This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

\jyear

2021

\equalcont

These authors contributed equally to this work. [2]\fnmRuJia \surYan \equalcontThese authors contributed equally to this work.

1]\orgdivHUA Loo-Keng Key Laboratory of Mathematics, Academy of Mathematics and Systems Science, \orgnameChinese Academy of Sciences, \orgaddress\streetNo.55, Zhongguancun East Road, \cityBeijing, \postcode100190, \countryChina [2]\orgdivAcademy of Mathematics and Systems Science, \orgnameChinese Academy of Sciences, \orgaddress\streetNo.55, Zhongguancun East Road, \cityBeijing, \postcode100190, \countryChina

The Stable Picard Group of A(n)A(n)

\fnmJianZhong \surPan [email protected]    [email protected] [ *
Abstract

In this paper, we showed that the Stable Picard group of A(n)A(n) for n2n\geq 2 is \mathbb{Z}\oplus\mathbb{Z} by considering the endotrivial modules over A(n)A(n). The proof relies on reductions from a Hopf algebra to its proper Hopf subalgebras.

keywords:
Picard group, Hopf algebra, Steenrod algebra, Endotrivial module

1 Introduction

Let HH be a cocommutative connected finite graded Hopf algebra over a base field kk with characteristic pp. The Picard group of Stab(H)Stab(H), denoted by Pic(H)Pic(H) is the group of graded stably \otimes-invertible HH-modules,

Pic(A):={MStab(H):N such that MN=k¯},Pic(A):=\{M\in Stab(H):\exists N\text{ such that }M\otimes N=\underline{k}\},

where k¯\underline{k} is the unit of the symmetric monoidal category (Stab(A),,k¯)(Stab(A),\otimes,\underline{k}). In this paper, we are interested in the determination of the Picard group of A(n)A(n), the Hopf subalgebra of 𝒜2\mathcal{A}_{2} generated by Sq1,Sq2,,Sq2nSq^{1},Sq^{2},\cdots,Sq^{2^{n}} (see section 2.3). The situation when n=1n=1 was calculated by Adams and Priddy in Adams and Priddy (1976) to show the uniqueness of BSOBSO in 1976. It says that

Pic(A(1))=/2.Pic(A(1))=\mathbb{Z}\oplus\mathbb{Z}\oplus\mathbb{Z}/2.

The \mathbb{C}-motivic version of the above result was given by Gheorghe, Isaksen and Ricka in Gheorghe et al. (2018), which is

Pic(A(1))=.Pic(A_{\mathbb{C}}(1))=\mathbb{Z}\oplus\mathbb{Z}\oplus\mathbb{Z}\oplus\mathbb{Z}.

In 2017, Bhattacharya and Ricka determined in Bhattacharya and Ricka (2017) that

Pic(A(2))=.Pic(A(2))=\mathbb{Z}\oplus\mathbb{Z}.

The authors of Bhattacharya and Ricka (2017) conjectured that

Pic(A(n))=Pic(A(n))=\mathbb{Z}\oplus\mathbb{Z}

for all n>2n>2, which is exactly the main result of this paper (section 4).

The ideal is to consider 𝕋(H)\mathbb{T}(H), the group of endotrivial modules over HH (See section 2.2). An HH-module is called endotrivial if Endk(M)=k(free)End_{k}(M)=k\oplus(free). By definition if MM is endotrivial, then MM is invertible since Endk(M)=MMEnd_{k}(M)=M^{*}\otimes M. On the other hand, since HH is cocommutative, the stable module category Stab(H)Stab(H) is a stable homotopy category (section 9.6 of Hovey et al. (1997)), and hence a closed symmetric monoidal category. Therefore Theorem A.2.8 of Margolis (1983) guarantees that if MM is invertible, then it is endotrivial. That is, the group of endotrivial modules of HH is isomorphic to the stable Picard group of HH.

The word endotrivialendotrivial comes from the representation theory of finite groups, since the group algebra kGkG of a finite pp-group GG is a well-defined Hopf algebra with diagonal coproduct. The group of endotrivial modules over kGkG has been completely figured out by joint efforts of many mathematicians including Puig, Alperin, Dade, Carlson and Thevenaz, check Carlson and Thévenaz (2000) for references.

The main object of this paper is to give a proof of the following main result:

Theorem 1.1 (Corollary 4.8).

Given n2n\geq 2, the morphism of groups

f:Pic(A(n))f:\mathbb{Z}\oplus\mathbb{Z}\to Pic(A(n))

which sends (m,l)(m,l) to [σ(m)ΩA(n)l(/2)][\sigma(m)\Omega_{A(n)}^{l}(\mathbb{Z}/2)] is an isomorphism of groups.

This article is organized as follows. Section 2 introduces some preliminaries and conventions. In section 3, we showed that the problem of classifying endotrivial HH-modules can somehow be reduced to the one about each EE-modules for certain Hopf subalgebras EE (See Theorem 3.14). With help of this reduction theorem, the Picard group of A(n)A(n) was given in section 4.

Throughout this paper, kk will denote a field with positive character pp. Every algebraic structure is implicitly over the base field kk, and tensor products are taken over kk. The Hopf algebras under consideration in this paper are cocommutative, unless explicitly specified otherwise.

2 Preliminaries

A Hopf algebra is a bialgebra over field kk (of characteristic p>0p>0) with a compatible antipode map SS (Montgomery (1993)). We will adapt the following notation: If HH is a Hopf algebra, denote :HHH\triangledown:H\otimes H\to H as its product; :HHH\triangle:H\to H\otimes H as its coproduct; ϵ:Hk\epsilon:H\to k and η:kH\eta:k\to H are counit and unit respectively.

Suppose M,NM,N are HH-modules, then MNM\otimes N is an HH-module defined by h(mn)=iaimbinh(m\otimes n)=\sum_{i}a_{i}m\otimes b_{i}n, where hHh\in H, mMm\in M and nNn\in N with (h)=iaibi\triangle(h)=\sum_{i}a_{i}\otimes b_{i}. For an HH-module MM, the dual MM^{*} is defined by M=Homk(H,k)M^{*}=Hom_{k}(H,k), with the natural HH-module structure

(hf)(v)=f(S(h)v)h,vH;fM(h\cdot f)(v)=f(S(h)v)\quad\forall h,v\in H\mathchar 24635\relax\;\ f\in M^{*}

2.1 Finite connected graded Hopf algebra

If the underlying algebra of HH is decomposed into a direct sum of vector spaces:

H=n=0Hn=H0H1H2H=\bigoplus_{n=0}^{\infty}H_{n}=H_{0}\oplus H_{1}\oplus H_{2}\oplus\cdots

such that HiHjHi+jH_{i}H_{j}\subset H_{i+j} for all nonnegative ii and jj, we say that HH is a graded algebra.

Moreover, a graded HH-module MM is an HH-module which has a decomposition as direct sum of vector spaces:

M=MnM=\bigoplus_{-\infty}^{\infty}M_{n}

such that HiMjMi+jH_{i}\cdot M_{j}\subset M_{i+j}. When MM, NN are graded HH-module, both MNM\otimes N and MM^{*} defined above are still graded modules. In the rest of the paper, all algebras and modules are assumed to be graded.

An algebra (or module) is finite if it is a finite dimensional vector space over the base field.

Let grHgrH be the category with objects finitely generated graded HH-modules, and morphisms graded HH-homomorphisms. The ii th shift functor, σ(i):grHgrH\sigma(i):grH\to grH, is defined by σ(i)X=σ(i)(Xn)=Y=Yn\sigma(i)X=\sigma(i)(\oplus X_{n})=Y=\oplus Y_{n}, where Yn=XniY_{n}=X_{n-i}, and the obvious way on morphisms.

Suppose HH is a finite graded Hopf algebra over a field kk, then there is a minimal n+n\in\mathbb{Z}^{+} such that Hj=0H_{j}=0 if j>nj>n, and we say that the top degree of HH is nn and denote as |H|=n|H|=n.

The trivial HH-module kk can be viewed as a graded HH-module k¯\underline{k} where k¯0=k,k¯j=0\underline{k}_{0}=k,\underline{k}_{j}=0 for j0j\neq 0. We still use kk to denote k¯\underline{k} if there is no confusion.

Recall that a graded algebra Λ\Lambda is connected if Λ0=k\Lambda_{0}=k, and we have

Definition 2.1.

A finite graded connected algebra Λ\Lambda is a Poincare algebra if there is a map of graded kk-modules e:Λσ(n)ke:\Lambda\to\sigma(n)k for some nn such that the pairing ΛqΛnqΛn𝑒k\Lambda_{q}\otimes\Lambda_{n-q}\xrightarrow{\triangledown}\Lambda_{n}\xrightarrow{e}k is non-singular.

In particular, this implies that if Λ\Lambda is a Poincare algebra, then Λi=0\Lambda_{i}=0 for i>ni>n and Λnk\Lambda_{n}\cong k.

Theorem 2.2 (Theorem 12.2.5 & 12.2.9 of Margolis (1983)).

If HH is a finite graded connected Hopf algebra, then HH is a Poincare algebra.

So if HH is a finite graded connected Hopf algebra and |H|=n|H|=n, HH is a Poincare algebra and thus Hσ(n)HH\cong\sigma(n)H^{*}. Namely, H0HnkH_{0}\cong H_{n}\cong k.

Theorem 2.3 (Proposition 12.2.8 of Margolis (1983)).

Let Λ\Lambda be a Poincare algebra and let MM be a Λ\Lambda-module. Then the followings are equivalent:

(1) MM is free

(2) MM is projective

(3) MM is flat

(4) MM is injective.

Another fact worth to be known is that a Hopf algebra of our interest is always free over its Hopf subalgebras:

Theorem 2.4 (Theorem 1.3 of Aguiar and Lauve (2013)).

Let HH be a finite graded connected Hopf algebra. If ZHZ\subset H is a Hopf subalgebra, then HH is a free left (and right) ZZ-module. Moreover,

HZ(H/Z+H)H\cong Z\otimes(H/Z^{+}H)

as left ZZ-modules and as graded vector spaces.

Remark 2.5.

The proof of Aguiar and Lauve (2013) applies for arbitrary characteristic.

In the rest of this article, all Hopf algebras are assumed to be finite graded connected Hopf algebras.

2.2 Endotrivial modules

An HH-module MM is called endotrivial if there is an isomorphism of HH-modules Endk(M)k¯FEnd_{k}(M)\cong\underline{k}\oplus F, where FF is a free HH-module. As an HH-module, Endk(M)MMEnd_{k}(M)\cong M\otimes M^{*}, thus MM is endotrivial iff MMkFM\otimes M^{*}\cong k\oplus F for some free module FF. Note that MNM\otimes N and MM^{*} are endotrivial if MM and NN are endotrivial.

Two endotrivial HH-modules MM and NN are said to be equivalent if

eitherMN(free)orNM(free).either\quad M\cong N\oplus(free)\quad or\quad N\cong M\oplus(free).

Denote the equivalent class of an endotrivial module MM as [M][M], and define

𝕋(H)=[M]| M is an endotrivial H-module.\mathbb{T}(H)=[M]|\text{ M is an endotrivial H-module}.

𝕋(H)\mathbb{T}(H) is an abelian group with unit kk, and [M]+[N]=[MN][M]+[N]=[M\otimes N].

Lemma 2.6.

If an endotrivial HH-module MM decomposes as M=M1M2M=M_{1}\oplus M_{2}, then one of the summands is free and the other one is endotrivial.

Proof: By Theorem 2.3, HH is indecomposable as an HH-module. By definition,

kFEndk(M1)Endk(M2)Homk(M1,M2)Homk(M2,M1).k\oplus F\cong End_{k}(M_{1})\oplus End_{k}(M_{2})\oplus Hom_{k}(M_{1},M_{2})\oplus Hom_{k}(M_{2},M_{1}).

Krull-Schmidt Theorem shows that all the summands on the right hand side are free except one. Denote it as LL, one has dimL=1dimL=1 mod dimHdimH. On the other hand, dim(Homk(M1,M2))=dim(Homk(M2,M1))=dimM1dimM2dim(Hom_{k}(M_{1},M_{2}))=dim(Hom_{k}(M_{2},M_{1}))=dimM_{1}dimM_{2}, thus LHomk(M2,M1)L\neq Hom_{k}(M_{2},M_{1}) and LHomk(M1,M2)L\neq Hom_{k}(M_{1},M_{2}). Without loss of generality, we assume L=Endk(M1)=k(free)L=End_{k}(M_{1})=k\oplus(free) and Endk(M2)=(free)End_{k}(M_{2})=(free). Therefore Endk(M2)M2=M2M2M2End_{k}(M_{2})\otimes M_{2}=M_{2}\otimes M_{2}^{*}\otimes M_{2}, is free and has a direct summand M2M_{2}. By Theorem 2.3, M2M_{2} is free.

Therefore, in the equivalence class [M][M], there is, up to isomorphism, a unique indecomposable module M0M_{0} and every module in the class is isomorphic to M0(free)M_{0}\oplus(free).

Definition 2.7 (Definition 2.15 of Gheorghe et al. (2018)).

Consider the projective resolution of kk:

ker(ϵ)Hk.ker(\epsilon)\to H\to k.

The operator ΩH\Omega_{H} is given by

ΩHM=ker(ϵ)M,\Omega_{H}M=ker(\epsilon)\otimes M,

where ϵ:Hk\epsilon:H\to k is the augmentation of HH. For m0m\geq 0, define ΩHmM\Omega^{m}_{H}M inductively to be ΩH(ΩHm1M)\Omega_{H}(\Omega_{H}^{m-1}M). For m<0m<0, define ΩHmM\Omega_{H}^{m}M to be (ΩHm(M))(\Omega_{H}^{-m}(M^{*}))^{*}.

As showed in section 2.3 of Gheorghe et al. (2018), ΩHm(k)\Omega_{H}^{m}(k) is endotrivial and different choices of resolution will give the same equivalence class in 𝕋(H)\mathbb{T}(H), so the choice we made is harmless.

When H=kGH=kG, the group algebra of a finite pp-group GG, we know well about endotrivial modules on kGkG (Carlson and Thévenaz (2000)). With help of this example, we get:

Proposition 2.8 (Prop3.2 of Bhattacharya and Ricka (2017)).

Given an elementray Hopf algebra BB over kk generated by {x1,x2,,xt}\{x_{1},x_{2},\cdots,x_{t}\} where t2t\geq 2. Then f:𝕋(B)f:\mathbb{Z}\oplus\mathbb{Z}\to\mathbb{T}(B) with f(m,n)=[σ(m)ΩBn(k)]f(m,n)=[\sigma(m)\Omega_{B}^{n}(k)] is an isomorphism of groups.

Remark 2.9.

A Hopf algebra HH over kk is called elementary if it is bicommunicative and has xp=0x^{p}=0 for all xkerϵx\in ker\epsilon.

Remark 2.10.

The proof of Bhattacharya and Ricka (2017) still works for p>2p>2.

2.3 Mod 2 Steenrod algebra

We now turn to the algebra of central interest to us, the mod 22 Steenrod algebra 𝒜2\mathcal{A}_{2} and its subalgebra A(n)A(n). The background knowledge about 𝒜2\mathcal{A}_{2} in this section mainly comes from 15.1 of Margolis (1983).

The mod 22 Steenrod algebra 𝒜2\mathcal{A}_{2} is a graded vector space over /2\mathbb{Z}/2 with basis all formal symbols Sq(r1,r2,)Sq(r_{1},r_{2},\cdots), called Milnor basis, where ri0r_{i}\geq 0 and ri>0r_{i}>0 only finitely often. If rj=0r_{j}=0 for j>ij>i we will also write Sq(r1,,ri)Sq(r_{1},\cdots,r_{i}) and |Sq(r1,r2,)||Sq(r_{1},r_{2},\cdots)|=(2i1)ri\sum(2^{i}-1)r_{i}. The product of 𝒜2\mathcal{A}_{2} are defined as follows:

Sq(r1,r2,)Sq(s1,s2,)=Xβ(X)Sq(t1,t2,)Sq(r_{1},r_{2},\cdots)\cdot Sq(s_{1},s_{2},\cdots)=\sum_{X}\beta(X)Sq(t_{1},t_{2},\cdots)

the summation being over all matrices

X=[x01x02x10x11x20]X=\left[\begin{matrix}*&x_{01}&x_{02}&\cdots\\ x_{10}&x_{11}&\cdots&\\ x_{20}&\vdots&&\\ \vdots&&&\end{matrix}\right]

with nonnegative xijx_{ij} satisfying ixij=sj\sum_{i}x_{ij}=s_{j} and j2jxij=ri\sum_{j}2^{j}x_{ij}=r_{i}, and we will call such matrix allowable.

Then take tk=i+j=kxijt_{k}=\sum_{i+j=k}x_{ij} and β(X)=k(xk0,x(k1)1,x0k)/2\beta(X)=\prod_{k}(x_{k0},x_{(k-1)1},\cdots x_{0k})\in\mathbb{Z}/2, where (n1,,nr)=(n1+nr)!/n1!nr!(n_{1},\cdots,n_{r})=(n_{1}+\cdots n_{r})!/n_{1}!\cdots n_{r}! reduced mod 22. In particular, there is always an allowable matrix with xi0=rix_{i0}=r_{i} and x0j=sjx_{0j}=s_{j} and all the other entries are zero, we call such allowable matrix trivial.

The coproduct is defined as follow:

Sq(r1,r2,)=ri=si+tiSq(s1,s2,)Sq(t1,t2,).\triangle Sq(r_{1},r_{2},\cdots)=\sum_{r_{i}=s_{i}+t_{i}}Sq(s_{1},s_{2},\cdots)\otimes Sq(t_{1},t_{2},\cdots).

Here is a useful Lemma for the calculation of (n1,,nr)(n_{1},\cdots,n_{r}).

Lemma 2.11.

Suppose the unique dyadic expansion of the natural number nn is n=jaj2jn=\sum_{j}a_{j}2^{j}, then we say 2in2^{i}\in n if ai=1a_{i}=1; and 2in2^{i}\notin n if ai=0a_{i}=0.
Then (n1,,nr)=1(n_{1},\cdots,n_{r})=1 if and only if 2inj2^{i}\in n_{j} implies 2ink2^{i}\notin n_{k} for kjk\neq j.

We denote the Milnor basis element Sq(0,,0,r)Sq(0,\cdots,0,r) with rr at the tt-th position as Pt(r)P_{t}(r), and denote Pts=Pt(2s)P^{s}_{t}=P_{t}(2^{s}). The homogeneous primitives of 𝒜2\mathcal{A}_{2} are precisely the Pt0P^{0}_{t} for all t1t\geq 1.

Lemma 2.12.

Here are some direct consequences about the product in 𝒜2\mathcal{A}_{2}:

(1) If r<2vr<2^{v} and u<2tu<2^{t} then [Pt(r),Pv(u)]=0[P_{t}(r),P_{v}(u)]=0.

(2) If r<2tr<2^{t} then Pt(r)Pt(s)=(r,s)Pt(r+s)P_{t}(r)P_{t}(s)=(r,s)P_{t}(r+s), in particular Pt(r)2=0P_{t}(r)^{2}=0.

(3) PttPtt=Pt(2t1)P2t0P^{t}_{t}\cdot P^{t}_{t}=P_{t}(2^{t}-1)P^{0}_{2t}.

(4) If 1r<2t1\leq r<2^{t} then [Ptt,Pt(r)]=Pt(r1)P2t0[P^{t}_{t},P_{t}(r)]=P_{t}(r-1)P^{0}_{2t}

(5) If 1r2i2t1\leq r\leq 2^{i}\leq 2^{t} and r<2tr<2^{t}, then [Pi(r),Pti]=Pi(r1)Pt+i0[P_{i}(r),P^{i}_{t}]=P_{i}(r-1)P_{t+i}^{0}

(6) If 2ir2t\leq 2^{i}\leq r\leq 2^{t}, then [Pi0,Pt(r)]=Pt(r2i)Pt+i0[P_{i}^{0},P_{t}(r)]=P_{t}(r-2^{i})P_{t+i}^{0}

Proof: (1)-(4) comes from Lemma 1.3 of Adams and Margolis (1992).

By definition, if X=(xij)X=(x_{ij}) is an allowable matrix for the product Sq(r1,r2)Sq(s1,s2)Sq(r_{1},r_{2}\cdots)\cdot Sq(s_{1},s_{2}\cdots), then the entries satisfies

ixij=sj\displaystyle\sum_{i}x_{ij}=s_{j}
j2jxij=ri.\displaystyle\sum_{j}2^{j}x_{ij}=r_{i}.

As for (5) Pi(r)PtiP_{i}(r)P^{i}_{t} only has trivial allowable matrix, while beside the trivial matrix, PtiPi(r)P^{i}_{t}P_{i}(r) has another allowable matrix X=(xuv)X=(x_{uv}) where x0i=r1x_{0i}=r-1, xti=1x_{ti}=1 and all the other entries are zero. Note that β(X)=1\beta(X)=1, therefore [Pti,Pi(r)]=Sq(r1,r2)[P^{i}_{t},P_{i}(r)]=Sq(r_{1},r_{2}\cdots) with ri=r1r_{i}=r-1, rt+i=1r_{t+i}=1 and all the other coordinates are zero. On the other hand Pi(r1)Pt+i0P_{i}(r-1)P_{t+i}^{0} only has trivial allowable matrix, thus Pi(r1)Pt+i0=Sq(r1,r2)=[Pti,Pi(r)]P_{i}(r-1)P_{t+i}^{0}=Sq(r_{1},r_{2}\cdots)=[P^{i}_{t},P_{i}(r)].

As for (6), Pi0Pt(r)P^{0}_{i}P_{t}(r) only has trivial allowable matrix, while beside the trivial matrix, Pt(r)Pi0P_{t}(r)P^{0}_{i} has another allowable matrix X=(xuv)X=(x_{uv}) where xt0=r2ix_{t0}=r-2^{i}, xti=1x_{ti}=1 and all the other entries are zero. Note that β(X)=1\beta(X)=1, therefore [Pi0,Pt(r)]=Sq(r1,r2)[P^{0}_{i},P_{t}(r)]=Sq(r_{1},r_{2}\cdots) with rt=r2ir_{t}=r-2^{i}, rt+i=1r_{t+i}=1 and all the other components are zero. Meanwhile, Pt(r2i)Pt+i0=Sq(r1,r2)P_{t}(r-2^{i})P_{t+i}^{0}=Sq(r_{1},r_{2}\cdots)with rt=r2ir_{t}=r-2^{i}, rt+i=1r_{t+i}=1 and all the other components are zero, since the product only has trivial allowable matrix, we are done.

Lemma 2.13.

Suppose r=i=0tαi2ir=\sum_{i=0}^{t}\alpha_{i}2^{i} is the dyadic decomposition of rr and r<2t+1r<2^{t+1}, then Pt(r)=αi=1PtiP_{t}(r)=\prod_{\alpha_{i}=1}P^{i}_{t}.

Proof: The lemma follows from (2) of Lemma 2.12 and Lemma 2.11.

We define the excess of a Milnor basis element by

ex(Sq(r1,r2,))=r1+r2+.ex(Sq(r_{1},r_{2},\cdots))=r_{1}+r_{2}+\cdots.

This relates well to the product:

Lemma 2.14.

[Lemma 15.1.2 of Margolis (1983)]

Sq(r1,r2,)Sq(s1,s2,)=bSq(r1+s1,r2+s2,)+ΣSq(t1,t2,)Sq(r_{1},r_{2},\cdots\cdot)Sq(s_{1},s_{2},\cdots)=bSq(r_{1}+s_{1},r_{2}+s_{2},\cdots)+\Sigma Sq(t_{1},t_{2},\cdots)

with b=(ri,si)b=\prod(r_{i},s_{i}) and ex(Sq(t1,t2,))<ex(Sq(r1+s1,r2+s2,))ex(Sq(t_{1},t_{2},\cdots))<ex(Sq(r_{1}+s_{1},r_{2}+s_{2},\cdots))

Definition 2.15.

Let BB be a Hopf subalgebra of 𝒜2\mathcal{A}_{2} and define the profile function of BB, hB:{1,2,}{0,1,,}h_{B}:\{1,2,\cdots\}\to\{0,1,\cdots,\infty\} by

hB(t)=min{s|rt<2sforallSq(r1,r2,)B},h_{B}(t)=\min\{s|r_{t}<2^{s}\ for\ all\ Sq(r_{1},r_{2},\cdots)\in B\},

and hB(t)=h_{B}(t)=\infty if no such ss exists.

The importance of profile functions and the PtsP^{s}_{t}’s can be seen in the following classification theorem:

Theorem 2.16 (Theorem 15.1.6 of Margolis (1983)).

Given a Hopf subalgebra BB of 𝒜2\mathcal{A}_{2}, then

(a) BB is spanned by the Milnor basis elements in it; precisely, BB has a /2\mathbb{Z}/2-basis {Sq(r1,r2,)|rt<2hB(t)}\{Sq(r_{1},r_{2},\cdots)|r_{t}<2^{h_{B}(t)}\}.

(b) BB is generated as an algebra by {Pts|s<hB(t)}\{P^{s}_{t}|s<h_{B}(t)\}.

(c) h:{1,2,}{0,1,,}h:\{1,2,\cdots\}\to\{0,1,\cdots,\infty\} is the profile function of a Hopf subalgebra if and only if for all u,v1u,v\geq 1, h(u)v+h(u+v)h(u)\leq v+h(u+v) or h(v)h(u+v)h(v)\leq h(u+v). Moreover, the algebra being normal if the latter condition is always satisfied.

Remark 2.17.

Suppose hA(t)h_{A}(t) and hB(t)h_{B}(t) are two profile functions. Then AA is a Hopf subalgebra of BB if and only if hA(t)hB(t)h_{A}(t)\leq h_{B}(t) for all tt.

Remark 2.18.

Suppose hA(t)h_{A}(t) and hB(t)h_{B}(t) are two profile functions for graded Hopf subalgebras. Then m(t)=min{hA(t),hB(t)}m(t)=min\{h_{A}(t),h_{B}(t)\} is the profile function of ABA\cap B.

Definition 2.19.

Here are two examples:

(1) For each tt define a Hopf subalgebra J(t)J(t) by the profile function h(u)=0,uth(u)=0,u\neq t and h(t)=th(t)=t. Then J(t)J(t) is an exterior algebra on generators Pt0,,Ptt1P^{0}_{t},\cdots,P^{t-1}_{t}.

(2) A(n)A(n) is defined by the profile function h(t)=max{n+2t,0}h(t)=max\{n+2-t,0\}. It has a minimal generating set {P10,P11,P1n}.\{P^{0}_{1},P^{1}_{1},\cdots P^{n}_{1}\}.

Remark 2.20.

A(n)A(n) is the Hopf subalgebra generated by Sq1,Sq2,,Sq2nSq^{1},Sq^{2},\cdots,Sq^{2^{n}}.

Corollary 2.21.

Suppose BB is a finite Hopf subalgebra of 𝒜2\mathcal{A}_{2}, then

t=1(s=0hB(t)1Pts)0.\prod_{t=1}\left(\prod_{s=0}^{h_{B}(t)-1}P^{s}_{t}\right)\neq 0.

Proof: Iterated applications of Lemma 2.14 tell that

t=1(s=0hB(t)1Pts)=Sq(r1,r2,)+(terms with smaller excess)0\prod_{t=1}\left(\prod_{s=0}^{h_{B}(t)-1}P^{s}_{t}\right)=Sq(r_{1},r_{2},\cdots)+(\text{terms with smaller excess})\neq 0

where ri=2hB(i)1r_{i}=2^{h_{B}(i)}-1. Next, we introduce conditions for a function to be a profile function.

Lemma 2.22.

If h2h_{2} is a profile function, h1:{1,2,}{0,1,2}h_{1}:\{1,2,\cdots\}\to\{0,1,2\cdots\} is a function such that

(1) h1(t)h2(t)h_{1}(t)\leq h_{2}(t), for all t1t\geq 1.

(2) t01\exists t_{0}\geq 1 such that h1(t)=h2(t)h_{1}(t)=h_{2}(t) for tt0t\geq t_{0} .

Then h1h_{1} is a profile function if h1(u)v+h1(u+v)h_{1}(u)\leq v+h_{1}(u+v) or h1(v)h1(u+v)h_{1}(v)\leq h_{1}(u+v), for all u,v1u,v\geq 1 and u+vt0u+v\leq t_{0}.

Proof: If u+v>t0u+v>t_{0} and h1(v)>h1(u+v)=h2(u+v)h_{1}(v)>h_{1}(u+v)=h_{2}(u+v) then h2(v)h1(v)>h2(u+v)h_{2}(v)\geq h_{1}(v)>h_{2}(u+v). By assumption, h2h_{2} is a profile function, thus h2(u)v+h2(u+v)h_{2}(u)\leq v+h_{2}(u+v). Therefore h1(u)h2(u)v+h2(u+v)=v+h1(u+v)h_{1}(u)\leq h_{2}(u)\leq v+h_{2}(u+v)=v+h_{1}(u+v).

Lemma 2.23.

Suppose hh is a profile function, h1:{1,2,}{0,1,2}h_{1}:\{1,2,\cdots\}\to\{0,1,2\cdots\} is a function such that

h1(t)={h(t)tt00t=t0h_{1}(t)=\begin{cases}h(t)&t\neq t_{0}\\ 0&t=t_{0}\end{cases}

Then h1(t)h_{1}(t) is a profile function if for 1u<t01\leq u<t_{0}, h1(t0u)>0h_{1}(t_{0}-u)>0 implies h1(u)t0uh_{1}(u)\leq t_{0}-u.

Proof: Since h1(t0)=0h_{1}(t_{0})=0, h1(t0)h1(t0+u)u+h1(t0+u)h_{1}(t_{0})\leq h_{1}(t_{0}+u)\leq u+h_{1}(t_{0}+u).

By assumption, for 1u<t01\leq u<t_{0}, h1(t0u)>0h_{1}(t_{0}-u)>0 implies h1(u)t0uh_{1}(u)\leq t_{0}-u. Therefore, h1(u)t0u+h1(t0)h_{1}(u)\leq t_{0}-u+h_{1}(t_{0}) or h1(t0u)h1(t0)h_{1}(t_{0}-u)\leq h_{1}(t_{0}).

For the rest situation, namely 1u,v1\leq u,v such that ut0u\neq t_{0}, vt0v\neq t_{0} and u+vt0u+v\neq t_{0}, one has h1(u)=h(u)h_{1}(u)=h(u), h1(v)=h(v)h_{1}(v)=h(v) and h1(u+v)=h(u+v)h_{1}(u+v)=h(u+v). Since hh is a profile function, h1(u)v+h1(u+v)h_{1}(u)\leq v+h_{1}(u+v) or h1(v)h1(u+v)h_{1}(v)\leq h_{1}(u+v), for all such u,vu,v.

Suppose A𝒜2A\subset\mathcal{A}_{2} is a Hopf subalgebra and BAB\subset A is normal Hopf subalgebra of AA. In order to describe A//BA//B, it’s helpful to know more about AB+=B+AAB^{+}=B^{+}A, the ideal generated by BB.

Lemma 2.24.

Suppose A𝒜2A\subset\mathcal{A}_{2} is a Hopf subalgebra and B=CAB=C\cap A where CC is normal Hopf subalgebra of 𝒜2\mathcal{A}_{2}. If s,t\forall s,t such that 0s<hC(t)0\leq s<h_{C}(t), one has 2srt2^{s}\notin r_{t}, then Sq(r1,r2,)Sq(r_{1},r_{2},\cdots) is not a summand of any xAB+=B+Ax\in AB^{+}=B^{+}A.

Proof: By the proof of Theorem 15.1.6 (c) of Margolis (1983), we have that

𝒜2C+=C+𝒜2=span{Sq(u1,u2)|0s<hC(t) such that 2sut}.\mathcal{A}_{2}C^{+}=C^{+}\mathcal{A}_{2}=span\{Sq(u_{1},u_{2}\cdots)|\exists 0\leq s<h_{C}(t)\text{ such that }2^{s}\in u_{t}\}.

Thus by assumption, Sq(r1,r2,)Sq(r_{1},r_{2},\cdots) is not a summand of any x𝒜2C+=C+𝒜2x\in\mathcal{A}_{2}C^{+}=C^{+}\mathcal{A}_{2} and thus Sq(r1,r2,)Sq(r_{1},r_{2},\cdots) is not a summand of any xAB+=B+Ax\in AB^{+}=B^{+}A.

We state the following two lemmas here for later argument.

Lemma 2.25.

Given a finite Hopf subalgebra AA of 𝒜2\mathcal{A}_{2}. If BB is a maximal elementary Hopf subalgebra of AA then B=BiAB=B_{i}\cap A for some ii, where

hBi(t)={0t<iitih_{B_{i}}(t)=\begin{cases}0&t<i\\ i&t\geq i\end{cases}

Proof: By example A.6. of Palmieri (1997), BiB_{i} are all the maximal elementary Hopf subalgebras of 𝒜2\mathcal{A}_{2}. If BA𝒜2B\subset A\subset\mathcal{A}_{2} is an elementary subHopf algebra, then BBiAB\subset B_{i}\cap A for some ii. On the other hand, BiAB_{i}\cap A themselves are elementary Hopf subalgebra, thus if BB is maximal, then B=BiAB=B_{i}\cap A for some ii.

Remark 2.26.

The maximal elementary Hopf subalgebras of A(n)A(n) are exactly {BiA(n)}\{B_{i}\cap A(n)\} where 1i[n/2]+11\leq i\leq[n/2]+1, since BiA(n)B[n/2]+1A(n)B_{i}\cap A(n)\subset B_{[n/2]+1}\cap A(n) if i[n/2]+1i\geq[n/2]+1 and BiA(n)BjA(n)B_{i}\cap A(n)\nsubseteq B_{j}\cap A(n) if i,j[n/2]+1i,j\leq[n/2]+1 and iji\neq j.

Let E(n)A(n)E(n)\subset A(n) be the sub algebra generated by {Pt0|tn+1}\{P^{0}_{t}|t\leq n+1\}. Then E(n)E(n) is an exterior algebra and is a normal Hopf subalgebra of A(n)A(n).

Lemma 2.27 (Proposition 15.3.29 of Margolis (1983)).

There is a map η:A(n1)A(n)//E(n)\eta:A(n-1)\to A(n)//E(n) which doubles degree and which is an algebra isomorphism.

In general a map (resp. isomorphism) of algebras induces a functor (resp. isomorphism) of the module categories. Here η\eta introduces a small technicality.

Proposition 2.28 (Proposition 15.3.30 of Margolis (1983)).

η\eta induces an isomorphism

η:gr(A(n)//E(n))evgrA(n1)\eta^{*}:gr(A(n)//E(n))^{ev}\to grA(n-1)

Here gr(A(n)//E(n))evgr(A(n)//E(n))^{ev} denotes the full subcategory of gr(A(n)//E(n))gr(A(n)//E(n)) of modules concentrated in even degree.

3 Reduction to Hopf subalgebras

In this section we hope to reduce the calculation of 𝕋(H)\mathbb{T}(H) to the calculation of 𝕋(E)\mathbb{T}(E), where EE is some proper Hopf subalgebra of HH (Theorem 3.14). Then in some sense we can do an induction (see next section). This section is devoted to prove Theorem 3.14.

Recall that by assumption HH is a Poincare algebra and thus Hσ(n)HH\cong\sigma(n)H^{*}. Namely, H0HnkH_{0}\cong H_{n}\cong k. We denote the unique basis (up to a nonzero scale multiplication) of the 11-dimensional vector space HnH_{n} as t1Ht^{H}_{1}.

Definition 3.1.

Let MM be an HH-module and ZZ be a normal Hopf subalgebra of HH. Define

MZ:={xM|hx=ϵ(h)x,hZ},M1Z:={t1Zm|mM}.M^{Z}:=\{x\in M|hx=\epsilon(h)x,\forall h\in Z\},\quad M^{Z}_{1}:=\{t^{Z}_{1}\cdot m|\forall m\in M\}.

MZM^{Z} is called the (left) invariant ZZ-submodule of MM.

Remark 3.2.

MZM^{Z} is a well defined H//ZH//Z-module where αm=hm,αH//Z,mMZ,\alpha\cdot m=h\cdot m,\quad\forall\alpha\in H//Z,m\in M^{Z}, and hh is an arbitrary preimage of α\alpha of the quotient map π:HH//Z\pi:H\to H//Z.

It’s easy to verify that HomH(k,M){xM|hx=ϵ(h)x,hH}Hom_{H}(k,M)\cong\{x\in M|hx=\epsilon(h)x,\forall h\in H\} as graded vector spaces. Notice that HH itself is an HH-module and we can also consider the invariant HH-submodule of HH.

Lemma 3.3.

Suppose |H|=n|H|=n, then HH=HnH^{H}=H_{n}

Proof: Let x=i=0nxiHHx=\sum_{i=0}^{n}x_{i}\in H^{H}, with xiHix_{i}\in H_{i}. And let xjx_{j} be the first nonzero summand. Then since HH is a Poincare algebra, there is ynjHnjy_{n-j}\in H_{n-j} such that ynjxj=t1H0y_{n-j}x_{j}=t^{H}_{1}\neq 0. Therefore, ynjx=t1Hy_{n-j}x=t^{H}_{1} for grading reason. On the other hand, since xHHx\in H^{H}, ynjx=ϵ(ynj)xy_{n-j}x=\epsilon(y_{n-j})x, which leads to ϵ(ynj)x=t1H\epsilon(y_{n-j})x=t^{H}_{1}. This is true only if x=at1Hx=at^{H}_{1} and y0=1/ay_{0}=1/a for some nonzero aka\in k. Thus HHHnH^{H}\subset H_{n}. By definition, hH+\forall h\in H^{+}, |ht1H|>n|ht^{H}_{1}|>n, hence ht1H=0=ϵ(h)t1hht^{H}_{1}=0=\epsilon(h)t^{h}_{1}. Therefore, HnHHH_{n}\subset H^{H}.

By the same argument, (σ(m)H)H=σ(m)Hn(\sigma(m)H)^{H}=\sigma(m)H_{n}, the top degree of σ(m)H\sigma(m)H. Another observation is that (σ(m)H)1H=σ(m)Hn(\sigma(m)H)^{H}_{1}=\sigma(m)H_{n} , since t1Hσ(m)Hiσ(m)Hn+i=0t^{H}_{1}\cdot\sigma(m)H_{i}\subset\sigma(m)H_{n+i}=0 for i>0i>0 and t1Hσ(m)H0=σ(m)Hnt^{H}_{1}\sigma(m)H_{0}=\sigma(m)H_{n}.

In general, M1HM^{H}_{1} and MHM^{H} are not equal. However, this is true when MM is free.

Lemma 3.4.

When MM is a free HH module, MH=M1HM^{H}=M^{H}_{1}.

Proof: Suppose |H|=n|H|=n. hH,mM\forall h\in H,m\in M, ht1Hm=ϵ(h)t1Hmh\cdot t^{H}_{1}m=\epsilon(h)t^{H}_{1}m by definition, thus M1HMHM^{H}_{1}\subset M^{H}.

On the other hand, suppose {m1,ms}\{m_{1},\cdots m_{s}\} is an HH-basis for MM as free module, then {t1H(m1),,t1H(ms)}\{t^{H}_{1}(m_{1}),\cdots,t^{H}_{1}(m_{s})\} is a basis for M1HM^{H}_{1} as vector space, since t1Hσ(m)H=σ(m)Hnt^{H}_{1}\sigma(m)H=\sigma(m)H_{n} MHHomH(k,M)HomH(k,i=1sσ(αi)H)i=1s(σ(αi)H)Hi=1sσ(αi)Hn=i=1sσ(αi+n)k.M^{H}\cong Hom_{H}(k,M)\cong Hom_{H}(k,\oplus_{i=1}^{s}\sigma(\alpha_{i})H)\cong\oplus_{i=1}^{s}(\sigma(\alpha_{i})H)^{H}\cong\oplus_{i=1}^{s}\sigma(\alpha_{i})H_{n}=\oplus_{i=1}^{s}\sigma(\alpha_{i+n})k. Thus MH=M1HM^{H}=M^{H}_{1} since they have same dimension as vector spaces over kk.

The element t1Ht^{H}_{1} can be used to detect the freeness of an HH-module.

Lemma 3.5 (Lemma 12.2.6 of Margolis (1983)).

If MM is an HH-module, then multiplication by t1Ht^{H}_{1} induces a map f:kHMMf:k\otimes_{H}M\to M (of degree |t1H||t^{H}_{1}|). Furthermore, ff is a monomorphism if and only if MM is free.

Lemma 3.6.

Let MM be an HH-module. Suppose t1Hm1,t1Hm2,,t1Hmnt^{H}_{1}m_{1},t^{H}_{1}m_{2},\cdots,t^{H}_{1}m_{n} are kk-linearly independent. Then m1,m2,,mnm_{1},m_{2},\cdots,m_{n} generate a free HH-submodule LML\subset M.

Proof: By assumption, kHLk\otimes_{H}L is a kk vector space spanned by 1m1,1m2,,1mn1\otimes m_{1},1\otimes m_{2},\cdots,1\otimes m_{n}. Consider the map f:kHLLf:k\otimes_{H}L\to L defined in Lemma3.5:

f(i=1naimi)=i=1nait1Hmif(\sum_{i=1}^{n}a_{i}\otimes m_{i})=\sum_{i=1}^{n}a_{i}t^{H}_{1}m_{i}

where aika_{i}\in k. Since t1Hm1,,t1Hmnt^{H}_{1}m_{1},\cdots,t^{H}_{1}m_{n} are kk-linearly independent, ff is injective.

By Lemma 3.5, we are done.

In fact, t1Ht^{H}_{1} is just the integral of the Hopf algebra HH (Montgomery (1993)), and it has some kind of transitivity:

Lemma 3.7.

Suppose ZHZ\subset H is a normal Hopf subalgebra. Denote the generator of Z|Z|Z_{|Z|} and (H//Z)|H//Z|(H//Z)_{|H//Z|} as t1Zt^{Z}_{1} and tZH¯\overline{t^{H}_{Z}} respectively. Then tZHt1Zt^{H}_{Z}t^{Z}_{1} is the generator of H|H|H_{|H|} where tZHt^{H}_{Z} is a preimage of tZH¯\overline{t^{H}_{Z}} under the quotient map.

Proof: Show that tZHt1Zt^{H}_{Z}t^{Z}_{1} is well-defined first:

aHZ+,(tZH+a)t1Z=tZHt1Z+hzt1Z=tZHt1Z\forall a\in HZ^{+},\ \ (t^{H}_{Z}+a)t^{Z}_{1}=t^{H}_{Z}t^{Z}_{1}+hzt^{Z}_{1}=t^{H}_{Z}t^{Z}_{1}

where a=hza=hz for some hHh\in H and zZ+z\in Z^{+}.

Recall that HH is a Poincare algebra. By definition, xH|H||Z|\exists x\in H_{|H|-|Z|} such that xt1Z0xt^{Z}_{1}\neq 0. By Theorem 2.4, xH|H//Z|x\in H_{|H//Z|}. Since H|H|=kH_{|H|}=k, is a one dimensional vector space generated by t1Ht^{H}_{1}, we can assume that

xt1Z=t1Hxt^{Z}_{1}=t^{H}_{1}

after a scalar multiplication if necessary.

Now that xt1Z0xt^{Z}_{1}\neq 0, one has xHZ+x\notin HZ^{+} and hence the quotient image of xx in H//ZH//Z is nonzero. Therefore xx is a preimage of tZH¯\overline{t^{H}_{Z}} since |x|=|H//Z||x|=|H//Z| and (H//Z)|H//Z|=k(H//Z)_{|H//Z|}=k, a one dimensional vector space.

Thanks to the transitivity, we have the following two technical lemmas:

Lemma 3.8.

Let ZHZ\subset H be a normal Hopf subalgebra. Suppose MM is a finitely generated HH-module such that M1ZM^{Z}_{1} is a free H//ZH//Z module, and MZHk(free)M\downarrow^{H}_{Z}\cong k\oplus(free). Then Mk(free)M\cong k\oplus(free) as HH-modules.

Proof: Let t1Zm1,,t1Zmnt^{Z}_{1}m_{1},\cdots,t^{Z}_{1}m_{n} be a basis of M1ZM^{Z}_{1} as free H//ZH//Z-module. Then by Lemma 3.7 and Lemma 3.4, we have

tZHt1Zm1=t1Hm1,,tZHt1Zmn=t1Hmnt^{H}_{Z}t^{Z}_{1}m_{1}=t^{H}_{1}m_{1},\cdots,t^{H}_{Z}t^{Z}_{1}m_{n}=t^{H}_{1}m_{n}

are kk-linearly independent. Therefore {mi}\{m_{i}\} generate a free HH-submodule of MM by Lemma 3.6, named as LL. Therefore, LL is also a free ZZ-module, which implies LZ=L1Z=M1ZL^{Z}=L^{Z}_{1}=M^{Z}_{1}, generated by t1Zm1,,t1Zmnt^{Z}_{1}m_{1},\cdots,t^{Z}_{1}m_{n} as free H//ZH//Z-module. On the other hand, M1Z=F1Z=FZM^{Z}_{1}=F^{Z}_{1}=F^{Z}, where by assumption MZHkFM\downarrow^{H}_{Z}\cong k\oplus F for some FF free as ZZ-module. Therefore LZ=FZL^{Z}=F^{Z}, hence dim(L)=dim(F)dim(L)=dim(F) since both LL and FF are free ZZ-modules. Thus dim(L)=dim(M)1dim(L)=dim(M)-1 and since LL is free, hence injective, as an HH-module, we get MLkM\cong L\oplus k, as was to be shown.

Corollary 3.9.

Let ZZ be a normal Hopf subalgebra. Then

HZσ(|Z|)H//ZH^{Z}\cong\sigma(|Z|)H//Z

as H//ZH//Z-modules.

Proof: By Theorem 2.4, and Lemma 3.4, as graded kk-vector spaces,

HZ=t1ZZH//Z=k{t1Z}H//Zσ(|Z|)H//Z.H^{Z}=t^{Z}_{1}\cdot Z\otimes H//Z=k\{t^{Z}_{1}\}\otimes H//Z\cong\sigma(|Z|)H//Z.

On the other hand,

tZH¯HZ=tZHt1HH=t1HH=k{t1H}.\overline{t^{H}_{Z}}H^{Z}=t^{H}_{Z}t^{H}_{1}H=t^{H}_{1}H=k\{t^{H}_{1}\}.

Then by Lemma 3.6, HZH^{Z} has a free H//ZH//Z-submodule. Therefore, HZσ(|Z|)H//ZH^{Z}\cong\sigma(|Z|)H//Z as H//ZH//Z- modules for dimensional reason.

Definition 3.10.

Suppose AA is an algebra and MM is an AA-module. Let ={Bi|iI}\mathcal{B}=\{B_{i}|i\in I\} be a collection of nontrivial subalgebras of AA. If MM is free iff MM restricted to every subalgebra BB\in\mathcal{B} is free, then we say AA has detect property in grAgrA, with detecting set \mathcal{B}.

Here BB is nontrivial means BAB\neq A and BkB\neq k.

Theorem 3.11 (Theorem 1.3, Example A.6 of Palmieri (1997)).

Suppose EE is a finite Hopf subalgebra of the mod 22 Steenrod algebra 𝒜2\mathcal{A}_{2}, MM is an EE-module. Then MM is projective if and only if MM restricted to BB is projective for every elementary Hopf subalgebra BB of EE.

Remark 3.12.

Here EE and BB are all graded, and the grading was given by that of 𝒜2\mathcal{A}_{2}.

Remark 3.13.

In particular, if EE is not elementary, EE has detect property in grEgrE with detecting set {B|B is a maximal elementary Hopf subalgebra }\{B|B\text{ is a maximal elementary Hopf subalgebra }\}.

Now we are ready to prove our main theorem:

Theorem 3.14.

Suppose HH has a nontrivial normal Hopf subalgebra ZZ such that H//ZH//Z has detect property in gr(H//Z)gr(H//Z) with detecting set ={Bi|iI}\mathcal{B}=\{B_{i}|i\in I\}. Then

Res:𝕋(H)E𝕋(E).Res:\mathbb{T}(H)\longrightarrow\prod_{E\in\mathcal{E}}\mathbb{T}(E).

is injective, where ={Ei|Ei//Z=Bi,iI}\mathcal{E}=\{E_{i}|E_{i}//Z=B_{i},i\in I\}.

Proof: By assumption there is a nontrivial normal Hopf subalgebra ZZ of HH such that H//ZH//Z has detect property with detecting set \mathcal{B}.

Let MM be an endotrivial HH-module whose class in 𝕋(H)\mathbb{T}(H) is in the kernel of the restriction map ResRes. This means MkFM\cong k\oplus F as EE-modules, for EE\in\mathcal{E}, where FF is a free EE-module. Since ZZ is finite nontrivial connected Hopf algebra, |t1Z|0|t^{Z}_{1}|\neq 0 and thus t1ZZ+t^{Z}_{1}\in Z^{+}. If follows that

MZ=kFZ,M1Z=F1Z.M^{Z}=k\oplus F^{Z},\quad M^{Z}_{1}=F^{Z}_{1}.

Since FF is a free EE-module, FF is also a free ZZ module, thus by Lemma 3.4, F1Z=FZF^{Z}_{1}=F^{Z} . Moreover, FZF^{Z} is free over E//ZE//Z since FF is free over EE, see Corollary 3.9.

Therefore, M1ZM^{Z}_{1} is a free E//ZE//Z module, for every EE\in\mathcal{E}. Using the fact that H//ZH//Z has detect property with detecting set ={Bi|iI}={E//Z|E}\mathcal{B}=\{B_{i}|i\in I\}=\{E//Z|E\in\mathcal{E}\}, we have M1ZM^{Z}_{1} is a free H//ZH//Z-module. By Lemma 3.8, we are done.

4 Endotrivial group of A(n)

Consider the Hopf algebra A(n),n2A(n),n\geq 2. From now on, let a=[(n1)/2]a=[(n-1)/2], ii be an integer with 1ia+11\leq i\leq a+1 and we will work on the field /2\mathbb{Z}/2.

By Lemma 2.27, there is an isomorphism of algebras η:A(n1)A(n)//E(n)\eta:A(n-1)\to A(n)//E(n) which doubles the grading. Use the same symbol as in the proof of Lemma 2.25, the maximal elementary Hopf subalgebras of A(n1)A(n-1) are exactly B1A(n1),,Ba+1A(n1)B_{1}\cap A(n-1),\cdots,B_{a+1}\cap A(n-1), with which A(n1)A(n-1) has detect property. Define

hBi(n)(t)={hBiA(n1)(t)+11tn+10t>n+1h_{B_{i}^{\prime}(n)}(t)=\begin{cases}h_{B_{i}\cap A(n-1)}(t)+1&1\leq t\leq n+1\\ 0&t>n+1\end{cases}

By Lemma 2.22, Bi(n)B_{i}^{\prime}(n) is a Hopf subalgebra, since hBi(n)(t)h_{B_{i}^{\prime}(n)}(t) is an increasing function when tit\leq i and hBi(n)(t)=hA(n)(t)h_{B_{i}^{\prime}(n)}(t)=h_{A(n)}(t) when tit\geq i.

Again by Lemma 2.22, for each Bi(n)B_{i}^{\prime}(n) there are Hopf subalgebras Di(n)D_{i}(n) defined by

hDi(n)(t)={0t<ihBi(n)(t)ti.h_{D_{i}(n)}(t)=\begin{cases}0&t<i\\ h_{B_{i}^{\prime}(n)}(t)&t\geq i\end{cases}.

Given n2n\geq 2, recall that B1(n)=D1(n)B_{1}^{\prime}(n)=D_{1}(n).

There are injective homomorphisms between the group of endotrivial modules for these Hopf subalgebras:

𝕋(A(n))i=11+a𝕋(Bi(n))\mathbb{T}(A(n))\to\prod_{i=1}^{1+a}\mathbb{T}(B_{i}^{\prime}(n)) (Proposition 4.1)

𝕋(Bi(n))𝕋()×𝕋(Di(n)),2ia+1\mathbb{T}(B_{i}^{\prime}(n))\to\mathbb{T}(\ast)\times\mathbb{T}(D_{i}(n)),2\leq i\leq a+1 (Proposition 4.4).

𝕋(Di(n))𝕋()×𝕋(),1ia+1\mathbb{T}(D_{i}(n))\to\mathbb{T}(\ast)\times\mathbb{T}(\ast),1\leq i\leq a+1 (Proposition 4.5 4.6)

All the above maps are the restriction map induced by inclusion of Hopf subalgebras, \ast means certain elementary Hopf subalgebras.

Finally, we define Oi=<Pt0|itn+1>O_{i}=<P^{0}_{t}|i\leq t\leq n+1>.

The rest of this section will be devoted to the proof of Theorem 1.1.

Proposition 4.1.

n2\forall n\geq 2,

Res:𝕋(A(n))i=1a+1𝕋(Bi(n))Res:\mathbb{T}(A(n))\longrightarrow\prod_{i=1}^{a+1}\mathbb{T}(B^{\prime}_{i}(n))

is injective.

Proof: Given a finitely generated A(n)//E(n)A(n)//E(n)-module MM, assume that M|Bi(n)//E(n)M|_{B_{i}^{\prime}(n)//E(n)} is free in gr(Bi(n)//E(n))gr\left(B_{i}^{\prime}(n)//E(n)\right) for each 1ia+11\leq i\leq a+1, we will show that MM is a free A(n)//E(n)A(n)//E(n)-module.

The trick is that there is a natural splitting M=M1σ(1)M2M=M_{1}\oplus\sigma(1)M_{2} where M1M_{1} and M2M_{2} are all concentrated in even degree. This is because that A(n)//E(n)A(n)//E(n) is concentrated in even degree. Proposition 2.28 guarantees that (ηMj)|BiA(n1)(\eta^{*}M_{j})|_{B_{i}\cap A(n-1)} is free for j=1,2j=1,2; and then the detect property of A(n1)A(n-1) with detecting set {BiA(n1)|1ia+1}\{B_{i}\cap A(n-1)|1\leq i\leq a+1\} implies ηMj\eta^{*}M_{j} is a free A(n1)A(n-1)-module. Again, since η\eta^{*} is an isomorphism between the two module categories, MjM_{j} is a free A(n)//E(n)A(n)//E(n)-module, j=1,2j=1,2. Namely, M=M1σ(1)M2M=M_{1}\oplus\sigma(1)M_{2} is a free A(n)//E(n)A(n)//E(n)-module.

The above argument shows that A(n)//E(n)A(n)//E(n) has detect property with detecting sets {Bi(n)//E(n)|1ia+1}\{B_{i}^{\prime}(n)//E(n)|1\leq i\leq a+1\}. By Theorem 3.14, we are done.

Lemma 4.2.

Given two Hopf subalgebras E1E_{1}, E2E_{2} of HH and suppose MM is a finitely generated HH-module such that M|EjM|_{E_{j}} is a free EjE_{j}-module for j=1,2j=1,2.

M is a free H-module, if

(1)t10E1,t20E2\exists t_{1}\neq 0\in E_{1},t_{2}\neq 0\in E_{2} such that |t1t2|=|H||t_{1}t_{2}|=|H|, and

(2)for all homogeneous y0E2+,|y|>|t1|y\neq 0\in E^{+}_{2},|y|>|t_{1}| or y=zy=z with z2=0,0zt1E1+z^{2}=0,0\neq zt_{1}\in E^{+}_{1}.

Proof: Suppose f:j=1qHMf:\oplus_{j=1}^{q}H\to M with f(h1,h2,,hq)=j=1qhjmjf(h_{1},h_{2},\cdots,h_{q})=\sum_{j=1}^{q}h_{j}m_{j} is a surjective HH-module homomorphism with qq is minimal. Such ff and qq exist since MM is finitely generated. Furthermore, we assume that |mj||mj+1||m_{j}|\leq|m_{j+1}|. We show that m1m_{1} generates a free HH-submodule first.

Suppose m1=lxlblm_{1}=\sum_{l}x_{l}b_{l} for some homogeneous xlE1+,blMx_{l}\in E_{1}^{+},b_{l}\in M. Then for each ll, one has |xlbl|=|xl|+|bl|>|m1||x_{l}b_{l}|=|x_{l}|+|b_{l}|>|m_{1}|, contradiction. Hence xwxw is not a summand of m1m_{1}, xE1+,wM\forall x\in E_{1}^{+},w\in M. By assumption MM is a free E1E_{1} module, thus xm10xm_{1}\neq 0. In particular, t1m10t_{1}m_{1}\neq 0.

Suppose t1m1=lylclt_{1}m_{1}=\sum_{l}y_{l}c_{l} for some homogeneous ylE2+,clMy_{l}\in E_{2}^{+},c_{l}\in M and yl0z\exists y_{l_{0}}\neq z. Then |yl0cl0|=|yl0|+|cl0|>|t1|+|cl0||t1m1||y_{l_{0}}c_{l_{0}}|=|y_{l_{0}}|+|c_{l_{0}}|>|t_{1}|+|c_{l_{0}}|\geq|t_{1}m_{1}|, which is impossible. Thus t1m1=zct_{1}m_{1}=z\cdot c where c=lclc=\sum_{l}c_{l}, then zt1m1=(z)2c=0z\cdot t_{1}m_{1}=(z)^{2}c=0. On the other hand, by condition (2), 0zt1E10\neq zt_{1}\in E_{1}, which is a contradiction since we showed xm10xm_{1}\neq 0, for all nonzero xE1x\in E_{1}. In conclusion, ywyw is not a summand of t1m1t_{1}m_{1} for all yE2+y\in E_{2}^{+} and wMw\in M. Again since MM is a free E2E_{2} module, yt1m10yt_{1}m_{1}\neq 0.

In summary, yt1m10,yE2+yt_{1}m_{1}\neq 0,\forall y\in E_{2}^{+}. In particular, t2t1m10t_{2}t_{1}m_{1}\neq 0, which implies that t2t10t_{2}t_{1}\neq 0. By assumption, |t2t1|=|H||t_{2}t_{1}|=|H|. Therefore, t2t1=t1Ht_{2}t_{1}=t^{H}_{1} since HH is a finite connected Hopf algebra and t2t10t_{2}t_{1}\neq 0. By Lemma 3.6, m1m_{1} generates a free HH-submodule L1ML_{1}\subset M.

Since qq is minimal, mjL1m_{j}\notin L_{1} for j2j\geq 2. Define M2=M/L1M_{2}=M/L_{1}, then f2:j=1q1HM2f_{2}:\oplus_{j=1}^{q-1}H\to M_{2} with f(h1,,hq1)=j=1q1hjmj+1¯f(h_{1},\cdots,h_{q-1})=\sum_{j=1}^{q-1}h_{j}\overline{m_{j+1}} is a surjective HH-module homomorphism with q1q-1 is minimal , where |mj¯||mj+1¯||\overline{m_{j}}|\leq|\overline{m_{j+1}}|. By an induction on the number of generators, one may assume that M2M_{2} is free. Therefore, M=L1M2M=L_{1}\oplus M_{2} is a free HH-module. The proof of Proposition 4.4, Proposition 4.5, and Proposition 4.6 rely heavily on the above Lemma. To apply this Lemma, one must know well about the degree of the related algebraic generators PtsP^{s}_{t}.

Lemma 4.3.

For s,u>0s,u>0 and t,v1t,v\geq 1,

(1) If s+t=u+vs+t=u+v and tvt\geq v then |Pts||Pvu||P^{s}_{t}|\geq|P^{u}_{v}|;

(2) If s+t>u+vs+t>u+v, then |Pts|>|Pvu||P^{s}_{t}|>|P^{u}_{v}|.

Proof: (1) By definition, |Pts|=2s(2t1)=2s+t2s|P^{s}_{t}|=2^{s}(2^{t}-1)=2^{s+t}-2^{s} and |Pvu|=2u(2v1)=2u+v2u|P^{u}_{v}|=2^{u}(2^{v}-1)=2^{u+v}-2^{u}. By assumption usu\geq s and hence |Pts||Pvu||P^{s}_{t}|\geq|P^{u}_{v}|.

(2) By assumption, s+t>1s+t>1. By (1), since Ps+t0=2s+t1>2s+t1>2u+v1=P1u+v1P^{0}_{s+t}=2^{s+t}-1>2^{s+t-1}>2^{u+v-1}=P^{u+v-1}_{1}, we are done.

Proposition 4.4.

For each 2ia+12\leq i\leq a+1, n2n\geq 2,

Res:𝕋(Bi(n))𝕋(E(n))×𝕋(Di(n))Res:\mathbb{T}(B_{i}^{\prime}(n))\longrightarrow\mathbb{T}(E(n))\times\mathbb{T}(D_{i}(n))

is injective.

Proof: By Lemma 2.12, Pn+10P^{0}_{n+1} commutes with all the PtsBi(n)P^{s}_{t}\in B_{i}^{\prime}(n), thus <Pn+10><P^{0}_{n+1}> is a normal Hopf subalgebra and Bi(n)//<Pn+10>B_{i}^{\prime}(n)//<P^{0}_{n+1}> is a connected Hopf algebra, denote it as HH. Denote E(n)//<Pn+10>E(n)//<P^{0}_{n+1}> and Di(n)//<Pn+10>D_{i}(n)//<P^{0}_{n+1}> as E1E_{1} and E2E_{2} respectively, then E1E_{1} and E2E_{2} are connected Hopf subalgebra of HH. We will show that HH has detect property in grHgrH with detecting set {E1,E2}\{E_{1},E_{2}\}.

As an algebra, E1E_{1} is generated by ={Pt0¯|1tn}\mathcal{F}=\{\overline{P^{0}_{t}}|1\leq t\leq n\} , E2E_{2} is generated by 𝒢={Pts¯|itn,0shDi(n)(t)}\mathcal{G}=\{\overline{P^{s}_{t}}|i\leq t\leq n,0\leq s\leq h_{D_{i}(n)}(t)\} and HH is generated by 𝒢\mathcal{F}\cup\mathcal{G}.

Consider the element t1=j=1i1Pj0¯t_{1}=\prod_{j=1}^{i-1}\overline{P^{0}_{j}} and t2=t=in(s=0hBi(n)(t)1Pts¯)t_{2}=\prod_{t=i}^{n}\left(\prod_{s=0}^{h_{B_{i}^{\prime}(n)}(t)-1}\overline{P^{s}_{t}}\right) where t1E1t_{1}\in E_{1} and t2E2t_{2}\in E_{2}. By Corollary 2.21 and Lemma 2.24, they are nonzero. Next, we show they satisfy the assumption of Lemma 4.2.

Since |t1|=2ii|t_{1}|=2^{i}-i and |Pi0¯|=2i1|\overline{P^{0}_{i}}|=2^{i}-1, by Lemma 4.3, yE2+\forall y\in E^{+}_{2}, |y|>|t1||y|>|t_{1}|.

Notice that the Milnor basis of highest degree in Bi(n)B_{i}^{\prime}(n) is Sq(r1,r2,,rn+1)Sq(r_{1},r_{2},\cdots,r_{n+1}) with rt=2hBi(n)(t)1r_{t}=2^{h_{B_{i}^{\prime}(n)}(t)}-1, one has

|Bi(n)|=t=1n+1(2hBi(n)(t)1)(2t1).|B_{i}^{\prime}(n)|=\sum_{t=1}^{n+1}\left(2^{h_{B_{i}^{\prime}(n)}(t)}-1\right)\left(2^{t}-1\right).

Since |Pn+10|=2n+11|P^{0}_{n+1}|=2^{n+1}-1 and hBi(n)(n+1)=1h_{B_{i}^{\prime}(n)}(n+1)=1, by Theorem 2.4,

|H|=|Bi(n)||<Pn+10>|=t=1n(2hBi(n)(t)1)(2t1).|H|=|B_{i}^{\prime}(n)|-|<P^{0}_{n+1}>|=\sum_{t=1}^{n}\left(2^{h_{B_{i}^{\prime}(n)}(t)}-1\right)\left(2^{t}-1\right).

On the other hand

|t1|+|t2|=t=1n(2hBi(n)(t)1)(2t1),|t_{1}|+|t_{2}|=\sum_{t=1}^{n}\left(2^{h_{B_{i}^{\prime}(n)}(t)}-1\right)\left(2^{t}-1\right),

Therefore, we have |H|=|t1|+|t2||H|=|t_{1}|+|t_{2}|. By Lemma 4.2(1), HH has detect property in grHgrH with detecting set {E1,E2}\{E_{1},E_{2}\}. By Theorem 3.14, we are done.

Proposition 4.5.

For each 2ia+12\leq i\leq a+1, n2n\geq 2

Res:𝕋(Di(n))𝕋(BiDi(n))×𝕋(Bi+1Di(n))Res:\mathbb{T}(D_{i}(n))\longrightarrow\mathbb{T}(B_{i}\cap D_{i}(n))\times\mathbb{T}(B_{i+1}\cap D_{i}(n))

is injective.

Proof: Consider Yi(n)Di(n)Y_{i}(n)\subset D_{i}(n) with profile function

hYi(n)(t)={2i+1ti+1t2i10otherwise.h_{Y_{i}(n)}(t)=\begin{cases}2i+1-t&i+1\leq t\leq 2i-1\\ 0&otherwise\end{cases}.

Since n2n\geq 2 and ia+1i\leq a+1, Yi(n)A(2i1)Y_{i}(n)\subset A(2i-1) is a Hopf subalgebra of Di(n)D_{i}(n) by Lemma 2.22 and Lemma 2.23. Define Xi(n)Di(n)X_{i}(n)\subset D_{i}(n) by the function

hXi(n)(t)={i+1t=i1t=2i0otherwise.h_{X_{i}(n)}(t)=\begin{cases}i+1&t=i\\ 1&t=2i\\ 0&otherwise\end{cases}.

Directly by Theorem 2.16, Xi(n)X_{i}(n) is a Hopf subalgebra of Di(n)D_{i}(n).

By Lemma 2.12 (3) [x,y]=0[x,y]=0, xDi(n),yYi(n)\forall x\in D_{i}(n),y\in Y_{i}(n). Namely, Yi(n)Y_{i}(n) is a normal Hopf subalgebra of Di(n)D_{i}(n), and Di(n)//Yi(n)D_{i}(n)//Y_{i}(n) is a connected Hopf algebra, denoted as HH. Define E1=(BiDi(n))//Yi(n)E_{1}^{\prime}=\left(B_{i}\cap D_{i}(n)\right)//Y_{i}(n) and E2=(Bi+1Di(n))//Yi(n)E_{2}=\left(B_{i+1}\cap D_{i}(n)\right)//Y_{i}(n). We will show that HH has detect property in grHgrH with detecting sets {E1,E2}\{E_{1}^{\prime},E_{2}\}.

Suppose MM is a finitely generated HH-module, and M|E1M|_{E_{1}^{\prime}}, M|E2M|_{E_{2}} are free. We want to show that MM is a free HH-module.

By Lemma 2.24, Xi(n)(Yi(n)+Di(n))=0X_{i}(n)\cap\left(Y_{i}(n)^{+}D_{i}(n)\right)=0, namely π|Xi(n)\pi|_{X_{i}(n)} is an isomorphism of graded algebras from Xi(n)X_{i}(n) to πXi(n)\pi X_{i}(n), where π:Di(n)H\pi:D_{i}(n)\to H is the quotient map. By Lemma 3.11 and Lemma 2.25, Xi(n)X_{i}(n) has detect property in grXi(n)grX_{i}(n) with detecting set {BiXi(n)}\{B_{i}\cap X_{i}(n)\}, hence πXi(n)\pi X_{i}(n) has detect property in gr(πXi(n))gr\left(\pi X_{i}(n)\right) with detecting set {π(BiXi(n))}\{\pi\left(B_{i}\cap X_{i}(n)\right)\}. Now that MM is a free E1E_{1}^{\prime}-module and π(BiXi(n))\pi(B_{i}\cap X_{i}(n)) is a Hopf subalgebra of E1E_{1}^{\prime}, it is free over π(BiXi(n))\pi(B_{i}\cap X_{i}(n)) , and thus MM is a free πXi(n)\pi X_{i}(n)-module. Now, denote E1=πXi(n)E_{1}=\pi X_{i}(n).

Consider the element t1=j=0iPij¯t_{1}=\prod_{j=0}^{i}\overline{P^{j}_{i}} and t2=t=i+1n+1(s=hYi(n)(t)hDi(n)(t)1Pts¯)t_{2}=\prod_{t=i+1}^{n+1}\left(\prod_{s=h_{Y_{i}(n)}(t)}^{h_{D_{i}(n)}(t)-1}\overline{P^{s}_{t}}\right) where t1E1t_{1}\in E_{1} and t2E2t_{2}\in E_{2}. By Corollary 2.21 and Lemma 2.24, they are nonzero. Next, we show they satisfy the assumption (2) of Lemma 4.2.

Take z=P2i0¯z=\overline{P^{0}_{2i}}, we know z2=0z^{2}=0. By definition of the profile functions, zE1E2z\in E_{1}\cap E_{2}. Since |t1|=(2i+11)(2i1)|t_{1}|=(2^{i+1}-1)(2^{i}-1) and |Pi+1i¯|=(2i1)2i|\overline{P^{i}_{i+1}}|=(2^{i}-1)2^{i}, by Lemma 4.3, yE2+\forall y\in E^{+}_{2}, |y|>|t1||y|>|t_{1}| as long as yzy\neq z. Moreover, zt1E1zt_{1}\in E_{1} because t1,zE1t_{1},z\in E_{1} and by Corollary 2.21 and 2.24, zt10zt_{1}\neq 0.

By definition,

|t1t2|=t=in+1(2hDi(n)(t)2hYi(n)(t))(2t1).|t_{1}t_{2}|=\sum_{t=i}^{n+1}\left(2^{h_{D_{i}(n)}(t)}-2^{h_{Y_{i}(n)}(t)}\right)\left(2^{t}-1\right).

By the definition of profile function,

|Di(n)|=t=in+1(2hDi(n)(t)1)(2t1).|D_{i}(n)|=\sum_{t=i}^{n+1}\left(2^{h_{D_{i}(n)}(t)}-1\right)\left(2^{t}-1\right).

and

|Yi(n)|=t=in+1(2hYi(n)(t)1)(2t1).|Y_{i}(n)|=\sum_{t=i}^{n+1}\left(2^{h_{Y_{i}(n)}(t)}-1\right)\left(2^{t}-1\right).

Therefore, we have

|H|=|Di(n)||Yi(n)|=|t1|+|t2|.|H|=|D_{i}(n)|-|Y_{i}(n)|=|t_{1}|+|t_{2}|.

By Lemma 4.2, MM is a free HH-module, namely, HH has detect property in grHgrH with detecting set {E1,E2}\{E_{1}^{\prime},E_{2}\}. By Theorem 3.14, we are done.

Proposition 4.6.

For each n2n\geq 2

Res:𝕋(D1(n))𝕋(E(n))×𝕋(B2D1(n))Res:\mathbb{T}(D_{1}(n))\longrightarrow\mathbb{T}(E(n))\times\mathbb{T}(B_{2}\cap D_{1}(n))

is injective.

Proof: By Lemma 2.12, Pn+10P^{0}_{n+1} commutes with all the PtsD1(n)P^{s}_{t}\in D_{1}(n), thus <Pn+10><P^{0}_{n+1}> is a normal Hopf subalgebra and D1(n)//<Pn+10>D_{1}(n)//<P^{0}_{n+1}> is a connected Hopf algebra, denote it as HH. Moreover, denote E(n)//<Pn+10>E(n)//<P^{0}_{n+1}> and B2D1(n)//<Pn+10>B_{2}\cap D_{1}(n)//<P^{0}_{n+1}> as E1E_{1}^{\prime} and E2E_{2} respectively, then E1E_{1}^{\prime} and E2E_{2} are connected Hopf subalgebra of HH, and we will show that HH has detect property in grHgrH with detecting set {E1,E2}\{E_{1}^{\prime},E_{2}\}.

Suppose MM is a finitely generated HH-module, and M|E1M|_{E^{\prime}_{1}}, M|E2M|_{E_{2}} are free. We want to show that MM is a free HH-module.

By assumption n2n\geq 2, thus |A(1)|=6<|Pn+10|=2n+11|A(1)|=6<|P^{0}_{n+1}|=2^{n+1}-1. Therefore A(1)<Pn+10>+D1(n)=0A(1)\cap<P^{0}_{n+1}>^{+}D_{1}(n)=0, which implies π|A(1)\pi|_{A(1)} is an isomorphism of graded algebras from A(1)A(1) to πA(1)\pi A(1), where π:D1(n)D1(n)//<Pn+10>\pi:D_{1}(n)\to D_{1}(n)//<P^{0}_{n+1}> is the quotient map. That is, πA(1)\pi A(1) has detect property in gr(πA(1))gr(\pi A(1)) with detecting set {πE(1)}\{\pi E(1)\}. Since MM is free on E1=πE(n)E_{1}^{\prime}=\pi E(n) and πE(1)\pi E(1) is a normal Hopf subalgebra of πE(n)\pi E(n), MM is a free πE(1)\pi E(1)-module by Theorem 2.4. It follows that MM is a free graded πA(1)\pi A(1)-module. Now we denote E1=πA(1)E_{1}=\pi A(1).

Consider the element t1=P10P11¯E1t_{1}=\overline{P^{0}_{1}P^{1}_{1}}\in E_{1} and t2=t=2nPt0Pt1¯E2t_{2}=\prod_{t=2}^{n}\overline{P^{0}_{t}P^{1}_{t}}\in E_{2}. By Corollary 2.21 and Lemma 2.24, they are nonzero. Next, we show they satisfy the assumption (2) of Lemma 4.2.

Take z=P20¯E1E2z=\overline{P^{0}_{2}}\in E_{1}\cap E_{2} we know z2=0z^{2}=0. Since |t1|=3|t_{1}|=3 and |P21¯|=6|\overline{P^{1}_{2}}|=6, by Lemma 4.3, yE2+\forall y\in E^{+}_{2}, |y|>|t1||y|>|t_{1}| as long as yzy\neq z. Moreover, by Corollary 2.21 and 2.24, zt10zt_{1}\neq 0.

By definition,

|t1t2|=t=1n3(2t1).|t_{1}t_{2}|=\sum_{t=1}^{n}3\left(2^{t}-1\right).

Notice that the Milnor basis of highest degree in D1(n)D_{1}(n) is that Sq(r1,r2,rn+1)Sq(r_{1},r_{2},\cdots r_{n+1}) with rt=3r_{t}=3 for tnt\leq n, and rn+1=1r_{n+1}=1, one has

|D1(n)|=2n+11+t=1n3(2t1).|D_{1}(n)|=2^{n+1}-1+\sum_{t=1}^{n}3\left(2^{t}-1\right).

Since |Pn+10|=2n+11|P^{0}_{n+1}|=2^{n+1}-1, by Theorem 2.4,

|H|=|D1(n)||<Pn+10>|=t=1n3(2t1).|H|=|D_{1}(n)|-|<P^{0}_{n+1}>|=\sum_{t=1}^{n}3\left(2^{t}-1\right).

Therefore, we have |H|=|t1|+|t2|.|H|=|t_{1}|+|t_{2}|. By Lemma 4.2, MM is a free HH-module, namely, HH has detect property in grHgrH with detecting set {E1,E2}\{E_{1}^{\prime},E_{2}\}. By Theorem 3.14, we are done.

Theorem 4.7.

Given n2n\geq 2, the morphism of groups

f:𝕋(A(n))f:\mathbb{Z}\oplus\mathbb{Z}\to\mathbb{T}(A(n))

which sends (m,l)(m,l) to [σ(m)ΩA(n)l(/2)][\sigma(m)\Omega_{A(n)}^{l}(\mathbb{Z}/2)] is an isomorphism of groups.

Proof: By Proposition 4.1 4.4 4.5 4.6, one gets

Res:𝕋(A(n))L𝕋(L)Res:\mathbb{T}(A(n))\longrightarrow\prod_{L}\mathbb{T}(L)

is an injective group homomorphism, where each LL is elementary and contains Oa+1O_{a+1}. By Proposition 2.8, ResOa+1L:𝕋(L)𝕋(Oa+1)Res^{L}_{O_{a+1}}:\mathbb{T}(L)\to\mathbb{T}(O_{a+1}) is an isomorphism. Compose ResRes with ResOa+1LRes^{L}_{O_{a+1}} on each component of L𝕋(L)\prod_{L}\mathbb{T}(L), one gets an injection

Res:𝕋(A(n))𝕋(Oa+1).Res:\mathbb{T}(A(n))\longrightarrow\mathbb{T}(O_{a+1}).

Recall that σ(m)ΩA(n)l(/2)𝕋(A(n))\sigma(m)\Omega_{A(n)}^{l}(\mathbb{Z}/2)\in\mathbb{T}(A(n)) and it maps to σ(m)ΩOa+1l(/2)\sigma(m)\Omega_{O_{a+1}}^{l}(\mathbb{Z}/2) under the restriction ResOa+1A(n)Res^{A(n)}_{O_{a+1}}, the above injection is in fact an isomorphism.

Recall that the group of endotrivial modules are the same as the stable Picard group, thus one has the following Corollary:

Corollary 4.8.

Given n2n\geq 2, the morphism of groups

f:Pic(A(n))f:\mathbb{Z}\oplus\mathbb{Z}\to Pic(A(n))

which sends (m,l)(m,l) to [σ(m)ΩA(n)l(/2)][\sigma(m)\Omega_{A(n)}^{l}(\mathbb{Z}/2)] is an isomorphism of groups.

References

  • \bibcommenthead
  • Adams and Priddy (1976) Adams, J.F., Priddy, S.: Uniqueness of bsobso. In: Mathematical Proceedings of the Cambridge Philosophical Society, vol. 80, pp. 475–509 (1976). Cambridge University Press
  • Adams and Margolis (1992) Adams, J., Margolis, H.: Modules over the steenrod algebra. The Selected Works of J. Frank Adams: Volume 2 2, 106 (1992)
  • Aguiar and Lauve (2013) Aguiar, M., Lauve, A.: Lagrange’s theorem for hopf monoids in species. Canadian Journal of Mathematics 65(2), 241–265 (2013)
  • Bhattacharya and Ricka (2017) Bhattacharya, P., Ricka, N.: The stable Picard group of A(2)A(2). Preprint at https://arxiv.org/abs/1702.01493 (2017)
  • Carlson and Thévenaz (2000) Carlson, J.F., Thévenaz, J.: Torsion endo-trivial modules. Algebras and representation Theory 3(4), 303–335 (2000)
  • Gheorghe et al. (2018) Gheorghe, B., Isaksen, D.C., Ricka, N.: The picard group of motivic 𝒜(1)\mathcal{A}_{\mathbb{C}}(1). Journal of Homotopy and Related Structures 13(4), 847–865 (2018)
  • Hovey et al. (1997) Hovey, M., Palmieri, J.H., Strickland, N.P.: Axiomatic Stable Homotopy Theory vol. 610. American Mathematical Soc., Washington, DC (1997)
  • Margolis (1983) Margolis, H.R.: Spectra and the Steenrod Algebra. North-Holland Mathematical Library, vol. 29. North-Holland Publishing Co., Amsterdam (1983). Modules over the Steenrod algebra and the stable homotopy category
  • Montgomery (1993) Montgomery, S.: Hopf Algebras and Their Actions on Rings. CBMS Regional Conference Series in Mathematics, vol. 82. American Mathematical Society, Washington, DC (1993)
  • Palmieri (1997) Palmieri, J.H.: A note on the cohomology of finite-dimensional cocommutative hopf algebras. Journal of Algebra 188(1), 203–215 (1997)