1. Introduction
Let $\mathop {\rm map}\nolimits (X,\,Y)$ denote the space of continuous maps $X\to Y$ where $Y$ has the homotopy type of a nilpotent CW complex, and let $l\colon Y\to Y_{(P)}$ denote localization at the set of primes $P$. The induced map
is known to be localization at $P$ on each path component under various assumptions on $X\!$. By the classical results of Hilton, Mislin, Roitberg and Steiner [Reference Hilton, Mislin, Roitberg and Steiner13, Theorems A and B], $X$ can be a homologically finite CW complex. The more recent results of Klein, Schochet, and Smith [Reference Klein, Schochet and Smith16] (and [Reference Schochet and Smith26]) show that this also holds whenever $X$ is a compact metric space and one knows, a priori, that the function spaces involved are nilpotent. Similar results hold for spaces of sections (see Møller [Reference Møller24] for the case of a CW domain, and [Reference Klein, Schochet and Smith16] for the case of a compact metric domain). Here we address the question of nilpotency and localization for function and section spaces with more general paracompact domains by a unified approach; the case of a locally compact subspace of a Euclidean space has served as a particular motivation.
To achieve our goal, we need a good grip on the homotopy groups of $\mathop {\rm map}\nolimits (X,\,Y)$. Federer [Reference Federer10] constructed a spectral sequence of groups, arising from an exact couple, converging to $\pi _*(\mathop {\rm map}\nolimits (X,\,Y))$ for a finite-dimensional CW complex $X$ and a simple CW complex $Y$ (see also [Reference Smith27] for a based rational version). Dyer [Reference Dyer8] generalized Federer's spectral sequence to the case of a paracompact Hausdorff space of finite covering dimension $X$. The method was to represent $X$ as an inverse limit of polyhedra and take the direct limit of Federer's exact couples. Analogous approaches were those of Dror Farjoun and Schochet [Reference Farjoun and Schochet7] and Klein, Schochet and Smith [Reference Klein, Schochet and Smith16].
Here, we take the dual approach (as in [Reference Hilton, Mislin, Roitberg and Steiner13] and [Reference Møller24]), and express the nilpotent CW complex $Y\!$ as a weak limit of the refined Postnikov tower of principal fibrations $Y_q\to Y_{q-1}$. We note that, similarly as in the case of a CW complex $X$, the associated tower of maps $\mathop {\rm map}\nolimits (X,\,Y_q)\to \mathop {\rm map}\nolimits (X,\,Y_{q-1})$ is also one of principal fibrations under reasonable conditions. Of course, there is the spectral sequence of Bousfield and Kan [Reference Bousfield and Kan5] associated to any tower of fibrations, but that involves sets and groups acting on sets, and is not ready-made for localization. We observe that Federer's method admits an abstract generalization to any tower of principal fibrations and obtain a spectral sequence of groups (arising from an exact couple) and a slightly tweaked version thereof that contains sufficient information to detect localization of the limiting homotopy groups at the $E^2$-level. It seems that these observations are novel. To state our main result, we first let $\mathop {\rm map}\nolimits _\phi (X,\,Y)$ denote the subspace of $\mathop {\rm map}\nolimits (X,\,Y)$ consisting of maps whose restriction to the closed subspace $A\subset X$ is a prescribed map $\phi \colon A\to Y$. Further let $X$ be a paracompact Hausdorff k-space; we will be using the shorthand $X\in {\rm PHK}$. (The obvious examples are metric spaces, compact Hausdorff spaces, and CW complexes.) Assume that $Y$ is a CW approximation to the inverse limit of successive principal fibrations with fibres $K(\pi ^q,\,n_q)$ where $1=n_1\leqslant n_2\leqslant \dots$. Let $\check H^*(X,\,A;\pi )$ denote Čech cohomology with coefficients in the abelian group $\pi$. Additionally, assume one of the following.
(i) $\phi =\mathop {\rm const}\nolimits _{y_0}$ where $y_0$ is nondegenerate; we use $\mathop {\rm map}\nolimits _{A\to y_0}(X,\,Y)$ in this case.
(ii) $A\hookrightarrow X$ is a closed Hurewicz cofibration (for example, $A$ may be empty).
(iii) $(X,\,A)$ is a proper pair (see definition 1.5) and $Y$ is an ANR, i.e., an absolute neighbourhood retract for metric spaces.
Theorem 1.1 Take $f\in \mathop {\rm map}\nolimits _\phi (X,\,Y)$. There exists an upper half-plane homology type spectral sequence of groups where $E^2_{-p,q}$ is isomorphic with $\check H^{n_q-q+p}(X,\,A;\pi ^q)$ for $1\leqslant q\geqslant p\geqslant 0$ and trivial otherwise. The $n$-th differential consists of morphisms $d^n_{-p,q}\colon E^n_{-p,q}\to E^n_{-p-n,q+n-1}$.
Assume also that $Y\!$ has only finitely many nontrivial homotopy groups or that $\mathop {\rm dim}\nolimits X\!$ (respectively $\mathop {\rm dim}\nolimits (X,\,A)$ under (iii) above) is finite. Then $\mathop {\rm map}\nolimits _\phi (X,\,Y)$ is nilpotent and the spectral sequence converges to $\pi _k(\mathop {\rm map}\nolimits _\phi (X,\,Y),\,f)$ for $k\geqslant 1$.
The spectral sequence is natural with respect to maps of pairs $(X,\,A)\to (X',\,A')$ over $Y$ and with respect to maps $Y\to Y'$ induced by maps of classifying spaces $K(\pi ^q_{Y},\,n_q+1)\to K(\pi ^q_{Y'},\,n_q+1)$. Let $E^n_{-p,q}\to E'^n_{-p,q}$ be the associated map in any of the two cases and let $\pi _k\to \pi _k'$ be the associated morphism of homotopy groups of mapping spaces. Assume that both $E^n$ and $E'^n$ converge and that
• $E^2_{-p,q}\to E'^2_{-p,q}$ P-localizes for $q>p\geqslant 0$, and
• the torsion in $E'^2_{-q,q}$ is P-local and $E^2_{-q,q}\to E'^2_{-q,q}$ is P-injective for $q>0$.
Then $\pi _k\to \pi _k'$ are P-localizations for all $k\geqslant 1$.
If $Y\!$ is connected and simple in all dimensions, we obtain a second-quadrant sequence with non-zero $E^2$-terms given by $E^2_{-p,q}\cong \check H^p(X,\,A;\pi _q(Y))$ for $q\geqslant p\geqslant 0$.
The theorem is proved in § 5 as a consequence of theorem 3.1 and the results in § 4; theorem 3.1 also contains a precise description of the first differential in terms of the $k$-invariants of $Y$. We emphasize that in the absolute case (i.e., $A=\emptyset$), the result always depends only on the homotopy type of $Y$, while if $A$ is nonempty and not cofibred in $X$, then the topology of $Y$ will play a role (§ 7 contains details for the ‘twisted’ relative case). For example, $Y$ is an ANR if it is a smooth manifold or a locally finite CW complex.
The nilpotency of $\mathop {\rm map}\nolimits _\phi (X,\,Y)$ and convergence of the spectral sequence are automatic if $Y$ is a Postnikov section; this is a feature of our approach. The applications of [Reference Klein, Schochet and Smith16] include cases where $Y=K_{(0)}$ is the rationalization of a finite CW complex $K$; if $K$ is rationally elliptic, then $Y$ will be a Postnikov section, guaranteeing convergence for any $X\in {\rm PHK}$. (Compare also with [Reference Farjoun and Schochet7, Theorem A].)
The following is a formal consequence of the naturality in theorem 1.1.
Corollary 1.2 Assume the conditions for convergence in theorem 1.1. If the canonical morphisms $\check H^i(X,\,A;\pi ^q)\to \check H^i(X,\,A;\pi ^q\otimes {\mathbb {Z}}_{(P)})$ are localizations at P for all $q$ and all $i\leqslant n_q$, then $\mathop {\rm map}\nolimits _\phi (X,\,Y)\to \mathop {\rm map}\nolimits _\phi (X,\,Y_{(P)})$ localizes homotopy groups at all basepoints. (Here, $Y_{(P)}$ is required to be an ANR if $Y\!$ is.)
Universal coefficients for Čech cohomology of compact spaces immediately imply
Corollary 1.3 Suppose that $X$ is compact Hausdorff and Y has the homotopy type of a nilpotent CW complex. If $\mathop {\rm dim}\nolimits X<\infty$ or $Y$ is a Postnikov section, the induced map (1.1) is localization on each path component.
Example 1.4 In [Reference Kahn15], D. S. Kahn constructed a compact metric space $X$ with trivial integral Čech cohomology groups $\check H^q(X;{\mathbb {Z}})$ for $q>1$ but with essential maps $X\to {\mathbb {S}}^3$. Let $Y=B{\mathbb {S}}^3$ be the classifying space of the sphere ${\mathbb {S}}^3$ with basepoint $y_0$. Then $\pi _1(\mathop {\rm map}\nolimits (X,\,Y),\,\mathop {\rm const}\nolimits _{y_0})\cong \pi _1(\mathop {\rm map}\nolimits _*(X,\,Y),\,\mathop {\rm const}\nolimits _{y_0})\cong [X,\,\Omega B{\mathbb {S}}^3]$ is nontrivial. Therefore, the spectral sequence associated to $\mathop {\rm map}\nolimits (X,\,Y)$ cannot converge to $\pi _*(\mathop {\rm map}\nolimits (X,\,Y))$. In particular, $X$ is infinite-dimensional (as noted by Kahn).
We need to impose restrictions on a noncompact $X$ for $\mathop {\rm map}\nolimits (X,\,Y)$ to be amenable to our methods. Presumably, it would be sufficient to have a regular complete lattice of zero sets on $X$ that determines the topology and is generated by compact zero sets (see [Reference Smrekar28, pp. 1 and 2] for the definition of a regular lattice). To avoid technicalities, we content ourselves with the ‘countable case’ as follows.
Definition 1.5 A topological pair $(X,\,A)$ will be called proper if $A\in {\rm PHK}$ and $X$ is the union of an ascending chain of subspaces $A=X_0\subset X_1\subset X_2\subset \dots$ (called an admissible chain) such that for all $i\geqslant 1$ there is a pushout
where $S_i$ is a zero set in the compact Hausdorff space $C_i$ and $X$ has the weak (colimit) topology with respect to $\left \{X_i\right \}$. Each $X_i$ is paracompact and it follows that $X\in {\rm PHK}$ by Michael [Reference Michael21, Theorem 8.2]. Also, the $X_i$ are zero sets in $X$.
We call $X$ a proper space if $(X,\,\emptyset )$ is a proper pair. In this case, the $X_i$ are compact Hausdorff. Examples include countable CW complexes and proper metric spaces (i.e. closed balls are compact). Also, locally compact subspaces of Euclidean spaces ${\mathbb {R}}^d$ are proper. More generally, if $X$ is a Hausdorff k-space that admits a proper function $\phi \colon X\to [0,\,\infty )$ (the preimages of compact sets are compact), then one sees that $X$ is proper by setting $X_i=\phi ^{-1}([0,\,i])$.
Observe that if $(X,\,A)$ is a proper pair with admissible chain $\left \{X_i\right \}$ and $B$ is a zero set in $X$, then $(X,\,A\cup B)$ is a proper pair with admissible chain $\left \{X_i\cup B\right \}$. In particular, if $X$ is a proper space, then $(X,\,B)$ is a proper pair for any zero set $B$.
A proper pair $(X,\,A)$ will be called locally finite-dimensional if the $C_i$ have finite covering dimension. Every countable relative CW complex is then a locally finite-dimensional proper pair. We write $\mathop {\rm dim}\nolimits (X,\,A)\leqslant d$ if $\mathop {\rm dim}\nolimits C_i\leqslant d$ for all $i$ and call such a pair finite-dimensional.
Any admissible chain of a compact proper space (or pair) is essentially finite. In particular, a locally finite-dimensional compact proper space is finite-dimensional.
Hilton, Mislin, Roitberg, and Steiner pointed out in [Reference Hilton, Mislin, Roitberg and Steiner13] that the natural class of CW complexes $X$ for which (1.1) always localizes is that of globally homologically finite ones; $X$ is such when $\oplus _{n=0}^\infty H_n(X)$ is finitely generated. From our point of view, that is precisely because a CW complex $X$ is globally homologically finite if and only if $\mathop {\rm map}\nolimits (X,\,Y)$ has the homotopy type of a CW complex for every nilpotent CW complex $Y$. (For sufficiency, use [Reference Smrekar28, Proposition 2.6.4] and [Reference Smrekar30, Corollary 1.2]. For necessity, use the proof of [Reference Smrekar28, c. of Theorem 4.5.3].) Thus in this case, (1.1) is actually localization in the category of nilpotent spaces of CW homotopy type.
The following is a generalization of Theorems A and B of [Reference Hilton, Mislin, Roitberg and Steiner13].
Theorem 1.6 Let $(X,\,A)$ be a proper pair and let $Y$ be a nilpotent CW complex. Fix a map $\phi \colon A\to Y$. If $\mathop {\rm map}\nolimits _\phi (X,\,Y)$ and $\mathop {\rm map}\nolimits _{l\phi }(X,\,Y_{(P)})$ have CW homotopy type, then they are nilpotent and (1.1) is CW localization at P on path components.
It turns out that the question of whether or not $\mathop {\rm map}\nolimits (X,\,Y)$ has CW homotopy type is intimately related to the behaviour of the spectral sequence of theorem 1.1.
Definition 1.7 Let $A$ be a closed subspace of $X\in {\rm PHK}$. The pair $(X,\,A)$ is quasi-finite, $(X,\,A)\in {\mathcal {Q}}{\mathcal {F}}$, if for each abelian group $G$ and all $n\geqslant 1$, the space $\mathop {\rm map}\nolimits _{A\to *}(X,\,K(G,\,n))$ has the homotopy type of a CW complex. (We assume $K(G,\,n)$ to be well-pointed.) We call $X$ quasi-finite if $(X,\,\emptyset )$ is a quasi-finite pair.
By [Reference Smrekar28, Theorem 4.5.3, a.], a CW complex belongs to ${\mathcal {Q}}{\mathcal {F}}$ if and only if all its homology groups are finitely generated.
Proposition 1.8 Let $(X,\,A)$ be a quasi-finite and locally finite-dimensional proper pair, and let $Y$ have the homotopy type of a nilpotent CW complex; if $A$ is nonempty and not cofibred in $X$, we assume that $Y$ is an ANR. Let $\phi \colon A\to Y$ be any map.
If $\mathop {\rm map}\nolimits _\phi (X,\,Y)$ has the homotopy type of a CW complex, then the spectral sequence of theorem 1.1 converges to $\pi _*(\mathop {\rm map}\nolimits _\phi (X,\,Y),\,f)$ for any $f\in \mathop {\rm map}\nolimits _\phi (X,\,Y)$.
We note that if $(X,\,A)$ is a skeleton-finite relative CW pair, then it is a quasi-finite and locally finite-dimensional proper pair.
For proper pairs, quasi-finiteness is actually a cohomological property as follows. Recall that a tower of groups $\dots \to G_2\to G_1$ is Mittag-Leffler if for each $i$, the images of $G_j\to G_i$ stabilize for all big enough $j\geqslant i$.
Proposition 1.9 Let $(X,\,A)$ be a proper pair with an admissible chain $\left \{X_i\right \}$. Then $(X,\,A)\in {\mathcal {Q}}{\mathcal {F}}$ if and only if for each abelian group $G$, the towers of abelian groups
are Mittag-Leffler for $0\leqslant k<\infty$ and, in addition, for each $k$ there exists $i$ such that restriction $\check {H}^k(X,\,A;G)\to \check {H}^k(X_i,\,A;G)$ is injective.
We turn to spaces of sections.
Let $X$ be connected and let $Q\to X$ be a Hurewicz fibration with space of sections $\Gamma (X,\,Q)\subset \mathop {\rm map}\nolimits (X,\,Q)$. Assume that $Q\to X$ is a pullback of a fibration $p\colon E\to B$ of spaces of CW homotopy type where $B$ is connected and the typical fibre of $p$, call it $Y$, is nilpotent. Let $P$ be a set of primes and let $\ell \colon E\to E_{(P)}$ be the fibrewise localization at $P$ over $B$. By pulling back, we obtain a fibrewise localization $\ell \colon Q\to Q_{(P)}$ over $X$. Note that $\ell$ induces a map
In [Reference Schochet and Smith26], the authors studied the effect of $\ell _*$ for compact $X$ provided $\Gamma (X,\,Q)$ is known, a priori, to be nilpotent. We give conditions for nilpotency and enhance their localization result [Reference Schochet and Smith26, Theorem 3] as follows.
Theorem 1.10 Suppose $X$ is a connected proper space and assume that, in addition, $X$ is locally finite-dimensional or $Y$ is a Postnikov section.
(a) If X is compact, then $\Gamma (X,\,Q)$ and $\Gamma (X,\,Q_{(P)})$ are nilpotent spaces of CW type and (1.3) is CW localization at $P$ on each path component.
(b) If $\Gamma (X,\,Q)$ has CW type, then $\Gamma (X,\,Q)$ and $\Gamma (X,\,Q_{(P)})$ are nilpotent and (1.3) localizes at $P$ on path components. If, in addition, Y is a Postnikov section, also $\Gamma (X,\,Q_{(P)})$ has CW type and (1.3) is CW localization at $P$.
A relative version for sections prescribed on $A\subset X$, which also generalizes the main results of [Reference Møller24], is stated in § 7.
Underlying categories and techniques. Our results are valid in the category ${\mathcal {K}}$ of k-spaces and in the category ${\mathcal {T}}\!\!{op}$ of all topological spaces (see [Reference May and Sigurdsson20, Sections 1.1 and 4.1]). There are the underlying Strø m closed model category structures on ${\mathcal {K}}$ and ${\mathcal {T}}\!\!{op}$ with Hurewicz fibrations, closed Hurewicz cofibrations, and homotopy equivalences (called h-equivalences in [Reference May and Sigurdsson20]). While for some of our results, Serre fibrations would be sufficient, we also lean on the results of Stasheff [Reference Stasheff33] and Brown and Heath [Reference Brown and Heath4] that require Hurewicz fibrations, so we stick to the latter. Our mapping spaces, which have the (k-ified) compact open topology, are homotopy metrizable in case of a proper domain, and remarks of [Reference Smrekar29, Section 2.3] apply. Some of our key technical results have been obtained by replacing the target space with a homotopy equivalent absolute neighbourhood retract; Milnor [Reference Milnor22] proved that ANRs and CW complexes contain the same homotopy types. The fact that ANRs together with homotopy equivalences and Hurewicz fibrations form a fibration category in the sense of Baues [Reference Baues2] (see Miyata [Reference Miyata23] as well as remark 7.1) is of importance.
2. A generalized Federer's spectral sequence
Assume a tower of Hurewicz fibrations
We consider a spectral sequence converging to the homotopy groups of the inverse limit, taking on the approach of Federer [Reference Federer10]. He considered only the special case where $Z_i=\mathop {\rm map}\nolimits (X^{(i)},\,Y)$ is the space of maps from the $i$-skeleton of a CW complex $X$ to a simple CW complex $Y$, and the maps $Z_i\to Z_{i-1}$ are restriction fibrations. However, his treatment applies to a fairly general setting as we proceed to explain.
Let the space $Z_\infty$, together with projections $P^i\colon Z_\infty \to Z_i$, be the (topological) inverse limit of (2.1). Let $R\colon Z\to Z_\infty$ be a map of another space $Z$ into $Z_\infty$ and denote $R^i=P^i\circ R\colon Z\to Z_i$. In our applications, $Z\to Z_\infty$ typically would not be a homotopy equivalence (i.e., it would not have a homotopy inverse), but it will induce isomorphisms on homotopy groups in favourable circumstances.
Finally, pick $\zeta \in Z$ and set $\zeta _i=R^i(\zeta )$ for all $i$. Let $Z_i'$ denote the path component of $\zeta _i$ in $Z_i$, and let $F_i'$ denote the fibre of $Z_i'\to Z_{i-1}'$ over $\zeta _{i-1}$. Note that $F_i'$ may be disconnected. In such a setting, Federer defines a generalized exact couple
where $A_{p,q}=\pi _p(Z_q',\,\zeta _q)$ are groups (generally noncommutative for $p=1$), and $C_{p,q}=\pi _p(F_q',\,\zeta _q)$ are abelian groups for all $p\geqslant 0$. When $p<0$ or $q<0$, everything is trivial. Morphisms $r$, $\partial$, and $i$ arise from the long exact sequences of fibrations
and are morphisms of groups throughout. Federer notes that the theory of ‘abelian’ exact couples carries over to this setting. We recall that the differential is $d=\partial \circ i$; we set $(C^{(0)},\,d^{(0)})=(C,\,d)$ and let $C^{(n)}$ be the homology of $(C^{(n-1)},\,d^{(n-1)})$.
In the following lemma we provide a sufficient condition for the existence of a Federer's exact couple in the setting of a tower of fibrations (2.1), and give an abstract identification of the groups $C_{p,q}$ and the first differential.
Lemma 2.1 Assume that for each $q\geqslant 0$, the fibration $Z_q\to Z_{q-1}$ is principal, obtained as the pullback of a fibration $EL_q\to BL_q$ along a classifying map $l_q\colon Z_{q-1}\to BL_q$ where $EL_q$ is contractible and $BL_q$ is a not necessarily connected space with abelian fundamental group (in each component). (For example, $BL_q$ may be an H-group.) By taking $A_{p,q}=\pi _p(Z_q',\,\zeta _q)$ and $C_{p,q}=\pi _p(F_q',\,\zeta _q)$ with morphisms (2.3), one obtains an exact couple (2.2) in the sense of Federer.
In the associated spectral sequence we have that $C_{p,q}\cong \pi _{p+1}(BL_q,\,l_q(\zeta _{q-1}))$ if $p,\,q\geqslant 1$, and that $C_{0,q}$ is isomorphic with a subgroup of $\pi _1(BL_q,\,l_q(\zeta _{q-1}))$ for $q>0$. If $q\leqslant 0$ or $p<0$, then $C_{p,q}=0$. Under those isomorphisms, the first differential $d\colon C_{p,q}\to C_{p-1,q+1}$ for $p\geqslant 1$ corresponds to the composite
where $\delta _q$ can be identified with the connecting morphism in the homotopy exact sequence of the fibration $EL_q\to BL_q$.
Remark 2.2 As $Z_{-1}=\left \{\mathrm {point}\right \}$, $Z_0$ is assumed to be a loop space of $BL_0$. Also, when $p=1$, the codomain of (2.4) has to be understood as the image of $(l_{q+1})_*$.
Remark 2.3 Under $\pi _{p+1}(BL_q,\,l_q(\zeta _{q-1}))\cong \pi _p(\Omega (BL_q,\,l_q(\zeta _{q-1})),\,\mathop {\rm const}\nolimits _{l_q(\zeta _{q-1})})$, the morphism $\delta _q$ from (2.4) is induced by a continuous map $\Omega (BL_q,\,l_q(\zeta _{q-1}))\to F_q$ (taking $\mathop {\rm const}\nolimits _{l_q(\zeta _{q-1})}$ to $\zeta _q$) if $EL_q\to BL_q$ is a regular fibration (i.e. stationary homotopies can be lifted to stationary homotopies).
Remark 2.4 The spectral sequence is natural with respect to maps of towers induced by maps of the classifying spaces $BL_q$. This holds also for the identification of the first differential (2.4).
Addendum 2.5 Starting with $C_{0,q}\leqslant \widetilde C_{0,q}=\pi _0(F_q,\,\zeta _q)\cong \pi _1(BL_q,\,l_q(\zeta _{q-1}))$ with incoming differential $\tilde d\colon C_{1,q-1}\xrightarrow {d}C_{0,q}\leqslant \widetilde C_{0,q}$, each derived group $C_{0,q}^{(n)}$ is a subgroup of the group $\widetilde C_{0,q}^{(n)}=\widetilde C_{0,q}^{(n-1)}/\mathop {\rm im}\nolimits \tilde d^{(n-1)}$ with incoming differential $\tilde d^{(n)}\colon C^{(n)}_{1,q-n-1}\xrightarrow {d^{(n)}}C^{(n)}_{0,q}\leqslant \widetilde C_{0,q}^{(n)}$. Moreover, $\widetilde C^{(n)}_{0,q}$ and $\tilde d^{(n)}$ are natural with respect to morphisms of towers induced by maps of classifying spaces $BL_q$. The exact couple $(A,\,C)$ together with natural subgroup inclusions $C^{(n)}_{0,q}\leqslant \widetilde C^{(n)}_{0,q}$ will be called an augmented Federer's exact couple and denoted $(A,\,C,\,\widetilde C)$.
Proof. As $Z_q\to Z_{q-1}\xrightarrow {l_q}BL_q$ is a homotopy fibration, $\partial \colon \pi _{p+1}(Z_{q-1},\,\zeta _{q-1})\to \pi _{p}(F_q',\,\zeta _q)=C_{p,q}$ can be viewed as $(l_q)_*\colon \pi _{p+1}(Z_{q-1},\,\zeta _{q-1})\to \pi _{p+1}(BL_q,\,l_q(\zeta _{q-1}))$. Moreover, $C_{0,q}$ can be identified with the image of $(l_q)_*$ in $\pi _1(BL_q,\,l_q(\zeta _{q-1}))$ which is abelian by our assumption on $BL_q$. This provides the identification of $C_{p,q}$ and that of $\partial$ in the differential $\partial \circ i\colon C_{p,q}\to C_{p-1,q+1}$. More precisely, as the fibration $Z_q\to Z_{q-1}$ is principal, $F_q$ is homotopy equivalent to the loop-space $\Omega (BL_q,\,l_q(\zeta _{q-1}))$, and the connecting morphism in the homotopy exact sequence of $EL_q\to BL_q$ induces the isomorphism $\delta _q\colon \pi _{p+1}(BL_q,\,l_q(\zeta _{q-1}))\to \pi _p(F_q,\,\zeta _q)$. Noting that $\partial =(\delta _q)^{-1}\circ (l_q)_*\colon \pi _{p+1}(Z_{q-1},\,\zeta _{q-1})\to \pi _p(F_q,\,\zeta _q)$ completes the proof of the main statement.
For remark 2.3, consider a general fibration $f\colon E\to B$ and let $f(e_0)=b_0$ be any coherent pair of basepoints. The connecting morphism $\pi _*(\Omega B,\,\mathop {\rm const}\nolimits _{b_0})\to \pi _{*}(F,\,e_0)$ is induced by a map $\Omega (B,\,b_0)\to F$ which is most easily constructed by means of a lifting function that continuously lifts loops at $b_0$ to paths in $E$ beginning at $e_0$ and takes the end point (in $F$). A regular fibration admits a lifting function sending constant paths to constant paths, forcing $\mathop {\rm const}\nolimits _{b_0}$ to map to $e_0$.
To prove the addendum by induction, note that if $\tilde d^{(n-1)}\colon C^{(n-1)}_{1,q-n}\to \widetilde C^{(n-1)}_{0,q}$ is just $d^{(n-1)}$ followed by the inclusion $C^{(n-1)}_{0,q}\leqslant \widetilde C^{(n-1)}_{0,q}$, the images of $\tilde d^{(n-1)}$ and $d^{(n-1)}$ in $\widetilde C^{(n-1)}_{0,q}$ coincide, yielding subgroup inclusion $C^{(n)}_{(0,q)}=C^{(n-1)}_{0,q}/\mathop {\rm im}\nolimits d^{(n-1)}\leqslant \widetilde C^{(n-1)}_{0,q}/\mathop {\rm im}\nolimits \tilde d^{(n-1)}=\widetilde C^{(n)}_{(0,q)}$.
Let $G_{p,q}$ denote the kernel of the induced morphism $R^q_{*}\colon \pi _p(Z,\,\zeta )\to A_{p,q}=\pi _p(Z_q',\,\zeta _q)$ and consider the normal chain for $\pi _p(Z,\,\zeta )$:
The question of convergence is covered by the following proposition which is a straightforward generalization of the argument of Federer (see [Reference Federer10, pp. 351–352]).
Proposition 2.6 Suppose that for each pair $(p,\,q)$, where $p\geqslant 1$, there exists an integer $j(p,\,q)\geqslant q$ for which the images of
coincide. Then for $n\geqslant \max \left \{j(p,\,q)+1-q,\,q\right \}$, the derived terms $C_{p,q}^{(n)}$ are stable:
Corollary 2.7 Fix $p\geqslant 1$. If, in addition to the assumption of proposition 2.6, $G_{p,q}$ is trivial for $q\geqslant Q=Q(p)$, then the normal chain (2.5) terminates and its successive quotients are given by (2.6) for all $n\geqslant \max \left \{j(p,\,Q),\,Q\right \}$. In this sense the spectral sequence converges classically to $\pi _p(Z,\,\zeta )$.
In practice, it is difficult to verify the assumptions of proposition 2.6 and corollary 2.7 (when they are not ‘automatic’). The entire § 4 is devoted to verifying those assumptions in a particular case, while proposition 1.8 shows they are implied by strong assumptions on homotopy type.
Routine diagram chasing coupled with exactness properties of localization of nilpotent groups yield the following theorem and its corollary.
Theorem 2.8 Assume an abstract morphism of augmented Federer's exact couples $(A,\,C,\,\widetilde C)\to (A',\,C',\,\widetilde C')$. Fix a set of primes $P$ (allowed to be the set of all primes). Assume that the following two conditions hold for some $n\geqslant 1$.
(i) The morphisms $C^{(n-1)}_{p,q}\to C'^{(n-1)}_{p,q}$ are localizations at $P$ for $p,\,q\geqslant 1$.
(ii) The torsion in $\widetilde C'^{(n-1)}_{0,q}$ is P-local and the morphisms $\widetilde C^{(n-1)}_{0,q}\to \widetilde C'^{(n-1)}_{0,q}$ are P-injective for $q>0$.
Then the same holds on level $n$.
3. The homotopy spectral sequence of a function space
In this section, we associate to the function space $\mathop {\rm map}\nolimits _\phi (X,\,Y)$ a tower of principal fibrations as studied above. We carry out the construction of the ensuing spectral sequence and the identification of the first differential. The questions of convergence and nilpotency will be addressed in the following sections.
Let $\check {H}^i(X,\,A;G)$ denote the $i$-th Čech cohomology group of the pair $(X,\,A)$ with coefficients in the discrete group $G$. We refer to Dowker [Reference Dowker6] and Bredon [Reference Bredon3] for properties of Čech cohomology.
Let $Y$ have the homotopy type of a connected nilpotent CW complex. Assume that $Y$ is approximated (up to weak homotopy type) with the inverse limit of successive principal fibrations with fibres $K(\pi ^q,\,n_q)$ where $1=n_1\leqslant n_2\leqslant \dots$. Precisely, we assume a tower of fibrations
where $Y_1=K(\pi ^1,\,1)=:K_1$ for an abelian group $\pi ^1$, and each map $\eta _q\colon Y_q\to Y_{q-1}$ is a principal fibration obtained from the $k$-invariant $k_q\colon Y_{q-1}\to K(\pi ^q,\,n_q+1)=BK_q$ where $\pi ^q$ is abelian. This is to say that $Y_q\to Y_{q-1}$ is obtained by pulling back the path fibration $\varepsilon \colon P(BK_q)\to BK_q$ along $k_q$. We also assume that the $BK_q$ and consequently $P(BK_q)$ are ANRs; it follows that so are $Y_q$ (see [Reference Miyata23] and remark 7.1). There are compatible maps $\eta ^q\colon Y\to Y_q$ whose connectivity tends to infinity with $q$. These induce a continuous map $\eta \colon Y\to Y_\infty$ where $Y_\infty$ is the inverse limit of the tower. The map $\eta$ is a weak homotopy equivalence.
Let $(X,\,A)$ be a pair where $X$ is a Hausdorff k-space and let $\phi \colon A\to Y$ be a map. Set $\phi _q=\eta ^q\circ \phi \colon A\to Y_q$; the maps $\phi _q$ induce a continuous map $\phi _\infty \colon A\to Y_\infty$. We may form an associated tower as follows.
Using the fact that the $\eta _q$ are regular fibrations, the proof of [Reference Smrekar29, Lemma A.4] may be amended to render (3.2) a tower of fibrations (without conditions on $A$). The maps $\eta ^q\colon Y\to Y_q$ induce compatible maps $R^q\colon \mathop {\rm map}\nolimits _\phi (X,\,Y)\to \mathop {\rm map}\nolimits _{\phi _q}(X,\,Y_q)$ and thus a limit map $R\colon \mathop {\rm map}\nolimits _\phi (X,\,Y)\to \lim _q\mathop {\rm map}\nolimits _{\phi _q}(X,\,Y_q)=\mathop {\rm map}\nolimits _{\phi _\infty }(X,\,Y_\infty )$. Set $Z_q=\mathop {\rm map}\nolimits _{\phi _q}(X,\,Y_q)$ and $Z=\mathop {\rm map}\nolimits _\phi (X,\,Y)$. Pick $\zeta \in Z$ and let $\zeta _q=R^q(\zeta )$ for all $q$.
Finally assume that, in addition, $A\hookrightarrow X$ is a closed cofibration or $(X,\,A)$ is a proper pair. One can use [Reference Smrekar29, Lemma A.2] in the first case and proposition 6.1 in the second to infer that restrictions $\mathop {\rm map}\nolimits (X,\,T)\to \mathop {\rm map}\nolimits (A,\,T)$ are fibrations for all $T$, respectively all $T$ which are ANRs.
Theorem 3.1 The tower $\left \{Z_i\right \}$ satisfies the assumptions of lemma 2.1. In the associated spectral sequence we have that $C_{p,q}\cong \check H^{n_q-p}(X,\,A;\pi ^q)$ if $p,\,q\geqslant 1$, and $\widetilde C_{0,q}\cong \check H^{n_q}(X,\,A;\pi ^q)$ for $q>0$. If $q\leqslant 0$ or $p<0$, then $C_{p,q}=\widetilde C_{0,q}=0$. These identifications are natural in $(X,\,A)$ (i.e. with respect to maps over $Y$) as well as with respect to maps of towers (3.1) induced by maps $K(\pi ^q_{(1)},\,n_q+1)\to K(\pi ^q_{(2)},\,n_q+1)$ for $q\geqslant 1$.
Proof. There are induced pullback diagrams as follows. (Amend [Reference Smrekar29, Lemma A.3].)
Here, $\psi _q$ denotes the composite $A\xrightarrow {\phi _q}Y_q\to P(BK_q)$. Also, let $\Psi _q$ denote the composite $X\xrightarrow {\zeta _q}Y_q\to P(BK_q)$. Note that $\varepsilon \psi _q=k_q\phi _{q-1}$ and $\varepsilon \Psi _q=k_q\zeta _{q-1}$. As mentioned above, the vertical arrows in (3.3) are fibrations. As $P(BK_q)$ is contractible, $\mathop {\rm map}\nolimits (X,\,P(BK_q))\to \mathop {\rm map}\nolimits (A,\,P(BK_q))$ is a fibration and homotopy equivalence. Thus $EL_q=\mathop {\rm map}\nolimits _{\psi _q}(X,\,P(BK_q))$, its fibre over $\psi _q$, is also contractible, rendering $(\eta _q)_*$ in diagram (3.3) a principal fibration. Set $BL_q=\mathop {\rm map}\nolimits _{\varepsilon \psi _q}(X,\,BK_q)$. Assuming an H-group multiplication on $BK_q$ with strict unit $\mathbb {1}$, we get induced H-group structures on $\mathop {\rm map}\nolimits (X,\,BK_q)$ and $\mathop {\rm map}\nolimits (A,\,BK_q)$ and, therefore, an induced map of fibrations as follows.
The horizontal arrows are homotopy equivalences and by coglueing homotopy equivalences (see [Reference Brown and Heath4, Corollary 1.5] or [Reference May and Ponto19, Lemma 2.2.4]), so are the maps on the fibres. In particular,
is a homotopy equivalence. This shows that, in fact, $BL_q$ is an H-group and so lemma 2.1 can be applied. The identification of $C_{p,q}$ as (a subgroup of) $\check H^{n_q-p}(X,\,A;\pi ^q)$ comes via (3.5) through the obvious homeomorphism of loop spaces
and representability of Čech cohomology over paracompact Hausdorff pairs (see [Reference Huber14] and [Reference Goto11]). The reader may check that this identification is natural with respect to maps of pairs $(X,\,A)\to (X',\,A')$ over $Y\!$.
For full naturality with respect to maps of towers (3.1), we assume that
are strict H-maps of ANR H-groups with strict units. To achieve that, we deloop a given homotopy representative for (3.6) to a based map $K(\pi ^q_{(1)},\,n_q+2)\to K(\pi ^q_{(2)},\,n_q+2)$ between ANRs and take the induced map between the associated Moore loop spaces (which will also be ANRs).
Addendum 3.2 The first differential $C_{p,q}\to C_{p-1,q+1}$ for $p\geqslant 1$ is obtained by applying $\pi _p$ to the composite of (based) continuous maps
The first map is induced by the self-homotopy equivalence $\mathop {\rm map}\nolimits _{A\to \mathbb {1}}(X,\,BK_q)\to \mathop {\rm map}\nolimits _{k_q\phi _{q-1}}(X,\,BK_q)$ sending the map $\mathop {\rm const}\nolimits _{\mathbb {1}}$ to $k_q\zeta _{q-1}$ that is in turn induced by an H-group multiplication on $BK_q$ with strict unit $\mathbb {1}$. The following 3 maps correspond to the sequence (2.4) in conjunction with remark 2.3. This identification of the differential is natural in $(X,\,A)$ and $Y\!$.
4. An auxiliary result for finite-dimensional domain
The results of this section will be used in § 5 to establish the convergence of the spectral sequence of theorem 1.1 in case of a finite dimensional domain.
In [Reference Klein, Schochet and Smith16], the authors establish natural isomorphisms $\mathop {\rm colim}\nolimits _j\pi _k(\mathop {\rm map}\nolimits (X_j,\,Y),\,f_j)\to \pi _k(\mathop {\rm map}\nolimits (X,\,Y),\,f)$ where $X$ is a compactum and $\left \{X_j\right \}$ an inverse system of compact polyhedra with limit $X$. They use a method of Spanier [Reference Spanier32, Theorem 13.4] and a clever trick to show validity for any map $f\colon X\to Y$ as basepoint.
We note that a generalization to any paracompact space $X$ (pair, even) is possible, using methods of Barratt (see [Reference Barratt1, Section 12]) and the same trick to allow for general basepoints in $\mathop {\rm map}\nolimits (X,\,Y)$. To begin, let $X$ be a paracompact Hausdorff space with closed subset $A$. For each locally finite open covering $\lambda$ of $X$, let $N_\lambda$ denote the nerve of $\lambda$ and $L_\lambda$ the subcomplex of $N_\lambda$ obtained by the embedding of the nerve of $\lambda \cap A$. A simplicial map $p_{\mu \lambda }\colon N_\mu \to N_\lambda$ exists whenever $\mu$ refines $\lambda$, and partitions of unity provide for maps $h_\lambda \colon X\to N_\lambda$. Finally, we assume that $(Y,\,y_0)$ has the homotopy type of a pointed CW complex and allow for an additional compact Hausdorff parameter space $T$. Our auxiliary result is the following.
Theorem 4.1 Product maps $h_\lambda \times \mathop {\rm id}\nolimits _T$ and $p_{\lambda \mu }\times \mathop {\rm id}\nolimits _T$ yield a natural bijection
Essentially, this follows by an application of [Reference Barratt1, Theorem 12.32] to the codomain $(\mathop {\rm map}\nolimits (T,\,Y),\,\mathop {\rm const}\nolimits _{y_0})$, which has the homotopy type of a pointed CW complex if $(Y,\,y_0)$ has, and the exponential law. Barratt's restriction to locally finite targets is not necessary (the straight-line homotopy used implicitly in his Lemma 12.31 should be replaced with one in the sense of [Reference Sakai25, Proposition 4.9.7]). Note that if $A$ is empty, we do not need a basepoint in $Y\!$.
To apply theorem 4.1, let $P^nY$ be the $n$-th Postnikov section (up to homotopy) together with the projection $p\colon Y\to P^nY$ (pointed if $Y$ is pointed).
Proposition 4.2 If $\mathop {\rm dim}\nolimits X\leqslant d< n$, then $p_*\colon \mathop {\rm map}\nolimits (X,\,Y)\to \mathop {\rm map}\nolimits (X,\,P^nY)$ and $p_*\colon \mathop {\rm map}\nolimits _{A\to y_0}(X,\,Y)\to \mathop {\rm map}\nolimits _{A\to y_0}(X,\,P^nY)$ are $(n-d+1)$-equivalences.
Proof. The dimension assumption guarantees a cofinal subfamily $\{\lambda \}$ with $\mathop {\rm dim}\nolimits N_\lambda \leqslant d$. Let $N_\lambda$ be such and let $T$ be a compact $k$-dimensional polyhedron. Consider
By relative cellular approximation, (4.1) is a surjection if $k+d\leqslant n+1$ and an injection if $k+d\leqslant n$. By theorem 4.1, $[X\times T,\,Y]\to [X\times T,\,P^nY]$ has the same properties. In particular, taking $T=\left \{*\right \}$ we infer that $[X,\,Y]\to [X,\,P^nY]$ is a bijection. Let $f\colon X\to Y$ be a basepoint in $\mathop {\rm map}\nolimits (X,\,Y)$ and let $*$ denote a basepoint in the sphere $S^k$. Making use of the projection retraction $X\times S^k\to X\times \left \{*\right \}\equiv X$ as in the proof of [Reference Klein, Schochet and Smith16, Theorem 6.4], we infer that
is a split short exact sequence of pointed sets. Here, $[X\times S^k,\,Y]_{f}$ is the set of homotopy classes of maps $X\times S^k\to Y$ that restrict to $f\colon X\times \left \{*\right \}\to Y$. Applying naturality of (4.2) to $p\colon Y\to P^nY$, a diagram chase shows that $p_*\colon [X\times S^k,\,Y]_f\to [X\times S^k,\,P^nY]_{pf}$, which can clearly be identified with $p_*\colon \pi _k(\mathop {\rm map}\nolimits (X,\,Y),\,f)\to \pi _k(\mathop {\rm map}\nolimits (X,\,P^nY),\,pf)$, is bijective for $k< n-d+1$ and surjective for $k=(n-d+1)$.
For $p_*\colon \mathop {\rm map}\nolimits _{A\to y_0}(X,\,Y)\to \mathop {\rm map}\nolimits _{A\to p(y_0)}(X,\,P^nY)$, use pairs everywhere.
We apply proposition 4.2 to the relative case as follows. Let $A$ be closed in $X$ and let $\phi \colon A\to Y$ be a map. Consider the induced map
Corollary 4.3 Assume that $n$ is big enough.
We call a map a weak $n$-equivalence if it induces an injection on $\pi _0$, isomorphisms on $\pi _k$ for $1\leqslant k< n$, and epimorphisms on $\pi _n$ for all choices of basepoint.
Proof. Case (i) follows from proposition 4.2 applied to the morphism of the homotopy exact sequences of fibrations $\mathop {\rm map}\nolimits (X,\,Y)\to \mathop {\rm map}\nolimits (A,\,Y)$ and $\mathop {\rm map}\nolimits (X,\,P^nY)\to \mathop {\rm map}\nolimits (A,\,P^nY)$. For (ii), consider the pullback diagrams induced by (1.2):
Using proposition 6.1, it follows that the vertical arrows in (4.4) are fibrations. Reasoning as in case (i), it follows that the mapping induced on the fibres of
is an $(n-d)$-equivalence. By naturality of (4.4) in $Y\!$,
is a weak $(n-d)$-equivalence for all $i$. Now $\mathop {\rm map}\nolimits _\phi (X,\,Y)$ is the inverse limit of $\left \{\mathop {\rm map}\nolimits _\phi (X_{i},\,Y)\right \}$ (see corollary 6.2) and there are the associated $\lim ^1--\lim$ short exact sequences (see [Reference Bousfield and Kan5, Chapter IX] and also [Reference Mardešić and Segal17, p. 178, Theorem 1]). By naturality in $Y$, we get morphisms of those exact sequences identifying the morphisms $\pi _k(\mathop {\rm map}\nolimits _\phi (X,\,Y),\,f)\to \pi _k(\mathop {\rm map}\nolimits _\phi (X,\,P^nY),\,pf)$. From those we infer (ii).
5. Convergence, nilpotency and localization
Definition 5.1 A topological space $Z$ (not necessarily of CW homotopy type) is nilpotent if, for any choice of basepoint $z_0$, the fundamental group $\pi _1(Z,\,z_0)$ is nilpotent and operates nilpotently on all higher homotopy groups $\pi _n(Z,\,z_0)$, $n\geqslant 2$.
Proof of theorem 1.1 Assume the notation of theorem 3.1. Diagram (3.3) exhibits $\mathop {\rm map}\nolimits _{\phi _q}(X,\,Y_q)$ as the homotopy fibre of $(k_q)_*\colon \mathop {\rm map}\nolimits _{\phi _{q-1}}(X,\,Y_{q-1})\to BL_q$ where $BL_q$ is an H-group and hence nilpotent. As $\mathop {\rm map}\nolimits _{\phi _{0}}(X,\,Y_{0})=\left \{*\right \}$, it follows by induction that $\mathop {\rm map}\nolimits _{\phi _q}(X,\,Y_q)$ are nilpotent for all $q$ (see [Reference Hilton, Mislin and Roitberg12, Theorem II.2.2] together with the final remark of the proof, as well as the proof of [Reference Hilton, Mislin, Roitberg and Steiner13, Theorem A]).
Case 1. $Y$ has finitely many nontrivial homotopy groups. Then $Y\to Y_q$ is a homotopy equivalence for some $q$. This finishes the proof if $A=\emptyset$ or if $\phi$ maps the entire $A$ to a nondegenerate $y_0\in Y$, for then we have homotopy equivalences $\mathop {\rm map}\nolimits (X,\,Y)\to \mathop {\rm map}\nolimits (X,\,Y_q)$ and $\mathop {\rm map}\nolimits _{A\to y_0}(X,\,Y)\to \mathop {\rm map}\nolimits _{A\to \eta ^q(y_0)}(X,\,Y_q)$. (See the proof of (1) in the discussion on page 7.)
Suppose that $A\hookrightarrow X$ is a cofibration or $(X,\,A)$ is a proper pair and $Y$ is an ANR. Then we have restriction fibrations $\mathop {\rm map}\nolimits (X,\,Y)\to \mathop {\rm map}\nolimits (A,\,Y)$ and $\mathop {\rm map}\nolimits (X,\,Y_q)\to \mathop {\rm map}\nolimits (A,\,Y_q)$ (note $Y_q$ is an ANR by construction), and a morphism between them induced by the homotopy equivalence $Y\to Y_q$. By [Reference Brown and Heath4, Corollary 1.5], we get the homotopy equivalence $\mathop {\rm map}\nolimits _\phi (X,\,Y)\to \mathop {\rm map}\nolimits _{\phi _q}(X,\,Y)$ between the fibres. Thus in Case 1, the assumptions of proposition 2.6 and corollary 2.7 are met trivially.
Case 2. $Y$ has infinitely many nontrivial homotopy groups. By proposition 4.2 and corollary 4.3, the connectivity of $\mathop {\rm map}\nolimits _\phi (X,\,Y)\to \mathop {\rm map}\nolimits _{\phi _q}(X,\,Y_q)$ tends to infinity with $q$, implying nilpotency of $\mathop {\rm map}\nolimits _\phi (X,\,Y)$ as well as the assumptions of proposition 2.6 and corollary 2.7.
Finally, we set $E^2_{-k,q}=C_{q-k,q}$ for $k< q$ and $E^2_{-q,q}=\widetilde C_{0,q}$ to obtain the upper half-plane spectral sequence with homology-type differentials as claimed. Localization on the level of $E^2$ implies localization in the limit in case of convergence by theorem 2.8.
Example 5.2 By way of example, Federer [Reference Federer10] studied the rationalized higher homotopy groups of $\mathop {\rm map}\nolimits _*(X,\,S^n)$ for a finite-dimensional CW complex $X$. To illustrate the usefulness of addendum 3.2, we also treat $\mathop {\rm map}\nolimits _*(X,\,S^n_{(0)})$ where $S^n_{(0)}$ is the rationalized sphere. Our treatment is valid for $X\in {\rm PHK}$; by corollary 1.3, $\mathop {\rm map}\nolimits _*(X,\,S^n)\to \mathop {\rm map}\nolimits _*(X,\,S^n_{(0)})$ is rationalization on path components when $X$ is compact of finite covering dimension.
The interesting case is of an even $n>2$. Then, $Y=S^n_{(0)}$ can be represented as the homotopy fibre of a single Postnikov invariant $k\colon K({\mathbb {Q}},\,n)\to K({\mathbb {Q}},\,2n)$ whose associated cohomology class in $H^{2n}(K({\mathbb {Q}},\,n);{\mathbb {Q}})$ is the square of the fundamental class $\iota _n\in H^n(K({\mathbb {Q}},\,n);{\mathbb {Q}})$.
First we consider $[X,\,Y]_*$. To this end, we investigate the Puppe sequence
Here, $\Rightarrow$ is employed to imply that the group $[X,\,K({\mathbb {Q}},\,2n-1)]_*$ acts on the set $[X,\,Y]_*$, and that $[X,\,Y]_*\to [X,\,K({\mathbb {Q}},\,n)]_*$ collapses precisely the orbits.
By definition, $k_\#[f]=[k\circ f]$ which we may view as $f^*(\iota _n^2)=(f^*(\iota _n))^2$. Therefore, $k_\#\colon [X,\,K({\mathbb {Q}},\,n)]_*\to [X,\,K({\mathbb {Q}},\,2n)]_*$ translates into the squaring operation $\check {H}^n(X,\,*;{\mathbb {Q}})\to \check {H}^{2n}(X,\,*;{\mathbb {Q}})$. This holds for any $X$, implying that $\Omega k_\#$ is trivial as it corresponds, by adjunction, to squaring in a suspension
This gives the short exact sequence
We consider the path component of a map $\zeta \colon X\to Y$. Diagram (3.3) reads
We assume that $K({\mathbb {Q}},\,n)$ and $K({\mathbb {Q}},\,2n)$ are well-pointed and that $k$ is a fibration between well-pointed spaces. Next we assume that $\mu \colon K({\mathbb {Q}},\,n)\times K({\mathbb {Q}},\,n)\to K({\mathbb {Q}},\,n)$ is an H-group multiplication with strict unit $\mathbb {1}$. By virtue of the homotopy lifting property, we may also assume that $kp\zeta$ is the constant map.
For our spectral sequence, we assume that $Y_i=*$ for $1\leqslant i\leqslant n-1$, next $Y_n=\Omega (K({\mathbb {Q}},\,n+1),\,*)=K({\mathbb {Q}},\,n)$, then $Y_i=Y_n$ for $n+1\leqslant i\leqslant 2n-2$, and finally $Y=Y_{2n-1}$ is the homotopy fibre of $k\colon Y_n\to K({\mathbb {Q}},\,2n)$. The only nontrivial differential is $d^{(n-1)}$. The fibre $F_n$ is equal to $\mathop {\rm map}\nolimits _*(X,\,\Omega (K({\mathbb {Q}},\,n+1),\,*))=\mathop {\rm map}\nolimits _*(X,\,Y_n)$ (with basepoint $\zeta _n=p\circ \zeta$). In this sense, $\delta _n$ in (2.4) can be viewed as the identity morphism, and therefore $d^{(n-1)}$ is obtained by applying $\pi _p(\_,\,p\circ \zeta )$ to $k_*\colon \mathop {\rm map}\nolimits _*(X,\,K({\mathbb {Q}},\,n))\to \mathop {\rm map}\nolimits _*(X,\,K({\mathbb {Q}},\,2n))$ (see also (3.7)). For our identification $C_{p,q}\cong \check {H}^{q-p}(X,\,*;\pi _q)$, we need to precompose $k_*$ with the natural homotopy equivalence $\mathop {\rm map}\nolimits _*(X,\,K({\mathbb {Q}},\,n))\to \mathop {\rm map}\nolimits _*(X,\,K({\mathbb {Q}},\,n))$, induced by $\mu$, that sends the path component of $\mathop {\rm const}\nolimits _{\mathbb {1}}$ to that of $p\circ \zeta$. Thus, the nontrivial differential corresponds to the map induced on $\pi _p(\_,\,\mathop {\rm const}\nolimits _{\mathbb {1}})$ by
We view the latter as the morphism $[S^p\wedge X,\,K({\mathbb {Q}},\,n)]_*\to [S^p\wedge X,\,K({\mathbb {Q}},\,2n)]_*$ between sets of pointed homotopy classes. In light of the isomorphism $\check H^*(S^p\wedge X,\,*;{\mathbb {Q}})\cong \check H^*(S^p\times X,\,S^p\vee X;{\mathbb {Q}})$ and the split short exact sequence
we consider the composite
The cross product with a generator $E^p\in \check H^p(S^p,\,*;{\mathbb {Q}})=[S^p,\,K({\mathbb {Q}},\,p)]_*$ realizes the iterated suspension isomorphism $[X,\,K({\mathbb {Q}},\,n-p)]_*\to [(S^p\times X,\,S^p\vee X),\,(K({\mathbb {Q}},\,n),\,*)]$, hence we compute the effect of (5.2) on $E^p\vert \times x\vert$ where $E^p\vert \in \check H^p(S^p;{\mathbb {Q}})$ and $x\vert \in \check H^{n-p}(X;{\mathbb {Q}})$ are restrictions. For general $A\colon S^p\times X\to K({\mathbb {Q}},\,n)$ with $A(S^p\vee X)=\left \{\mathbb {1}\right \}$, the class of $A$ maps to $[k\circ \mu \circ (A,\,p\zeta \circ \mathop {\rm pr}\nolimits _X)]$; the latter can be viewed as
Here, $(p\zeta )^*(\iota _n^2)$ corresponds to $[k\circ p\zeta ]$ which is trivial. Finally, if $A^*(\iota _n)=E^p\vert \times x\vert$, we have $A^*(\iota _n^2)=(E^p\vert \times x\vert )^2=0$. Setting $I_n=p^*(\iota _n)\in \check H^n(Y;{\mathbb {Q}})$, we express
Thus, $d^{(n-1)}\colon \check H^{n-p}(X,\,*;{\mathbb {Q}})\to \check H^{2n-p}(X,\,*;{\mathbb {Q}})$ reads $\xi \mapsto 2\xi \cup \zeta ^*(I_n)$; termination after $d^{(n-1)}$ yields exact sequences
where all cohomology is based with rational coefficients and $\pi _p=\pi _p(\mathop {\rm map}\nolimits _*(X,\,Y);\zeta )$.
6. Proper domain
A useful approach to studying mapping spaces $\mathop {\rm map}\nolimits (X,\,Y)$ where $X$ is a CW complex is to view $\mathop {\rm map}\nolimits (X,\,Y)$ as the inverse limit of the system $\left \{\mathop {\rm map}\nolimits (K,\,Y)\right \}$ where $K$ ranges over the finite subcomplexes of $X$ and restriction mappings
are associated to subcomplex inclusions $K\leqslant L$. As the latter are closed cofibrations, the restriction mappings (6.1) are Hurewicz fibrations for any space $Y$.
We cannot expect a general paracompact domain space $X$ to be expressible as colimit of a suitable system of cofibrations. However, since we are interested in the homotopy type of $\mathop {\rm map}\nolimits (X,\,Y)$, it is enough to replace the target space $Y$ with a homotopy equivalent space in a way that renders restrictions (6.1) Hurewicz fibrations for sufficiently general compact pairs $(L,\,K)$. To this end, we replace $Y$ with a homotopy equivalent ANR; for example, we may take a simplicial complex with the metric topology. This can be done by Milnor [Reference Milnor22, Theorem 2].
We say that $A\subset X$ is $P$-embedded if continuous pseudo-metrics on $A$ extend to $X$. If, in addition, $A$ is a zero set, then it is called $P_0$-embedded. We refer to Stramaccia [Reference Stramaccia34] for more details; we need $P$-embeddings in the context of Morita's homotopy extension theorem used in the proof of proposition 6.1.
Recall that an absolute retract, with shorthand AR, is a contractible ANR.
Proposition 6.1 Let $(X,\,A)$ be a compact proper pair and let $Y$ be an ANR.
(i) Restriction $R\colon \mathop {\rm map}\nolimits (X,\,Y)\to \mathop {\rm map}\nolimits (A,\,Y)$ is a Hurewicz fibration.
(ii) Further let $B$ be an AR, $f\colon X\to B$ a given map, and $p\colon Y\to B$ a Hurewicz fibration. For $L\subset X$ denote $\Gamma (L)=\left \{s\colon L\to Y\,\vert \,p\circ s=f\vert _L\right \}\subset \mathop {\rm map}\nolimits (L,\,Y)$. Then $R\vert _{\Gamma (X)}\colon \Gamma (X)\to \Gamma (A)$ is a Hurewicz fibration.
Proof. As $R$ and $R\vert _{\Gamma (X)}$ are maps between metrizable spaces, it suffices to prove the homotopy lifting property for metrizable spaces (see [Reference Dugundji9, XX.2.3]).
To prove (i), we take a metric space $Z$. The solid part of the lifting diagram
induces a map $\eta \colon Z\times X\times \left \{0\right \}\cup Z\times A\times [0,\,1]\to Y$. Being the product of a paracompact and a compact space, $Z\times X$ is paracompact (Barratt [Reference Barratt1, Lemma 12.43]) and hence collection-wise normal. Therefore its $P_0$-embedded subsets are precisely its zero sets ([Reference Stramaccia34, Examples 1.2]). Thus, $Z\times A$ is $P_0$-embedded in $Z\times X$, and Morita's extension theorem (see [Reference Stramaccia34, Theorem 2.2]) yields a map $Z\times X\times [0,\,1]\to Y$ extending $\eta$. Its adjoint is the lifting sought in (6.2).
We turn to (ii). The lifting problem analogous to (6.2) but with $\mathop {\rm map}\nolimits (X,\,Y)$ and $\mathop {\rm map}\nolimits (A,\,Y)$ replaced by, respectively, $\Gamma (X)$ and $\Gamma (A)$, is equivalent to
By the proof of (i), $\eta$ extends to a map $\bar \eta \colon Z\times X\times [0,\,1]\to Y$. We remedy the possible difference between $p\circ \bar \eta$ and $f\circ \mathop {\rm pr}\nolimits _X$ (where $\mathop {\rm pr}\nolimits _X$ is projection onto $X$) as follows. Let a partial map into $B$ be defined by $p\circ \bar \eta$ on $Z\times X\times [0,\,1]\times \left \{0\right \}$ and $f\circ \mathop {\rm pr}\nolimits _X$ on $Z\times X\times [0,\,1]\times \left \{1\right \}\cup (Z\times X\times \left \{0\right \}\sqcup Z\times A\times [0,\,1])\times [0,\,1]$. This partial map is defined on a zero (hence $P_0$-embedded) set in $Z\times X\times [0,\,1]\times [0,\,1]$ and since $B$ is an AR, there exists an extension over all $Z\times X\times [0,\,1]\times [0,\,1]$ (see [Reference Stramaccia34, Theorem 2.1]). We consider the lifting problem
By construction, the homotopy into $B$ is stationary on $Z\times X\times \left \{0\right \}\cup Z\times A\times [0,\,1]$. By regularity of $p$ (see [Reference Dugundji9, XX.2.4]), the lifting into $Y$ can also be assumed to be stationary there. Level 1 of that lifting is a map $Z\times X\times [0,\,1]\to Y$ that extends $\eta$ and projects to $f\circ \mathop {\rm pr}\nolimits _X$ by $p$, and thus it constitutes the lifting sought in (6.3).
Corollary 6.2 Let $(X,\,A)$ be a proper pair with an admissible chain $A=X_0\subset X_1\subset X_2\subset \dots$, let $Y$ be an ANR, and let $\phi \colon A\to Y$ be a map. The induced tower
is one of Hurewicz fibrations and the terms $\mathop {\rm map}\nolimits _\phi (X_i,\,Y)$ have CW homotopy type. Restrictions $R^i\colon \mathop {\rm map}\nolimits _\phi (X,\,Y)\to \mathop {\rm map}\nolimits _\phi (X_i,\,Y)$, which are also Hurewicz fibrations, exhibit $\mathop {\rm map}\nolimits _\phi (X,\,Y)$ as the inverse limit.
Proof. The canonical inverse limit is the subspace $Z'\subset \prod _{i=1}^\infty \mathop {\rm map}\nolimits _\phi (X_i,\,Y)$ consisting of compatible sequences of maps $f_i\colon X_i\to Y$ with $f_i\vert _A=\phi$. Restrictions $R^i$ furnish the obvious homeomorphism $R\colon \mathop {\rm map}\nolimits _\phi (X,\,Y)\to Z'$; it is an embedding since the $X_i$ dominate compact sets in $X$ and onto because $X=\mathop {\rm colim}\nolimits _iX_i$. In diagram (4.4), restriction $\mathop {\rm map}\nolimits (C_i,\,Y)\to \mathop {\rm map}\nolimits (S_i,\,Y)$ is a fibration by (i) of proposition 6.1, and hence so is the pullback restriction $\mathop {\rm map}\nolimits _\phi (X_i,\,Y)\to \mathop {\rm map}\nolimits _\phi (X_{i-1},\,Y)$. The spaces $\mathop {\rm map}\nolimits (C_i,\,Y)$ and $\mathop {\rm map}\nolimits (S_i,\,Y)$ have CW homotopy type by [Reference Milnor22, Theorem 3]. Now $\mathop {\rm map}\nolimits _\phi (X_0,\,Y)$ is a point and it follows from [Reference Smrekar29, Proposition 4.2] that all $\mathop {\rm map}\nolimits _\phi (X_i,\,Y)$ have CW homotopy type.
By way of composition, $\mathop {\rm map}\nolimits _\phi (X_j,\,Y)\to \mathop {\rm map}\nolimits _\phi (X_i,\,Y)$ are fibrations for all $j\geqslant i$. The universal property of the limit implies that also the $R^i$ are fibrations.
Proof of theorem 1.6 Corollary 1.3 guarantees that the maps
localize homotopy groups. Corollary 6.2 identifies the inverse limit of maps (6.4) as $\mathop {\rm map}\nolimits _\phi (X,\,Y)\to \mathop {\rm map}\nolimits _{l\phi }(X,\,Y_{(P)})$. By [Reference Smrekar31, Theorem 6.1], the spaces $\mathop {\rm map}\nolimits (X,\,Y)$ and $\mathop {\rm map}\nolimits (X,\,Y_{(P)})$ are nilpotent and this latter map localizes homotopy groups.
Proof of proposition 1.8 Assume the notation of § 3. Diagram (3.3) exhibits $Z_q=\mathop {\rm map}\nolimits _{\phi _q}(X,\,Y_q)$ as the homotopy fibre of $(k_q)_*\colon Z_{q-1}\to \mathop {\rm map}\nolimits _{A\to *}(X,\,BK_q)$ where $BK_q$ is an Eilenberg–MacLane space. Now $\mathop {\rm map}\nolimits _{A\to *}(X,\,BK_q)$ has CW type since $(X,\,A)$ is quasi-finite, and as $Z_0=\left \{*\right \}$, an inductive application of [Reference Stasheff33, Proposition 0] shows that all $Z_q$ have CW homotopy type.
Set $Z^i=\mathop {\rm map}\nolimits _\phi (X_i,\,Y)$, $\zeta ^i=\zeta \vert _{X_i}$, $Z^i_q=\mathop {\rm map}\nolimits _\phi (X_i,\,Y_q)$, and $\zeta ^i_q=\zeta _q\vert _{X_i}$. By corollary 6.2, $Z=\mathop {\rm map}\nolimits _\phi (X,\,Y)$ is the limit of $Z^i$ and $Z_q$ is the limit of $Z^i_q$ for all $q$, and all $Z^i$ and $Z^i_q$ have CW homotopy type. To establish the (additional) assumption of corollary 2.7, we show that for every $p\geqslant 1$, restrictions $\pi _p(Z,\,\zeta )\to \pi _p(Z_q,\,\zeta _q)$ are injective for all big enough $q$. By [Reference Smrekar29, Corollary 3.4,(i)], applied to $Z=\lim _iZ^i$, there exists an $i$ such that $\pi _p(Z,\,\zeta )\to \pi _p(Z^i,\,\zeta ^i)$ are injective for all $p\geqslant 1$. Fix some $p\geqslant 1$. By corollary 4.3, $\pi _p(Z^i,\,\zeta ^i)\to \pi _p(Z^i_q,\,\zeta ^i_q)$ is an isomorphism for all big enough $q$. Thus, $\pi _p(Z,\,\zeta )\to \pi _p(Z_q,\,\zeta _q)$ must be injective for those $q$.
To establish the assumptions of proposition 2.6, fix some $q$. By another application of [Reference Smrekar29, Corollary 3.4,(i)], this time to $Z_q=\lim _iZ^i_q$, there exists an $i$ such that $\pi _p(Z_q,\,\zeta _q)\to \pi _p(Z^i_q,\,\zeta ^i_q)$ are injective for all $p\geqslant 1$. By [Reference Smrekar29, Corollary 3.4,(ii)], applied to $Z=\lim _iZ^i$, there exists a $j\geqslant i$ such that the image of $\pi _p(Z,\,\zeta )\to \pi _p(Z^i,\,\zeta ^i)$ coincides with that of $\pi _p(Z^j,\,\zeta ^j)\to \pi _p(Z^i,\,\zeta ^i)$ for all $p$. Finally, we fix some $p\geqslant 1$. By corollary 4.3, there exists a $t$ such that $\pi _p(Z^j,\,\zeta ^j)\to \pi _p(Z^j_t,\,\zeta ^j_t)$ is an isomorphism. A straightforward diagram chase shows that every element in the image of $\pi _p(Z_t,\,\zeta _t)\to \pi _p(Z_q,\,\zeta _q)$ is also in the image of $\pi _p(Z,\,\zeta )\to \pi _p(Z_q,\,\zeta _q)$.
Proof of proposition 1.9 By homotopy invariance of $\mathop {\rm map}\nolimits _{A\to *}(X,\,Y)$ for a well-pointed $Y$, we may assume that $Y=K(G,\,n)$ is an H-group with strict unit $\mathbb {1}$ and an ANR. If $\left \{X_i\right \}$ is any admissible chain for $(X,\,A)$, then by corollary 6.2, $Z^i=\mathop {\rm map}\nolimits _{A\to \mathbb {1}}(X_i,\,Y)$ defines a tower of fibrations whose terms have CW homotopy type and whose limit is $Z=\mathop {\rm map}\nolimits _{A\to \mathbb {1}}(X,\,Y)$. Homotopical representation of Čech cohomology directly implies that restriction $Z\to Z^i$ induces the cohomological restriction $\check H^n(X,\,A;G)\to \check H^n(X_i,\,A;G)$.
Let $\zeta \colon X\to Y$ be a map with restrictions $\zeta ^i=\zeta \vert _{X_i}$. Pointwise multiplications by $\zeta$, respectively $\zeta ^i$, comprise a homotopy auto-equivalence of the tower $\left \{Z^i\right \}$ and its limit that maps the components of $\mathop {\rm const}\nolimits _{\mathbb {1}}$ to those of $\zeta$, respectively $\zeta ^i$. Thus we can identify the morphisms $\pi _k(Z^j,\,\zeta ^j)\to \pi _k(Z^i,\,\zeta ^i)$ with $\check H^{n-k}(X_j,\,A;G)\to \check H^{n-k}(X_i,\,A;G)$. An application of [Reference Smrekar31, Theorem 4.1] concludes the proof.
Example 6.3 Assume the notation of definition 1.5 for a proper space $X$ and set
We claim that if $c(i)\geqslant n+2$ for $i\geqslant i_0$, then restriction $\mathop {\rm map}\nolimits (X,\,K(G,\,n))\to \mathop {\rm map}\nolimits (X_{i_0},\,K(G,\,n))$ is a homotopy equivalence for every abelian group $G$. To verify this claim, note first that if $i\geqslant i_0$, universal coefficients imply $\check H^k(C_{i+1},\,S_i;G)=0$ for any $G$ and all $k\leqslant n+1$. Next, Wallace's theorem implies cohomology isomorphisms $\check H^*(X_{i+1},\,X_i;G)\to \check H^*(C_{i+1},\,S_i;G)$. Arguing as in the above proof, we use the cohomology exact sequence of the pair $(X_{i+1},\,X_i)$ to infer that if $i\geqslant i_0$, then restriction $\mathop {\rm map}\nolimits (X_{i+1},\,K(G,\,n))\to \mathop {\rm map}\nolimits (X_{i},\,K(G,\,n))$ is a homotopy equivalence. It follows that so also is $\mathop {\rm map}\nolimits (X,\,K(G,\,n))\to \mathop {\rm map}\nolimits (X_{i_0},\,K(G,\,n))$ (see [Reference Smrekar29, Corollary 3.7]), rendering $\mathop {\rm map}\nolimits (X,\,K(G,\,n))$ homotopy equivalent to a CW complex. In particular, if $\liminf _{i\to \infty }c(i)=\infty$, then $\mathop {\rm map}\nolimits (X,\,K(G,\,n))$ has CW homotopy type for every $G$ and $n$. For a simple-minded concrete example of a proper quasi-finite space that does not have the homotopy type of a CW complex, one can take the infinite wedge $X=\vee _{n=1}^\infty {\mathbb {H}}_n$ where ${\mathbb {H}}_n$ is the $n$-dimensional Hawaiian earring.
7. Spaces of sections
Let $X\in {\rm PHK}$ be connected and let $\pi \colon Q\to X$ be a Hurewicz fibration with typical fibre of the homotopy type of a nilpotent CW complex. Let $\Gamma (\pi )$ denote the space of sections of $\pi$. We assume that $\pi =f^*(p')$ is the pullback of a fibration $p'\colon E'\to B$ of spaces of CW homotopy type along the continuous map $f\colon X\to B$. We can change $\pi$ up to fibre homotopy equivalence (over $X$) and keep the homotopy type of $\Gamma (\pi )$. Thus we assume that $B$ is in fact a simplicial complex with the strong topology; $B$ is then an ANR with a covering of contractible open subsets (which are ANRs), see Milnor [Reference Milnor22]. The corresponding total space $E''$ will still have CW homotopy type by results of Stasheff [Reference Stasheff33]. Choosing a homotopy equivalent ANR $E'''$ we split the resulting map $E'''\to B$ into the composite of a homotopy equivalence and a fibration $p\colon E\to B$ where $E$ is an ANR. The ‘usual’ splitting in the sense of the Hurewicz model structure has this property; see also Miyata [Reference Miyata23]. Consequently, each fibre of $p$ is an ANR (being cofibred in $E$ by [Reference Strøm35, Theorem 12]) and we will use $Y$ to denote the fibre over a point understood from the context. As $\pi$ is fibre homotopy equivalent to $f^*(p)$, we identify $\Gamma (\pi )$ with the space of maps $X\to E$ whose composite with $p$ equals $f$.
Remark 7.1 In the fibration category on ANRs studied by [Reference Miyata23], the fibrations are what we call regular Hurewicz fibrations (and the author calls maps with the strong homotopy lifting property). As remarked above, every Hurewicz fibration $E\to B$ between ANRs is regular on the ground of metrizability of $B$; this remark makes certain parts of [Reference Miyata23] superfluous.
More generally, assume that $(X,\,A)$ is a proper pair with an admissible chain $\left \{X_i\right \}$ and that $\phi \colon A\to E$ is a map over $f\vert _A$. For $A\subset L\subset X$, we set
viewed as a subspace of $\mathop {\rm map}\nolimits (L,\,E)$. For $A\subset K\subset L$ we have a restriction map $\Gamma _\phi (L,\,E)\to \Gamma _\phi (K,\,E)$. As noted above, if $A=\emptyset$, then the homotopy type of $\Gamma (X,\,E)$ is an invariant of the homotopy type of $p\colon E\to B$. Invariance in the relative version will be discussed after the following result.
Theorem 7.2
(i) Restriction maps $\Gamma _\phi (X_i,\,E)\to \Gamma _\phi (X_{i-1},\,E)$ are fibrations between spaces of CW homotopy type.
(ii) Let $l\colon E\to E_{(P)}$ be a fibrewise localization at the set of primes $P$ where we assume that $E_{(P)}$ is an ANR. If $(X,\,A)$ is locally finite dimensional or $Y$ is a Postnikov section, then $\Gamma _\phi (X_{i},\,E)$ are nilpotent and $\Gamma _\phi (X_{i},\,E)\to \Gamma _\phi (X_{i},\,E_{(P)})$ are localizations at $P$ on path components.
Proof. (i) By the assumption on $B$ and normality of the $C_i$ in (1.2), we may refine each step $X_{i-1}\subset X_i$ into a finite chain of adjunctions to assume, after reindexing, that $f(\gamma _i(C_i))\subset K_i$ for a contractible open set $K_i$ in $B$. We get induced pullbacks
By proposition 6.1, the vertical arrow on the right is a fibration, and hence so is the one on the left. Consider $p_*\colon \mathop {\rm map}\nolimits (C_i,\,E\vert _{K_i})\to \mathop {\rm map}\nolimits (C_i,\,K_i)$. As $C_i$ is compact, this is a fibration between spaces of CW homotopy type. In fact $\mathop {\rm map}\nolimits (C_i,\,K_i)$ is contractible and hence fibre inclusion $\Gamma (C_i,\,E\vert _{K_i})\hookrightarrow \mathop {\rm map}\nolimits (C_i,\,E\vert _{K_i})$ is a homotopy equivalence. On the other hand, $Y\hookrightarrow E\vert _{K_i}$ is also a homotopy equivalence, hence $\mathop {\rm map}\nolimits (C_i,\,Y)\simeq \Gamma (C_i,\,E\vert _{K_i})$ is nilpotent (under the additional hypothesis) by theorem 1.1. The same holds for $S_i$ in place of $C_i$ and since $\Gamma _\phi (X_0,\,E)$ is a one point-space, it follows by induction that all $\Gamma _\phi (X_{i},\,E)$ are nilpotent (see May and Ponto [Reference May and Ponto19, Proposition 4.4.3]) of CW type (see Stasheff [Reference Stasheff33, Propositions 0 and 12]).
(ii) We apply the same reasoning to $\Gamma _{l\phi }(X_i,\,E_{(P)})$ and note that $\Gamma (C_i,\,E\vert _{K_i})\to \Gamma (C_i,\,E_{(P)}\vert _{K_i})$ and $\Gamma (S_i,\,E\vert _{K_i})\to \Gamma (S_i,\,E_{(P)}\vert _{K_i})$ may be identified with, respectively, $\mathop {\rm map}\nolimits (C_i,\,Y)\to \mathop {\rm map}\nolimits (C_i,\,Y_{(P)})$ and $\mathop {\rm map}\nolimits (S_i,\,Y)\to \mathop {\rm map}\nolimits (S_i,\,Y_{(P)})$ that localize by corollary 1.3. Thus if $\Gamma _\phi (X_{i-1},\,E)\to \Gamma _\phi (X_{i-1},\,E_{(P)})$ localizes (on path components), then so does $\Gamma _\phi (X_{i},\,E)\to \Gamma _\phi (X_{i},\,E_{(P)})$ by [Reference May and Ponto19, Proposition 6.2.5].
We turn to the question of homotopy invariance of $\Gamma _\phi (X,\,E)$ under changes of $p\colon E\to B$ for general $X\in {\rm PHK}$ with closed subspace $A$. To this end, assume that $p$ is homotopy equivalent to $p'\colon E'\to B'$, i.e., there are homotopy equivalences $\eta \colon E\to E'$ and $\beta \colon B\to B'$ for which $p'\eta =\beta p$. Think of
as the restriction to the fibre of the transformation $p_*\to p'_*$:
Assume that $p_*$ and $p_*'$ are fibrations. If $\eta _*$ and $\beta _*$ are homotopy equivalences, so is (7.2) by coglueing [Reference Brown and Heath4, Corollary 1.5]. This will be true in the following cases.
(i) $\phi =\mathop {\rm const}\nolimits _{e_0}$ (forcing $f\vert _A=\mathop {\rm const}\nolimits _{b_0}$ where $b_0=p(e_0)$) and $p,\,p',\,\eta,\,\beta$ are pointed maps of well-pointed spaces,
(ii) $\phi =\mathop {\rm const}\nolimits _{e_0}$ and $p$ and $p'$ are regular fibrations (for example if $B$ and $B'$ are metrizable) and $\eta$ and $\beta$ are homotopy equivalences in the pointed category,
(iii) $E,\,E',\,B,\,B'$ are ANRs and $(X,\,A)$ is a proper pair,
(iv) $A$ is cofibred (in particular, $A$ may be empty).
For (1), the proof of [Reference Smrekar29, Lemma A.4] shows that $p_*$ and $p_*'$ are fibrations, while if $\eta '\colon E'\to E$ is a pointed inverse for $\eta$ and $\eta '\eta \simeq \mathop {\rm id}\nolimits _E$ and $\eta \eta '\simeq \mathop {\rm id}\nolimits _{E'}$ are pointed homotopies, then the standard proof that $\eta _*\colon \mathop {\rm map}\nolimits (X,\,E)\to \mathop {\rm map}\nolimits (X,\,E')$ is a homotopy equivalence (as for example [Reference Maunder18, Theorem 6.2.25]) will show its restriction $\mathop {\rm map}\nolimits (X,\,E)_{A\to e_0}\to \mathop {\rm map}\nolimits (X,\,E')_{A\to \eta (e_0)}$ is also a homotopy equivalence.
For (2), the proof of [Reference Smrekar29, Lemma A.4] may be amended using regular lifting functions (as in Definition 2.1 in [Reference Dugundji9, Chapter XX]) to show that $p_*$ and $p_*'$ are fibrations, while the proof that $\eta _*$ and $\beta _*$ are homotopy equivalences is as above.
For (3), we note that $p_*$ and $p_*'$ are fibrations because $p$ and $p'$ are regular on the ground of metrizability of $B$. We consider $\eta _*\colon \mathop {\rm map}\nolimits _\phi (X,\,E)\to \mathop {\rm map}\nolimits _{\eta \phi }(X,\,E')$ as the map between the fibres induced by the transformation of restrictions:
The horizontal arrows in (7.4) are homotopy equivalences ([Reference Maunder18, ibid.]) and since $E$ and $E'$ are ANRs, the restrictions (i.e., the vertical arrows) are fibrations by proposition 6.1. Thus, the coglueing theorem will guarantee that the maps between the fibres will be homotopy equivalences as well. The same argument applies to $\beta _*$, using the fact that $B$ and $B'$ are ANRs.
For (4), the argument that $p_*$ and $p_*'$ are Hurewicz fibrations is standard. In ${\mathcal {K}}$, we simply use the exponential law in ${\mathcal {K}}$, the fact that inclusion $X\times \left \{0\right \}\cup A\times [0,\,1]\hookrightarrow X\times [0,\,1]$ is a closed cofibration and the lifting properties of $p_*$ and $p_*'$. For ${\mathcal {T}}\!\!{op}$, we extend the proof of [Reference Smrekar29, Lemma A.4]. When $A$ is cofibred in $X$, the vertical arrows in (7.4) are fibrations for all $E$ and $E'$.
Proof of theorem 1.10 The theorem deals with absolute sections ($A=\emptyset$), and we may take a representative for $p$ as specified prior to the statement of theorem 7.2. Thus, statement (a) follows directly from theorem 7.2 as then $X=X_i$ for some $i$.
For (b), note that $\Gamma (X,\,E)\to \Gamma (X,\,E_{(P)})$ is the limit of maps $\Gamma (X_i,\,E)\to \Gamma (X_i,\,E_{(P)})$ (which localize by theorem 7.2). Apply [Reference Smrekar31, Theorems 6.1 and 6.2].
We remark that theorem 7.2 also contains [Reference Møller24, Theorem 5.3] as a special case.
Acknowledgements
The author was supported in part by the Slovenian Research Agency grants N1-0237 and P1-0291.