΛ‰)βŠ—s.(\bar{\mathcal{A}})^{\otimes s}. We'll denote this dual as (AΛ‰βŠ—s)βˆ—(\bar{\mathcal{A}}^{\otimes s})^* and will denote most other duals with Vβˆ—V^*-notation, however, we'll be making a special exception for the Steenrod algebra itself which is denoted Aβˆ—\mathcal{A}_* and we'll explain why later (spoiler alert: it's because A\mathcal{A} is a Hopf algebra). Altogether, we have the following isomorphism:

HomA(Ps,Z2)β‰…(AΛ‰βŠ—s)βˆ—\text{Hom}_\mathcal{A}(P_s, \mathbb{Z}_2) \cong (\bar{\mathcal{A}}^{\otimes s})^*

We also get the following induced differentials βˆ‚sβˆ—\partial^*_s which map from HomA(Psβˆ’1,Z2)\text{Hom}_\mathcal{A}(P_{s-1}, \mathbb{Z}_2) to HomA(Ps,Z2)\text{Hom}_\mathcal{A}(P_{s}, \mathbb{Z}_2) (reversed due to contravariant functor), i.e.

βˆ‚sβˆ—:(AΛ‰βŠ—sβˆ’1)βˆ—β†’(AΛ‰βŠ—s)βˆ—\partial_s^*: (\bar{\mathcal{A}}^{\otimes s-1})^* \to (\bar{\mathcal{A}}^{\otimes s})^*

We denote this entire dual bar complex object as ((AΛ‰βŠ—s)βˆ—,βˆ‚sβˆ—). \left( (\bar{\mathcal{A}}^{\otimes s})^*, \partial_s^* \right) . Before in our general HomA(βˆ’,N)\text{Hom}_\mathcal{A}(-,N) case we took the cohomology of the complex (HomA(Pβˆ—,N),βˆ‚sβˆ—),\left( \text{Hom}_\mathcal{A}(P_*, N), \partial_s^* \right), but now we'll be taking the cohomology of ((AΛ‰βŠ—s)βˆ—,βˆ‚sβˆ—)\left( (\bar{\mathcal{A}}^{\otimes s})^*, \partial_s^* \right) to get ExtA(Z2,Z2)\text{Ext}_\mathcal{A}(\mathbb{Z}_2, \mathbb{Z}_2) [5].

Let's compute ExtA1(Z2,Z2).\text{Ext}_\mathcal{A}^1(\mathbb{Z}_2, \mathbb{Z}_2). Looking at the cohomology of ((AΛ‰βŠ—s)βˆ—,βˆ‚sβˆ—), \left( (\bar{\mathcal{A}}^{\otimes s})^*, \partial_s^* \right), we get:

ExtA1(Z2,Z2)β‰…ker⁑(βˆ‚2βˆ—)im(βˆ‚1βˆ—)\text{Ext}_\mathcal{A}^1(\mathbb{Z}_2, \mathbb{Z}_2) \cong \frac{\ker( \partial_2^* )}{ \text{im}( \partial_1^* )}

Let's first look at βˆ‚1βˆ—:(AΛ‰βŠ—0)βˆ—β†’(AΛ‰βŠ—1)βˆ—.\partial_1^*: (\bar{\mathcal{A}}^{\otimes 0})^* \to (\bar{\mathcal{A}}^{\otimes 1})^*. Note that this dual (AΛ‰βŠ—0)βˆ—β‰…HomZ2(AΛ‰βŠ—0,Z2),(\bar{\mathcal{A}}^{\otimes 0})^* \cong \text{Hom}_{\mathbb{Z}_2}( \bar{\mathcal{A}}^{\otimes 0}, \mathbb{Z}_2 ), and since AΛ‰\bar{\mathcal{A}} is a Z2\mathbb{Z}_2 vector space then AΛ‰βŠ—0β‰…Z2,\bar{\mathcal{A}}^{\otimes 0} \cong \mathbb{Z}_2, so we get:

HomZ2(Z2,Z2)β‰…Z2\text{Hom}_{\mathbb{Z}_2}( \mathbb{Z}_2, \mathbb{Z}_2 ) \cong \mathbb{Z}_2

so we just get the following differential:

βˆ‚1βˆ—:Z2β†’AΛ‰βˆ—\partial_1^*: \mathbb{Z}_2 \to \bar{\mathcal{A}}^*

This is dual to βˆ‚1:AβŠ—AΛ‰β†’A\partial_1: \mathcal{A} \otimes \bar{\mathcal{A}} \to \mathcal{A} such that βˆ‚1(aβŠ—[a1])=aβ‹…a1.\partial_1(a \otimes [a_1]) = a\cdot a_1. Let Ο•βˆˆHomA(P0,Z2)β‰…Z2,\phi \in \text{Hom}_\mathcal{A}(P_0, \mathbb{Z}_2) \cong \mathbb{Z}_2, where this space is spanned by the augmentation map Ο΅:Aβ†’Z2.\epsilon: \mathcal{A} \to \mathbb{Z}_2. Consider the pull-back βˆ‚1βˆ—(Ο•)=Ο•βˆ˜βˆ‚1\partial_1^*(\phi) = \phi \circ \partial_1 where we use inputs from AβŠ—AΛ‰.\mathcal{A} \otimes \bar{\mathcal{A}}. Let aβŠ—[a1]∈AβŠ—AΛ‰,a \otimes [a_1] \in \mathcal{A} \otimes \bar{\mathcal{A}}, then consider:

Ο•βˆ˜βˆ‚1(aβŠ—[a1])=Ο•(aβ‹…a1)=Ο΅(a)Ο•(a1)\begin{aligned} \phi \circ \partial_1( a \otimes [a_1] ) &= \phi ( a \cdot a_1 ) \\ &= \epsilon(a)\phi(a_1) \end{aligned}

Note that since Ο•βˆˆZ2\phi \in \mathbb{Z}_2 then Ο•(a1)=0\phi(a_1) = 0 if Ο•\phi is the zero map, so suppose Ο•=Ο΅.\phi = \epsilon. Note that Ο•(a1)=Ο΅(a1)=0\phi(a_1) = \epsilon(a_1) = 0 for all a1a_1 in AΛ‰.\bar{\mathcal{A}}. This means that Ο΅(a)Ο΅(a1)=0, \epsilon(a)\epsilon(a_1) = 0, hence

βˆ‚1βˆ—(Ο•)=Ο•βˆ˜βˆ‚1(aβŠ—[a1])=0\partial_1^*(\phi) = \phi \circ \partial_1( a \otimes [a_1] ) = 0

for any aβŠ—[a1]∈AβŠ—AΛ‰. a \otimes [a_1] \in \mathcal{A} \otimes \bar{\mathcal{A}}. This means that βˆ‚1βˆ—\partial_1^* is the zero map, and hence:

im(βˆ‚1βˆ—)=0\text{im}(\partial_1^*) = 0

Therefore, we get a major simplification for Ext1\text{Ext}^1 where we have:

ExtA1(Z2,Z2)β‰…ker⁑(βˆ‚2βˆ—)\text{Ext}_\mathcal{A}^1(\mathbb{Z}_2, \mathbb{Z}_2) \cong \ker(\partial_2^*)

Let's look at this differential βˆ‚2βˆ—:AΛ‰βˆ—β†’(AΛ‰βŠ—AΛ‰)βˆ—.\partial_2^*: \bar{\mathcal{A}}^* \to (\bar{\mathcal{A}} \otimes \bar{\mathcal{A}})^*. Note that AΛ‰βˆ—β‰…HomZ2(AΛ‰,Z2)\bar{\mathcal{A}}^* \cong \text{Hom}_{\mathbb{Z}_2}(\bar{\mathcal{A}}, \mathbb{Z}_2) and ker⁑(βˆ‚2βˆ—)βŠ‚AΛ‰βˆ—.\ker(\partial_2^*) \subset \bar{\mathcal{A}}^*. This means that for ker⁑(βˆ‚2βˆ—),\ker(\partial_2^*), we're looking for all the AΛ‰β†’Z2\bar{\mathcal{A}} \to \mathbb{Z}_2 map homomorphisms, i.e. let f∈HomZ2(AΛ‰,Z2),f \in \text{Hom}_{\mathbb{Z}_2}(\bar{\mathcal{A}}, \mathbb{Z}_2), such that βˆ‚2βˆ—(f)=0\partial_2^*(f) = 0 in (AΛ‰βŠ—AΛ‰)βˆ—.(\bar{\mathcal{A}} \otimes \bar{\mathcal{A}})^*. Consider the pull-back:

βˆ‚2βˆ—(f)=fβˆ˜βˆ‚2=0,\partial_2^*(f) = f \circ \partial_2 = 0,

where βˆ‚2βˆ—\partial_2^* is the dual of βˆ‚2\partial_2 such that βˆ‚2:AβŠ—AΛ‰βŠ—AΛ‰β†’AβŠ—AΛ‰.\partial_2: \mathcal{A} \otimes \bar{\mathcal{A}} \otimes \bar{\mathcal{A}} \to \mathcal{A} \otimes \bar{\mathcal{A}}.

Let aβŠ—[a1∣a2]∈AβŠ—AΛ‰βŠ—AΛ‰a \otimes [a_1 \mid a_2] \in \mathcal{A} \otimes \bar{\mathcal{A}} \otimes \bar{\mathcal{A}}, and note that

βˆ‚2(aβŠ—[a1∣a2])= aβ‹…a1βŠ—[a2]+aβŠ—[a1a2]\begin{aligned} \partial_2(a \otimes [a_1 \mid a_2]) =\,&a\cdot a_1 \otimes [a_2] \\ + &a \otimes [a_1 a_2] \end{aligned}

Then when we apply ff we get

f(aβ‹…a1βŠ—[a2])+f(aβŠ—[a1a2])f(a\cdot a_1 \otimes [a_2]) + f(a \otimes [a_1 a_2])

Note that this term f(aβ‹…a1βŠ—[a2])=0f(a\cdot a_1 \otimes [a_2]) = 0 since we have the term Ο΅(aβ‹…a1)=Ο΅(a)Ο΅(a1)=0.\epsilon(a\cdot a_1) = \epsilon(a)\epsilon(a_1) = 0. For the second term we get:

f(aβŠ—[a1a2])=Ο΅(a)f(1βŠ—[a1a2])f(a \otimes [a_1 a_2]) = \epsilon(a)f(1 \otimes [a_1 a_2])

Which means that fβˆ˜βˆ‚2=0f \circ \partial_2 = 0, hence we want f(1βŠ—[a1a2])=0.f(1 \otimes [a_1 a_2]) = 0. This is equivalent to looking for maps f:AΛ‰β†’Z2f: \bar{\mathcal{A}} \to \mathbb{Z}_2 such that f([a1a2])=0f([a_1a_2]) = 0 for all a1,a2∈AΛ‰a_1,a_2 \in \bar{\mathcal{A}} since we can just factor out the (1βŠ—βˆ’)(1 \otimes -) term. Additionally, another way to see this is that we're looking for maps ff such that for any elements [a]∈AΛ‰[a] \in \bar{\mathcal{A}} that can be written as a=a1a2a = a_1a_2 for a1,a2∈AΛ‰a_1,a_2 \in \bar{\mathcal{A}} we get f([a])=0.f([a]) = 0.

Let's take a tiny step back and then tie this back to f([a])=0f([a]) = 0 for a=a1a2.a = a_1a_2. Elements in the Steenrod algebra A\mathcal{A} are either decomposable or indecomposable. A simple example is Sq3Sq^3 which can be rewritten as Sq3=Sq1∘Sq2Sq^3 = Sq^1 \circ Sq^2 via the Adem relations. Since we can write Sq3=Sq1∘Sq2,Sq^3 = Sq^1 \circ Sq^2, this means that Sq3Sq^3 is decomposable [6]. Examples of indecomposable elements are Sq1Sq^1, Sq2Sq^2, and Sq4.Sq^4. Spoiler alert: it turns out that indecomposable elements are of the form Sq2iSq^{2^i} which becomes relevant very soon. Recall that AΛ‰\bar{\mathcal{A}} is spanned by {Sqi}i>0,\{ Sq^i \}_{i > 0}, so Sq3∈AΛ‰Sq^3 \in \bar{\mathcal{A}} such that we can write Sq3=a1a2Sq^3 = a_1a_2 where a1=Sq1a_1 = Sq^1 and a2=Sq2.a_2 = Sq^2. Note that this means that Sq3Sq^3 is such that f([Sq3])=0f([Sq^3]) = 0 for f∈ker⁑(βˆ‚2βˆ—).f \in \ker(\partial_2^*).

So we want our maps ff to vanish for any decomposable element in AΛ‰.\bar{\mathcal{A}}. Note that when we write these decomposable elements like a1a2,a_1a_2, these are elements in AΛ‰β‹…AΛ‰βŠ‚AΛ‰.\bar{\mathcal{A}}\cdot\bar{\mathcal{A}} \subset \bar{\mathcal{A}}. This means that ker⁑(βˆ‚2βˆ—)\ker(\partial_2^*) is the set of all AΛ‰β†’Z2\bar{\mathcal{A}} \to \mathbb{Z}_2 maps that are zero at AΛ‰β‹…AΛ‰.\bar{\mathcal{A}}\cdot\bar{\mathcal{A}}. Note the following equivalence:

{TheΒ setΒ ofΒ AΛ‰β†’Z2Β mapsthatΒ vanishΒ atΒ AΛ‰β‹…AΛ‰βŠ‚AΛ‰}⟺{TheΒ setΒ ofΒ allΒ mapsAΛ‰/(AΛ‰β‹…AΛ‰)β†’Z2}\left\{ \begin{array}{c} \text{The set of $\bar{\mathcal{A}} \to \mathbb{Z}_2$ maps} \\ \text{that vanish at $\bar{\mathcal{A}}\cdot \bar{\mathcal{A}} \subset \bar{\mathcal{A}}$} \end{array} \right\} \Longleftrightarrow \left\{ \begin{array}{c} \text{The set of \emph{all} maps} \\ \text{$\bar{\mathcal{A}} / (\bar{\mathcal{A}} \cdot \bar{\mathcal{A}}) \to \mathbb{Z}_2$} \end{array} \right\}

And then note that the set of maps from Aˉ/(Aˉ⋅Aˉ)→Z2\bar{\mathcal{A}} / (\bar{\mathcal{A}} \cdot \bar{\mathcal{A}}) \to \mathbb{Z}_2 is just the dual:

HomZ2(AΛ‰/(AΛ‰β‹…AΛ‰),Z2)β‰…(AΛ‰/(AΛ‰β‹…AΛ‰))β€‰β£βˆ—\text{Hom}_{\mathbb{Z}_2}( \bar{\mathcal{A}} / (\bar{\mathcal{A}} \cdot \bar{\mathcal{A}}), \mathbb{Z}_2 ) \cong \left( \bar{\mathcal{A}} / (\bar{\mathcal{A}} \cdot \bar{\mathcal{A}})\right)^{\!*}

Putting everything together, we get that

ExtA1(Z2,Z2)β‰…ker⁑(βˆ‚2βˆ—)β‰…(AΛ‰/(AΛ‰β‹…AΛ‰))β€‰β£βˆ—\begin{aligned} \text{Ext}_{\mathcal{A}}^1(\mathbb{Z}_2, \mathbb{Z}_2) &\cong \ker(\partial_2^*) \\ &\cong \left( \bar{\mathcal{A}} / (\bar{\mathcal{A}} \cdot \bar{\mathcal{A}})\right)^{\!*} \end{aligned}

Importantly, note that since we're modding out by all the decomposable elements in Aˉ\bar{\mathcal{A}} then we're left with the indecomposable elements.

Theorem: The vector space of indecomposable elements of A,\mathcal{A}, i.e. AΛ‰/(AΛ‰β‹…AΛ‰)βŠ‚A,\bar{\mathcal{A}} / (\bar{\mathcal{A}} \cdot \bar{\mathcal{A}}) \subset \mathcal{A}, denoted as Q(A),Q(\mathcal{A}), has a Z2\mathbb{Z}_2-basis given by the elements {Sq2i}iβ‰₯0.\{ Sq^{2^i}\}_{i\geq 0}. Note that ExtA1(Z2,Z2)\text{Ext}^1_{\mathcal{A}}(\mathbb{Z}_2, \mathbb{Z}_2) is the dual of Q(A)Q(\mathcal{A}) which means that Ext1\text{Ext}^1 must have a dual basis, and this is dual to the indecomposable elements in AΛ‰/(AΛ‰β‹…AΛ‰).\bar{\mathcal{A}} / (\bar{\mathcal{A}} \cdot \bar{\mathcal{A}}).

The dual basis of Ext1\text{Ext}^1 is the set of generators we call hih_i:

With this, we've constructed the s=1{s=1} column of the Adams E2E_2 page. It's just an infinite column of generators hi,h_i, each corresponding to the indecomposable elements Sq2i.Sq^{2^i}.

Next, we'll be building the s=2{s=2} column of ExtA2(Z2,Z2)\text{Ext}_{\mathcal{A}}^2(\mathbb{Z}_2, \mathbb{Z}_2) using the hih_i generators from the s=1{s=1} column. These are of the form hihjh_{i}h_{j} such that:

Warning: For h02=h0h0,h_0^2 = h_0h_0, one may be tempted to look at its corresponding (dual) Sq1∘Sq1Sq^1 \circ Sq^1 which is just zero due to the Adem relations in A,\mathcal{A}, and then incorrectly conclude that Sq1∘Sq1=0Sq^1 \circ Sq^1 = 0 implies that h02=0h_0^2 = 0 in Ext2.\text{Ext}^2. This turns out to not be the case. Since we're looking at these hihjh_ih_j elements in

ExtA2(Z2,Z2)β‰…ker⁑(βˆ‚3βˆ—)/im(βˆ‚2βˆ—)\text{Ext}_{\mathcal{A}}^2(\mathbb{Z}_2, \mathbb{Z}_2) \cong \ker(\partial_3^*) / \text{im}(\partial_2^*)

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 πŸ‘», but turns out to make life in Ext2\text{Ext}^2 a lot easier. Since A\mathcal{A} is a Hopf algbra, then we can use the following theorem:

Hβˆ—(Bar⁑(A))β‰…Hβˆ—(Cobar⁑(Aβˆ—)),H^*(\operatorname{Bar}(\mathcal{A})) \cong H^*(\operatorname{Cobar}(\mathcal{A}_*)),

where the cohomology of the bar complex of the Steenrod algebra A\mathcal{A} is isomorphic to the cohomology of the cobar complex of the dual for A\mathcal{A} which we denote as Aβˆ—\mathcal{A}_* [7, 8]. This is the bar-cobar isomorphism. So far we've been using the dual of the reduced bar complex:

((AΛ‰βŠ—s)βˆ—,βˆ‚sβˆ—)\left( (\bar{\mathcal{A}}^{\otimes s})^*, \partial_s^* \right)

where we have the differentials βˆ‚sβˆ—:(AΛ‰βŠ—sβˆ’1)βˆ—β†’(AΛ‰βŠ—s)βˆ—.\partial_s^*: (\bar{\mathcal{A}}^{\otimes s-1})^* \to (\bar{\mathcal{A}}^{\otimes s})^*. When we switch to this new cobar complex, we get the following:

((AΛ‰βˆ—)βŠ—s,ds)\left( (\bar{\mathcal{A}}^*)^{\otimes s} , d^s \right)

where we have these new differentials ds:(AΛ‰βˆ—)βŠ—sβ†’(AΛ‰βˆ—)βŠ—s+1.d^s: (\bar{\mathcal{A}}^*)^{\otimes s} \to (\bar{\mathcal{A}}^*)^{\otimes s+1}. Importantly, note that these are not isomorphic when s>1s > 1:

(AΛ‰βˆ—)βŠ—s≆(AΛ‰βŠ—s)βˆ—(\bar{\mathcal{A}}^*)^{\otimes s} \ncong (\bar{\mathcal{A}}^{\otimes s})^*

So we build this new cobar complex from the dual algebra Aβˆ—\mathcal{A}_* which is a graded-commutative polynomial algebra which we denote:

Aβˆ—=Z2[ΞΎ1,ΞΎ2,ΞΎ3,…]\mathcal{A}_* = \mathbb{Z}_2[\xi_1, \xi_2, \xi_3, \ldots]

where these generators ΞΎk\xi_k have internal degree deg⁑(ΞΎk)=2kβˆ’1.\deg(\xi_k) = 2^k-1. Similar to the A\mathcal{A} case, we look at the augmentation ideal AΛ‰βˆ—\bar{\mathcal{A}}^* of Aβˆ—\mathcal{A}_* to get Cs=(AΛ‰βˆ—)βŠ—sC^s = (\bar{\mathcal{A}}^*)^{\otimes s} where our differentials are maps ds:Csβ†’Cs+1.d^s: C^{s} \to C^{s+1}. Furthermore, note that ss in this case is the same ss-degree in Exts,βˆ—,\text{Ext}^{s,*}, and we also denote the degree tt with a subscript Cts.C^s_t. Let's use this new cobar framework to view our previous results for Ext1\text{Ext}^1 again:

In general, hi∈Ext1,2ih_i \in \text{Ext}^{1, 2^i} is represented by the cocycle ΞΎ12i∈C2i1\xi_1^{2^i} \in C_{2^i}^1 with such that deg⁑(ΞΎ12i)=2i.\deg(\xi_1^{2^i}) = 2^i. Note that this degree is different from the degree of ΞΎk\xi_k as a generator which is 2kβˆ’1.2^k-1.

We take the cohomology of this cobar complex ((AΛ‰βˆ—)βŠ—s,ds)=(Cs,ds)\left( (\bar{\mathcal{A}}^*)^{\otimes s} , d^s \right) = (C^s, d^s) to get Ext:\text{Ext}:

ExtAs,t(Z2,Z2)β‰…ker⁑(ds:Ctsβ†’Cts+1)im(dsβˆ’1:Ctsβˆ’1β†’Cts)\text{Ext}_{\mathcal{A}}^{s, t}(\mathbb{Z}_2, \mathbb{Z}_2) \cong \frac{\ker(d^{s}: C^{s}_t \to C^{s+1}_t )}{\text{im}(d^{s-1}: C^{s-1}_t \to C^{s}_t )}

h02h_0^2: We're finally ready to compute h02.h_0^2. Recall that h02=h0h0h_0^2 = h_0h_0 has bidegree (s,t)=(2,2),{(s, t) = (2, 2)}, so we're computing ExtA2,2β‰…ker⁑(d2:C22β†’C23)/im(d1:C21β†’C22).\text{Ext}_{\mathcal{A}}^{2, 2} \cong \ker(d^2: C_2^{2} \to C_2^{3} ) / \text{im}(d^1: C_2^{1} \to C_2^{2} ). Furthermore, h0h_0 is represented by the 11-cocycle ΞΎ1∈C11,\xi_1 \in C_1^1, so h02=h0h0h_0^2 = h_0h_0 is represented by the 22-cochain ΞΎ1βŠ—ΞΎ1∈C22.\xi_1 \otimes \xi_1 \in C_2^2. Side note: we call general elements x∈Csx \in C^s cochains, and if x∈ker⁑(ds)x \in \ker(d^s) then it's a cocycle, and if x∈im(dsβˆ’1)x \in \text{im}(d^{s-1}) then it's a coboundary. Additionally, as we do a deep dive into ker⁑(ds)\ker(d^s) and im(dsβˆ’1),\text{im}(d^{s-1}), note that the differential dsd^s 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 ΞΎ1βŠ—ΞΎ1∈ker⁑(d2),\xi_1 \otimes \xi_1 \in \ker(d^2), i.e. to check if ΞΎ1βŠ—ΞΎ1\xi_1 \otimes \xi_1 is a cocycle:

d2(ΞΎ1βŠ—ΞΎ1)= d1(ΞΎ1)βŠ—ΞΎ1+ (βˆ’1)deg⁑(ΞΎ1)ΞΎ1βŠ—d1(ΞΎ1)\begin{aligned} d^2(\xi_1 \otimes \xi_1) =\,&d^1(\xi_1) \otimes \xi_1 \\ + \,&(-1)^{\deg(\xi_1)}\xi_1 \otimes d^1(\xi_1) \end{aligned}

Note that (βˆ’1)deg⁑(ΞΎk)=1(-1)^{\deg(\xi_k)} = 1 for any kk since we're working in Z2\mathbb{Z}_2 (mod 2). Note that our d1d^1 operator is the same as the reduced coproduct Ξ”Λ‰:AΛ‰βˆ—β†’AΛ‰βˆ—βŠ—AΛ‰βˆ—.\bar{\Delta}: \bar{\mathcal{A}}^* \to \bar{\mathcal{A}}^* \otimes \bar{\mathcal{A}}^*. This map is derived from the full coproduct Ξ”\Delta on the dual Steenrod algebra Aβˆ—\mathcal{A}_* by the relation:

Ξ”Λ‰(x)=Ξ”(x)βˆ’(xβŠ—1)βˆ’(1βŠ—x),\bar{\Delta}(x) = \Delta(x) - (x\otimes 1) - (1 \otimes x),

where Ξ”Λ‰\bar{\Delta} captures the non-trivial part of the coproduct by subtracting the trivial parts xβŠ—1x\otimes 1 and 1βŠ—x,1 \otimes x, etc. With this, we get:

d2(ΞΎ1βŠ—ΞΎ1)=d1(ΞΎ1)βŠ—ΞΎ1+ΞΎ1βŠ—d1(ΞΎ1)=Ξ”Λ‰(ΞΎ1)βŠ—ΞΎ1+ΞΎ1βŠ—Ξ”Λ‰(ΞΎ1)=0βŠ—ΞΎ1+ΞΎ1βŠ—0=0\begin{aligned} d^2(\xi_1 \otimes \xi_1) &= d^1(\xi_1) \otimes \xi_1 + \xi_1 \otimes d^1(\xi_1) \\ &= \bar{\Delta}(\xi_1) \otimes \xi_1 + \xi_1 \otimes \bar{\Delta}(\xi_1) \\ &= 0\otimes \xi_1 + \xi_1 \otimes 0 \\ &= 0 \end{aligned}

where Ξ”Λ‰(ΞΎ1)=0\bar{\Delta}(\xi_1) = 0 since Ξ”(ΞΎ1)=ΞΎ1βŠ—1+1βŠ—ΞΎ1.\Delta(\xi_1) = \xi_1 \otimes 1 + 1\otimes \xi_1. Therefore, ΞΎ1βŠ—ΞΎ1∈ker⁑(d2),\xi_1 \otimes \xi_1 \in \ker(d^2), so it's a cocycle. Next, let's see if it's a coboundary, i.e. if ΞΎ1βŠ—ΞΎ1∈im(d1:C21β†’C22).\xi_1 \otimes \xi_1 \in \text{im}( d^1: C_2^1 \to C_2^2 ). Looking at the source space C21,C_2^1, note that this has degree 2, and we have the following available generators from AΛ‰βˆ—\bar{\mathcal{A}}^*: ΞΎ1,ΞΎ2,…\xi_1, \xi_2, \ldots such that deg⁑(ΞΎ1)=1,\deg(\xi_1) = 1, deg⁑(ΞΎ2)=3,\deg(\xi_2) = 3, etc. There's no single degree 2 generator, so the only way to get a generator for C21C_2^1 is to consider ΞΎ12\xi_1^2 with degree deg⁑(ΞΎ12)=2.\deg(\xi_1^2) = 2. We don't have anything else (since even considering ΞΎ1ΞΎ2\xi_1\xi_2 has degree 4), so our source space C21C_2^1 is spanned by the single element ΞΎ12,\xi_1^2, i.e.

C21=span{ΞΎ12}C_2^1 = \text{span}\{\xi_1^2\}

Let's see how d1=Ξ”Λ‰d^1 = \bar{\Delta} maps ΞΎ12\xi_1^2 to C22.C_2^2. First, let's look at Ξ”(ΞΎ12):\Delta(\xi_1^2):

Ξ”(ΞΎ12)=Ξ”(ΞΎ1)2=(ΞΎ1βŠ—1+1βŠ—ΞΎ1)2=ΞΎ1βŠ—12+12βŠ—ΞΎ12=ΞΎ12βŠ—1+1βŠ—ΞΎ12\begin{aligned} \Delta(\xi_1^2) &= \Delta(\xi_1)^2 = (\xi_1 \otimes 1 + 1 \otimes \xi_1)^2 \\ &= \xi_1\otimes1^2 + 1^2 \otimes \xi_1^2\\ &= \xi_1^2 \otimes 1 + 1 \otimes \xi_1^2 \end{aligned}

where we get (ΞΎ1βŠ—1+1βŠ—ΞΎ1)2=ΞΎ12βŠ—1+1βŠ—ΞΎ12(\xi_1 \otimes 1 + 1 \otimes \xi_1)^2 = \xi_1^2 \otimes 1 + 1 \otimes \xi_1^2 since we're working over Z2\mathbb{Z}_2 (characteristic 2). Hence, when we look at d1,d^1, we get:

d1(ΞΎ12)=Ξ”Λ‰(ΞΎ12)=Ξ”(ΞΎ12)βˆ’(ΞΎ12βŠ—1)βˆ’(1βŠ—ΞΎ12)=ΞΎ12βŠ—1+1βŠ—ΞΎ12βˆ’(ΞΎ12βŠ—1)βˆ’(1βŠ—ΞΎ12)=0\begin{aligned} d^1(\xi_1^2) &= \bar{\Delta}(\xi_1^2) = \Delta(\xi_1^2) - (\xi_1^2 \otimes 1) - (1 \otimes \xi_1^2) \\ &= \xi_1^2 \otimes 1 + 1 \otimes \xi_1^2 - (\xi_1^2 \otimes 1) - (1 \otimes \xi_1^2) \\ &= 0 \end{aligned}

Since C21=span{ΞΎ12}C_2^1 = \text{span}\{\xi_1^2\} and d1(ΞΎ12)=0,d^1(\xi_1^2) = 0, then this means that im(d1)={0}.\text{im}(d^1) = \{0\}. We conclude that ΞΎ1βŠ—ΞΎ1\xi_1 \otimes \xi_1 is not a coboundary, and get:

ExtA2,2(Z2,Z2)=span{ΞΎ1βŠ—ΞΎ1}{0}β‰…Z2\text{Ext}_{\mathcal{A}}^{2,2}(\mathbb{Z}_2, \mathbb{Z}_2) = \frac{\text{span}\{\xi_1 \otimes \xi_1\}}{\{ 0 \}} \cong \mathbb{Z}_2

Thus, h02h_0^2 is a non-zero generator of Ext2,2.\text{Ext}^{2,2}.

h0h1h_0h_1: As an additional example, let's look at h0h1.h_0h_1. Similar to before, recall that h0h1h_0h_1 has bidegree (s,t)=(2,3),(s,t) = (2,3), so we're looking at

ExtA2,3(Z2,Z2)β‰…ker⁑(d2:C32β†’C33)im(d1:C31β†’C32)\text{Ext}_{\mathcal{A}}^{2,3}(\mathbb{Z}_2, \mathbb{Z}_2) \cong \frac{\ker( d^2: C_3^2 \to C_3^{3} )}{\text{im}( d^1: C_3^1 \to C_3^2 )}

Let's look at ker⁑(d2).\ker(d^2). We have the source space C32C_3^2 with elements from AΛ‰βˆ—βŠ—AΛ‰βˆ—.\bar{\mathcal{A}}^* \otimes \bar{\mathcal{A}}^*. Note that since 1βˆ‰AΛ‰βˆ—1 \notin \bar{\mathcal{A}}^* then we can't have elements like 1βŠ—ΞΎ21 \otimes \xi_2 or ΞΎ2βŠ—1.\xi_2 \otimes 1. Since any variation of ΞΎ2βŠ—ΞΎ1\xi_2\otimes \xi_1 will have degree β‰₯4\geq 4 then we can only span C32C_3^2 using ΞΎ1.\xi_1. We get:

C32=span{ΞΎ12βŠ—ΞΎ1,ΞΎ1βŠ—ΞΎ12},C_3^2 = \text{span}\{\xi_1^2 \otimes \xi_1, \xi_1 \otimes \xi_1^2\},

where the degree of both ΞΎ12βŠ—ΞΎ1\xi_1^2 \otimes \xi_1 and ΞΎ1βŠ—ΞΎ12\xi_1 \otimes \xi_1^2 is 3. Let's see how d2d^2 maps these. Let's start with ΞΎ12βŠ—ΞΎ1:\xi_1^2 \otimes \xi_1:

d2(ΞΎ12βŠ—ΞΎ1)=d1(ΞΎ12)βŠ—ΞΎ1+ΞΎ12βŠ—d1(ΞΎ1)=0βŠ—ΞΎ1+ΞΎ12βŠ—0=0\begin{aligned} d^2(\xi_1^2 \otimes \xi_1) &= d^1(\xi_1^2)\otimes \xi_1 + \xi_1^2 \otimes d^1(\xi_1) \\ &= 0\otimes \xi_1 + \xi_1^2 \otimes 0\\ &= 0 \end{aligned}

where d1(ΞΎ12)=0d^1(\xi_1^2) = 0 from our previous h02h_0^2 calculation, and d1(ΞΎ1)=0d^1(\xi_1) = 0 since ΞΎ1\xi_1 is primitive. Note that d2(ΞΎ1βŠ—ΞΎ12)d^2(\xi_1 \otimes \xi_1^2) is a similar story where we get d2(ΞΎ1βŠ—ΞΎ12)=0.d^2(\xi_1 \otimes \xi_1^2) = 0. Since d2d^2 maps everything from the source space to zero, then we get:

ker⁑(d2:C32β†’C33)β‰…C32β‰…span{ΞΎ12βŠ—ΞΎ1,ΞΎ1βŠ—ΞΎ12}\begin{aligned} \ker( d^2: C_3^2 \to C_3^{3}) &\cong C_3^2 \\ &\cong \text{span}\{\xi_1^2 \otimes \xi_1, \xi_1 \otimes \xi_1^2\} \end{aligned}

Next, let's look at the coboundaries im(d1).\text{im}( d^1 ). In this case, since we're getting elements of C31C_3^1 from AΛ‰βˆ—\bar{\mathcal{A}}^* then we get the following:

C31=span{ΞΎ13,ΞΎ2}C_3^1 = \text{span}\{\xi_1^3, \xi_2\}

Let's apply d1=Ξ”Λ‰d^1 = \bar{\Delta} to this basis. First, let's look at Ξ”\Delta for ΞΎ13\xi_1^3 and then just remove the trivial terms ΞΎ13βŠ—1\xi_1^3 \otimes 1 and 1βŠ—ΞΎ13.1 \otimes \xi_1^3. Note that we'll be using the (mod 2) expansion of (a+b)3=a3+a2b+ab2+b3:(a+b)^3 = a^3 + a^2b + ab^2 + b^3:

Ξ”(ΞΎ13)=Ξ”(ΞΎ1)3=(ΞΎ1βŠ—1+1βŠ—ΞΎ1)3=ΞΎ13βŠ—1+(ΞΎ12βŠ—ΞΎ1)+(ΞΎ1βŠ—ΞΎ12)+1βŠ—ΞΎ13\begin{aligned} \Delta(\xi_1^3) &= \Delta(\xi_1)^3 = (\xi_1 \otimes 1 + 1 \otimes \xi_1)^3 \\ &= \xi_1^3 \otimes 1 + ( \xi_1^2 \otimes \xi_1 ) \\ &+ (\xi_1 \otimes \xi_1^2 ) + 1\otimes \xi_1^3 \end{aligned}

where we get ΞΎ12βŠ—ΞΎ1\xi_1^2 \otimes \xi_1 from (ΞΎ12βŠ—1)βŠ—(1βŠ—ΞΎ1),(\xi_1^2 \otimes 1) \otimes(1 \otimes \xi_1), and ΞΎ1βŠ—ΞΎ12\xi_1 \otimes \xi_1^2 from (ΞΎ1βŠ—1)βŠ—(1βŠ—ΞΎ12).(\xi_1\otimes 1)\otimes (1 \otimes \xi_1^2). We can remove the trivial terms (ΞΎ13βŠ—1\xi_1^3 \otimes 1 and 1βŠ—ΞΎ131\otimes \xi_1^3) from this expression to get:

d1(ΞΎ13)=Ξ”Λ‰(ΞΎ13)=ΞΎ12βŠ—ΞΎ1+ΞΎ1βŠ—ΞΎ12d^1( \xi_1^3 ) = \bar{\Delta}(\xi_1^3) = \xi_1^2 \otimes \xi_1 + \xi_1 \otimes \xi_1^2

For the next element in the basis, ΞΎ2,\xi_2, we need a bit of help from Milnor.

Theorem: The coproduct Ξ”\Delta of the kk-th generator ΞΎk\xi_k is:

Ξ”(ΞΎk)=βˆ‘i=0kΞΎkβˆ’i2iβŠ—ΞΎi\Delta(\xi_k) = \sum_{i=0}^k \xi_{k-i}^{2^i} \otimes \xi_i

Note that this may feel reminiscent of the Cartan formula for the Steenrod squares in A.\mathcal{A}. Using this, we get:

Ξ”(ΞΎ2)=βˆ‘i=02ΞΎ2βˆ’i2iβŠ—ΞΎi=ΞΎ21βŠ—ΞΎ0+ΞΎ12βŠ—ΞΎ1+ΞΎ04βŠ—ΞΎ2=ΞΎ2βŠ—1+ΞΎ12βŠ—ΞΎ1+1βŠ—ΞΎ2\begin{aligned} \Delta(\xi_2) &= \sum_{i=0}^2 \xi_{2-i}^{2^i} \otimes \xi_i \\ &= \xi_{2}^1 \otimes \xi_0 + \xi_1^2\otimes \xi_1 + \xi_0^4 \otimes \xi_2 \\ &= \xi_2 \otimes 1 + \xi_1^2 \otimes \xi_1 + 1 \otimes \xi_2 \end{aligned}

Next, similar to before, we simply remove the trivial terms ( ΞΎ2βŠ—1\xi_2 \otimes 1 and 1βŠ—ΞΎ21 \otimes \xi_2 ) to get:

d1(ΞΎ2)=Ξ”Λ‰(ΞΎ2)=ΞΎ12βŠ—ΞΎ1d^1( \xi_2 ) = \bar{\Delta}(\xi_2) = \xi_1^2 \otimes \xi_1

Putting things together, we see that d1d^1 maps the basis of C31C_3^1 to the following space:

im(d1)=span{ΞΎ12βŠ—ΞΎ1,ΞΎ12βŠ—ΞΎ1+ΞΎ1βŠ—ΞΎ12}\text{im}(d^1) = \text{span}\{ \xi_1^2 \otimes \xi_1, \xi_1^2 \otimes \xi_1 + \xi_1 \otimes \xi_1^2 \}

Note that this space is equivalent to span{ΞΎ12βŠ—ΞΎ1,ΞΎ1βŠ—ΞΎ12}\text{span}\{ \xi_1^2 \otimes \xi_1, \xi_1 \otimes \xi_1^2 \} since the generating sets span the same space. We get the following for Ext2,3:\text{Ext}^{2,3}:

ExtA2,3β‰…span{ΞΎ12βŠ—ΞΎ1,ΞΎ1βŠ—ΞΎ12}span{ΞΎ12βŠ—ΞΎ1,ΞΎ1βŠ—ΞΎ12}β‰…{0}\begin{aligned} \text{Ext}_{\mathcal{A}}^{2,3} &\cong \frac{ \text{span}\{ \xi_1^2 \otimes \xi_1, \xi_1 \otimes \xi_1^2 \} }{ \text{span}\{ \xi_1^2 \otimes \xi_1, \xi_1 \otimes \xi_1^2 \} } \\ &\cong \{ 0 \} \end{aligned}

This implies that h0h1h_0h_1 is a coboundary, thus h0h1=0.h_0h_1 = 0. Overall, we could continue analyzing ExtA2,βˆ—\text{Ext}_{\mathcal{A}}^{2,*} to fully characterize the s=2{s=2} column (but we won't do that here). So far we've only scratched the surface of the Adams E2E_2 page, but this feels like a good spot to wrap things up. πŸ‘»


[1] N. E. Steenrod, Products of cocycles and extensions of mappings, Ann. of Math. (2) 48 (1947), 290–320.

[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), 1–2.

[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), 150–171.

[7] J.F. Adams, On the cobar construction, Proc. Nat. Acad. Sci. U.S.A. 42 (1956), 409–412.

[8] J.F. Adams, On the structure and applications of the Steenrod algebra, Comment. Math. Helv. 32 (1958), 180–214.