Hostname: page-component-78c5997874-lj6df Total loading time: 0 Render date: 2024-11-05T05:41:59.245Z Has data issue: false hasContentIssue false

The mod-p homology of the classifying spaces of certain gauge groups

Published online by Cambridge University Press:  19 September 2022

Daisuke Kishimoto
Affiliation:
Faculty of Mathematics, Kyushu University, Fukuoka, 819-0395, Japan ([email protected])
Stephen Theriault
Affiliation:
Mathematical Sciences, University of Southampton, Southampton SO17 1BJ, United Kingdom ([email protected])
Rights & Permissions [Opens in a new window]

Abstract

Let $G$ be a compact connected simple Lie group of type $(n_{1},\,\ldots,\,n_{l})$, where $n_{1}<\cdots < n_{l}$. Let $\mathcal {G}_k$ be the gauge group of the principal $G$-bundle over $S^{4}$ corresponding to $k\in \pi _3(G)\cong \mathbb {Z}$. We calculate the mod-$p$ homology of the classifying space $B\mathcal {G}_k$ provided that $n_{l}< p-1$.

Type
Research Article
Copyright
Copyright © The Author(s), 2022. Published by Cambridge University Press on behalf of The Royal Society of Edinburgh

1. Introduction

Let $G$ be a Lie group and $P\to X$ be a principal $G$-bundle over a manifold $X$. Automorphisms of $P$ are by definition $G$-equivariant self-maps of $P$ covering the identity map of $X$. The topological group of automorphisms of $P$ is called the gauge group of $P$; we denote this group by $\mathcal {G}(P)$. The classifying space of $\mathcal {G}(P)$ is denoted by $B\mathcal {G}(P)$.

Gauge groups are fundamental in modern physics and geometry. Since the classifying space $B\mathcal {G}(P)$ is homotopy equivalent to the moduli space of connections on $P$ as in [Reference Atiyah and Bott1], the topology of gauge groups over $4$-manifolds and their classifying spaces has proved to be of immense value in studying diffeomorphism structures on $4$-manifolds [Reference Donaldson5], Yang–Mills theory [Reference Atiyah and Jones2], and invariants of $3$-manifolds [Reference Floer6]. Donaldson famously used the rational cohomology of $B\mathcal {G}(P)$ in the case when $G=SU(2)$ and $X$ is a simply-connected $4$-manifold in order to construct polynomial invariants to distinguish diffeomorphism types. Ever since, an important problem has been to calculate the mod-$p$ (co)-homology of $B\mathcal {G}(P)$ when $G=SU(2)$ and $X$ is a simply-connected $4$-manifold for a prime $p$, in the hope of finding new polynomial invariants of diffeomorphism types.

A certain subring of the mod-$2$ cohomology of $B\mathcal {G}(P)$ was studied by Masbaum [Reference Masbaum16] when $G=SU(2)$ and $X$ is a simply-connected closed $4$-manifold, but otherwise nothing else is known. In terms of other Lie groups, Choi [Reference Choi3] has some partial results on the mod-2 homology for $G=Sp(n)$ and $X=S^{4}$. In this paper we make significant progress, completely calculating the mod-$p$ homology of $B\mathcal {G}(P)$ for a family of Lie groups $G$ when $X=S^{4}$. In particular, this includes the pivotal case of $G=SU(2)$ for $p\geq 5$.

To state our results, we need some notation. Let $G$ be a compact connected simple Lie group. Principal $G$-bundles over $S^{4}$ are classified by $\pi _3(G)$, where $\pi _3(G)\cong \mathbb {Z}$ since $G$ is simple. Let $\mathcal {G}_k$ denote the gauge group of a principal $G$-bundle over $S^{4}$ corresponding to $k\in \mathbb {Z}\cong \pi _3(G)$. Let $\operatorname {Map}_{k}(S^{4},\,BG)$ be the component of the space of continuous (not necessarily pointed) maps from $S^{4}$ to $BG$ which are of degree $k$, and similarly define $\operatorname {Map}^{*}_{k}(S^{4},\,BG)$ with respect to pointed maps. There is a fibration

(1.1)\begin{equation} \operatorname{Map}_k^{*}(S^{4},BG)\to \operatorname{Map}_k(S^{4},BG)\xrightarrow{ev}BG \end{equation}

where $ev$ evaluates a map at the basepoint of $S^{4}$. Let $G\langle 3\rangle$ be the three-connected cover of $G$. For each $k\in \mathbb {Z}$, the space $\operatorname {Map}^{*}_{k}(S^{4},\,BG)$ is homotopy equivalent to $\Omega ^{3} G\langle 3\rangle$. By [Reference Atiyah and Bott1, Reference Gottlieb7], there is a homotopy equivalence $B\mathcal {G}_k\simeq \operatorname {Map}_{k}(S^{4},\,BG)$. Thus there is a homotopy fibration

(1.2)\begin{equation} \Omega^{3} G\langle 3\rangle\to B\mathcal{G}_k\xrightarrow{ev}{BG}. \end{equation}

The Lie group $G$ has type $(n_{1},\,\ldots,\,n_{l})$ if the rational cohomology of $G$ is generated by elements in degrees $2n_{1}-1,\,\ldots,\,2n_{l}-1$, where $n_{1}<\cdots < n_{l}$. Unless otherwise indicated, homology is assumed to be with mod-$p$ coefficients.

Theorem 1.1 Let $G$ be a compact connected simple Lie group of type $(n_{1},\,\ldots,\,n_{l})$ and let $p$ be a prime. If $n_{l}< p-1$ then there is an isomorphism of $\mathbb {Z}/p\mathbb {Z}$-vector spaces

\[ H_*(B\mathcal{G}_k)\cong H_*(BG)\otimes H_*(\Omega^{3}G\langle 3\rangle). \]

Under the assumption of theorem 1.1 the Lie group $G$ is $p$-locally homotopy equivalent to the product $\prod _{i=1}^{l} S^{2n_i-1}$, so we can also calculate the Poincaré series of $H_*(B\mathcal {G}_k)$ by theorem 1.1 (corollary 4.5). Remarkably, theorem 1.1 implies that the mod-$p$ homology of $B\mathcal {G}_k$ is independent of $k$ for $p$ large, whereas there is more than one $p$-local homotopy type in the family $\{B\mathcal {G}_k\}_{k\in \mathbb {Z}}$ for $p$ large as was proved in [Reference Kishimoto and Tsutaya12]. The approach to theorem 1.1 is to consider the Serre spectral sequence applied to the fiberwise coproduct of (1.2). Control is obtained over the differentials by showing that the first nontrivial differential is a transgression on a certain element, which is not obvious, and then atomicity-style arguments (cf. [Reference Selick18]) are used to show the spectral sequence must collapse at the $E^{2}$-term.

Of key interest is when $G=SU(2)$. Theorem 1.1 holds if $p\geq 5$ in this case. We also obtain partial results for the prime $3$, which are of a different flavour than those in theorem 1.1. Let $(m,\,n)$ denote the gcd of integers $m$ and $n$. For $(k,\,3)=1$, it is the Serre spectral sequence for the homotopy fibration $SU(2)\to \Omega ^{3}SU(2)\langle 3\rangle \to B\mathcal {G}_k$ induced from (1.2) that collapses at the $E^{2}$-term.

Theorem 1.2 Let $G=SU(2)$ and $p=3$. If $(k,\,3)=1$ then there is an isomorphism of $\mathbb {Z}/3\mathbb {Z}$-vector spaces

\[ H_*(B\mathcal{G}_k)\cong H_*(\Omega^{3} S^{3}\langle 3\rangle)/(x_{3}) \]

where $x_3$ is a generator of $H_3(\Omega ^{3} S^{3}\langle 3\rangle )\cong \mathbb {Z}/3\mathbb {Z}$ and $(x_{3})$ is the ideal generated by $x_{3}$.

The case $(k,\,3)=3$ is still open. The difference of the mod-$p$ homology of $B\mathcal {G}_k$ for $G=SU(2)$ in Theorems 1.1 and 1.2 comes from the homotopy commutativity of $SU(2)$; it is homotopy commutative if and only if $p\ge 5$ as in [Reference McGibbon17]. This is notable because the product decomposition $\mathcal {G}_k\simeq G\times \Omega ^{4}G\langle 3\rangle$ as $A_n$-spaces is guaranteed by the higher homotopy commutativity of $G$ as in [Reference Kishimoto and Kono11, Reference Kishimoto and Tsutaya12], whereas theorem 1.1 shows the homological product decomposition as $A_\infty$-spaces.

2. Serre spectral sequence

Consider a homotopy fibration

(2.1)\begin{equation} F\to E\to B \end{equation}

over a path-connected base $B$ such that $F$ is an H-space and there is a fiberwise action of $F$ on $E$ which restricts to the multiplication of $F$. We assume that $\pi _1(B)$ acts trivially on $H_*(F)$. Let $(E^{r},\,d^{r})$ denote the associated homology Serre spectral sequence.

Lemma 2.1 There is a coalgebra map

\[ \mu\colon E^{r}_{p,q}\otimes H_{q'}(F)\to E^{r}_{p,q+q'} \]

having the following properties.

  1. (1) The map

    \[ \mu\colon H_p(B)\otimes H_q(F)=E^{2}_{p,0}\otimes H_q(F)\to E^{2}_{p,q} \]
    coincides with the canonical isomorphism;
  2. (2) For $x\in E^{r}$ and $y\in H_*(F),$

    \[ d^{r}(\mu(x\otimes y))=\mu(d^{r}(x)\otimes y). \]

Proof. Since the action of $F$ on $E$ is fiberwise, there is a homotopy commutative diagram

Since rows of this diagram are homotopy fibrations, there is a map between associated homology Serre spectral sequences. Since the homology Serre spectral sequences of the top row is isomorphic with $(E^{r}\otimes H_*(F),\,d^{r}\otimes 1)$, we obtain the map $\mu$. By definition, $\mu$ is a coalgebra map. The second statement holds because the differential on $H_*(F)$ in $E^{r}\otimes H_*(F)$ is trivial. Let $p\colon E\to B$ denote the projection. For each $y\in E$, $\pi ^{-1}(\pi (y))$ is homotopy equivalent to the orbit space $y\cdot F$ including $y$. Therefore, by the construction of the Serre spectral sequence, the first statement holds.

Corollary 2.2 If $d^{2}=\cdots =d^{r-1}=0$, then the map $\mu \colon E^{r}_{p,0}\otimes H_q(F)\to E^{r}_{p,q}$ coincides with the canonical isomorphism

\[ E^{r}_{p,q}\cong H_p(B)\otimes H_q(F). \]

Let $\overline {F}\to \overline {E}\to B$ be a homotopy fibration which is a homotopy retract of (2.1). Let $(\overline {E}^{r},\,\overline {d}^{r})$ denote the associated homology Serre spectral sequence.

Lemma 2.3 If $\overline {d}^{2}=\cdots =\overline {d}^{r-1}=0$ and $\overline {d}^{r}\ne 0,$ then the following statements hold:

  1. (1) $\overline {d}^{r}x\ne 0$ for some $x\in H_*(B)=\overline {E}^{r}_{*,0};$

  2. (2) If $y$ is an element of least degree in $H_*(B)=\overline {E}^{r}_{*,0}$ with $\overline {d}^{r}y\ne 0$, then $y$ is transgressive and $\overline {d}^{r}(y)$ is a primitive element of $H_{*-1}(F)=\widehat {E}^{r}_{0,*-1}$.

Proof. Let $i\colon \overline {E}^{2}\to E^{2}$ and $q\colon E^{2}\to \overline {E}^{2}$ denote the inclusion and the retraction, respectively. Since

\[ d^{2}(x)=d^{2}(i_*(x))=i_*(\overline{d}^{2}(x))=0 \]

the hypothesis that $\overline {d}^{2}=0$ implies that $d^{2}=0$. Therefore $\overline {E}^{3}=\overline {E}^{2}$ is a retract of $E^{3}=E^{2}$. Iterating this argument, since $\overline {d}_{3}=\cdots =\overline {d}^{r-1}=0$, we also obtain $d^{2}=\cdots =d^{r-1}=0$ and therefore $\overline {E}^{r}$ is a retract of $E^{r}$. The inclusion and retraction at the $r^{th}$-stage may still be denoted by $i\colon \overline {E}^{2}\to E^{2}$ and $q\colon E^{2}\to \overline {E}^{2}$. Suppose that $\overline {d}^{r}(x)=0$ for all $x\in H_p(B)=\widehat {E}^{r}_{p,0}$. Then

\[ d^{r}(x)=d^{r}(i_*(x))=i_*(\overline{d}^{r}(x))=0, \]

implying $d^{r}=0$ by lemma 2.1 and corollary 2.2. This is a contradiction, so the first statement is proved.

Let $y$ be an element of least degree element in $H_*(B)=\overline {E}^{r}_{*,0}$ such that $\overline {d}^{r}y\ne 0$. Let $\Delta$ denote comultiplication, and set

\[ \Delta(y)=y\otimes 1+1\otimes y+\sum_iy_i'\otimes y_i'' \]

where $|y_i'|<|y|$ and $|y_i''|<|y|$ for each $i$. Since $y$ is an element of least degree in $H_*(B)=E^{r}_{*,0}$ with $d^{r}y\ne 0$, we have $d^{r}y_i'=0$ and $d^{r}y_i''=0$ for each $i$. Therefore

\[ \Delta(d^{r}(y))=d^{r}(\Delta(y))=d^{r}\left(y\otimes 1+1\otimes y+\sum_iy_i'\otimes y_i''\right)=d^{r}(y)\otimes 1+1\otimes d^{r}(y) \]

so $d^{r}(y)$ is primitive.

If $r<|y|$, then by corollary 2.2, $d^{r}(y)=\mu (a\otimes b)$ for some non-trivial $a\in H_{|y|-r}(B)$ and $b\in H_{r-1}(F)$. This is impossible because $\mu (a\otimes b)$ is not primitive by lemma 2.1. Thus $r\ge |y|$. Clearly, $r\le |y|$ since $r>|y|$ implies $d^{r}(y)$ lands in the second quadrant of the spectral sequence, which is zero. Therefore $r=|y|$, implying that $y$ is transgressive. Moreover, since

\[ \overline{d}^{r}(y)=\overline{d}^{r}(q_*(y))=q_*(d^{r}(y)) \]

and $d^{r}y$ is primitive, $\overline {d}^{r}y$ is also primitive. Therefore the second statement is proved.

Next, we consider a family of homotopy fibrations $F_n\to E_n\to B$ with a common base $B$ for $n\in \mathbb {Z}$. Let $(E^{r}_n,\,d^{r}_n)$ denote the associated homology Serre spectral sequence. We can form a homotopy fibration

\[ \coprod_{n\in\mathbb{Z}}F_n\to\coprod_{n\in\mathbb{Z}}E_n\to B. \]

Let $(\widehat {E}^{r},\,\widehat {d}^{r})$ denote the associated homology spectral sequence.

Lemma 2.4 If $(\widehat {E}^{r},\,\widehat {d}^{r})$ collapses at the second term, then so does $(E^{r}_n,\,d^{r}_n)$ for each $n\in \mathbb {Z}$.

Proof. Since there are isomorphisms

\[ (E^{2}_n)_{p,q}\cong H_p(B)\otimes H_q(F_n)\quad\text{and}\quad\widehat{E}^{2}_{p,q}\cong H_p(B)\otimes\left(\bigoplus_{n\in\mathbb{Z}}H_q(F_n)\right), \]

the inclusion $E_n\to E$ induces an injection $E^{2}_n\to \widehat {E}^{2}$. Then the statement is proved by induction on $r$.

3. The mod-$p$ homology of $\Omega ^{3} G\langle 3\rangle$ when $G$ is $p$-regular

Localize at an odd prime $p$ and take homology with mod-$p$ coefficients. If $G$ is $p$-regular of type $(n_{1},\,\ldots,\, n_{l})$ then there is a homotopy equivalence $G\simeq \prod _{i=1}^{l} S^{2n_{i}-1}$. Therefore

(3.1)\begin{equation} \Omega^{3} G\langle 3\rangle\simeq\Omega^{3} S^{3}\langle 3\rangle\times\prod_{i=2}^{l}\Omega^{3} S^{2n_{i}-1}. \end{equation}

This is an equivalence of H-spaces and so induces an isomorphism of Hopf algebras in homology. In this section we record a property of $\Omega ^{3} G\langle 3\rangle$ which will be important later. This begins with a general definition.

In general, for a path-connected space $X$ of finite type, let $PH_*(X)$ be the subspace of primitive elements in $H_*(X)$. Let

\[ \mathcal{P}^{n}_*\colon H_{q}(X)\to H_{q-2(p-1)n}(X) \]

be the dual of the Steenrod operation $\mathcal {P}^{n}$. For $r\geq 1$, let $\beta ^{r}$ be the $r^{\text{th}}$-Bockstein. Let $MH_*(X)$ be the subspace of $PH_*(X)$ defined by

\[ MH_*(X)=\{x\in PH_*(X)\mid\,\mathcal{P}^{n}_*(x)=0\text{ for all }n>0\text{ and }\beta^{r}(x)=0\text{ for all }r\geq 1\}. \]

Since $MH_*(X)$ records information about the primitive elements in $H_*(X)$, if $X\simeq A\times B$ then

\[ MH_*(X)=MH_*(A\times B)=MH_*(A)\oplus MH_*(B). \]

Now consider $MH_*(\Omega ^{3} G\langle 3\rangle )$. The product decomposition (3.1) implies that

(3.2)\begin{equation} MH_*(\Omega^{3}\langle G\rangle)=MH_*(\Omega^{3} S^{3}\langle 3\rangle)\oplus\left(\bigoplus_{i=2}^{l} MH_*(\Omega^{3} S^{2n_{i}-1})\right). \end{equation}

By [Reference Cohen, Lada and May4] there is an isomorphism of Hopf algebras

(3.3)\begin{align} & H_*(\Omega^{3} S^{2n+1})\cong\bigotimes_{k\geq 1, j\geq 0}\Lambda(a_{2(np^{k}-1)p^{j}-1})\nonumber\\ & \quad\otimes\bigotimes_{k\geq 1, j\geq 1}\mathbb{Z}/p\mathbb{Z}[b_{2(np^{k}-1)p^{j}-2}]\otimes\bigotimes_{k\geq 0}\mathbb{Z}/p\mathbb{Z}[c_{2n^{k}-2}] \end{align}

such that $|a_i|=|b_i|=i$. Here, the generators are primitive and many are related by the action of the dual Steenrod algebra. Selick [Reference Selick18] determined $MH_*(\Omega ^{3} S^{2n+1})$ in full, we record only the subset of elements of odd degree.

Lemma 3.1 If $n>1$ then $MH_\text {odd}(\Omega ^{3} S^{2n+1})=\mathbb {Z}/p\mathbb {Z}\{a_{2np-3}\}$.

In the references that follow for $\Omega ^{3} S^{3}\langle 3\rangle$, the statements in [Reference Theriault19] are in terms of Anick spaces, but by [Reference Gray and Theriault8] the space $\Omega S^{3}\langle 3\rangle$ is homotopy equivalent to the Anick space $T^{2p+1}(p)$ for $p\geq 3$. By [Reference Theriault19, proposition 4.1], for $p$ odd there is an isomorphism of Hopf algebras

(3.4)\begin{equation} H_*(\Omega^{3} S^{3}\langle 3\rangle)\cong H_*(\Omega^{2} S^{2p-1})\otimes H_*(\Omega^{3} S^{2p+1}) \end{equation}

which respects the action of the dual Steenrod operations and the Bockstein operations. This can be phrased in terms of a generating set using the isomorphism of Hopf algebras

(3.5)\begin{equation} H_*(\Omega^{2} S^{2p-1})\cong\bigotimes_{k=0}^{\infty}\Lambda(\bar{a}_{2(p-1)p^{k}-1})\otimes\bigotimes_{k=1}^{\infty}\mathbb{Z}/p\mathbb{Z}[\bar{b}_{2(p-1)p^{k}-2}] \end{equation}

proved in [Reference Cohen, Lada and May4], where $|\bar {a}_i|=|\bar {b}_i|=i$, and the $n=p$ case of (3.3). Again, the generators are primitive and many are related by the action of the dual Steenrod algebra. A description of $MH_*(\Omega ^{3}\langle 3\rangle )$ in full was given in [Reference Theriault19, lemma 4.2], but again we need to only record the subset of elements of odd degree.

Lemma 3.2 $MH_\text {odd}(\Omega ^{3} S^{3}\langle 3\rangle )=\mathbb {Z}/p\mathbb {Z}\{\bar {a}_{2p-3}\}$.

Combining (3.2), lemmas 3.1 and 3.2 we obtain the following.

Lemma 3.3 If $G$ is $p$-regular of type $(n_{1},\,\ldots,\,n_{l})$ then

\[ MH_\text{odd}(\Omega^{3} G\langle 3\rangle)=\mathbb{Z}/p\mathbb{Z}\{\bar{a}_{2p-3},a_{2n_{2}p-3},\ldots,a_{2n_{l}p-3}\}. \]

4. The proof of theorem 1.1

We continue to localize at an odd prime $p$ and take homology with mod-$p$ coefficients. Let $G$ be a compact connected simple Lie group of type $(n_1,\,\ldots,\,n_l)$. Consider the homotopy fibration sequence

\[ G\xrightarrow{\partial_k}\Omega^{3}G\langle 3\rangle\xrightarrow{g_k}B\mathcal{G}_k\xrightarrow{ev}BG. \]

First, we show properties of $\partial _k\colon G\to \Omega ^{3}G\langle 3\rangle$.

Lemma 4.1 Suppose that $G$ is $p$-regular. Then $\partial _k$ are null homotopic for all $k$ if and only if $n_l< p-1$.

Proof. Let $\epsilon _i\colon S^{2n_i-1}\to G$ be the inclusion for $i=1,\,\ldots,\,l$, where $G\simeq \prod _{i=1}^{l} S^{2n_i-1}$. By [Reference Lang15] $\partial _k$ corresponds to the Samelson product $\langle k\epsilon _1,\,1_G\rangle$ through the adjoint congruence $[G,\,\Omega ^{3}G\langle 3\rangle ]\cong [S^{3}\wedge G,\,G]$, where $\Omega ^{3}G\langle 3\rangle$ is homotopy equivalent to the component of $\Omega ^{3}G$ containing the basepoint. By the linearity of Samelson products, $\langle k\epsilon _1,\,1_G\rangle =k\langle \epsilon _1,\,1_G\rangle$. Thus we aim to get a condition that guarantees the triviality of $\langle \epsilon _1,\,1_G\rangle$. Arguing as in [Reference Kaji and Kishimoto10], we can see that $\langle \epsilon _1,\,1_G\rangle$ is trivial if and only if $\langle \epsilon _1,\,\epsilon _i\rangle$ is trivial for all $i$. It is shown that $\langle \epsilon _1,\,\epsilon _i\rangle$ is trivial for all $i$ if and only if $n_l< p-1$ in [Reference Kishimoto and Tsutaya13] when $G$ is a classical group and in [Reference Hasui, Kishimoto and Ohsita9] when $G$ is an exceptional group. Thus the proof is complete.

Since $\mathcal {G}_k$ is homotopy equivalent to the homotopy fibre of $\partial _k$, the following is immediate from proposition 4.1.

Corollary 4.2 If $n_l< p-1$ then there is a homotopy equivalence $\mathcal {G}_k\simeq G\times \Omega ^{4}G\langle 3\rangle$.

Next, we consider the map $g_k\colon \Omega ^{3}G\langle 3\rangle \to B\mathcal {G}_k$.

Lemma 4.3 If $n_l< p-1$ then the restriction of $(g_k)_*$ to $MH_\text {odd}(\Omega ^{3} G\langle 3\rangle )$ is an injection.

Proof. First, for $n\geq 2$, consider the homotopy fibration $\Omega ^{4} S^{2n+1}\to \ast \to \Omega ^{3} S^{2n+1}$. We claim that the element $a_{2np-3}\in H_*(\Omega ^{3} S^{2n+1})$ transgresses to a nonzero element in $H_*(\Omega ^{4} S^{2n+1})$. To see this, let $E^{2}\colon S^{2n-1}\to \Omega ^{2} S^{2n+1}$ be the double suspension. Let $W_{n}$ be the homotopy fibre of $E^{2}$. Then there is a homotopy fibration $\Omega S^{2n+1}\xrightarrow {\Omega E^{2}}\Omega ^{3} S^{2n+1}\to W_{n}$. By [Reference Cohen, Lada and May4] this fibration induces an isomorphism of Hopf algebras

\[ H_*(\Omega^{3} S^{2n+1})\cong H_*(\Omega S^{2n+1})\otimes H_*(W_{n}). \]

In particular, since $W_{n}$ is $(2np-4)$-connected, the element $a_{2np-3}\in H_*(\Omega ^{3} S^{2n+1})$ corresponds to an element $c\in H_{2np-3}(W_{n})$. On the other hand, $W_{n}$ has a single cell in dimension $2np-3$, so $c$ represents the inclusion of the bottom cell. Now consider the homotopy fibration diagram

As $c$ represents the bottom cell in $H_{2np-3}(W_{n})$, it transgresses nontrivially to a class $d\in H_{2np-4}(\Omega W_{n})$. Since $a_{2np-3}$ maps to $c$, the naturality of the transgression implies that $a_{2np-3}$ must transgress to a nontrivial class in $H_{2np-4}(\Omega ^{4} S^{2n+1})$.

Next, consider the homotopy fibration diagram

(4.1)

By corollary 4.2, the map $\Omega g_k$ has a left homotopy inverse. In particular, $(\Omega g_k)_{\ast }$ is an injection. Since $G$ is $p$-regular for $n_l< p$, the left column in the fibration diagram is a product of the homotopy fibrations $\Omega ^{4} S^{3}\langle 3\rangle \to \ast \to \Omega ^{3} S^{3}\langle 3\rangle$ and, for $2\leq i\leq n_{l}$, $\Omega ^{4} S^{2n_{i}-1}\to \ast \to \Omega ^{3} S^{2n_{i}-1}$. Therefore, by the argument in the first paragraph of the proof, the element $a_{2n_{i}-1}\in MH_*(\Omega ^{3} G\langle 3\rangle )$ transgresses to a nontrivial element in $H_{2n_{i}p-4}(\Omega ^{4} G\langle 3\rangle )$. As this element injects into $H_{2n_{i}p-4}(\mathcal {G}_k)$, the naturality of the transgression in (4.1) implies that $(g_k)_*(a_{2n_{i}p-3})$ must be nontrivial.

Finally, consider the element $\bar {a}_{2p-3}\in H_*(\Omega ^{3} G\langle 3\rangle )$. It comes from an element $x\in H_{2p-3}(\Omega ^{3} S^{3}\langle 3\rangle )$. The description of $H_*(\Omega ^{3} S^{3}\langle 3\rangle )$ in (3.4) implies that $\Omega ^{3} S^{3}\langle 3\rangle$ is $(2p-4)$-connected and has a single cell in dimension $2p-3$. Thus $x$ represents the inclusion of the bottom cell. Therefore, $x$ transgresses to a nontrivial element in $H_{2p-4}(\Omega ^{4} S^{3}\langle 3\rangle )$, and so $\bar {a}_{2p-3}$ transgresses to a nontrivial element in $H_{2p-4}(\Omega ^{4} G\langle 3\rangle )$. Since $(\Omega g_k)_*$ is an injection, the naturality of the transgression in (4.1) implies that $(g_k)_*(a_{2p-3})$ is nontrivial.

Consider the fibration

(4.2)\begin{equation} \operatorname{Map}^{*}(S^{4},BG)\to\operatorname{Map}(S^{4},BG)\xrightarrow{ev}BG \end{equation}

where $ev$ is the evaluation map at the basepoint of $S^{4}$. Since $G$ is $p$-regular and we are localizing at the prime $p$, $S^{3}$ is a homotopy retract of $G$. Therefore (4.2) is a homotopy retract of the fibration

(4.3)\begin{equation} \operatorname{Map}^{*}(\Sigma G,BG)\to\operatorname{Map}(\Sigma G,BG)\xrightarrow{ev}BG. \end{equation}

We may identify $\operatorname {Map}^{*}(\Sigma G,\,BG)$ with $\operatorname {Map}^{*}(G,\,G)$ by the adjoint congruence. $\operatorname {Map}^{*}(G,\,G)$ is an H-space by the composite of maps, and there is a fiberwise action of $\operatorname {Map}^{*}(G,\,G)$ on $\operatorname {Map}(\Sigma G,\,BG)$ given by

\begin{align*} \operatorname{Map}(\Sigma G,BG)\times\operatorname{Map}^{*}(G,G)\to\operatorname{Map}(\Sigma G,BG),\quad(f,g)\mapsto f\circ\Sigma g. \end{align*}

Thus lemma 2.3 applies to the fibration (4.2). The inclusion of the fibre in (4.2) may be identified with the coproduct of the maps $\operatorname {Map}^{*}_k(S^{4},\,BG)\to \operatorname {Map}_k(S^{4},\,BG)$ for all $k\in \mathbb {Z}$. Equivalently, this is the coproduct of the maps $g_k\colon \Omega ^{3}G\langle 3\rangle \to B\mathcal {G}_k$ for all $k\in \mathbb {Z}$. Each $(g_{k})_{\ast }$ is injective on $MH_{odd}(\operatorname {Map}^{*}(S^{4},\,BG))$ by lemma 4.3, so the coproduct is as well. This leads to the mod-$p$ homology Serre spectral sequence for (4.2) collapsing at the $E^{2}$-term.

Proposition 4.4 Let $G$ be a compact connected simple Lie group of type $(n_1,\,\ldots,\,n_l)$, let $p$ be a prime, and suppose that $n_l< p-1$. Then the mod-$p$ homology Serre spectral sequence for (4.2) is totally nonhomologous to zero.

Proof. Let $(E^{r},\,d^{r})$ denote the mod-$p$ homology Serre spectral sequence for (4.2). Let $z\in H_{m}(\operatorname {Map}^{*}(S^{4},\,BG))$ be an element in the kernel of $g_*$ of least dimension, where $g\colon \operatorname {Map}^{*}(S^{4},\,BG)\to \operatorname {Map}(S^{4},\,BG)$ is the fibre inclusion of (4.2). We begin by establishing some properties of $z$.

Property 1: $z$ is primitive. If not, then $\overline {\Delta }(z)=\Sigma _{\alpha \in A} z'_{\alpha }\otimes z''_{\alpha }$ for some elements $z'_{\alpha },\,z''_{\alpha }$ of degrees $< m$ such that $\{z'_\alpha \otimes z''_\alpha \}_{\alpha \in A}$ is linearly independent, where $\overline {\Delta }$ is the reduced diagonal. The reduced diagonal is natural for any map of spaces, so $(g_*\otimes g_*)\circ \overline {\Delta }=\overline {\Delta }\circ g_*$. Now

\begin{align*} (g_*\otimes g_*)\circ\overline{\Delta}(z)=(g_*\otimes g_*)(\Sigma_{\alpha} z''_{\alpha}\otimes z''_{\alpha})=\Sigma_{\alpha}g_*(z'_{\alpha})\otimes g_*(z''_{\alpha}) \end{align*}

is a sum of linearly independent elements since $z'_{\alpha },\,z''_{\alpha }$ are of degrees $< m$ while the element of least degree in $\mathrm {Ker}\,g_*$ is of degree $m$. On the other hand,

\begin{align*} \overline{\Delta}\circ g_*(z)=\overline{\Delta}(0)=0. \end{align*}

This contradiction implies that $\overline {\Delta }(z)$ must be $0$; that is, $z$ is primitive.

Property 2: $\mathcal {P}^{n}_*(z)=0$ for every $n\ge 1$. Suppose $\mathcal {P}^{n}_{\ast }(z)=y$ for some $n\ge 1$, where $y$ is nonzero. The (dual) Steenrod operations are natural for any map of spaces, so $g_*\mathcal {P}^{n}_*=\mathcal {P}^{n}_*g_*$. Now

\begin{align*} g_*\mathcal{P}^{n}_{{\ast}}(z)=g_*(y) \end{align*}

is nonzero, since $|y|< m$ while the element of least degree in $\mathrm {Ker}\,g_*$ is of degree $m$. On the other hand,

\[ \mathcal{P}^{n}_{{\ast}}g_*(z)=\mathcal{P}^{n}_*(0)=0. \]

This contradiction implies that $\mathcal {P}^{n}_{\ast }(z)=0$ for every $n\ge 1$.

Property 3: $\beta ^{r}(z)=0$ for every $r\geq 1$. The reasoning is exactly as in the proof of property 2.

Property 4: $z$ is in the image of the transgression for the mod- $p$ homology Serre spectral sequence for the homotopy fibration (4.2). By assumption, the first nontrivial differential $d^{r}$ in the mod-$p$ homology Serre spectral sequence for (4.2) satisfies the properties of lemma 2.3. In particular, $d^{r}$ is determined by how it acts on the elements in $H_*(BG)$. The $E^{2}$-term of the spectral sequence is $H_*(BG)\otimes H_*(\operatorname {Map}^{*}(S^{4},\,BG))$. Since $z$ is an element of least degree in the kernel of $(g_k)_*$ and is of degree $m$, the map $(g_k)_*$ is an injection in degrees $< m$. Thus the spectral sequence is totally non-homologous to zero in this degree range, implying that it collapses at $E^{2}$. In particular, the differential $d^{r}$ on an element of $H_*(BG)$ of degree $\leq m$ is zero, and so as $d^{r}$ satisfies the properties of lemma 2.3, $d^{r}(x\otimes y)=0$ for $x\in H_t(BG)$ with $t\leq m$ and $y\in H_*(\operatorname {Map}^{*}(S^{4},\,BG))$. Hence, for degree reasons, if $z$ is in the image of $d^{r}$ then the only possibility is that $z=d^{r}(x)$ and $r=m+1$ for some $x\in H_{m+1}(BG)$. That is, $z$ is in the image of the transgression. On the other hand, as $z\in \mathrm {Ker}\,g_*$, it cannot survive the spectral sequence and so must be hit by some differential.

Property 5: $z$ has odd degree. By property 4, in the mod-$p$ homology Serre spectral sequence for (4.2), we have $z=d^{m+1}(x)$ for some $x\in H_{m+1}(BG)$. As $H_*(BG)$ is concentrated in even degrees, the degree of $z$ must be odd.

Let's examine the consequences of properties 1 to 5. Collectively, Properties 1 to 3 imply that $z\in MH_*(\operatorname {Map}^{*}(S^{4},\,BG))$. Property 5 then implies that $z\in MH_{odd}(\operatorname {Map}^{*}(S^{4},\,BG))$. But $g_*$ is an injection on $MH_{odd}(\operatorname {Map}^{*}(S^{4},\,BG))$ as we saw above, implying in turn that $z\notin \mathrm {Ker}\,g_*$, a contradiction. Thus $g_*$ is an injection, and this completes the proof.

We are ready to prove theorem 1.1.

Proof of theorem 1.1. By proposition 4.4, the mod-$p$ homology Serre spectral sequence for (4.2) collapses at the $E^{2}$-term. Thus, as (4.2) is the fiberwise coproduct of the fibrations $\operatorname {Map}^{*}_k(S^{4},\,BG)\to \operatorname {Map}_k(S^{4},\,BG)\xrightarrow {ev} BG$ in (1.1) for all $k\in \mathbb {Z}$, it follows from lemma 2.4 that the mod-$p$ homology Serre spectral sequence for each fibration collapses at the $E^{2}$-term. Equivalently, the mod-$p$ homology Serre spectral sequence for $\Omega ^{3} G\langle 3\rangle \to B\mathcal {G}_k\xrightarrow {ev} BG$ collapses at the $E^{2}$-term for all $k\in \mathbb {Z}$.

We calculate the Poincaré series of the mod-$p$ homology of $B\mathcal {G}_k$. Let $P_t(X)$ denote the Poincaré series of the mod-$p$ homology of a space $X$. Let

\begin{align*} P(t)& =\prod_{k\ge 0}(1+t^{2(p-1)p^{k}-1})\prod_{k\ge 1}\frac{1}{1-t^{2(p-1)p^{k}-2}}\\ Q_n(t)& =\prod_{k\ge 1,j\ge 0}(1+t^{2(np^{k}-1)p^{j}-1})\prod_{k\ge 1,j\ge 1}\frac{1}{1-t^{2(np^{k}-1)p^{j}-2}}\prod_{k\ge 0}\frac{1}{1-t^{2n^{k}-2}}. \end{align*}

Corollary 4.5 Under the same hypothesis of theorem 1.1, the Poincaré series of the mod-$p$ homology of $B\mathcal {G}_k$ is given by

\[ P_{t}(B\mathcal{G}_k)=P(t)Q_{p-1}(t)\prod_{i=2}^{l} Q_{n_i-1}(t)\prod_{i=1}^{l}\frac{1}{1-t^{2n_i}}. \]

Proof. By theorem 1.1, $P_t(B\mathcal {G}_k)=P_t(BG)P_t(\Omega ^{3}G\langle 3\rangle )$. Since $n_l< p-1$, the mod-$p$ cohomology of $BG$ is a polynomial algebra with generators in dimension $2n_1,\,\ldots,\,2n_l$, and so $P_t(BG)=\prod _{i=1}^{l}\frac {1}{1-t^{2n_i}}$. By (3.1), $P_t(\Omega ^{3}G\langle 3\rangle )=P_t(\Omega ^{3}S^{3}\langle 3\rangle )\prod _{i=2}^{l} P_t(\Omega ^{3}S^{2n_i-1})$. By (3.3), (3.4) and (3.5), $P_t(\Omega ^{3}S^{3}\langle 3\rangle )=P(t)Q_{p-1}(t)$ and $P_t(\Omega ^{3}S^{2n_i-1})=P_{n_i-1}(t)$, completing the proof.

5. The mod-$3$ homology of $SU(2)$-gauge groups

The mod-$p$ homology of $B\mathcal {G}_k$ for $G=SU(2)$ and $p\geq 5$ is given by theorem 1.1. Since $SU(2)$-gauge groups are pivotal in Donaldson Theory, in this section we expand beyond the statement of theorem 1.1 by also calculating the mod-$3$ homology.

Since there is a homeomorphism $SU(2)\cong S^{3}$, we phrase what follows in terms of $S^{3}$. By (1.2) there is a homotopy fibration sequence

\[ S^{3}\xrightarrow{\partial_k}\Omega^{3} S^{3}\langle 3\rangle\xrightarrow{g_k}B\mathcal{G}_k\xrightarrow{ev}BS^{3}. \]

By [Reference Kono14] we have the following.

Lemma 5.1 The map $\partial _1\colon S^{3}\to \Omega ^{3} S^{3}\langle 3\rangle$ has order $12$.

In the proof of lemma 4.1 it is shown that $\partial _k=k\circ \partial _1$. In particular, if we localize at $p=3$ then $\partial _k$ has order $3/(k,\,3)$. Observe that for $p=3$, the description of $H_*(\Omega ^{3} S^{3}\langle 3\rangle )$ in (3.4) implies that $\Omega ^{3} S^{3}\langle 3\rangle$ is $2$-connected with a single cell in dimension $3$.

Lemma 5.2 If $p=3$ and $(k,\,3)=1$ then the map $\partial _k\colon S^{3}\to \Omega ^{3} S^{3}\langle 3\rangle$ induces an injection in mod-$3$ homology.

Proof. By hypothesis, $(k,\,3)=1$ so the above discussion implies that $\partial _{k}$ is nontrivial. Thus it represents a nontrivial element in $\pi _{3}(\Omega ^{3} S^{3}\langle 3\rangle )\cong \pi _{6}(S^{3}\langle 3\rangle )\cong \pi _{6}(S^{3})$. On the other hand, by [Reference Toda20], the $3$-component of $\pi _{6}(S^{3})$ is isomorphic to $\mathbb {Z}/3\mathbb {Z}$. Therefore $\partial _{k}$ represents a generator of the $3$-component of $\pi _{3}(\Omega ^{3} S^{3}\langle 3\rangle )$. Next, since $\Omega ^{3}S^{3}\langle 3\rangle$ is $2$-connected, the Hurewicz map $\pi _3(\Omega ^{3}S^{3}\langle 3\rangle )\otimes \mathbb {Z}/3\mathbb {Z}\to H_*(\Omega ^{3}S^{3}\langle 3\rangle )$ is an isomorphism, where we take homology with mod-$3$ coefficients. By the naturality of Hurewicz maps there is a commutative diagram

where the vertical maps are the Hurewicz maps. We saw that the top and the right maps are isomorphisms. The left arrow is an isomorphism by the Hurewicz theorem. Thus the bottom arrow is an isomorphism, completing the proof.

Next, by (3.4) together with (3.3) and (3.5), $H_{3}(\Omega ^{3} S^{3}\langle 3\rangle )\cong \mathbb {Z}/3\mathbb {Z}$ and if $x_{3}$ is its generator, then $\Lambda (x_{3})$ is a tensor product factor of $H_*(\Omega ^{3} S^{3}\langle 3\rangle )$.

Proof of theorem 1.2 Consider the homotopy fibration $S^{3}\xrightarrow {\partial _k}\Omega ^{3} S^{3}\langle 3\rangle \xrightarrow {g_k}B\mathcal {G}_k$. By lemma 5.2, $(\partial _k)_*$ is an injection. Therefore the homology Serre spectral sequence for this homotopy fibration is totally non-homologous to zero, implying that it collapses at the $E^{2}$-term, so $H_*(\Omega ^{3} S^{3}\langle 3\rangle )\cong H_*(S^{3})\otimes H_*(B\mathcal {G}_k)$. Since $H_*(S^{3})\cong \Lambda (y_{3})$ for $|y_3|=3$ such that $y_3$ corresponds to $x_3$ and as observed above, $\Lambda (x_{3})$ is a tensor product factor of $H_*(\Omega ^{3} S^{3}\langle 3\rangle )$, we obtain the isomorphism asserted in the statement of the theorem.

Acknowledgments

The authors would like to thank the referee for the helpful comments. The first author was partly supported by JSPS KAKENHI (No. 17K05248).

References

Atiyah, M. F. and Bott, R.. The Yang-Mills equations over Riemann surfaces. Philos. Trans. Roy. Soc. London Ser. A 308 (1983), 523615.Google Scholar
Atiyah, M. F. and Jones, J. D. S.. Topological aspects of Yang-Mills theory. Commun. Math. Phys. 61 (1978), 97118.CrossRefGoogle Scholar
Choi, Y.. Homology of the classifying space of $Sp(n)$ gauge groups. Israel J. Math. 151 (2006), 167177.CrossRefGoogle Scholar
Cohen, F. R., Lada, T. J. and May, J. P.. The homology of iterated loop spaces, Lecture Notes in Math. Vol. 533 (Springer-Verlag, 1976).CrossRefGoogle Scholar
Donaldson, S. K.. Connections, cohomology and the intersection forms on $4$-manifolds. J. Differ. Geom. 24 (1986), 275341.Google Scholar
Floer, A.. An instanton invariant for $3$-manifolds. Commun. Math. Phys. 118 (1988), 215240.CrossRefGoogle Scholar
Gottlieb, D. H.. Applications of bundle map theory. Trans. Amer. Math. Soc. 171 (1972), 2350.CrossRefGoogle Scholar
Gray, B. and Theriault, S.. An elementary construction of Anick's fibration. Geom. Topol. 14 (2010), 243276.CrossRefGoogle Scholar
Hasui, S., Kishimoto, D. and Ohsita, A.. Samelson products in $p$-regular exceptional Lie groups. Topology Appl. 178 (2014), 1729.CrossRefGoogle Scholar
Kaji, S. and Kishimoto, D.. Homotopy nilpotency in $p$-regular loop spaces. Math. Z. 264 (2010), 209224.CrossRefGoogle Scholar
Kishimoto, D. and Kono, A.. Splitting of gauge groups. Trans. Amer. Math. Soc. 362 (2010), 67156731.CrossRefGoogle Scholar
Kishimoto, D. and Tsutaya, M.. Infiniteness of $A_\infty$-types of gauge groups. J. Topol. 9 (2016), 181191.CrossRefGoogle Scholar
Kishimoto, D. and Tsutaya, M.. Samelson products in $p$-regular $\mathrm {SO}(2n)$ and its homotopy normality. Glasg. Math. J. 60 (2018), 165174.CrossRefGoogle Scholar
Kono, A.. A note on the homotopy type of certain gauge groups. Proc. Roy. Soc. Edinburgh Sect. A 117 (1991), 295297.CrossRefGoogle Scholar
Lang, G. E.. The evaluation map and $EHP$ sequences. Pacific J. Math. 44 (1973), 201210.CrossRefGoogle Scholar
Masbaum, G.. On the cohomology of the classifying space of the gauge group over some $4$-complexes. Bull. Soc. Math. France 119 (1991), 131.CrossRefGoogle Scholar
McGibbon, C. A.. Homotopy commutativity in localized groups. Amer. J. Math. 106 (1984), 665687.CrossRefGoogle Scholar
Selick, P. S.. A reformulation of the Arf invariant one mod-$p$ problem and applications to atomic spaces. Pacific J. Math. 108 (1983), 431450.CrossRefGoogle Scholar
Theriault, S. D.. Atomicity for Anick's spaces. J. Pure Appl. Alg. 219 (2015), 23462358.CrossRefGoogle Scholar
Toda, H.. Composition methods in homotopy groups of spheres, Annals of Math. Studies No. Vol. 49 (Princeton University Press, Princeton NJ, 1962).Google Scholar