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

Higher homotopy normalities in topological groups

Mitsunobu Tsutaya Faculty of Mathematics, Kyushu University, Fukuoka 819-0395, Japan [email protected]
Abstract.

The purpose of this paper is to introduce Nk()N_{k}(\ell)-maps (1k,1\leq k,\ell\leq\infty), which describe higher homotopy normalities, and to study their basic properties and examples. An Nk()N_{k}(\ell)-map is defined with higher homotopical conditions. It is shown that a homomorphism is an Nk()N_{k}(\ell)-map if and only if there exists fiberwise maps between fiberwise projective spaces with some properties. Also, the homotopy quotient of an Nk(k)N_{k}(k)-map is shown to be an HH-space if its LS category is not greater than kk. As an application, we investigate when the inclusions SU(m)SU(n)\operatorname{SU}(m)\to\operatorname{SU}(n) and SO(2m+1)SO(2n+1)\operatorname{SO}(2m+1)\to\operatorname{SO}(2n+1) are pp-locally Nk()N_{k}(\ell)-maps.

2010 Mathematics Subject Classification:
55P45 (primary), 55R70 (secondary)
The author was supported by JSPS KAKENHI 19K14535.

1. Introduction

A normal subgroup HGH\subset G of a topological group GG is defined to be a subgroup closed under the conjugation by GG. A crossed module is a natural generalization of a normal subgroup, which plays crucial roles in homotopy theory.

Definition 1.1.

A (topological) crossed module is topological groups HH and GG equipped with continuous group homomorphisms

f:HGandρ:GAut(H)f\colon H\to G\quad\text{and}\quad\rho\colon G\to\operatorname{Aut}(H)

satisfying the following conditions:

  1. (1)

    ρ(f(h))(x)=hxh1\rho(f(h))(x)=hxh^{-1} for any h,xHh,x\in H,

  2. (2)

    f(ρ(g)(x))=gf(x)g1f(\rho(g)(x))=gf(x)g^{-1} for any xHx\in H and gGg\in G.

In this paper, we propose and investigate higher homotopy variants of crossed module, which sit between homomorphisms (without any special property) and crossed modules.

Ordinary and higher homotopy normalities have been extensively studied. McCarty [McC64] defined a homotopy normality as follows: a subgroup HGH\subset G is homotopy normal if the conjugation G×HGG\times H\to G, (g,h)ghg1(g,h)\mapsto ghg^{-1} is homotopic to a map into HH through a homotopy of maps between topological pairs (G×H,H×H)(G,H)(G\times H,H\times H)\to(G,H). James [Jam67] defined a weaker homotopy normality as follows: a subgroup HGH\subset G is homotopy normal if the conjugation G×HGG\times H\to G is homotopic to a map into HH through an ordinary homotopy of continuous maps. These conditions are recognized to be “the first order” homotopy normality in our viewpoint. But these homotopy normalities do not guarantee the existence of a multiplicative structure on the quotient G/HG/H.

There have been several works on investigating these kinds of homotopy normalities of Lie groups. The methods adopted so far have been applications of the (relative) Samelson products [McC64, Jam67, Jam71, Jam76, Kac82, Fur85, Fur87, KT18] and the Hopf algebra structures on (generalized) cohomology groups [Fur95, KY98, KY01, KN03, Nis06].

For higher homotopy normality, it has been focused on where a given homomorphism is placed in a homotopy fiber sequence [FS10, Pre12]. In general, a homomorphism f:HGf\colon H\to G induces the homotopy fiber sequence

(1) H𝑓GEH×fGBHBfBG,\displaystyle\cdots\to H\xrightarrow{f}G\to EH\times_{f}G\to BH\xrightarrow{Bf}BG,

where EH×fGEH\times_{f}G is the Borel construction by the action of HH on GG through ff. If we suppose HGH\subset G is a normal subgroup, then the above sequence can be extended as follows:

HGG/HBHBGB(G/H).\cdots\to H\to G\to G/H\to BH\to BG\to B(G/H).

Actually, this is obtained by applying the construction of the homotopy fiber sequence (1) to the homomorphism GG/HG\to G/H. This construction is generalized to crossed modules by Farjoun and Segev [FS10]. Although their construction is for simplicial groups, it also works for topological groups. Conversely, they called f:HGf\colon H\to G a homotopy normal map if the homotopy fiber sequence (1) can be extended as follows by some map BGWBG\to W:

H𝑓GEH×fGBHBfBGW.\displaystyle\cdots\to H\xrightarrow{f}G\to EH\times_{f}G\to BH\xrightarrow{Bf}BG\to W.

In particular, the homomorphism HH\to\ast to the trivial group is homotopy normal in the sense of Farjoun and Segev if and only if BHBH is a loop space.

The main objective of this paper is to introduce another class of higher homotopy normality called Nk()N_{k}(\ell)-maps for 1k,1\leq k,\ell\leq\infty (Definition 6.1). In particular, we see that a homomorphism is an N1(1)N_{1}(1)-map if and only if it is homotopy normal in the sense of McCarty (Remark 6.5). Also, we show that a homomorphism HH\to\ast to the trivial group is an Nk()N_{k}(\ell)-map if and only if HH is a C(k,)C(k,\ell)-space (Theorem 7.3), which is introduced in [KK10]. As a consequence, HH is a C(,)C(\infty,\infty)-space if and only if BHBH is an HH-space. This implies that our homotopy normality is much weaker than that of Farjoun and Segev.

The advantage of our homotopy normality is a good connection with fiberwise homotopy theory. Actually, our main theorem is roughly stated as follows: a homomorphism HGH\to G is an Nk()N_{k}(\ell)-map if and only if the induced fiberwise homomorphism of group bundles EkH×conjHEkG×conjGE_{k}H\times_{\operatorname{conj}}H\to E_{k}G\times_{\operatorname{conj}}G associated to the conjugation action of a group on itself admits a factorization as a fiberwise AA_{\ell}-map

EkH×conjHEEkG×conjGE_{k}H\times_{\operatorname{conj}}H\to E\to E_{k}G\times_{\operatorname{conj}}G

through some fiberwise AA_{\ell}-space EE over BkGB_{k}G with fiber equivalent to HH (Theorem 6.2). The latter condition can be checked by the obstruction theory of fiberwise projective spaces, which provides us a method to study Nk()N_{k}(\ell)-maps (Sections 9 and 10).

We also discuss when the homotopy quotient X=EH×fGX=EH\times_{f}G of a homomorphism f:HGf\colon H\to G is an HH-space. We show that XX is an HH-space if ff is an Nk(k)N_{k}(k)-map and the LS-category catX\operatorname{cat}X of XX is estimated as catXk\operatorname{cat}X\leq k (Theorem 8.2).

Our argument is based on the category of topological monoids and AnA_{n}-maps between them constructed in [Tsu16]. This is just a choice of a model among possible higher categorical settings for AnA_{n}-maps between AA_{\infty}-spaces. Our definitions and results in the present work should be valid in other settings.

This paper is constructed as follows. In Section 2, we recall the category of topological monoids and AnA_{n}-maps between them introduced in [Tsu16]. In Sections 3 and 4, we reformulate AnA_{n}-space and AnA_{n}-map from an AnA_{n}-space to a topological monoid. Our composition of AnA_{n}-maps between topological monoids with an AnA_{n}-map from an AnA_{n}-space to a topological monoid is defined to be associative. In Section 5, we recall the classification theorem of fiberwise AnA_{n}-spaces established in [Tsu12, Tsu15], which plays a central role in the proof of the main theorem. In Section 6, we introduce Nk()N_{k}(\ell)-maps and prove our main result Theorem 6.2. In Section 7, we discuss the relation between C(k,)C(k,\ell)-space and Nk()N_{k}(\ell)-map. In Section 8, we discuss the HH-structure on homotopy quotient. In Section 9, we recall the basic properties of fiberwise projective spaces and give some criterion for non-Nk()N_{k}(\ell)-maps (Theorem 9.5). In Section 10, we investigate when the inclusions SU(m)SU(n)\operatorname{SU}(m)\to\operatorname{SU}(n) and SO(2m+1)SO(2n+1)\operatorname{SO}(2m+1)\to\operatorname{SO}(2n+1) are pp-locally Nk()N_{k}(\ell)-maps.

In this paper, we always work in the category of compactly generated spaces 𝐂𝐆\mathbf{CG}. So the adjunction between products and mapping spaces is always available.

2. Category of topological monoids and AnA_{n}-maps

We begin by recalling the construction of the topologically enriched category of topological monoids and AnA_{n}-maps introduced in [Tsu16, Section 4]. Proofs will be omitted here. Let [0,][0,\infty] be the one point compactification of [0,)[0,\infty), which is homeomorphic to the unit interval [0,1][0,1].

Definition 2.1.

Let f:GGf\colon G\to G^{\prime} be a pointed map between topological monoids. A pair ({fi}i,L)(\{f_{i}\}_{i},L) of a family of maps {fi:[0,]i1×GiG}i=1n\{f_{i}\colon[0,\infty]^{i-1}\times G^{i}\to G^{\prime}\}_{i=1}^{n} and L[0,]L\in[0,\infty] is called an AnA_{n}-form on ff if the following conditions hold:

  1. (1)

    for any xGx\in G, f1(x)=f(x)f_{1}(x)=f(x),

  2. (2)

    for any 2in2\leq i\leq n, 1ki11\leq k\leq i-1, t1,,ti1[0,]t_{1},\ldots,t_{i-1}\in[0,\infty] and x1,,xiGx_{1},\ldots,x_{i}\in G,

    fi(t1,,ti1;x1,,xi)\displaystyle f_{i}(t_{1},\ldots,t_{i-1};x_{1},\ldots,x_{i})
    ={fi1(t1,,tk1,tk+1,,ti1;x1,,xk1,xkxk+1,xk+2,,xi)if tk=0,fk(t1,,tk1;x1,,xk)fik(tk+1,,ti1;xk+1,,xi)if tkL,\displaystyle=\begin{cases}f_{i-1}(t_{1},\ldots,t_{k-1},t_{k+1},\ldots,t_{i-1};x_{1},\ldots,x_{k-1},x_{k}x_{k+1},x_{k+2},\ldots,x_{i})&\text{if $t_{k}=0$},\\ f_{k}(t_{1},\ldots,t_{k-1};x_{1},\ldots,x_{k})f_{i-k}(t_{k+1},\ldots,t_{i-1};x_{k+1},\ldots,x_{i})&\text{if $t_{k}\geq L$},\end{cases}
  3. (3)

    for any 2in2\leq i\leq n, 1ki1\leq k\leq i, t1,,ti1[0,]t_{1},\ldots,t_{i-1}\in[0,\infty] and x1,,xi1Gx_{1},\ldots,x_{i-1}\in G,

    fi(t1,,ti1;x1,,xk1,,xk,,xi1)\displaystyle f_{i}(t_{1},\ldots,t_{i-1};x_{1},\ldots,x_{k-1},\ast,x_{k},\ldots,x_{i-1})
    ={fi1(t2,,ti1;x1,,xi1)if k=1,fi1(t1,,tk2,max{tk1,tk},tk+1,,ti1;x1,,xk1,xk,,xi1)if 1<k<i,fi1(t1,,ti2;x1,,xi1)if k=i.\displaystyle=\begin{cases}f_{i-1}(t_{2},\ldots,t_{i-1};x_{1},\ldots,x_{i-1})&\text{if $k=1$},\\ f_{i-1}(t_{1},\ldots,t_{k-2},\max\{t_{k-1},t_{k}\},t_{k+1},\ldots,t_{i-1};x_{1},\ldots,x_{k-1},x_{k},\ldots,x_{i-1})&\text{if $1<k<i$},\\ f_{i-1}(t_{1},\ldots,t_{i-2};x_{1},\ldots,x_{i-1})&\text{if $k=i$}.\end{cases}

A pair (f,({fi}i,L))(f,(\{f_{i}\}_{i},L)) is called an AnA_{n}-map. The space of AnA_{n}-maps between topological monoids GG and GG^{\prime} is denoted by 𝒜n(G,G)\mathcal{A}_{n}(G,G^{\prime}).

The composition is given as follows.

Definition 2.2.

The composition (h,{hi}i,L+L)=gf(h,\{h_{i}\}_{i},L+L^{\prime})=g\circ f of f=(f,{fi}i,L)𝒜n(G,G)f=(f,\{f_{i}\}_{i},L)\in\mathcal{A}_{n}(G,G^{\prime}) and g=(g,{gi}i,L)𝒜n(G,G′′)g=(g,\{g_{i}\}_{i},L^{\prime})\in\mathcal{A}_{n}(G^{\prime},G^{\prime\prime}) defined as follows: for any r,i1,,ir1r,i_{1},\ldots,i_{r}\geq 1 with i1++irni_{1}+\cdots+i_{r}\leq n, t1,,tr1[0,]t_{1},\ldots,t_{r-1}\in[0,\infty], 𝐬k[0,]ik1\mathbf{s}_{k}\in[0,\infty]^{i_{k}-1}, 𝐱kGik\mathbf{x}_{k}\in G^{i_{k}}, we have

hi1++ir(𝐬1,t1+L,𝐬2,t2+L,,tr1+L,𝐬r;𝐱1,,𝐱k)=gr(t1,,tr1;fi1(𝐬1;𝐱1),,fir(𝐬r;𝐱r)).\displaystyle h_{i_{1}+\cdots+i_{r}}(\mathbf{s}_{1},t_{1}+L,\mathbf{s}_{2},t_{2}+L,\ldots,t_{r-1}+L,\mathbf{s}_{r};\mathbf{x}_{1},\ldots,\mathbf{x}_{k})=g_{r}(t_{1},\ldots,t_{r-1};f_{i_{1}}(\mathbf{s}_{1};\mathbf{x}_{1}),\ldots,f_{i_{r}}(\mathbf{s}_{r};\mathbf{x}_{r})).

A homomorphism f:GHf\colon G\to H can be seen as the AnA_{n}-map equipped with the standard AnA_{n}-form ({fi}i,0)(\{f_{i}\}_{i},0) given as

fi(t1,,ti1;g1,,gi)=f(g1gi)=f(g1)f(gi).f_{i}(t_{1},\ldots,t_{i-1};g_{1},\ldots,g_{i})=f(g_{1}\cdots g_{i})=f(g_{1})\cdots f(g_{i}).

In particular, the identity map idG𝒜n(G,G)\operatorname{id}_{G}\in\mathcal{A}_{n}(G,G) acts as an identity for the above composition.

The following is proved in [Tsu16, Section 4].

Theorem 2.3.

Topological monoids and the spaces of AnA_{n}-maps between them form a topologically enriched category 𝒜n\mathcal{A}_{n}. Moreover, the category of topological monoids and homomorphisms can be embedded in 𝒜n\mathcal{A}_{n} with the standard AnA_{n}-forms.

The categories of left and right actions of topological monoids on topological spaces and AnA_{n}-equivariant maps are similarly defined.

Definition 2.4.

Let GG and GG^{\prime} be topological monoids acting from the left on XX and XX^{\prime}, respectively, and f=(f,({fi}i,L)):GGf=(f,(\{f_{i}\}_{i},L))\colon G\to G^{\prime} be an AnA_{n}-map between topological monoids. A family of maps {ϕi:[0,]i×Gi×XX}i=0n\{\phi_{i}\colon[0,\infty]^{i}\times G^{i}\times X\to X^{\prime}\}_{i=0}^{n} is an AnA_{n}-form on a map ϕ:XX\phi\colon X\to X^{\prime} if the following conditions hold:

  1. (1)

    for any xXx\in X, ϕ0(x)=ϕ(x)\phi_{0}(x)=\phi(x),

  2. (2)

    for any 1in1\leq i\leq n, 1ki1\leq k\leq i, t1,,ti[0,]t_{1},\ldots,t_{i}\in[0,\infty], g1,,giGg_{1},\ldots,g_{i}\in G and xXx\in X,

    ϕi(t1,,ti;g1,,gi;x)\displaystyle\phi_{i}(t_{1},\ldots,t_{i};g_{1},\ldots,g_{i};x)
    ={ϕi1(t1,,tk1,tk+1,,ti;g1,,gk1,gkgk+1,gk+2,,gi;x)if tk=0 and k<i,ϕi1(t1,,ti1;g1,,gi1,gix)if ti=0,fk(t1,,tk1;g1,,gk)ϕik(tk+1,,ti;gk+1,,gi;x)if tkL,\displaystyle=\begin{cases}\phi_{i-1}(t_{1},\ldots,t_{k-1},t_{k+1},\ldots,t_{i};g_{1},\ldots,g_{k-1},g_{k}g_{k+1},g_{k+2},\ldots,g_{i};x)&\text{if $t_{k}=0$ and $k<i$},\\ \phi_{i-1}(t_{1},\ldots,t_{i-1};g_{1},\ldots,g_{i-1},g_{i}x)&\text{if $t_{i}=0$},\\ f_{k}(t_{1},\ldots,t_{k-1};g_{1},\ldots,g_{k})\phi_{i-k}(t_{k+1},\ldots,t_{i};g_{k+1},\ldots,g_{i};x)&\text{if $t_{k}\geq L$},\end{cases}
  3. (3)

    for any 1in1\leq i\leq n, 1ki1\leq k\leq i, t1,,ti[0,]t_{1},\ldots,t_{i}\in[0,\infty], g1,,gi1Gg_{1},\ldots,g_{i-1}\in G and xXx\in X,

    ϕi(t1,,ti;g1,,gk1,,gk,,gi1;x)\displaystyle\phi_{i}(t_{1},\ldots,t_{i};g_{1},\ldots,g_{k-1},\ast,g_{k},\ldots,g_{i-1};x)
    ={ϕi1(t2,,ti;g1,,gi1;x)if k=1,ϕi1(t1,,tk2,max{tk1,tk},tk+1,,ti;g1,,gk1,gk,,gi1;x)if 1<ki.\displaystyle=\begin{cases}\phi_{i-1}(t_{2},\ldots,t_{i};g_{1},\ldots,g_{i-1};x)&\text{if $k=1$},\\ \phi_{i-1}(t_{1},\ldots,t_{k-2},\max\{t_{k-1},t_{k}\},t_{k+1},\ldots,t_{i};g_{1},\ldots,g_{k-1},g_{k},\ldots,g_{i-1};x)&\text{if $1<k\leq i$}.\end{cases}

A triple (f,({fi}i,L),{ϕi}i)(f,(\{f_{i}\}_{i},L),\{\phi_{i}\}_{i}) is called an AnA_{n}-equivariant map. The space of AnA_{n}-equivariant maps between a left GG-space XX and a left GG^{\prime}-space XX^{\prime} is denoted by 𝒜nL((G,X),(G,X))\mathcal{A}^{\mathrm{L}}_{n}((G,X),(G^{\prime},X^{\prime})). Moreover, the topologically enriched category 𝒜nL\mathcal{A}^{\mathrm{L}}_{n} of spaces with left actions of topological monoids and AnA_{n}-equivariant maps is similarly defined. Equivariant AnA_{n}-maps between spaces with right actions of topological monoids are similarly defined. The category and a mapping space in it are denoted by 𝒜nR\mathcal{A}^{\mathrm{R}}_{n} and 𝒜nR((X,G),(X,G))\mathcal{A}^{\mathrm{R}}_{n}((X,G),(X^{\prime},G^{\prime})), respectively.

Our bar construction functor is defined on the category of AnA_{n}-equivariant maps. We take a model of the ii-dimensional simplex Δi\Delta^{i} as

Δi={(t0,,ti)[0,]itk= for some k}.\Delta^{i}=\{(t_{0},\ldots,t_{i})\in[0,\infty]^{i}\mid\text{$t_{k}=\infty$ for some $k$}\}.

The face k:Δi1Δi\partial_{k}\colon\Delta^{i-1}\to\Delta^{i} and degeneracy ϵk:Δi+1Δi\epsilon_{k}\colon\Delta^{i+1}\to\Delta^{i} (k=0,,ik=0,\ldots,i) are given by

k(t0,,ti1)\displaystyle\partial_{k}(t_{0},\ldots,t_{i-1}) =(t0,,tk1,0,tk,,ti1),\displaystyle=(t_{0},\ldots,t_{k-1},0,t_{k},\ldots,t_{i-1}),
ϵk(t0,,ti+1)\displaystyle\epsilon_{k}(t_{0},\ldots,t_{i+1}) =(t0,,tk1,max{tk,tk+1},tk+2,,ti+1).\displaystyle=(t_{0},\ldots,t_{k-1},\max\{t_{k},t_{k+1}\},t_{k+2},\ldots,t_{i+1}).

Consider the fiber product category 𝒜nR×𝒜n𝒜nL\mathcal{A}_{n}^{\mathrm{R}}\times_{\mathcal{A}_{n}}\mathcal{A}_{n}^{\mathrm{L}} in the obvious sense, where an object is a triple (X,G,Y)(X,G,Y) of a topological monoid GG, a right GG-space XX and a left GG-space YY and a morphism is an AnA_{n}-equivariant map between them.

Definition 2.5.

For a triple (X,G,Y)𝒜nR×𝒜n𝒜nL(X,G,Y)\in\mathcal{A}_{n}^{\mathrm{R}}\times_{\mathcal{A}_{n}}\mathcal{A}_{n}^{\mathrm{L}}, the space Bn(X,G,Y)B_{n}(X,G,Y) is defined to be the quotient space

Bn(X,G,Y)=(0inΔi×X×Gi×Y)/B_{n}(X,G,Y)=\left(\coprod_{0\leq i\leq n}\Delta^{i}\times X\times G^{i}\times Y\middle)\right/{\sim}

by the usual simplicial relation. For a morphism

(ϕ,f,ψ)=(ϕ,{ϕi}i,f,{fi}i,ψ,{ψi}i,L):(X,G,Y)(X,G,Y)(\phi,f,\psi)=(\phi,\{\phi_{i}\}_{i},f,\{f_{i}\}_{i},\psi,\{\psi_{i}\}_{i},L)\colon(X,G,Y)\to(X^{\prime},G^{\prime},Y^{\prime})

in 𝒜nR×𝒜n𝒜nL\mathcal{A}_{n}^{\mathrm{R}}\times_{\mathcal{A}_{n}}\mathcal{A}_{n}^{\mathrm{L}}, the induced map

Bn(ϕ,f,ψ):Bn(X,G,Y)Bn(X,G,Y)B_{n}(\phi,f,\psi)\colon B_{n}(X,G,Y)\to B_{n}(X^{\prime},G^{\prime},Y^{\prime})

is defined by

Bn(ϕ,f,ψ)[𝐬1,t1+L,𝐬2,t2+L,,tr1+L,𝐬r;x,𝐠1,,𝐠r,y]\displaystyle B_{n}(\phi,f,\psi)[\mathbf{s}_{1},t_{1}+L,\mathbf{s}_{2},t_{2}+L,\ldots,t_{r-1}+L,\mathbf{s}_{r};x,\mathbf{g}_{1},\ldots,\mathbf{g}_{r},y]
=[t1,,tr1;ϕi1(𝐬1;x,𝐠1),fi2(𝐬2;𝐠2),,fir1(𝐬r1;𝐠r1),ψir(𝐬r;𝐠r,y)].\displaystyle=[t_{1},\ldots,t_{r-1};\phi_{i_{1}}(\mathbf{s}_{1};x,\mathbf{g}_{1}),f_{i_{2}}(\mathbf{s}_{2};\mathbf{g}_{2}),\ldots,f_{i_{r-1}}(\mathbf{s}_{r-1};\mathbf{g}_{r-1}),\psi_{i_{r}}(\mathbf{s}_{r};\mathbf{g}_{r},y)].

for i=i1++irni=i_{1}+\cdots+i_{r}\leq n, t1,,tr1[0,]t_{1},\ldots,t_{r-1}\in[0,\infty], 𝐬k[0,L]ik1\mathbf{s}_{k}\in[0,L]^{i_{k}-1}, xXx\in X, 𝐠kGik\mathbf{g}_{k}\in G^{i_{k}}, yYy\in Y. This construction defines a continuous functor

Bn:𝒜nR×𝒜n𝒜nL𝐂𝐆.B_{n}\colon\mathcal{A}_{n}^{\mathrm{R}}\times_{\mathcal{A}_{n}}\mathcal{A}_{n}^{\mathrm{L}}\to\mathbf{CG}.

In particular, the correspondence GBnG=Bn(,G,)G\mapsto B_{n}G=B_{n}(\ast,G,\ast) defines the nn-th projective space functor

Bn:𝒜n𝐂𝐆.B_{n}\colon\mathcal{A}_{n}\to\mathbf{CG}_{\ast}.

For (X,G,Y)𝒜R×𝒜𝒜L(X,G,Y)\in\mathcal{A}_{\infty}^{\mathrm{R}}\times_{\mathcal{A}_{\infty}}\mathcal{A}_{\infty}^{\mathrm{L}}, let

ιn:Bn(X,G,Y)B(X,G,Y)=B(X,G,Y)\iota_{n}\colon B_{n}(X,G,Y)\to B(X,G,Y)=B_{\infty}(X,G,Y)

denote the natural inclusion.

Note that our bar construction functor coincides with the bar construction for usual equivariant maps through the obvious embedding into the category 𝒜nR×𝒜n𝒜nL\mathcal{A}_{n}^{\mathrm{R}}\times_{\mathcal{A}_{n}}\mathcal{A}_{n}^{\mathrm{L}}.

The following is a technical lemma found in [May75, Theorem 7.6].

Lemma 2.6.

Let GG be a grouplike topological monoid with basepoint having the homotopy extension property and (X,G,Y)𝒜nR×𝒜n𝒜nL(X,G,Y)\in\mathcal{A}_{n}^{\mathrm{R}}\times_{\mathcal{A}_{n}}\mathcal{A}_{n}^{\mathrm{L}}. Then the maps

Bn(X,G,Y)Bn(X,G,)andBn(X,G,Y)Bn(,G,Y)B_{n}(X,G,Y)\to B_{n}(X,G,\ast)\quad\text{and}\quad B_{n}(X,G,Y)\to B_{n}(\ast,G,Y)

are quasifibrations.

The following is the main theorem in [Tsu16].

Theorem 2.7.

Let GG be a topological monoid, of which the underlying space is a CW complex, and GG^{\prime} be a grouplike topological monoid, of which the basepoint has the homotopy extension property. Then the following composite is a weak homotopy equivalence:

𝒜n(G,G)BnMap(BnG,BnG)(ιn)Map(BnG,BG).\mathcal{A}_{n}(G,G^{\prime})\xrightarrow{B_{n}}\operatorname{Map}_{\ast}(B_{n}G,B_{n}G^{\prime})\xrightarrow{(\iota_{n})_{\sharp}}\operatorname{Map}_{\ast}(B_{n}G,BG^{\prime}).

This homotopy equivalence establishes the one-to-one correspondence between the homotopy classes of AnA_{n}-maps GGG\to G^{\prime} and the basepoint-preserving maps BnGBGB_{n}G\to BG^{\prime}. The homotopy classes of an AnA_{n}-map and the corresponding basepoint-preserving map are said to be adjoint to each other. The reason for this is the homotopy equivalence induces the adjunction

π0(𝒜n(G,ΩMX))π0(Map(BnG,X))\pi_{0}(\mathcal{A}_{n}(G,\Omega^{\mathrm{M}}X))\cong\pi_{0}(\operatorname{Map}_{\ast}(B_{n}G,X))

between the projective space functor BnB_{n} and the Moore based loop space functor ΩM\Omega^{\mathrm{M}} in the homotopy categories.

3. Planar rooted trees and associahedra

Let us recall the construction of associahedra in [BV73]. But our construction is slightly different from it since we need to make the composition of AnA_{n}-maps associative.

We set some notions related to planar rooted trees.

  • A rooted tree in our sense is a contractible finite graph with distinguished vertex called the root such that the root is of degree 11 and no vertex is of degree 22.

  • A planar rooted tree is an equivalence class of embeddings of a rooted tree into the upper half plane {(x,y)2y0}\{(x,y)\in\mathbb{R}^{2}\mid y\geq 0\} such that the root is mapped to the origin. Two such embeddings of rooted tree are said to be equivalent if they are isotopic through an isomorphism between rooted trees.

  • In a planar rooted tree, a vertex of degree 11 different from the root is called a leaf. We will write 𝒯n\mathcal{T}_{n} for the set of planar rooted trees with nn leaves. For τ𝒯n\tau\in\mathcal{T}_{n}, we assign numbers 1,,n1,\ldots,n to the leaves of τ\tau from left to right and call the leaf corresponding to kk the kk-th leaf.

  • We will call the edge connected to the root the root edge, an edge connected to a leaf a leaf edge and other edges internal edges. We will write I(τ)I(\tau) for the set of internal edges of a planar rooted tree τ\tau.

  • Along the shortest paths from leaves to the root, an orientation is assigned to each edge in a planar rooted tree.

Consider the space

𝒯n=τ𝒯n{τ}×[0,]I(τ),\mathcal{LT}_{n}=\coprod_{\tau\in\mathcal{T}_{n}}\{\tau\}\times[0,\infty]^{I(\tau)},

where [0,]I(τ)[0,\infty]^{I(\tau)} is the set of maps I(τ)[0,]I(\tau)\to[0,\infty]. An elementary collapse of a planar rooted tree τ\tau at eI(τ)e\in I(\tau) is a planar rooted tree obtained by just collapsing ee to a point. Define the equivalence relation on 𝒯n\mathcal{LT}_{n} generated by (τ,)(τ,)(\tau,\ell)\sim(\tau^{\prime},\ell^{\prime}) such that τ\tau^{\prime} is an elementary collapse of τ\tau at eI(τ)e\in I(\tau), (e)=0\ell(e)=0, and, considering I(τ)I(τ)I(\tau^{\prime})\subset I(\tau), the function \ell restricts to \ell^{\prime} on I(τ)I(\tau^{\prime}). Denote by 𝒦n\mathcal{K}_{n} the quotient space of 𝒯n\mathcal{LT}_{n}, called the nn-th associahedron. In the rest of the paper, an element of 𝒯n\mathcal{LT}_{n} or 𝒦n\mathcal{K}_{n} will be denoted simply by τ\tau. The function I(τ)[0,]I(\tau)\to[0,\infty] will be denoted by \ell for any τ\tau. Such a convention does not cause a confusion.

Refer to caption
Figure 1. The equivalence by an elementary collapse.

Let ρ𝒯r\rho\in\mathcal{T}_{r}, σ𝒯s\sigma\in\mathcal{T}_{s} and 1kr1\leq k\leq r (kk\in\mathbb{Z}). Then the grafting k(ρ,σ)𝒯r+s1\partial_{k}(\rho,\sigma)\in\mathcal{T}_{r+s-1} of σ\sigma to ρ\rho at the kk-th leaf is obtained by identifying the root edge of σ\sigma and the kk-th leaf edge respecting the orientation. Here we note:

I(k(ρ,σ))=I(ρ)I(σ){the new internal edge},I(\partial_{k}(\rho,\sigma))=I(\rho)\sqcup I(\sigma)\sqcup\{\text{the new internal edge}\},

where the new internal edge is the identified edge in the construction. We can extend the grafting construction to

kL:𝒯r×𝒯s𝒯r+s1andkL:𝒦r×𝒦s𝒦r+s1\partial_{k}^{L}\colon\mathcal{LT}_{r}\times\mathcal{LT}_{s}\to\mathcal{LT}_{r+s-1}\quad\text{and}\quad\partial_{k}^{L}\colon\mathcal{K}_{r}\times\mathcal{K}_{s}\to\mathcal{K}_{r+s-1}

for L[0,]L\in[0,\infty] by defining :I(kL(ρ,σ))[0,]\ell\colon I(\partial_{k}^{L}(\rho,\sigma))\to[0,\infty] as

(e)={(e)if eI(ρ)I(σ),Lif e is the new internal edge.\ell(e)=\begin{cases}\ell(e)&\text{if $e\in I(\rho)\sqcup I(\sigma)$},\\ L&\text{if $e$ is the new internal edge}.\end{cases}

When L=L=\infty, we will simply write k=k\partial_{k}=\partial_{k}^{\infty}.

Let τ𝒯n\tau\in\mathcal{T}_{n} (n3n\geq 3) and 1kn1\leq k\leq n. Let us construct the degeneracy sk(τ)𝒯n1s_{k}(\tau)\in\mathcal{T}_{n-1} as follows. Let sk(τ)s_{k}(\tau) be a planar rooted tree obtained from just removing the kk-th leaf edge of τ\tau if the resulting tree does not have a vertex of degree 22 (i.e. the kk-th leaf edge is not connected to a vertex of degree 33). When it has a vertex vv of degree 22, define a planar rooted tree sk(τ)s_{k}(\tau) identifying its edges connected to vv respecting the orientation. To extend the degeneracy to the maps

sk:𝒯n𝒯n1andsk:𝒦n𝒦n1,s_{k}\colon\mathcal{LT}_{n}\to\mathcal{LT}_{n-1}\quad\text{and}\quad s_{k}\colon\mathcal{K}_{n}\to\mathcal{K}_{n-1},

we define (e0)=max{(e0),(e0′′)}\ell(e_{0})=\max\{\ell(e_{0}^{\prime}),\ell(e_{0}^{\prime\prime})\} for the edge e0I(sk(τ))e_{0}\in I(s_{k}(\tau)) obtained by identifying e0,e0′′I(τ)e_{0}^{\prime},e_{0}^{\prime\prime}\in I(\tau) if it exists and is internal. The same value (e)\ell(e) as in τ\tau is assigned for any other internal edge eI(sk(τ))e\in I(s_{k}(\tau)).

Refer to caption
Figure 2. Degeneracy s2s_{2} on 𝒯4\mathcal{LT}_{4}.

4. AnA_{n}-spaces and AnA_{n}-maps

In this section, we recall the definitions of AnA_{n}-spaces and of AnA_{n}-maps from AnA_{n}-spaces to topological monoids.

Definition 4.1.

Let HH be a pointed space. A family of maps {mi:𝒦i×HiH}i=2n\{m_{i}\colon\mathcal{K}_{i}\times H^{i}\to H\}_{i=2}^{n} is called an AnA_{n}-form on HH if the following conditions hold:

  1. (1)

    for any r,s2r,s\geq 2 with r+s1nr+s-1\leq n, 1kr1\leq k\leq r, ρ𝒦r\rho\in\mathcal{K}_{r}, σ𝒦s\sigma\in\mathcal{K}_{s} and x1,,xr+s1Hx_{1},\ldots,x_{r+s-1}\in H,

    mr+s1(k(ρ,σ);x1,,xr+s1)=mr(ρ;x1,,xk1,ms(σ;xk,,xk+s1),xk+s,,xr+s1),\displaystyle m_{r+s-1}(\partial_{k}(\rho,\sigma);x_{1},\ldots,x_{r+s-1})=m_{r}(\rho;x_{1},\ldots,x_{k-1},m_{s}(\sigma;x_{k},\ldots,x_{k+s-1}),x_{k+s},\ldots,x_{r+s-1}),
  2. (2)

    for any xHx\in H,

    m2(;x,)=m2(;,x)=x,m_{2}(\ast;x,\ast)=m_{2}(\ast;\ast,x)=x,
  3. (3)

    for any 3in3\leq i\leq n, 1ki1\leq k\leq i, τ𝒦i\tau\in\mathcal{K}_{i} and x1,,xi1Hx_{1},\ldots,x_{i-1}\in H,

    mi(τ;x1,,xk1,,xk,,xi1)=mi1(sk(τ);x1,,xk1,xk,,xi1).\displaystyle m_{i}(\tau;x_{1},\ldots,x_{k-1},\ast,x_{k},\ldots,x_{i-1})=m_{i-1}(s_{k}(\tau);x_{1},\ldots,x_{k-1},x_{k},\ldots,x_{i-1}).

A pair (H,{mi}i)(H,\{m_{i}\}_{i}) of a pointed space and an AnA_{n}-form on it is called an AnA_{n}-space.

Similarly, we can define an AnA_{n}-form {μi:𝒦i+1×Hi×XX}i=1n\{\mu_{i}\colon\mathcal{K}_{i+1}\times H^{i}\times X\to X\}_{i=1}^{n} of AnA_{n}-actions on XX by HH. Taking the adjoint, this can be considered as an “AnA_{n}-map” from HH to Map(X,X)\operatorname{Map}(X,X). This idea is extended to define an AnA_{n}-map from an AnA_{n}-map to a topological monoid as follows while our definition includes the parameter LL not appearing in the one by Stasheff [Sta70].

Definition 4.2.

Let (H,{mi}i)(H,\{m_{i}\}_{i}) be an AnA_{n}-space, GG a topological monoid and f:HGf\colon H\to G a pointed map. A pair ({fi}i,L)(\{f_{i}\}_{i},L) of a family of maps {fi:𝒦i+1×HiG}i=1n\{f_{i}\colon\mathcal{K}_{i+1}\times H^{i}\to G\}_{i=1}^{n} and L[0,]L\in[0,\infty] is called an AnA_{n}-form on ff if the following conditions hold:

  1. (1)

    for any r1r\geq 1, s2s\geq 2 with r+s1nr+s-1\leq n, 1kr1\leq k\leq r, ρ𝒦r+1\rho\in\mathcal{K}_{r+1}, σ𝒦s\sigma\in\mathcal{K}_{s} and x1,,xr+s1Hx_{1},\ldots,x_{r+s-1}\in H,

    fr+s1(k(ρ,σ);x1,,xr+s1)=fr(ρ;x1,,xk1,ms(σ;xk,,xk+s1),xk+s,,xr+s1),\displaystyle f_{r+s-1}(\partial_{k}(\rho,\sigma);x_{1},\ldots,x_{r+s-1})=f_{r}(\rho;x_{1},\ldots,x_{k-1},m_{s}(\sigma;x_{k},\ldots,x_{k+s-1}),x_{k+s},\ldots,x_{r+s-1}),
  2. (2)

    for any r,s1r,s\geq 1 with r+snr+s\leq n, ρ𝒦r+1\rho\in\mathcal{K}_{r+1}, σ𝒦s+1\sigma\in\mathcal{K}_{s+1}, x1,,xr+sHx_{1},\ldots,x_{r+s}\in H and L\ell\geq L,

    fr+s(r+1(ρ,σ);x1,,xr+s)=fr(ρ;x1,,xr)fs(σ;xr+1,,xr+s),f_{r+s}(\partial_{r+1}^{\ell}(\rho,\sigma);x_{1},\ldots,x_{r+s})=f_{r}(\rho;x_{1},\ldots,x_{r})f_{s}(\sigma;x_{r+1},\ldots,x_{r+s}),
  3. (3)

    for any xHx\in H,

    f1(;x)=f(x),f_{1}(\ast;x)=f(x),
  4. (4)

    for any 2in2\leq i\leq n, 1ki1\leq k\leq i, τ𝒦i+1\tau\in\mathcal{K}_{i+1} and x1,,xi1Hx_{1},\ldots,x_{i-1}\in H,

    fi(τ;x1,,xk1,,xk,,xi1)=fi1(sk(τ);x1,,xk1,xk,,xi1).\displaystyle f_{i}(\tau;x_{1},\ldots,x_{k-1},\ast,x_{k},\ldots,x_{i-1})=f_{i-1}(s_{k}(\tau);x_{1},\ldots,x_{k-1},x_{k},\ldots,x_{i-1}).

A pair (f,({fi}i,L))(f,(\{f_{i}\}_{i},L)) of a pointed map and an AnA_{n}-form on it is called an AnA_{n}-map. The space of AnA_{n}-maps from an AnA_{n}-space HH to a topological monoid GG is denoted by

𝒜n(H,G)(i=1nMap(𝒦i+1×Hi,G))×[0,].\mathcal{A}^{\prime}_{n}(H,G)\subset\left(\prod_{i=1}^{n}\operatorname{Map}(\mathcal{K}_{i+1}\times H^{i},G)\right)\times[0,\infty].

Note that we have two types of AnA_{n}-maps between topological monoids since any topological monoid GG is equipped with the standard AnA_{n}-form {mi}i\{m_{i}\}_{i} given by

mi(τ;x1,,xi)=x1xi.m_{i}(\tau;x_{1},\ldots,x_{i})=x_{1}\cdots x_{i}.

We will see in Proposition 4.5 that 𝒜n(G,G)\mathcal{A}^{\prime}_{n}(G,G^{\prime}) and 𝒜n(G,G)\mathcal{A}_{n}(G,G^{\prime}) are naturally homotopy equivalent. So the two types of AnA_{n}-maps are not essentially different in the homotopical sense.

Now we define the composition between 𝒜n(H,G)\mathcal{A}^{\prime}_{n}(H,G) and 𝒜n(G,G)\mathcal{A}_{n}(G,G^{\prime}). Let τ𝒯n\tau\in\mathcal{LT}_{n} and L[0,]L\in[0,\infty]. Consider the subset I(τ)I(τ)I^{\prime}(\tau)\subset I(\tau) of the internal edges that lie in the shortest path between the nn-th leaf and the root and

IL′′(τ)={eI(τ)(e)L}.I^{\prime\prime}_{L}(\tau)=\{e\in I^{\prime}(\tau)\mid\ell(e)\geq L\}.

Cutting τ\tau at the internal edges in IL′′(τ)I^{\prime\prime}_{L}(\tau), we obtain the planar rooted trees τ1,,τr\tau_{1},\ldots,\tau_{r} when IL′′(τ)=r1\sharp I^{\prime\prime}_{L}(\tau)=r-1 such that τj\tau_{j} is closer to the root than τj\tau_{j^{\prime}} for j<jj<j^{\prime}.

Refer to caption
Figure 3. Cutting τ\tau at the edges in IL′′(τ)I^{\prime\prime}_{L}(\tau) when l3,l4Ll_{3},l_{4}\geq L.
Definition 4.3.

Let (H,{mi}i)(H,\{m_{i}\}_{i}) be an AnA_{n}-space and G,GG,G^{\prime} be topological monoids. The composition (h,({hi}i,L+L))=gf(h,(\{h_{i}\}_{i},L+L^{\prime}))=g\circ f of f=(f,({fi}i,L))𝒜n(H,G)f=(f,(\{f_{i}\}_{i},L))\in\mathcal{A}^{\prime}_{n}(H,G) and g=(g,({gi}i,L))𝒜n(G,G)g=(g,(\{g_{i}\}_{i},L^{\prime}))\in\mathcal{A}_{n}(G,G^{\prime}) is defined as follows: for any 1in1\leq i\leq n, τ𝒯i\tau\in\mathcal{LT}_{i} with τj𝒯ik\tau_{j}\in\mathcal{LT}_{i_{k}} (k=1,,rk=1,\ldots,r) obtained by cutting τ\tau as above and 𝐱kHik\mathbf{x}_{k}\in H^{i_{k}}, we define

hi(τ;𝐱1,,𝐱r)=gr((e1)L,,(er1)L;fi1(τ1;𝐱1),,fir(τr;𝐱r)),\displaystyle h_{i}(\tau;\mathbf{x}_{1},\ldots,\mathbf{x}_{r})=g_{r}(\ell(e_{1})-L,\ldots,\ell(e_{r-1})-L;f_{i_{1}}(\tau_{1};\mathbf{x}_{1}),\ldots,f_{i_{r}}(\tau_{r};\mathbf{x}_{r})),

where IL′′(τ)={e1,,er}I^{\prime\prime}_{L}(\tau)=\{e_{1},\ldots,e_{r}\} and eje_{j} is closer to the root than eje_{j^{\prime}} for j<jj<j^{\prime}.

By a similar argument to that in [Tsu16, Sections 3 and 4], the following theorem holds (see Figure 4).

Theorem 4.4.

Let (H,{mi}i)(H,\{m_{i}\}_{i}) be an AnA_{n}-space and G,G,G′′G,G^{\prime},G^{\prime\prime} be topological monoids. The composition

:𝒜n(G,G)×𝒜n(H,G)𝒜n(H,G)\circ\colon\mathcal{A}_{n}(G,G^{\prime})\times\mathcal{A}^{\prime}_{n}(H,G)\to\mathcal{A}^{\prime}_{n}(H,G^{\prime})

is continuous and the following associativity and unitality hold:

  1. (1)

    for any f𝒜n(H,G)f\in\mathcal{A}^{\prime}_{n}(H,G), g𝒜n(G,G)g\in\mathcal{A}_{n}(G,G^{\prime}) and g𝒜n(G,G′′)g^{\prime}\in\mathcal{A}_{n}(G^{\prime},G^{\prime\prime}), g(gf)=(gg)fg^{\prime}\circ(g\circ f)=(g^{\prime}\circ g)\circ f,

  2. (2)

    for any f𝒜n(H,G)f\in\mathcal{A}^{\prime}_{n}(H,G), idGf=f\operatorname{id}_{G}\circ f=f.

Refer to caption
Figure 4. The associativity of the composition ggfg^{\prime}\circ g\circ f in 𝒦4\mathcal{K}_{4}.

An AnA_{n}-map is said to be weak AnA_{n}-equivalence if the underlying map is a weak homotopy equivalence. Let 𝒜n(H,G)eq𝒜n(H,G)\mathcal{A}^{\prime}_{n}(H,G)_{\operatorname{eq}}\subset\mathcal{A}^{\prime}_{n}(H,G) and 𝒜n(G,G)eq𝒜n(G,G)\mathcal{A}_{n}(G,G^{\prime})_{\operatorname{eq}}\subset\mathcal{A}_{n}(G,G^{\prime}) denote the subspaces of weak AnA_{n}-equivalences. The following proposition can be shown by a proof similar to [Tsu16, Proposition 4.9].

Proposition 4.5.

Let HH be an AnA_{n}-space and G,GG,G^{\prime} be topological monoids. Assume that all of them have the homotopy extension property of the basepoint. The composition with weak AnA_{n}-equivalences f𝒜n(H,G)eqf\in\mathcal{A}_{n}^{\prime}(H,G)_{\operatorname{eq}} and g𝒜n(G,G)eqg\in\mathcal{A}_{n}(G,G^{\prime})_{\operatorname{eq}} induce the weak homotopy equivalences

f:𝒜n(G,G)𝒜n(H,G)andg:𝒜n(H,G)𝒜n(H,G).\displaystyle f^{\sharp}\colon\mathcal{A}_{n}(G,G^{\prime})\xrightarrow{\simeq}\mathcal{A}_{n}^{\prime}(H,G^{\prime})\quad\text{and}\quad g_{\sharp}\colon\mathcal{A}^{\prime}_{n}(H,G)\xrightarrow{\simeq}\mathcal{A}_{n}^{\prime}(H,G^{\prime}).

In particular, the canonical inclusion (idG):𝒜n(G,G)𝒜n(G,G)(\operatorname{id}_{G})^{\sharp}\colon\mathcal{A}_{n}(G,G^{\prime})\to\mathcal{A}_{n}^{\prime}(G,G^{\prime}) defined to be the composite with the identity is a weak homotopy equivalence for topological monoids GG and GG^{\prime}.

5. Fiberwise AnA_{n}-spaces and classification theorem

Let BB be a space. We follow the terminology of Crabb–James [CJ98] as follows. A fiberwise space is just a map π:EB\pi\colon E\to B called the projection and a fiberwise pointed space is a fiberwise space π:EB\pi\colon E\to B equipped with a section σ:BE\sigma\colon B\to E assigning the basepoint to each fiber. A fiberwise maps and a fiberwise pointed map are a map EEE\to E^{\prime} compatible with projections and sections in the obvious sense. Let MapB(E,E)\operatorname{Map}_{B}(E,E^{\prime}) and MapBB(E,E)\operatorname{Map}_{B}^{B}(E,E^{\prime}) denote the space of fiberwise and fiberwise pointed maps, respectively. The fiber product E×BEE\times_{B}E^{\prime} of EE and EE^{\prime} is exactly the categorical product with respect to fiberwise (pointed) maps.

A fiberwise AnA_{n}-space (E,{mi}i)(E,\{m_{i}\}_{i}) over BB is a pair of a fiberwise pointed space EE over BB and a fiberwise AnA_{n}-form {mi:𝒦i×EiE}i\{m_{i}\colon\mathcal{K}_{i}\times E^{i}\to E\}_{i}, where EiE^{i} means the ii-fold fiber product E×B×BEE\times_{B}\cdots\times_{B}E. Here, a fiberwise AnA_{n}-form is defined in the same way as in Section 4. A fiberwise AnA_{n}-map between a fiberwise AnA_{n}-space and a fiberwise topological monoid and between fiberwise topological monoids are also similarly defined.

We take a product-preserving functor 𝒬:𝐂𝐆𝐂𝐆\mathcal{Q}\colon\mathbf{CG}\to\mathbf{CG} as assigning a CW replacement. For example, it is sufficient to take 𝒬X\mathcal{Q}X to be the geometric realization of the singular simplices of a space XX.

Theorem 5.1.

Let EE be a fiberwise topological monoid over a connected pointed CW complex BB such that the projection EBE\to B is a Hurewicz fibration, the section BEB\to E has the homotopy extension property and the fiber over the basepoint is AnA_{n}-equivalent to a topological monoid GG where GG is a CW complex. Then there exists a map BB𝒬𝒜n(G,G)eqopB\to B\mathcal{Q}\mathcal{A}_{n}(G,G)_{\operatorname{eq}}^{\operatorname{op}}, which we will call a classifying map, unique up to homotopy such that the pullback fE~f^{\ast}\tilde{E} of the universal fiberwise AnA_{n}-space E~\tilde{E} is fiberwise AnA_{n}-equivalent to EE.

Remark 5.2.

We take the classifying space as B𝒬𝒜n(G,G)eqopB\mathcal{Q}\mathcal{A}_{n}(G,G)_{\operatorname{eq}}^{\operatorname{op}} rather than the usual one B𝒬𝒜n(G,G)eqB\mathcal{Q}\mathcal{A}_{n}(G,G)_{\operatorname{eq}}. The reason for this is that we defined only the composition of AnA_{n}-maps between topological monoids from the left.

Proof.

In the classification theorem in [Tsu12], fiberwise AnA_{n}-spaces are not assumed to be unital. But, as stated in [Tsu15, Section 7], our theorem is proved by a similar argument except the point that the classifying space there Mn(G)M_{n}(G) coincides with B𝒬𝒜n(G,G)eqopB\mathcal{Q}\mathcal{A}_{n}(G,G)_{\operatorname{eq}}^{\operatorname{op}} up to canonical weak homotopy equivalence. By the same argument as in [Tsu12, Section 5], the space

Prin(E~)op=bMn(G)𝒜n(E~b,G)eq\operatorname{Prin}^{\prime}(\tilde{E})^{\operatorname{op}}=\coprod_{b\in M_{n}(G)}\mathcal{A}^{\prime}_{n}(\tilde{E}_{b},G)_{\operatorname{eq}}

equipped with an appropriate topology is weakly contractible, where E~b\tilde{E}_{b} denotes the fiber over the point bMn(G)b\in M_{n}(G). Moreover, the composition in Theorem 4.4 defines a right action by the grouplike topological monoid 𝒜n(G,G)eqop\mathcal{A}_{n}(G,G)_{\operatorname{eq}}^{\operatorname{op}}, which is the topological monoid 𝒜n(G,G)eq\mathcal{A}_{n}(G,G)_{\operatorname{eq}} equipped with the opposite multiplication. Together with Proposition 4.5, this implies that Prin(E~)op\operatorname{Prin}^{\prime}(\tilde{E})^{\operatorname{op}} is a universal principal fibration for the grouplike topological monoid 𝒜n(G,G)eqop\mathcal{A}_{n}(G,G)_{\operatorname{eq}}^{\operatorname{op}}. Thus the space Mn(G)M_{n}(G) is weakly homotopy equivalent to B𝒬𝒜n(G,G)eqopB\mathcal{Q}\mathcal{A}_{n}(G,G)_{\operatorname{eq}}^{\operatorname{op}}. ∎

As in [Tsu12, Section 2], locally sliceable fiberwise spaces E,EE,E^{\prime} admit the fiberwise mapping space mapB(E,E)\operatorname{map}_{B}(E,E^{\prime}), where mapB(E,)\operatorname{map}_{B}(E,-) is the right adjoint functor to E×BE\times_{B}-. Since it is compatible with pullback by maps of base spaces, one can see that the projection mapB(E,E)\operatorname{map}_{B}(E,E^{\prime}) is a Hurewicz fibration when EBE\to B and EBE^{\prime}\to B are Hurewicz fibrations. Together with it, the fiberwise mapping spaces

𝒜n(E,E)=bB𝒜n(Eb,Eb)\mathscr{A}^{\prime}_{n}(E,E^{\prime})=\coprod_{b\in B}\mathcal{A}^{\prime}_{n}(E_{b},E^{\prime}_{b})

between a fiberwise AnA_{n}-space EE and a fiberwise topological monoid EE^{\prime} over BB is naturally topologized and becomes a fiberwise space over BB. Also, if the projections EBE\to B and EBE^{\prime}\to B are Hurewicz fibrations and the sections BEB\to E and BEB\to E^{\prime} have the homotopy extension properties, then the projection 𝒜n(E,E)B\mathscr{A}^{\prime}_{n}(E,E^{\prime})\to B is a Hurewicz fibration.

The following proposition describes the classifying map of the fiberwise mapping space 𝒜n(E,E)\mathscr{A}^{\prime}_{n}(E,E^{\prime}). Note that when EE^{\prime} is a fiberwise topological monoid with fibers AnA_{n}-equivalent to a topological monoid GG, the classifying map is the one BB𝒬𝒜n(G,G)eqB\to B\mathcal{Q}\mathcal{A}_{n}(G,G)_{\operatorname{eq}} of the principal fibration

Prin(E)=bB𝒜n(G,Eb)eq\operatorname{Prin}(E^{\prime})=\coprod_{b\in B}\mathcal{A}_{n}(G,E^{\prime}_{b})_{\operatorname{eq}}

equipped with the usual right action 𝒜n(G,G)eq\mathcal{A}_{n}(G,G)_{\operatorname{eq}}.

Proposition 5.3.

Let EE be a fiberwise AnA_{n}-space over BB with fibers AnA_{n}-equivalent to a topological monoid HH and EE^{\prime} be a fiberwise topological monoid over BB with fibers AnA_{n}-equivalent to a topological monoid GG. Suppose B,G,HB,G,H are CW complexes and the projections of EE and EE^{\prime} are Hurewicz fibrations and the sections of EE and EE^{\prime} have the homotopy extension property. If EE is classified by a map α:BB𝒬𝒜n(H,H)eqop\alpha\colon B\to B\mathcal{Q}\mathcal{A}_{n}(H,H)_{\operatorname{eq}}^{\operatorname{op}} and EE^{\prime} is classified by a map β:BB𝒬𝒜n(G,G)eq\beta\colon B\to B\mathcal{Q}\mathcal{A}_{n}(G,G)_{\operatorname{eq}}, then the fiberwise mapping space 𝒜n(E,E)\mathscr{A}^{\prime}_{n}(E,E^{\prime}) is classified by the composite

B(α,β)B𝒬𝒜n(H,H)eqop×B𝒬𝒜n(G,G)eqB𝒬ΘB𝒬Map(𝒬𝒜n(H,G),𝒬𝒜n(H,G))eqB\xrightarrow{(\alpha,\beta)}B\mathcal{Q}\mathcal{A}_{n}(H,H)_{\operatorname{eq}}^{\operatorname{op}}\times B\mathcal{Q}\mathcal{A}_{n}(G,G)_{\operatorname{eq}}\xrightarrow{B\mathcal{Q}\Theta}B\mathcal{Q}\operatorname{Map}(\mathcal{Q}\mathcal{A}_{n}(H,G),\mathcal{Q}\mathcal{A}_{n}(H,G))_{\operatorname{eq}}

where Map(X,Y)eqMap(X,Y)\operatorname{Map}(X,Y)_{\operatorname{eq}}\subset\operatorname{Map}(X,Y) denotes the subset of weak equivalences and the homomorphism

Θ:𝒜n(H,H)eqop×𝒜n(G,G)eqMap(𝒜n(H,G),𝒜n(H,G))eq\Theta\colon\mathcal{A}_{n}(H,H)_{\operatorname{eq}}^{\operatorname{op}}\times\mathcal{A}_{n}(G,G)_{\operatorname{eq}}\to\operatorname{Map}(\mathcal{A}_{n}(H,G),\mathcal{A}_{n}(H,G))_{\operatorname{eq}}

is given by Θ(f,g)(h)=ghf\Theta(f,g)(h)=g\circ h\circ f.

Proof.

Let E~B𝒬𝒜n(H,H)eqop\tilde{E}\to B\mathcal{Q}\mathcal{A}_{n}(H,H)_{\operatorname{eq}}^{\operatorname{op}} and E~B𝒬𝒜n(G,G)eq\tilde{E}^{\prime}\to B\mathcal{Q}\mathcal{A}_{n}(G,G)_{\operatorname{eq}} be the universal fiberwise AnA_{n}-spaces. We define 𝒢=𝒜n(H,H)eqop×𝒜n(G,G)eq\mathcal{G}=\mathcal{A}_{n}(H,H)_{\operatorname{eq}}^{\operatorname{op}}\times\mathcal{A}_{n}(G,G)_{\operatorname{eq}} and =B𝒬𝒢=B𝒬𝒜n(H,H)eqop×B𝒬𝒜n(G,G)eq\mathcal{B}=B\mathcal{QG}=B\mathcal{Q}\mathcal{A}_{n}(H,H)_{\operatorname{eq}}^{\operatorname{op}}\times B\mathcal{Q}\mathcal{A}_{n}(G,G)_{\operatorname{eq}} for simplicity. Consider the commutative diagram

𝒜n(E~,E~)\textstyle{\mathscr{A}^{\prime}_{n}(\tilde{E},\tilde{E}^{\prime})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}B(Prin(E~)op×Prin(E~),𝒬𝒢,𝒬𝒜n(H,G))\textstyle{B(\operatorname{Prin}^{\prime}(\tilde{E})^{\operatorname{op}}\times\operatorname{Prin}(\tilde{E}^{\prime}),\mathcal{QG},\mathcal{Q}\mathcal{A}_{n}(H,G))\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}B(,𝒬𝒢,𝒬𝒜n(H,G))\textstyle{B(\ast,\mathcal{QG},\mathcal{Q}\mathcal{A}_{n}(H,G))\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\textstyle{\mathcal{B}}B(Prin(E~)op×Prin(E~),𝒬𝒢,)\textstyle{B(\operatorname{Prin}^{\prime}(\tilde{E})^{\operatorname{op}}\times\operatorname{Prin}(\tilde{E}^{\prime}),\mathcal{QG},\ast)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}B(,𝒬𝒢,)\textstyle{B(\ast,\mathcal{QG},\ast)}

where the left arrows are induced from the compositions of AnA_{n}-maps and the squares are homotopy pullback by Lemma 2.6. Since the vertical maps are quasifibrations and the bottom arrows are weak homotopy equivalences, the top horizontal arrows are also weak homotopy equivalences. Inverting the bottom left arrow, we obtain the composite

=B𝒬𝒢B(Prin(E~)op×Prin(E~),𝒬𝒢,)B𝒬𝒢,\mathcal{B}=B\mathcal{QG}\to B(\operatorname{Prin}^{\prime}(\tilde{E})^{\operatorname{op}}\times\operatorname{Prin}(\tilde{E}^{\prime}),\mathcal{QG},\ast)\to B\mathcal{QG},

which is indeed homotopic to the identity map since it is the classifying map of the universal principal fibration. Consider the homotopy pullback square

B(,𝒬𝒢,𝒬𝒜n(H,G))\textstyle{B(\ast,\mathcal{QG},\mathcal{Q}\mathcal{A}_{n}(H,G))\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}B(,𝒬Map(𝒬𝒜n(H,G),𝒬𝒜n(H,G))eq,𝒬𝒜n(H,G))\textstyle{B(\ast,\mathcal{Q}\operatorname{Map}(\mathcal{Q}\mathcal{A}_{n}(H,G),\mathcal{Q}\mathcal{A}_{n}(H,G))_{\operatorname{eq}},\mathcal{Q}\mathcal{A}_{n}(H,G))\ignorespaces\ignorespaces\ignorespaces\ignorespaces}B𝒬𝒢\textstyle{B\mathcal{QG}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}B𝒬Θ\scriptstyle{B\mathcal{Q}\Theta}B𝒬Map(𝒬𝒜n(H,G),𝒬𝒜n(H,G))eq.\textstyle{B\mathcal{Q}\operatorname{Map}(\mathcal{Q}\mathcal{A}_{n}(H,G),\mathcal{Q}\mathcal{A}_{n}(H,G))_{\operatorname{eq}}.}

This implies that the Hurewicz fibration 𝒜n(E~,E~)\mathscr{A}^{\prime}_{n}(\tilde{E},\tilde{E}^{\prime})\to\mathcal{B} is classified by B𝒬ΘB\mathcal{Q}\Theta. Since 𝒜n(E,E)\mathscr{A}^{\prime}_{n}(E,E^{\prime}) is the pullback of 𝒜n(E~,E~)\mathscr{A}^{\prime}_{n}(\tilde{E},\tilde{E}^{\prime})\to\mathcal{B} by the map (α,β):B(\alpha,\beta)\colon B\to\mathcal{B}, the proposition follows. ∎

6. Nk()N_{k}(\ell)-map

We introduce Nk()N_{k}(\ell)-map mimicking Definition 1.1 of crossed module as follows.

Definition 6.1.

Let f:HGf\colon H\to G be a homomorphism between topological groups and k,k,\ell be positive integers or infinity. We say the homomorphism ff is an Nk()N_{k}(\ell)-map if an AkA_{k}-map ρ:G𝒜(H,H)\rho\colon G\to\mathcal{A}_{\ell}(H,H) is given and the following conditions hold:

  1. (1)

    the composite of AkA_{k}-maps

    H𝑓G𝜌𝒜(H,H)H\xrightarrow{f}G\xrightarrow{\rho}\mathcal{A}_{\ell}(H,H)

    is homotopic to conj:H𝒜(H,H)\operatorname{conj}\colon H\to\mathcal{A}_{\ell}(H,H) (conj(h)(x)=hxh1\operatorname{conj}(h)(x)=hxh^{-1}) as an AkA_{k}-map,

  2. (2)

    the pair of a map

    point𝒜(H,G),f\text{point}\to\mathcal{A}_{\ell}(H,G),\quad\ast\mapsto f

    and the AkA_{k}-map

    θ:GMap(𝒜(H,G),𝒜(H,G))eq,θ(g)(α)=conj(g)αρ(g1).\theta\colon G\to\operatorname{Map}(\mathcal{A}_{\ell}(H,G),\mathcal{A}_{\ell}(H,G))_{\operatorname{eq}},\quad\theta(g)(\alpha)=\operatorname{conj}(g)\circ\alpha\circ\rho(g^{-1}).

    extends to an AkA_{k}-equivariant map (G,)(Map(𝒜(H,G),𝒜(H,G))eq,𝒜(H,G))(G,\ast)\to(\operatorname{Map}(\mathcal{A}_{\ell}(H,G),\mathcal{A}_{\ell}(H,G))_{\operatorname{eq}},\mathcal{A}_{\ell}(H,G)),

  3. (3)

    the composite of the AkA_{k}-form of the previous AkA_{k}-equivariant map with f:(H,)(G,)f\colon(H,\ast)\to(G,\ast) is homotopic to the trivial one coming from the equality conj(f(h))fconj(h1)=f\operatorname{conj}(f(h))\circ f\circ\operatorname{conj}(h^{-1})=f for any hHh\in H.

This definition could be extended to an AnA_{n}-map ff. To do so, it is necessary to determine how good coherence can be guaranteed for the action of HH on Map(𝒜(H,G),𝒜(H,G))eq\operatorname{Map}(\mathcal{A}_{\ell}(H,G),\mathcal{A}_{\ell}(H,G))_{\operatorname{eq}} in the condition (3). We avoid this problem and concentrate on homomorphisms here.

If a homomorphism f:HGf\colon H\to G is an Nk()N_{k}(\ell)-map and kk,k^{\prime}\leq k,\ell^{\prime}\leq\ell, then ff is obviously an Nk()N_{k^{\prime}}(\ell^{\prime})-map.

If f:HGf\colon H\to G is a crossed module, then ff is obviously an N()N_{\infty}(\infty)-map. This is the only case when we can verify the condition of Definition 6.1 directly since it contains very complicated higher homotopical conditions in the present work. The following theorem provides us the way to check the conditions from the obstruction theoretic point of view.

Theorem 6.2.

Let f:HGf\colon H\to G be a homomorphism between topological groups G,HG,H, of which the underlying spaces are CW complexes, and k,k,\ell are non-negative integers. Then the homomorphism ff is an Nk()N_{k}(\ell)-map if and only if there exists a fiberwise AA_{\ell}-space EE over BkGB_{k}G and fiberwise AA_{\ell}-maps ϕ:(Bkf)EEkH×conjH\phi\colon(B_{k}f)^{\ast}E\to E_{k}H\times_{\operatorname{conj}}H over BkHB_{k}H and ψ:EEkG×conjG\psi\colon E\to E_{k}G\times_{\operatorname{conj}}G over BkGB_{k}G satisfying the following conditions:

  1. (1)

    ϕ\phi restricts to a weak AA_{\ell}-equivalence on each fiber,

  2. (2)

    for the restrictions to the fiber over the basepoint ϕ,ψ\phi_{\ast},\psi_{\ast}, the composite fϕ:EGf\circ\phi_{\ast}\colon E_{\ast}\to G, where EE_{\ast} is the fiber of EE over the basepoint, is homotopic to ψ\psi_{\ast} as an AA_{\ell}-map,

  3. (3)

    the composite of ϕ\phi and the induced map EkH×conjHEkH×conjfGE_{k}H\times_{\operatorname{conj}}H\to E_{k}H\times_{\operatorname{conj}\circ f}G of ff is homotopic to (Bkf)ψ(B_{k}f)^{\ast}\psi as a fiberwise AA_{\ell}-map.

Actually, the condition (2) is immediately implied by (3). But we dare to write in this way in order to clarify the correspondence of the conditions between this theorem and Definition 6.1.

Before proving the proof, let us see the following example to illustrate the meaning of the fiberwise AA_{\ell}-space EE in the theorem.

Example 6.3.

Suppose that HGH\subset G is a closed normal subgroup. The fiberwise topological group

E=EG×conjHE=EG\times_{\operatorname{conj}}H

is constructed by the conjugation action of GG on HH. Then we have the obvious factorization

EH×conjHEEG×conjGEH\times_{\operatorname{conj}}H\to E\to EG\times_{\operatorname{conj}}G

of the natural inclusion EH×conjHEG×conjGEH\times_{\operatorname{conj}}H\to EG\times_{\operatorname{conj}}G. Let ϕ:E|BHEH×conjH\phi\colon E|_{BH}\to EH\times_{\operatorname{conj}}H be the isomorphism of fiberwise topological groups over BHBH from EE restricted to the subspace BHBGBH\subset BG and ψ:EEG×conjG\psi\colon E\to EG\times_{\operatorname{conj}}G be the inclusion. This factorization satisfies the conditions in the theorem for k==k=\ell=\infty.

Proof of Theorem 6.2.

Suppose the condition (1) in the theorem. Let ρ0:BkGB𝒬𝒜(H,H)eqop\rho_{0}^{\prime}\colon B_{k}G\to B\mathcal{Q}\mathcal{A}_{\ell}(H,H)_{\operatorname{eq}}^{\operatorname{op}} be the classifying map of Prin(E)op\operatorname{Prin}^{\prime}(E)^{\operatorname{op}} as in Theorem 5.1. By the assumption (1), the composite ρ0Bkf\rho_{0}^{\prime}\circ B_{k}f is homotopic to ιkBk(conj(inversion)):BkHB𝒬𝒜(H,H)eq\iota_{k}\circ B_{k}(\operatorname{conj}\circ\text{(inversion)})\colon B_{k}H\to B\mathcal{Q}\mathcal{A}_{\ell}(H,H)_{\operatorname{eq}}. Let ρ:G𝒜(H,H)eqop\rho\colon G\to\mathcal{A}_{\ell}(H,H)_{\operatorname{eq}}^{\operatorname{op}} be the composite of the inversion and the AkA_{k}-map adjoint to ρ0\rho_{0}^{\prime}. Then the composite ρf\rho\circ f is homotopic to conj:H𝒜(H,H)\operatorname{conj}\colon H\to\mathcal{A}_{\ell}(H,H) as an AkA_{k}-map. This is the condition (1) in Definition 6.1.

Conversely, suppose the condition (1) in Definition 6.1. Let EE be the pullback of the universal fiberwise AnA_{n}-space by the composite

BkGBk(inversion)BkGopBkρBk𝒬𝒜(H,H)eqopιkB𝒬𝒜(H,H)eqop.B_{k}G\xrightarrow{B_{k}\text{(inversion)}}B_{k}G^{\operatorname{op}}\xrightarrow{B_{k}\rho}B_{k}\mathcal{Q}\mathcal{A}_{\ell}(H,H)_{\operatorname{eq}}^{\operatorname{op}}\xrightarrow{\iota_{k}}B\mathcal{Q}\mathcal{A}_{\ell}(H,H)_{\operatorname{eq}}^{\operatorname{op}}.

Then the pullback (Bkf)E(B_{k}f)^{\ast}E is weakly fiberwise AA_{\ell}-equivalent to the fiberwise topological group EkH×conjHE_{k}H\times_{\operatorname{conj}}H. This is the condition (1) in the theorem.

Suppose the condition (1), (2) and (3) in the theorem. The fiberwise AA_{\ell}-map ψ\psi defines a section of the fiberwise mapping space 𝒜(E,EkG×conjG)\mathscr{A}^{\prime}_{\ell}(E,E_{k}G\times_{\operatorname{conj}}G) over BkGB_{k}G. By the map ϕ\phi, we can identify the fiber of 𝒜(E,EkG×conjG)\mathscr{A}_{\ell}(E,E_{k}G\times_{\operatorname{conj}}G) over the base point with 𝒜(H,G)\mathcal{A}_{\ell}(H,G) of which the basepoint is ff. By Proposition 5.3, the fiberwise space 𝒜(E,EkG×conjG)\mathscr{A}^{\prime}_{\ell}(E,E_{k}G\times_{\operatorname{conj}}G) is classified by the composite

BkG(ρ0,ιkBkconj)B𝒬𝒜(H,H)eqop×B𝒬𝒜(G,G)eqB𝒬ΘB𝒬Map(𝒬𝒜(H,G),𝒬𝒜(H,G))eq.B_{k}G\xrightarrow{(\rho_{0}^{\prime},\iota_{k}\circ B_{k}\operatorname{conj})}B\mathcal{Q}\mathcal{A}_{\ell}(H,H)_{\operatorname{eq}}^{\operatorname{op}}\times B\mathcal{Q}\mathcal{A}_{\ell}(G,G)_{\operatorname{eq}}\xrightarrow{B\mathcal{Q}\Theta}B\mathcal{Q}\operatorname{Map}(\mathcal{Q}\mathcal{A}_{\ell}(H,G),\mathcal{Q}\mathcal{A}_{\ell}(H,G))_{\operatorname{eq}}.

The section induced from ψ\psi determines the lift

Ψ:BkGB𝒬Map(𝒬𝒜(H,G),𝒬𝒜(H,G))eq\Psi^{\prime}\colon B_{k}G\to B\mathcal{Q}\operatorname{Map}_{\ast}(\mathcal{Q}\mathcal{A}_{\ell}(H,G),\mathcal{Q}\mathcal{A}_{\ell}(H,G))_{\operatorname{eq}}

of the previous composite along the canonical map

B𝒬Map(𝒬𝒜(H,G),𝒬𝒜(H,G))eqB𝒬Map(𝒬𝒜(H,G),𝒬𝒜(H,G))eq.B\mathcal{Q}\operatorname{Map}_{\ast}(\mathcal{Q}\mathcal{A}_{\ell}(H,G),\mathcal{Q}\mathcal{A}_{\ell}(H,G))_{\operatorname{eq}}\to B\mathcal{Q}\operatorname{Map}(\mathcal{Q}\mathcal{A}_{\ell}(H,G),\mathcal{Q}\mathcal{A}_{\ell}(H,G))_{\operatorname{eq}}.

Let Ψ:GMap(𝒬𝒜(H,G),𝒬𝒜(H,G))eq\Psi\colon G\to\operatorname{Map}_{\ast}(\mathcal{Q}\mathcal{A}_{\ell}(H,G),\mathcal{Q}\mathcal{A}_{\ell}(H,G))_{\operatorname{eq}} be the AkA_{k}-map adjoint to Ψ\Psi^{\prime}. It defines the obvious AkA_{k}-equivariant map

Ψ:(G,)(Map(𝒬𝒜(H,G),𝒬𝒜(H,G))eq,𝒬𝒜(H,G)),\Psi\colon(G,\ast)\to(\operatorname{Map}(\mathcal{Q}\mathcal{A}_{\ell}(H,G),\mathcal{Q}\mathcal{A}_{\ell}(H,G))_{\operatorname{eq}},\mathcal{Q}\mathcal{A}_{\ell}(H,G)),

where the underlying map 𝒬𝒜(H,G)\ast\to\mathcal{Q}\mathcal{A}_{\ell}(H,G) is f\ast\mapsto f. By the argument so far, the underlying AkA_{k}-map of Ψ\Psi is homotopic to the composite of AkA_{k}-maps

G𝜃Map(𝒜(H,G),𝒜(H,G))eq𝒬Map(𝒬𝒜(H,G),𝒬𝒜(H,G))eq.G\xrightarrow{\theta}\operatorname{Map}(\mathcal{A}_{\ell}(H,G),\mathcal{A}_{\ell}(H,G))_{\operatorname{eq}}\xrightarrow{\mathcal{Q}}\operatorname{Map}(\mathcal{Q}\mathcal{A}_{\ell}(H,G),\mathcal{Q}\mathcal{A}_{\ell}(H,G))_{\operatorname{eq}}.

This homotopy determines an AkA_{k}-equivariant map

(G,)(Map(𝒜(H,G),𝒜(H,G))eq,𝒜(H,G))(G,\ast)\to(\operatorname{Map}(\mathcal{A}_{\ell}(H,G),\mathcal{A}_{\ell}(H,G))_{\operatorname{eq}},\mathcal{A}_{\ell}(H,G))

with the underlying AkA_{k}-map θ\theta. To be precise, it is determined up to homotopy. Also, the composite

BkHBkfBkGΨB𝒬Map(𝒬𝒜(H,G),𝒬𝒜(H,G))eqB_{k}H\xrightarrow{B_{k}f}B_{k}G\xrightarrow{\Psi^{\prime}}B\mathcal{Q}\operatorname{Map}_{\ast}(\mathcal{Q}\mathcal{A}_{\ell}(H,G),\mathcal{Q}\mathcal{A}_{\ell}(H,G))_{\operatorname{eq}}

is homotopic to the map adjoint to the AkA_{k}-map (actually a homomorphism)

H𝒬Map(𝒬𝒜(H,G),𝒬𝒜(H,G))eqH\to\mathcal{Q}\operatorname{Map}_{\ast}(\mathcal{Q}\mathcal{A}_{\ell}(H,G),\mathcal{Q}\mathcal{A}_{\ell}(H,G))_{\operatorname{eq}}

induced from the conjugation on HH and the conjugation through ff on GG by the assumption (3). Thus the condition (2) and (3) in Definition 6.1 is verified.

Conversely, suppose the conditions (1), (2) and (3) in Definition 6.1. By Proposition 5.3, BkθB_{k}\theta classifies the fiberwise mapping space 𝒜(E,EkG×conjG)\mathscr{A}^{\prime}_{\ell}(E,E_{k}G\times_{\operatorname{conj}}G). The bar construction of the AkA_{k}-equivariant map

(G,)(Map(𝒜(H,G),𝒜(H,G))eq,𝒜(H,G))(G,\ast)\to(\operatorname{Map}(\mathcal{A}_{\ell}(H,G),\mathcal{A}_{\ell}(H,G))_{\operatorname{eq}},\mathcal{A}_{\ell}(H,G))

defines a map

BkG=Bk(,G,)Bk(,𝒬Map(𝒬𝒜(H,G),𝒬𝒜(H,G))eq,𝒬𝒜(H,G))B_{k}G=B_{k}(\ast,G,\ast)\to B_{k}(\ast,\mathcal{Q}\operatorname{Map}(\mathcal{Q}\mathcal{A}_{\ell}(H,G),\mathcal{Q}\mathcal{A}_{\ell}(H,G))_{\operatorname{eq}},\mathcal{Q}\mathcal{A}_{\ell}(H,G))

of which the composition with the projection onto B𝒬Map(𝒬𝒜(H,G),𝒬𝒜(H,G))eqB\mathcal{Q}\operatorname{Map}(\mathcal{Q}\mathcal{A}_{\ell}(H,G),\mathcal{Q}\mathcal{A}_{\ell}(H,G))_{\operatorname{eq}} is homotopic to BkθB_{k}\theta. Then 𝒜(E,EkG×conjG)\mathscr{A}^{\prime}_{\ell}(E,E_{k}G\times_{\operatorname{conj}}G) admits a section which restricts to ff up to the canonical AkA_{k}-equivalence EHE_{\ast}\simeq H. This verifies the condition (2) in the theorem. Moreover, since the AkA_{k}-equivariant map restricts to the trivial one on HH as supposed, we can verify the condition (3). ∎

Let us consider the special case when k==1k=\ell=1. For pointed spaces A,BA,B with nondegenerate basepoints and a topological group GG, we have the natural split exact sequence

1[AB,G]q[A×B,G]i[AB,G]1,1\to[A\wedge B,G]\xrightarrow{q^{\ast}}[A\times B,G]\xrightarrow{i^{\ast}}[A\vee B,G]\to 1,

where i:ABA×Bi\colon A\vee B\to A\times B is the inclusion and q:A×BABq\colon A\times B\to A\wedge B is the quotient map. The Samelson product f,g[AB,G]\langle f,g\rangle\in[A\wedge B,G] of based maps f:AGf\colon A\to G and g:BGg\colon B\to G is the homotopy class determined by the commutator map

A×BG,(a,b)f(a)g(b)f(a)1g(b)1A\times B\to G,\quad(a,b)\mapsto f(a)g(b)f(a)^{-1}g(b)^{-1}

and the previous split exact sequence.

Theorem 6.4.

Let f:HGf\colon H\to G be a homomorphism between topological groups G,HG,H, of which the underlying spaces are CW complexes. The map ff is an N1(1)N_{1}(1)-map if and only if there exists a pointed map ρ:GHH\rho^{\prime}\colon G\wedge H\to H satisfying the following conditions:

  1. (1)

    the composite ρ(fidH):HHH\rho^{\prime}\circ(f\wedge\operatorname{id}_{H})\colon H\wedge H\to H is homotopic to the Samelson product idH,idH\langle\operatorname{id}_{H},\operatorname{id}_{H}\rangle,

  2. (2)

    the composite fρ:GHGf\circ\rho^{\prime}\colon G\wedge H\to G is homotopic to the Samelson product idG,f\langle\operatorname{id}_{G},f\rangle

  3. (3)

    the composite of above two homotopies is homotopic to the stationary homotopy on f,f\langle f,f\rangle.

HH\textstyle{H\wedge H\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}idH,idH\scriptstyle{\langle\operatorname{id}_{H},\operatorname{id}_{H}\rangle}fidH\scriptstyle{f\wedge\operatorname{id}_{H}}H\textstyle{H\ignorespaces\ignorespaces\ignorespaces\ignorespaces}f\scriptstyle{f}GH\textstyle{G\wedge H\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ρ\scriptstyle{\rho^{\prime}}idG,f\scriptstyle{\langle\operatorname{id}_{G},f\rangle}G\textstyle{G}
Proof.

For given ρ:GHH\rho^{\prime}\colon G\wedge H\to H, define

ρ:GMap(H,H),ρ(g)(h)=ρ(g,h)h.\rho\colon G\to\operatorname{Map}_{\ast}(H,H),\quad\rho(g)(h)=\rho^{\prime}(g,h)h.

Then the condition (1) is equivalent to the one that ρf\rho\circ f is homotopic to conj:HMap(H,H)\operatorname{conj}\colon H\to\operatorname{Map}_{\ast}(H,H). The latter condition is nothing but the condition (1) in Definition 6.1. The condition (2) is equivalent to the one that the map (g,h)f(ρ(g)(h))(g,h)\mapsto f(\rho(g)(h)) is homotopic to the map (g,h)gf(h)g1(g,h)\mapsto gf(h)g^{-1}. The latter condition is equivalent to the condition (2) in Definition 6.1. The equivalence between the condition (3) and the condition (3) in Definition 6.1 follows similarly. ∎

Remark 6.5.

From this result a homomorphism f:HGf\colon H\to G is an N1(1)N_{1}(1)-map if and only if ff is homotopy normal in the sense of McCarty [McC64]. The condition (2) is exactly the homotopy normality of James [Jam67].

7. C(k,)C(k,\ell)-space

A C(k,)C(k,\ell)-space is a topological monoid with certain higher homotopy commutativity introduced in [KK10]. We do not recall the precise definition but instead quote the following characterization [Tsu16, Theorem 8.3], where the equivalence with fourth condition follows from Theorem 5.1.

Theorem 7.1.

Let GG be a topological group, of which the underlying space is a CW complex. Then the following conditions are equivalent:

  1. (1)

    GG is a C(k,)C(k,\ell)-space,

  2. (2)

    the wedge sum of the inclusions (ιk,ι):BkGBGBG(\iota_{k},\iota_{\ell})\colon B_{k}G\vee B_{\ell}G\to BG extends over the product BkG×BGB_{k}G\times B_{\ell}G,

  3. (3)

    the homomorphism conj:𝒜(G,G)\operatorname{conj}\colon\mathcal{A}_{\ell}(G,G) is homotopic to the constant map to the identity as AkA_{k}-map,

  4. (4)

    the fiberwise topological group EkG×conjGE_{k}G\times_{\operatorname{conj}}G over BkGB_{k}G is fiberwise AA_{\ell}-equivalent to the trivial fiberwise topological group BkG×GB_{k}G\times G.

Note that, in particular, GG is a C(,)C(\infty,\infty)-space if and only if BGBG is an HH-space.

The following is an easy application of the previous theorem.

Theorem 7.2.

Let HH and GG be topological groups of which the underlying spaces are CW complexes and f:HGf\colon H\to G be a homomorphism. If HH and GG are C(k,)C(k,\ell)-spaces for some k,nk,\ell\leq n, then ff is an Nk()N_{k}(\ell)-map.

Proof.

By assumption and Theorem 7.1, the fiberwise topological groups EkH×conjHE_{k}H\times_{\operatorname{conj}}H and EkG×conjGE_{k}G\times_{\operatorname{conj}}G are fiberwise AA_{\ell}-equivalent to the trivial bundles. Consider the trivial fiberwise topological group E=BkG×HE=B_{k}G\times H. Then we can verify the conditions in Theorem 6.2. Thus ff is an Nk()N_{k}(\ell)-map. ∎

For the homomorphism HH\to\ast to the trivial group, we obtain the following.

Theorem 7.3.

Let HH be a topological group of which the underlying space is a CW complex. The homomorphism HH\to\ast to the trivial group is an Nk()N_{k}(\ell)-map if and only if HH is a C(k,)C(k,\ell)-space.

Proof.

Note that Bk=B_{k}\ast=\ast for any kk. Then by Theorem 6.2, the homomorphism HH\to\ast is an Nk()N_{k}(\ell)-map if and only if the fiberwise topological monoid EkH×conjHE_{k}H\times_{\operatorname{conj}}H is fiberwise AA_{\ell}-equivalent to the trivial fiberwise topological group BkH×HB_{k}H\times H. Thus the theorem follows from Theorem 7.1. ∎

8. HH-structure on quotient space

Since the homotopy quotient of a crossed module is known to be a topological monoid, we expect a similar result for Nk()N_{k}(\ell)-maps. Let us investigate the existence of an HH-structure on the homotopy quotient.

Proposition 8.1.

Let f:HGf\colon H\to G be a homomorphism between topological groups of which the underlying spaces are CW complexes. Suppose ff is an Nk()N_{k}(\ell)-map. Then, the topological group FF with classifying space BFB(,H,G)=EH×fGBF\simeq B(\ast,H,G)=EH\times_{f}G is a C(k,)C(k,\ell)-space.

Note that the topological group FF is AA_{\infty}-equivalent to the Moore based loop space of BFBF. This is a key proposition in our study of HH-structures on EH×fGEH\times_{f}G.

Proof.

We may suppose that ff is a Hurewicz fibration and F=kerfF=\ker f. Let i:FHi\colon F\to H denote the inclusion. Note that the results in [Tsu15, Sections 3 and 4] for AnA_{n}-spaces are similarly verified for fiberwise AnA_{n}-spaces. By [Tsu15, Proposition 4.1], we may suppose that there exists a fiberwise AA_{\ell}-space EE over BkGB_{k}G, a fiberwise AA_{\ell}-equivalence ϕ:(Bkf)EEkH×conjH\phi\colon(B_{k}f)^{\ast}E\to E_{k}H\times_{\operatorname{conj}}H and a fiberwise AA_{\ell}-homomorphism ψ:EEkG×conjG\psi\colon E\to E_{k}G\times_{\operatorname{conj}}G as in Theorem 6.2. Let EE^{\prime} be the fiberwise homotopy fiber of ψ\psi over BkGB_{k}G. Then by [Tsu15, Theorem 3.1] (the pullback of AnA_{n}-homomorphisms), EE^{\prime} admits the canonical structure of a fiberwise AA_{\ell}-space. Applying [Tsu15, Theorem 3.3] (the universal property of the pullback of AnA_{n}-homomorphisms), a fiberwise AA_{\ell}-map χ:(Bkf)EEkH×HF\chi\colon(B_{k}f)^{\ast}E^{\prime}\to E_{k}H\times_{H}F is induced as in the following diagram:

(Bkf)E\textstyle{(B_{k}f)^{\ast}E^{\prime}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}χ\scriptstyle{\chi}(Bkf)E\textstyle{(B_{k}f)^{\ast}E\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ψBkH\scriptstyle{\psi_{B_{k}H}}ϕ\scriptstyle{\phi}EkH×conjfG\textstyle{E_{k}H\times_{\operatorname{conj}\circ f}G\ignorespaces\ignorespaces\ignorespaces\ignorespaces}EkH×HF\textstyle{E_{k}H\times_{H}F\ignorespaces\ignorespaces\ignorespaces\ignorespaces}EkH×conjH\textstyle{E_{k}H\times_{\operatorname{conj}}H\ignorespaces\ignorespaces\ignorespaces\ignorespaces}EkH×conjfG\textstyle{E_{k}H\times_{\operatorname{conj}\circ f}G}

We can see that χ\chi is a fiberwise AA_{\ell}-equivalence. Pulling buck along the map Bki:BkFBkHB_{k}i\colon B_{k}F\to B_{k}H, we obtain the fiberwise AA_{\ell}-equivalence

(Bki)χ:(Bk(if))EEkF×conjF.(B_{k}i)^{\ast}\chi\colon(B_{k}(i\circ f))^{\ast}E^{\prime}\xrightarrow{\simeq}E_{k}F\times_{\operatorname{conj}}F.

Since ifi\circ f is the constant homomorphism, EkF×HFE_{k}F\times_{H}F is fiberwise AA_{\ell}-equivalent to the trivial fiberwise topological group BkF×FB_{k}F\times F. Thus, it follows from Theorem 7.1 that FF is a C(k,)C(k,\ell)-space. ∎

Theorem 8.2.

Let f:HGf\colon H\to G be a homomorphism between topological groups of which the underlying spaces are CW complexes and k1k\geq 1. Suppose that ff is an Nk(k)N_{k}(k)-map and the LS category of the homotopy quotient EH×fGEH\times_{f}G is estimated as cat(EH×fG)k\operatorname{cat}(EH\times_{f}G)\leq k. Then EH×fGEH\times_{f}G is an HH-space.

Proof.

Take FF as in the proof of Proposition 8.1 so that the homotopy equivalence BFEH×fGBF\simeq EH\times_{f}G and the estimate catBFk\operatorname{cat}BF\leq k hold. Then, as in [Iwa98], the inclusion BkFBFB_{k}F\to BF admits a homotopy section s:BFBkFs\colon BF\to B_{k}F. Since FF is a C(k,k)C(k,k)-space, it follows from the existence of ss and Theorem 7.1 that BFBF is an HH-space. ∎

Remark 8.3.

As a consequence of Theorem 7.3, one cannot expect that Nk()N_{k}(\ell)-map implies any higher homotopy associativity of the homotopy quotient HH-space. In order to guarantee higher homotopy associativity, we need even higher homotopy normality.

Remark 8.4.

Although Theorem 8.2 seems theoretically important, we do not have its good application at this point. For example, we shall give a sufficient condition in Section 10 below that the inclusion SU(m)SU(n)\operatorname{SU}(m)\to\operatorname{SU}(n) is pp-locally an Nk()N_{k}(\ell)-map. But it will not give a new result on the existence of an HH-structure since the quotient SU(n)/SU(m)\operatorname{SU}(n)/\operatorname{SU}(m) is pp-locally homotopy equivalent to a product of odd dimensional spheres under the assumption there.

We have only shown the existence of an HH-structure on the homotopy quotient in the previous theorem. We do not try to answer the following natural question in the present work.

Problem 8.5.

Suppose a homomorphism f:HGf\colon H\to G is an Nk(k)N_{k}(k)-map and cat(EH×fG)k\operatorname{cat}(EH\times_{f}G)\leq k. Construct a canonical HH-structure on EH×fGEH\times_{f}G. If it can be done, is the natural map GEH×fGG\to EH\times_{f}G an HH-map?

9. Fiberwise projective spaces

Let us recall the fiberwise projective spaces of topological monoids.

Definition 9.1.

The fiberwise nn-th projective space nE\mathscr{B}_{n}E of a fiberwise topological monoid EBE\to B is defined to be the quotient

nE=(0inΔi×Ei)/\mathscr{B}_{n}E=\left(\coprod_{0\leq i\leq n}\Delta^{i}\times E^{i}\middle)\right/{\sim}

by the usual simplicial relation, where EiE^{i} denotes the ii-fold fiber product. In particular, E=E\mathscr{B}E=\mathscr{B}_{\infty}E is called the fiberwise classifying space. Let ιn:nEE\iota_{n}\colon\mathscr{B}_{n}E\to\mathscr{B}E denote the canonical inclusion. As in Sections 2, the fiberwise projective space induce the fiberwise map

n:𝒜n(E,E)mapBB(nE,nE)\mathscr{B}_{n}\colon\mathscr{A}_{n}(E,E^{\prime})\to\operatorname{map}_{B}^{B}(\mathscr{B}_{n}E,\mathscr{B}_{n}E^{\prime})

to the fiberwise mapping space of fiberwise pointed maps for fiberwise topological monoids EE and EE^{\prime}.

The fiberwise nn-th projective space for a fiberwise AnA_{n}-space is similarly constructed as in [Sak10].

Definition 9.2.

The fiberwise nn-th projective space n𝒦E\mathscr{B}^{\mathcal{K}}_{n}E of a fiberwise AnA_{n}-space EBE\to B is the quotient

n𝒦E=(0in𝒦i+2×Ei)/\mathscr{B}^{\mathcal{K}}_{n}E=\left(\coprod_{0\leq i\leq n}\mathcal{K}_{i+2}\times E^{i}\middle)\right/{\sim}

constructed as in [Sta63, Construction 8].

The following is a variant of Theorem 2.7 in the category of fiberwise spaces.

Theorem 9.3.

Let EE and EE^{\prime} be fiberwise topological monoids over a CW complex BB and suppose that their fiberwise basepoints of EE and EE^{\prime} have the fiberwise homotopy extension property, their projections are Hurewicz fibrations and EE^{\prime} is grouplike on each fiber. Then, the composite of fiberwise maps

𝒜n(E,E)nmapBB(nE,nE)(ιn)mapBB(nE,E)\mathscr{A}_{n}(E,E^{\prime})\xrightarrow{\mathscr{B}_{n}}\operatorname{map}_{B}^{B}(\mathscr{B}_{n}E,\mathscr{B}_{n}E^{\prime})\xrightarrow{(\iota_{n})_{\sharp}}\operatorname{map}_{B}^{B}(\mathscr{B}_{n}E,\mathscr{B}E^{\prime})

is a weak homotopy equivalence in the category of spaces.

Proof.

In this setting, the projections 𝒜n(E,E)B\mathscr{A}_{n}(E,E^{\prime})\to B and mapBB(nE,E)B\operatorname{map}_{B}^{B}(\mathscr{B}_{n}E,\mathscr{B}E^{\prime})\to B are Hurewicz fibrations. Thus the given composite is a weak homotopy equivalence since it restricts to a weak homotopy equivalence on each fiber by Theorem 2.7. ∎

Although the following result is well-known, we give a proof here.

Proposition 9.4.

Let GG be a topological group. Then, the fiberwise classifying space EG×GBGEG\times_{G}BG is homeomorphic to BG×BGBG\times BG through a fiberwise pointed map over BGBG, where the fiberwise basepoint of BG×BGBG\times BG is given by the diagonal map BGBG×BGBG\to BG\times BG.

Proof.

The homeomorphism EG×GBGBG×BGEG\times_{G}BG\to BG\times BG is induced from the geometric realization of the simplicial map {Gi×GiGi×Gi}i\{G^{i}\times G^{i}\to G^{i}\times G^{i}\}_{i} given by the homeomorphisms

(g1,g2,,gi;x1,x2,,xi)\displaystyle(g_{1},g_{2},\ldots,g_{i};x_{1},x_{2},\ldots,x_{i})
(g1,g2,,gi;(g1gi)x1(g2gi)1,(g2gi)x2(g3gi)1,,gixi),\displaystyle\mapsto(g_{1},g_{2},\ldots,g_{i};(g_{1}\cdots g_{i})x_{1}(g_{2}\cdots g_{i})^{-1},(g_{2}\cdots g_{i})x_{2}(g_{3}\cdots g_{i})^{-1},\ldots,g_{i}x_{i}),

where EG×GBGEG\times_{G}BG and BG×BGBG\times BG are cosidered to be the geometric realizations of appropriate simplicial spaces {Gi×Gi}i\{G^{i}\times G^{i}\}_{i} with different simplicial structures. ∎

The following theorem provides a criterion for non-Nk()N_{k}(\ell)-map.

Theorem 9.5.

Let f:HGf\colon H\to G be a homomorphism between topological groups of which the underlying spaces are CW complexes and k,1k,\ell\geq 1. Then, if ff is an Nk()N_{k}(\ell)-map, then there exist fiberwise pointed maps

EkH×HBHΦΨEG×GBGE_{k}H\times_{H}B_{\ell}H\xrightarrow{\Phi}\mathcal{E}\xrightarrow{\Psi}EG\times_{G}BG

for some fiberwise pointed space \mathcal{E} over BkGB_{k}G, where Φ\Phi is precisely a fiberwise pointed map as a map EkH×HBH(Bkf)E_{k}H\times_{H}B_{\ell}H\to(B_{k}f)^{\ast}\mathcal{E}, satisfying the following conditions:

  1. (1)

    Φ\Phi restricts to a weak homotopy equivalence on each fiber,

  2. (2)

    the restriction of the composite ΨΦ\Psi\circ\Phi to the fiber over the basepoint is homotopic to ιBf:BHBG\iota_{\ell}\circ B_{\ell}f\colon B_{\ell}H\to BG,

  3. (3)

    the composite ((Bkf)Ψ)Φ:EkH×HBHEkH×HBG((B_{k}f)^{\ast}\Psi)\circ\Phi\colon E_{k}H\times_{H}B_{\ell}H\to E_{k}H\times_{H}BG of Φ\Phi and the pullback of Ψ\Psi by the map Bkf:BkHBkGB_{k}f\colon B_{k}H\to B_{k}G is homotopic to the map induced from ff as a fiberwise pointed map.

As in Theorem 6.2, the condition (2) which immediately follows from (3) is contained.

Remark 9.6.

The fiberwise pointed space \mathcal{E} is obtained as the fiberwise projective space of the fiberwise AnA_{n}-space in Theorem 6.2. So the converse of this theorem should hold as well. But, to verify it, we need to generalize Theorem 9.3 for fiberwise AnA_{n}-maps from fiberwise AnA_{n}-spaces in a way compatible with the composition of fiberwise AnA_{n}-maps between fiberwise topological monoids. We leave this problem for now.

Proof.

Suppose f:HGf\colon H\to G is an Nk()N_{k}(\ell)-map. Then we have a fiberwise AA_{\ell}-space EE over BkGB_{k}G and fiberwise AA_{\ell}-maps ϕ:EkH×conjH(Bkf)E\phi\colon E_{k}H\times_{\operatorname{conj}}H\to(B_{k}f)^{\ast}E and ψ:EEkG×conjG\psi\colon E\to E_{k}G\times_{\operatorname{conj}}G as in Theorem 6.2. Again by [Tsu15, Proposition 4.1], we may suppose that ϕ\phi and ψ\psi are fiberwise AA_{\ell}-homomorphisms. Moreover, we may suppose the composite of ϕ\phi and the fiberwise homomorphism EkH×conjHEkH×conjfGE_{k}H\times_{\operatorname{conj}}H\to E_{k}H\times_{\operatorname{conj}\circ f}G is homotopic to ψ\psi through fiberwise AA_{\ell}-homomorphisms by this argument. Let =𝒦E\mathcal{E}=\mathscr{B}^{\mathcal{K}}_{\ell}E. Note that a fiberwise AA_{\ell}-homomorphism induces a fiberwise pointed map between the fiberwise projective spaces in the obvious manner. Then we have fiberwise AA_{\ell}-maps Φ:𝒦(EkH×conjH)(Bkf)\Phi^{\prime}\colon\mathscr{B}^{\mathcal{K}}_{\ell}(E_{k}H\times_{\operatorname{conj}}H)\to(B_{k}f)^{\ast}\mathcal{E}, which is the homotopy inverse of the induced map of ϕ\phi, Ψ:𝒦(EG×conjG)\Psi^{\prime}\colon\mathcal{E}\to\mathscr{B}^{\mathcal{K}}_{\infty}(EG\times_{\operatorname{conj}}G), which is the induced map of ψ\psi, and F:𝒦(EkH×conjH)𝒦(EkH×conjfG)F^{\prime}\colon\mathscr{B}^{\mathcal{K}}_{\ell}(E_{k}H\times_{\operatorname{conj}}H)\to\mathscr{B}^{\mathcal{K}}_{\infty}(E_{k}H\times_{\operatorname{conj}\circ f}G), which is the induced map of ff. We also note that there exists a canonical homotopy commutative diagram of fiberwise pointed spaces

𝒦(EkH×conjH)\textstyle{\mathscr{B}^{\mathcal{K}}_{\ell}(E_{k}H\times_{\operatorname{conj}}H)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}F\scriptstyle{F^{\prime}}\scriptstyle{\simeq}𝒦(EkH×conjfG)\textstyle{\mathscr{B}^{\mathcal{K}}_{\infty}(E_{k}H\times_{\operatorname{conj}\circ f}G)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\simeq}EkH×HBH\textstyle{E_{k}H\times_{H}B_{\ell}H\ignorespaces\ignorespaces\ignorespaces\ignorespaces}F\scriptstyle{F}EkH×HBG\textstyle{E_{k}H\times_{H}BG}

where FF is the induced map of ff and the vertical maps are fiberwise pointed homotopy equivalences. This follows from the classification theorem of fiberwise principal bundles and the fact that both 𝒦(EkH×conjH)\mathscr{B}^{\mathcal{K}}_{\ell}(E_{k}H\times_{\operatorname{conj}}H) and EkH×HBHE_{k}H\times_{H}B_{\ell}H are characterized by the fiberwise variant of Ganea’s pullback-pushuout construction. Composing the above fiberwise pointed homotopy equivalences, we obtain the fiberwise pointed maps Φ\Phi and Ψ\Psi satisfying the desired conditions. ∎

10. pp-local normality of Lie groups

By the result of James [Jam67] and Remark 6.5, the inclusion SU(m)SU(n)\operatorname{SU}(m)\to\operatorname{SU}(n) (2m<n2\leq m<n) is not (22-locally) an N1(1)N_{1}(1)-map. But one might expect some normality when the spaces are localized at a large prime pp. In the rest of this section, we assume k1k\geq 1 and 1\ell\geq 1 and often omit the pp-localization symbol like G=G(p)G=G_{(p)}.

As is well-known, any subgroup of an abelian group is normal. But the analogue of this fact for a homomorphism between infinite loop spaces does not hold as in the following example.

Example 10.1.

Let f:HGf\colon H\to G be a homomorphism such that the delooping is homotopic to the map between the Eilenberg–MacLane spaces g:K(,2n)K(,4n)g\colon K(\mathbb{Q},2n)\to K(\mathbb{Q},4n) classified by the square u2H4n(K(,2n);)u^{2}\in H^{4n}(K(\mathbb{Q},2n);\mathbb{Q}) of the generator uH2n(K(,2n);)u\in H^{2n}(K(\mathbb{Q},2n);\mathbb{Q}). The homotopy fiber of gg is homotopy equivalent to the rationalized 2n2n-sphere S(0)2nS^{2n}_{(0)}, which is not an HH-space. Then, since we have catS(0)2n=1\operatorname{cat}S^{2n}_{(0)}=1, ff is not an N1(1)N_{1}(1)-map by Theorem 8.2. Similarly, the inclusion SO(2n)(0)SO(2n+1)(0)\operatorname{SO}(2n)_{(0)}\to\operatorname{SO}(2n+1)_{(0)} is not an N1(1)N_{1}(1)-map for n1n\geq 1.

But we have a normality result for some class of homomorphisms as follows.

Theorem 10.2.

Let f:HGf\colon H\to G be a homomorphism between compact connected semisimple Lie groups. Let 2m12m-1 and 2n12n-1 denote the largest degrees of generators of the rational cohomology algebra H(H;)H^{\ast}(H;\mathbb{Q}) and H(G;)H^{\ast}(G;\mathbb{Q}), respectively. Suppose that f:H(G;)H(H;)f^{\ast}\colon H^{\ast}(G;\mathbb{Q})\to H^{\ast}(H;\mathbb{Q}) is surjective and pkn+mp\geq kn+\ell m. Then the homomorphism ff is pp-locally an Nk()N_{k}(\ell)-map.

Proof.

We verify the conditions in Theorem 6.2 for the trivial fiberwise topological group E=BkG×HE=B_{k}G\times H. We employ the technique to reduce the projective spaces in [HKT19, Sections 3 and 4]. Since we assume pkn+mp\geq kn+\ell m, we have the AkA_{k}-equivalences

HS2m11××S2ms1andGS2n11××S2nr1H\simeq S^{2m_{1}-1}\times\cdots\times S^{2m_{s}-1}\quad\text{and}\quad G\simeq S^{2n_{1}-1}\times\cdots\times S^{2n_{r}-1}

for some sequences 2m1msm2\leq m_{1}\leq\cdots\leq m_{s}\leq m and 2n1nrn2\leq n_{1}\leq\cdots\leq n_{r}\leq n. Then we have the homotopy equivalences

GH×G/HandG/HS2q11××S2qrs1G\simeq H\times G/H\quad\text{and}\quad G/H\simeq S^{2q_{1}-1}\times\cdots\times S^{2q_{r-s}-1}

for some sequence 2q1qrsn2\leq q_{1}\leq\cdots\leq q_{r-s}\leq n since f:H(G;)H(H;)f^{\ast}\colon H^{\ast}(G;\mathbb{Q})\to H^{\ast}(H;\mathbb{Q}) is surjective. Note that the sequences m1,,msm_{1},\ldots,m_{s} and q1,,qrsq_{1},\ldots,q_{r-s} are subsequences of the sequence n1,,nrn_{1},\ldots,n_{r}. We can consider these equivalences are AkA_{k}-equivalences with an appropriate AkA_{k}-form on G/HG/H. Consider the space

Bk(i1,,it)=k1++kt=kBk1S2i11××BktS2it1.B_{k}(i_{1},\ldots,i_{t})=\bigcup_{k_{1}+\cdots+k_{t}=k}B_{k_{1}}S^{2i_{1}-1}\times\cdots\times B_{k_{t}}S^{2i_{t}-1}.

Then we have the homotopy commutative diagram

for maps

α:BkHBk(m1,,ms)andβ:Bk(G/H)Bk(q1,,qrs)\alpha\colon B_{k}H\to B_{k}(m_{1},\ldots,m_{s})\quad\text{and}\quad\beta\colon B_{k}(G/H)\to B_{k}(q_{1},\ldots,q_{r-s})

inducing injective homomorphisms of mod pp-cohomologies, where the homotopy commutativity of the left square follows from the construction of the left bottom arrow in [HKT19, Proposition 3.2]. Let γ:BkGBk(n1,,nr)\gamma\colon B_{k}G\to B_{k}(n_{1},\ldots,n_{r}) denote the composite of the bottom arrows. We also have the maps

α:Bk(m1,,ms)BHandγ:Bk(n1,,nr)BG\alpha^{\prime}\colon B_{k}(m_{1},\ldots,m_{s})\to BH\quad\text{and}\quad\gamma^{\prime}\colon B_{k}(n_{1},\ldots,n_{r})\to BG

such that the compositions αα\alpha^{\prime}\circ\alpha and γγ\gamma^{\prime}\circ\gamma are homotopic to the inclusions. Note that the same argument works for the \ell-th projective spaces. Consider the fiberwise pointed spaces

Bk(m1,,ms)Bk(m1,,ms)×B(m1,,ms)Bk(m1,,ms)and\displaystyle B_{k}(m_{1},\ldots,m_{s})\to B_{k}(m_{1},\ldots,m_{s})\times B_{\ell}(m_{1},\ldots,m_{s})\to B_{k}(m_{1},\ldots,m_{s})\quad\text{and}
Bk(n1,,nr)Bk(n1,,nr)×B(m1,,ms)Bk(n1,,nr)\displaystyle B_{k}(n_{1},\ldots,n_{r})\to B_{k}(n_{1},\ldots,n_{r})\times B_{\ell}(m_{1},\ldots,m_{s})\to B_{k}(n_{1},\ldots,n_{r})

with the first projections and the first inclusions. Thus we have the homotopy commutative diagram of fiberwise pointed spaces

(6)

Since we suppose pkn+mp\geq kn+\ell m, we have

π2i1(BH)=π2i1(BG)=0\pi_{2i-1}(BH)=\pi_{2i-1}(BG)=0

for i2kn+2mi\leq 2kn+2\ell m. This implies that there exists a commutative diagram of fiberwise pointed maps

(11)

Here the top and bottom arrows cover α\alpha^{\prime} and γ\gamma^{\prime}, respectively. Moreover, their restrictions to the fiber over the basepoint are α\alpha^{\prime} and the composite of the maps

B(m1,,ms)inclusionB(n1,,ns)γBG,B_{\ell}(m_{1},\ldots,m_{s})\xrightarrow{\text{inclusion}}B_{\ell}(n_{1},\ldots,n_{s})\xrightarrow{\gamma^{\prime}}BG,

respectively. Composing the squares (6) and (11), we get the homotopy commutative square of fiberwise pointed maps

BkH×BH\textstyle{B_{k}H\times B_{\ell}H\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}inclusionBkf×id\scriptstyle{B_{k}f\times\operatorname{id}}EH×HBH\textstyle{EH\times_{H}BH\ignorespaces\ignorespaces\ignorespaces\ignorespaces}induced map of ffBkG×BH\textstyle{B_{k}G\times B_{\ell}H\ignorespaces\ignorespaces\ignorespaces\ignorespaces}EG×GBG\textstyle{EG\times_{G}BG}

where the bottom arrow covers the inclusion BkGBGB_{k}G\to BG. Thus taking the adjoint to this diagram in the sense of Theorem 9.3, we obtain the homotopy commutative diagram of fiberwise AA_{\ell}-maps between fiberwise topological groups

EkH×conjH\textstyle{E_{k}H\times_{\operatorname{conj}}H\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}inclusionϕ\scriptstyle{\phi^{\prime}}EH×conjH\textstyle{EH\times_{\operatorname{conj}}H\ignorespaces\ignorespaces\ignorespaces\ignorespaces}induced map of ffE\textstyle{E\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ψ\scriptstyle{\psi}EG×conjG\textstyle{EG\times_{\operatorname{conj}}G}

where each arrow covers the same map as the corresponding arrow in the previous diagram. The map ϕ\phi^{\prime} can be inverted over BkGB_{k}G to be a fiberwise AA_{\ell}-equivalence ϕ:(Bkf)EEkH×conjH\phi\colon(B_{k}f)^{\ast}E\to E_{k}H\times_{\operatorname{conj}}H. Now we can verify the conditions in Theorem 6.2. Therefore, the homomorphism ff is pp-locally an Nk()N_{k}(\ell)-map. ∎

Remark 10.3.

In the proof of the theorem, the action G𝒜(H,H)G\to\mathcal{A}_{\ell}(H,H) of the resulting Nk()N_{k}(\ell)-map is null-homotopic since EE is trivial. Let us propose to call such an Nk()N_{k}(\ell)-map a Zk()Z_{k}(\ell)-map, where ZZ comes from the German word zentral. Although it seems stronger condition than being an Nk()N_{k}(\ell)-map, we have no example at this point for an Nk()N_{k}(\ell)-map but not a Zk()Z_{k}(\ell)-map between Lie groups.

Let us investigate the non-normality of the inclusion SU(m)SU(n)\operatorname{SU}(m)\to\operatorname{SU}(n).

Theorem 10.4.

Suppose max{knm,(k1)n+2}<pkn+(1)m\max\{kn-m,(k-1)n+2\}<p\leq kn+(\ell-1)m for some 2m<n2\leq m<n and k,1k,\ell\geq 1, then the inclusion SU(m)SU(n)\operatorname{SU}(m)\to\operatorname{SU}(n) is not pp-locally an Nk()N_{k}(\ell)-map.

Before proving the theorem, we recall the cohomology of projective spaces.

Lemma 10.5.

If the pp-local cohomology of a compact connected Lie group GG is given by

H(G;(p))=Λ(p)(x1,,xn),H^{\ast}(G;\mathbb{Z}_{(p)})=\Lambda_{\mathbb{Z}_{(p)}}(x_{1},\ldots,x_{n}),

then

H(BkG;(p))=(p)[y1,,yn]/(y1,,yn)k+1Sk,H^{\ast}(B_{k}G;\mathbb{Z}_{(p)})=\mathbb{Z}_{(p)}[y_{1},\ldots,y_{n}]/(y_{1},\ldots,y_{n})^{k+1}\oplus S_{k},

where yiy_{i} is the image of the transgression of xix_{i} in H(BG;(p))H^{\ast}(BG;\mathbb{Z}_{(p)}) and SkS_{k} is a free (p)\mathbb{Z}_{(p)}-module and mapped to 0 in H(Bk1G;(p))H^{\ast}(B_{k-1}G;\mathbb{Z}_{(p)}). Moreover, when the pp-local cohomology of a compact connected Lie group HH is given by

H(H;(p))=Λ(p)(x1,,xm)H^{\ast}(H;\mathbb{Z}_{(p)})=\Lambda_{\mathbb{Z}_{(p)}}(x^{\prime}_{1},\ldots,x^{\prime}_{m})

and a homomorphism f:HGf\colon H\to G induces a surjective map on pp-local cohomology, then the induced map

(Bkf):H(BkG;(p))H(BkH;(p))(B_{k}f)^{\ast}\colon H^{\ast}(B_{k}G;\mathbb{Z}_{(p)})\to H^{\ast}(B_{k}H;\mathbb{Z}_{(p)})

is also surjective.

Proof.

As in [Hem94, Iwa84], consider the spectral sequence converging to H(BG;(p))H^{\ast}(BG;\mathbb{Z}_{(p)}) induced from the filtration

=B0GB1GBkGBG.\ast=B_{0}G\subset B_{1}G\subset\cdots\subset B_{k}G\subset\cdots\subset BG.

The differential in the E1E_{1}-page coincides with the differential of the cobar construction of the exterior algebra H(G;(p))H^{\ast}(G;\mathbb{Z}_{(p)}) and then, as is well-known, the E2E_{2}-page is the polynomial algebra (p)[y1,,yn]\mathbb{Z}_{(p)}[y_{1},\ldots,y_{n}]. This implies that the spectral sequence collapses at the E2E_{2}-page. Truncate this spectral sequence to compute H(BkG;(p))H^{\ast}(B_{k}G;\mathbb{Z}_{(p)}), of which the E1E_{1}-page looks like

0E10,d1E11,d1d1E1k,0.0\to E_{1}^{0,\ast}\xrightarrow{d_{1}}E_{1}^{1,\ast}\xrightarrow{d_{1}}\cdots\xrightarrow{d_{1}}E_{1}^{k,\ast}\to 0.

When i<ki<k, the E2i,E_{2}^{i,\ast}-term of this spectral sequence is a free module of which the basis is given by monomials of length ii. Comparing with the spectral sequence for H(BG;(p))H^{\ast}(BG;\mathbb{Z}_{(p)}), the E2k,E_{2}^{k,\ast}-term is the direct sum of a similar module generated by monomials of length kk and the submodule SkS_{k} which is isomorphic to a submodule in the E1k+1,E_{1}^{k+1,\ast}-term of the spectral sequence for H(BG;(p))H^{\ast}(BG;\mathbb{Z}_{(p)}). This implies that SkS_{k} is free. Again comparing with the spectral sequence for H(BG;(p))H^{\ast}(BG;\mathbb{Z}_{(p)}), the one for H(BkG;(p))H^{\ast}(B_{k}G;\mathbb{Z}_{(p)}) also collapses at the E2E_{2}-page. Then we obtain H(BkG;(p))H^{\ast}(B_{k}G;\mathbb{Z}_{(p)}) as in the lemma. For the latter half, note that the homomorphism f:H(G;(p))H(H;(p))f^{\ast}\colon H^{\ast}(G;\mathbb{Z}_{(p)})\to H^{\ast}(H;\mathbb{Z}_{(p)}) is a map of coalgebra admitting a section. Thus the induced map on the E2E_{2}-page is surjective, completing the proof. ∎

The pp-local cohomology of BSU(n)B\operatorname{SU}(n) is given by

H(BSU(n);(p))=(p)[c2,,cn],H^{\ast}(B\operatorname{SU}(n);\mathbb{Z}_{(p)})=\mathbb{Z}_{(p)}[c_{2},\ldots,c_{n}],

where cic_{i} denotes the ii-th Chern class of degree 2i2i. We assume that the inclusion SU(m)SU(n)\operatorname{SU}(m)\to\operatorname{SU}(n) is pp-locally an Nk()N_{k}(\ell)-map and derive a contradiction. As in Theorem 9.5, we have a fiberwise projective space =E\mathcal{E}=\mathscr{B}_{\ell}E for some fiberwise AA_{\ell}-space EE over BkSU(n)B_{k}\operatorname{SU}(n) with fiber AA_{\ell}-equivalent to SU(m)\operatorname{SU}(m) and fiberwise pointed maps Φ:EkSU(m)×SU(m)BSU(m)BkSU(m)\Phi\colon E_{k}\operatorname{SU}(m)\times_{\operatorname{SU}(m)}B_{\ell}\operatorname{SU}(m)\to\mathcal{E}_{B_{k}\operatorname{SU}(m)} and Ψ:EkSU(n)×SU(n)BSU(m)\Psi\colon\mathcal{E}\to E_{k}\operatorname{SU}(n)\times_{\operatorname{SU}(n)}B\operatorname{SU}(m) satisfying the properties in Theorem 9.5.

Lemma 10.6.

The pp-local cohomology of \mathcal{E} is given by

H(;(p))\displaystyle H^{\ast}(\mathcal{E};\mathbb{Z}_{(p)})
=((p)[c2B,,cnB]/(c2B,,cnB)k+1SkB)((p){(c2F)i2(cmF)im0i2++im}SF)\displaystyle=(\mathbb{Z}_{(p)}[c_{2}^{B},\ldots,c_{n}^{B}]/(c_{2}^{B},\ldots,c_{n}^{B})^{k+1}\oplus S_{k}^{B})\otimes(\mathbb{Z}_{(p)}\{(c_{2}^{F})^{i_{2}}\cdots(c_{m}^{F})^{i_{m}}\mid 0\leq i_{2}+\cdots+i_{m}\leq\ell\}\oplus S_{\ell}^{F})

as a H(BkSU(n);(p))=(p)[c2B,,cnB]/(c2B,,cnB)k+1SkBH^{\ast}(B_{k}\operatorname{SU}(n);\mathbb{Z}_{(p)})=\mathbb{Z}_{(p)}[c_{2}^{B},\ldots,c_{n}^{B}]/(c_{2}^{B},\ldots,c_{n}^{B})^{k+1}\oplus S_{k}^{B}-module, where we write cjB=Ψ(cj×1)c_{j}^{B}=\Psi^{\ast}(c_{j}\times 1), cjF=Ψ(1×cj)c_{j}^{F}=\Psi^{\ast}(1\times c_{j}), SkBS_{k}^{B} and SFS_{\ell}^{F} are free modules and (p)S\mathbb{Z}_{(p)}S with a set SS denotes the free (p)\mathbb{Z}_{(p)}-module with basis SS.

Proof.

Since the rationalization SU(n)(0)\operatorname{SU}(n)_{(0)} is AA_{\infty}-equivalent to a topological abelian group, the rational cohomology Serre spectral sequence of EkSU(n)×SU(n)BSU(n)BkSU(n)E_{k}\operatorname{SU}(n)\times_{\operatorname{SU}(n)}B_{\ell}\operatorname{SU}(n)\to B_{k}\operatorname{SU}(n) collapses at the E2E_{2}-page. Then the pp-local cohomology spectral sequence E~\tilde{E}_{\ast} also collapses since its E2E_{2}-page is free by Lemma 10.5. Consider the pp-local cohomology Serre spectral sequence EE_{\ast} of BkSU(n)\mathcal{E}\to B_{k}\operatorname{SU}(n). The E2E_{2}-term is given as

E2=((p)[c2B,,cnB]/(c2B,,cnB)k+1SkB)((p)[c2F,,cF]/(c2F,,cmF)+1SF)E_{2}=(\mathbb{Z}_{(p)}[c_{2}^{B},\ldots,c_{n}^{B}]/(c_{2}^{B},\ldots,c_{n}^{B})^{k+1}\oplus S_{k}^{B})\otimes(\mathbb{Z}_{(p)}[c_{2}^{F},\ldots,c_{\ell}^{F}]/(c_{2}^{F},\ldots,c_{m}^{F})^{\ell+1}\oplus S_{\ell}^{F})

for some free modules SkBS_{k}^{B} and SFS_{\ell}^{F}. Again by Lemma 10.5, the induced map Ψ:E~2E2\Psi^{\ast}\colon\tilde{E}_{2}\to E_{2} is surjective. This implies that EE_{\ast} also collapses. Thus the lemma follows. ∎

By the inequalities

n+p1\displaystyle n+p-1 n+((k1)n+3)1=kn+2,\displaystyle\geq n+((k-1)n+3)-1=kn+2,
(m+1)+p1\displaystyle(m+1)+p-1 m+1+(kn+(1)m)1=kn+m,\displaystyle\leq m+1+(kn+(\ell-1)m)-1=kn+\ell m,

we can find integers 00\leq\ell^{\prime}\leq\ell, m+1inm+1\leq i\leq n and 2j<m2\leq j<m or j=0j=0 satisfying

i+p1=kn+m+jandm+jm.i+p-1=kn+\ell^{\prime}m+j\quad\text{and}\quad\ell^{\prime}m+j\leq\ell m.

By the assumption knm<pkn-m<p and this equality, one can see the inequalities

k<p3+1andp2.\displaystyle k<\frac{p}{3}+1\quad\text{and}\quad\ell^{\prime}\leq\frac{p}{2}.

From now on we shall work in the mod pp cohomology, where we use the same symbols as in the pp-local cohomology. Since the map Ψ:ESU(n)×SU(n)BSU(n)BSU(n)×BSU(n)\Psi\colon\mathcal{E}\to E\operatorname{SU}(n)\times_{\operatorname{SU}(n)}B\operatorname{SU}(n)\cong B\operatorname{SU}(n)\times B\operatorname{SU}(n) is fiberwise pointed and the composite ΨΦ:ESU(m)×SU(m)BSU(m)ESU(n)×SU(n)BSU(n)\Psi\circ\Phi\colon E\operatorname{SU}(m)\times_{\operatorname{SU}(m)}B\operatorname{SU}(m)\to E\operatorname{SU}(n)\times_{\operatorname{SU}(n)}B\operatorname{SU}(n) is homotopic to the induced map of the homomorphism f:SU(m)SU(n)f\colon\operatorname{SU}(m)\to\operatorname{SU}(n), we have

(12) Ψ(1×ci)=ciB+(a polynomial),\displaystyle\Psi^{\ast}(1\times c_{i})=c_{i}^{B}+(\text{a polynomial}),

where the polynomial consists of terms divisible by at least one of cm+1B,,ci1Bc_{m+1}^{B},\ldots,c_{i-1}^{B}. Let us compute 𝒫1(Ψ(1×ci))\mathcal{P}^{1}(\Psi^{\ast}(1\times c_{i})) the image of Steenrod operation 𝒫1\mathcal{P}^{1}.

Lemma 10.7.

The coefficient of cjcmcnkc_{j}c_{m}^{\ell^{\prime}}c_{n}^{k} (c0=1c_{0}=1 when j=0j=0) in 𝒫1ciH2i+2p2(BSU(n);𝔽p)\mathcal{P}^{1}c_{i}\in H^{2i+2p-2}(B\operatorname{SU}(n);\mathbb{F}_{p}) is nonzero.

Proof.

Recall the mod pp Wu formula [Sha77]

𝒫1ci=2i2++nin=i+p1\displaystyle\mathcal{P}^{1}c_{i}=\sum_{2i_{2}+\cdots+ni_{n}=i+p-1} (1)i2++in1(i2++in1)!i2!in!\displaystyle(-1)^{i_{2}+\cdots+i_{n}-1}\frac{(i_{2}+\cdots+i_{n}-1)!}{i_{2}!\cdots i_{n}!}
(i+p1j=2i1(i+p1j)iji2++in1)c2i2cnin.\displaystyle\cdot\left(i+p-1-\frac{\sum_{j=2}^{i-1}(i+p-1-j)i_{j}}{i_{2}+\cdots+i_{n}-1}\right)c_{2}^{i_{2}}\cdots c_{n}^{i_{n}}.

Let ′′=\ell^{\prime\prime}=\ell^{\prime} if j=0j=0 and ′′=+1\ell^{\prime\prime}=\ell^{\prime}+1 otherwise. Since we have

i+p1j=2i1(i+p1j)iji2++in1\displaystyle i+p-1-\frac{\sum_{j=2}^{i-1}(i+p-1-j)i_{j}}{i_{2}+\cdots+i_{n}-1}
=kn+m+j(kn+m+j)′′jmk+′′1\displaystyle=kn+\ell^{\prime}m+j-\frac{(kn+\ell^{\prime}m+j)\ell^{\prime\prime}-j-\ell^{\prime}m}{k+\ell^{\prime\prime}-1}
=k((k1)n+m+j)k+′′1,\displaystyle=\frac{k((k-1)n+\ell^{\prime}m+j)}{k+\ell^{\prime\prime}-1},

we can compute the coefficient of cjcmcnkc_{j}c_{m}^{\ell^{\prime}}c_{n}^{k} as

(1)k+′′1(k+′′1)!k!!k((k1)n+m+j)k+′′1\displaystyle(-1)^{k+\ell^{\prime\prime}-1}\frac{(k+\ell^{\prime\prime}-1)!}{k!\ell^{\prime}!}\cdot\frac{k((k-1)n+\ell^{\prime}m+j)}{k+\ell^{\prime\prime}-1}
=(1)k+′′1(k+′′2)!(k1)!!((k1)n+m+j).\displaystyle=(-1)^{k+\ell^{\prime\prime}-1}\frac{(k+\ell^{\prime\prime}-2)!}{(k-1)!\ell^{\prime}!}((k-1)n+\ell^{\prime}m+j).

Now we have

k+′′2<p3+1+p22<pand0<(k1)n+m+j=i+p1n<p\displaystyle k+\ell^{\prime\prime}-2<\frac{p}{3}+1+\frac{p}{2}-2<p\quad\text{and}\quad 0<(k-1)n+\ell^{\prime}m+j=i+p-1-n<p

and hence the coefficient of cjcmcnkc_{j}c_{m}^{\ell^{\prime}}c_{n}^{k} is nonzero. ∎

This lemma implies that the coefficient of (cnB)kcjF(cmF)(c_{n}^{B})^{k}c_{j}^{F}(c_{m}^{F})^{\ell^{\prime}} in 𝒫1Ψ(1×ci)\mathcal{P}^{1}\Psi^{\ast}(1\times c_{i}) is nonzero. But it cannot happen as follows.

Lemma 10.8.

The coefficient of (cnB)kcjF(cmF)(c_{n}^{B})^{k}c_{j}^{F}(c_{m}^{F})^{\ell^{\prime}} (c0F=1c^{F}_{0}=1 when j=0j=0) in 𝒫1Ψ(1×ci)\mathcal{P}^{1}\Psi^{\ast}(1\times c_{i}) is zero.

Proof.

Applying 𝒫1\mathcal{P}^{1} to the equality (12), the term (cnB)kcjF(cmF)(c_{n}^{B})^{k}c_{j}^{F}(c_{m}^{F})^{\ell^{\prime}} comes from (𝒫1csB)cjF(cmF)(\mathcal{P}^{1}c_{s}^{B})c_{j}^{F}(c_{m}^{F})^{\ell^{\prime}} with sns\neq n. Here we have m<s<nm<s<n. Since 𝒫1csB\mathcal{P}^{1}c_{s}^{B} must contain the term (cnB)k(c_{n}^{B})^{k}, ss is computed as s=kn(p1)s=kn-(p-1). But this contradicts to our assumption s>m>knps>m>kn-p in the theorem. Thus the coefficient of (cnB)kcjF(cmF)(c_{n}^{B})^{k}c_{j}^{F}(c_{m}^{F})^{\ell^{\prime}} is zero. ∎

This contradicts to the previous result, completing the proof of Theorem 10.4.

Let us examine the pp-local normality of the inclusion SU(2)SU(3)\operatorname{SU}(2)\to\operatorname{SU}(3) for small primes pp. By Theorems 10.2 and 10.4, we obtain Table 1, where ✓  and ✗  mean the inclusion is an Nk()N_{k}(\ell)-map and is not an Nk()N_{k}(\ell)-map, respectively, and ?? means we cannot determine the normality from these theorems.

kk 11 22 33 44 55
Nk(1)N_{k}(1)
Nk(2)N_{k}(2)
Nk(3)N_{k}(3)
Nk(4)N_{k}(4)
Nk(5)N_{k}(5)

p=3p=3

kk 11 22 33 44 55
Nk(1)N_{k}(1) ? ? ? ?
Nk(2)N_{k}(2)
Nk(3)N_{k}(3)
Nk(4)N_{k}(4)
Nk(5)N_{k}(5)

p=5p=5

kk 11 22 33 44 55
Nk(1)N_{k}(1) ? ? ? ?
Nk(2)N_{k}(2)
Nk(3)N_{k}(3)
Nk(4)N_{k}(4)
Nk(5)N_{k}(5)

p=7p=7

kk 11 22 33 44 55
Nk(1)N_{k}(1) ? ?
Nk(2)N_{k}(2)
Nk(3)N_{k}(3) ?
Nk(4)N_{k}(4)
Nk(5)N_{k}(5)

p=11p=11

Table 1. The pp-local higher homotopy normality of SU(2)SU(3)\operatorname{SU}(2)\subset\operatorname{SU}(3)

We leave the remaining cases for now. The first undetermined cases are as follows.

Problem 10.9.

Determine whether the inclusion SU(2)SU(3)\operatorname{SU}(2)\to\operatorname{SU}(3) is 55-locally an N2(1)N_{2}(1)-map or not. More generally, find methods to determine when the inclusion SU(m)SU(n)\operatorname{SU}(m)\to\operatorname{SU}(n) is pp-locally an Nk(1)N_{k}(1)-map.

We should note that E2SU(2)×conjSU(2)E_{2}\operatorname{SU}(2)\times_{\operatorname{conj}}\operatorname{SU}(2) is not 55-locally fiberwise based equivalent to the trivial bundle though its restriction E1SU(2)×conjSU(2)E_{1}\operatorname{SU}(2)\times_{\operatorname{conj}}\operatorname{SU}(2) over B1SU(2)B_{1}\operatorname{SU}(2) is 55-locally trivial.

We close this paper giving the corresponding results for SO(2m+1)SO(2n+1)\operatorname{SO}(2m+1)\to\operatorname{SO}(2n+1), which is pp-locally homotopy equivalent to Sp(m)Sp(n)\operatorname{Sp}(m)\to\operatorname{Sp}(n) for odd pp. The proof proceeds as Theorems 10.4 using the mod pp Wu formula

𝒫1pi=i1++nin=i+p12\displaystyle\mathcal{P}^{1}p_{i}=\sum_{i_{1}+\cdots+ni_{n}=i+\frac{p-1}{2}} (1)i1++in1+i+p12(i1++in1)!i1!in!\displaystyle(-1)^{i_{1}+\cdots+i_{n}-1+i+\frac{p-1}{2}}\frac{(i_{1}+\cdots+i_{n}-1)!}{i_{1}!\cdots i_{n}!}
(2i+p1j=1i1(2i+p12j)iji1++in1)p1i1pnin\displaystyle\cdot\left(2i+p-1-\frac{\sum_{j=1}^{i-1}(2i+p-1-2j)i_{j}}{i_{1}+\cdots+i_{n}-1}\right)p_{1}^{i_{1}}\cdots p_{n}^{i_{n}}

for the Pontryagin classes p1,,pnp_{1},\ldots,p_{n}.

Theorem 10.10.

Suppose max{2kn2m,2(k1)n+2}<p2kn+2(1)m1\max\{2kn-2m,2(k-1)n+2\}<p\leq 2kn+2(\ell-1)m-1 for some 1m<n1\leq m<n and k,1k,\ell\geq 1, then the inclusion SO(2m+1)SO(2n+1)\operatorname{SO}(2m+1)\to\operatorname{SO}(2n+1) is not pp-locally an Nk()N_{k}(\ell)-map.

References

  • [BV73] J. M. Boardman and R. M. Vogt. Homotopy invariant algebraic structures on topological spaces. Lecture Notes in Mathematics, Vol. 347. Springer-Verlag, Berlin-New York, 1973.
  • [CJ98] Michael Crabb and Ioan James. Fibrewise homotopy theory. Springer Monographs in Mathematics. Springer-Verlag London, Ltd., London, 1998.
  • [FS10] Emmanuel D. Farjoun and Yoav Segev. Crossed modules as homotopy normal maps. Topology Appl., 157(2):359–368, 2010.
  • [Fur85] Yasukuni Furukawa. Homotopy-normality of Lie groups. Quart. J. Math. Oxford Ser. (2), 36(141):53–56, 1985.
  • [Fur87] Yasukuni Furukawa. Homotopy-normality of Lie groups. II. Quart. J. Math. Oxford Ser. (2), 38(150):185–188, 1987.
  • [Fur95] Yasukuni Furukawa. Homotopy-normality of Lie groups. III. Hiroshima Math. J., 25(1):83–96, 1995.
  • [Hem94] Yutaka Hemmi. On exterior AnA_{n}-spaces and modified projective spaces. Hiroshima Math. J., 24(3):583–605, 1994.
  • [HKT19] Sho Hasui, Daisuke Kishimoto, and Mitsunobu Tsutaya. Higher homotopy commutativity in localized Lie groups and gauge groups. Homology Homotopy Appl., 21(1):107–128, 2019.
  • [Iwa84] Norio Iwase. On the KK-ring structure of XX-projective nn-space. Mem. Fac. Sci. Kyushu Univ. Ser. A, 38(2):285–297, 1984.
  • [Iwa98] Norio Iwase. Ganea’s conjecture on Lusternik-Schnirelmann category. Bull. London Math. Soc., 30(6):623–634, 1998.
  • [Jam67] I. M. James. On the homotopy theory of the classical groups. An. Acad. Brasil. Ci., 39:39–44, 1967.
  • [Jam71] I. M. James. Products between homotopy groups. Compositio Math., 23:329–345, 1971.
  • [Jam76] I. M. James. The topology of Stiefel manifolds. London Mathematical Society Lecture Note Series, No. 24. Cambridge University Press, Cambridge-New York-Melbourne, 1976.
  • [Kac82] Hideyuki Kachi. On the iterated Samelson product. Math. J. Okayama Univ., 24(1):37–44, 1982.
  • [KK10] Daisuke Kishimoto and Akira Kono. Splitting of gauge groups. Trans. Amer. Math. Soc., 362(12):6715–6731, 2010.
  • [KN03] Akira Kono and Osamu Nishimura. Homotopy normality of Lie groups and the adjoint action. J. Math. Kyoto Univ., 43(3):641–650, 2003.
  • [KT18] Daisuke Kishimoto and Mitsunobu Tsutaya. Samelson products in pp-regular SO(2n){\rm SO}(2n) and its homotopy normality. Glasg. Math. J., 60(1):165–174, 2018.
  • [KY98] Kenji Kudou and Nobuaki Yagita. Modulo odd prime homotopy normality for HH-spaces. J. Math. Kyoto Univ., 38(4):643–651, 1998.
  • [KY01] Kenji Kudou and Nobuaki Yagita. Note on homotopy normality and the nn-connected fiber space. Kyushu J. Math., 55(1):119–129, 2001.
  • [May75] J. Peter May. Classifying spaces and fibrations. Mem. Amer. Math. Soc., 1(1, 155):xiii+98, 1975.
  • [McC64] G. S. McCarty, Jr. Products between homotopy groups and the JJ-morphism. Quart. J. Math. Oxford Ser. (2), 15:362–370, 1964.
  • [Nis06] Osamu Nishimura. A note on homotopy normality of HH-spaces. J. Math. Kyoto Univ., 46(4):913–921, 2006.
  • [Pre12] Matan Prezma. Homotopy normal maps. Algebr. Geom. Topol., 12(2):1211–1238, 2012.
  • [Sak10] Michihiro Sakai. AA_{\infty}-spaces and L-S category in the category of fibrewise spaces. Topology Appl., 157(13):2131–2135, 2010.
  • [Sha77] P. Brian Shay. mod{\rm mod} pp Wu formulas for the Steenrod algebra and the Dyer-Lashof algebra. Proc. Amer. Math. Soc., 63(2):339–347, 1977.
  • [Sta63] James Dillon Stasheff. Homotopy associativity of HH-spaces. I, II. Trans. Amer. Math. Soc. 108 (1963), 275-292; ibid., 108:293–312, 1963.
  • [Sta70] James Stasheff. HH-spaces from a homotopy point of view. Lecture Notes in Mathematics, Vol. 161. Springer-Verlag, Berlin-New York, 1970.
  • [Tsu12] Mitsunobu Tsutaya. Finiteness of AnA_{n}-equivalence types of gauge groups. J. Lond. Math. Soc. (2), 85(1):142–164, 2012.
  • [Tsu15] Mitsunobu Tsutaya. Homotopy pullback of AnA_{n}-spaces and its applications to AnA_{n}-types of gauge groups. Topology Appl., 187:1–25, 2015.
  • [Tsu16] Mitsunobu Tsutaya. Mapping spaces from projective spaces. Homology Homotopy Appl., 18(1):173–203, 2016.