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

On the Whitehead theorem for nilpotent motivic spaces

Aravind Asok Asok was partially supported by National Science Foundation Awards DMS-1802060 and DMS-2101898    Tom Bachmann    Michael J. Hopkins Hopkins was partially supported by National Science Foundation Award DMS-1810917
Abstract

We improve some foundational connectivity results and the relative Hurewicz theorem in motivic homotopy theory, study functorial central series in motivic local group theory, establish the existence of functorial Moore–Postnikov factorizations for nilpotent morphisms of motivic spaces under a mild technical hypothesis and establish an analog of the Whitehead theorem for nilpotent motivic spaces. As an application, we deduce a surprising unstable motivic periodicity result.

1 Introduction

In classical algebraic topology, Whitehead proved that a map of simply connected spaces that induces an isomorphism on homology is a weak equivalence. This result was generalized in various directions: Dror introduced the notion of a nilpotent space and established that a map of nilpotent spaces that is a homology isomorphism is necessarily a weak equivalence [DF71, Theorem 3.1]. In modern terminology, these results establish that the stabilization functor (from unstable to stable homotopy theory) is conservative on the subcategory of nilpotent spaces.

In this work, we analyze versions of these results in motivic homotopy theory. Assume kk is a field and write Smk\mathrm{Sm}_{k} for the category of smooth kk-schemes. The \infty-category of motivic spaces Spc(k)\mathrm{Spc}({k}) is a (left Bousfield) localization of the \infty-category of presheaves of spaces on Smk\mathrm{Sm}_{k}. Following [AFH22, Definition 3.2.1], we say that a motivic space is nilpotent if its fundamental (Nisnevich) sheaf of groups is 𝔸1{\mathbb{A}}^{1}-nilpotent and acts 𝔸1{\mathbb{A}}^{1}-nilpotently on the higher homotopy sheaves (we called such spaces 𝔸1{\mathbb{A}}^{1}-nilpotent in [AFH22]; our terminology differs slightly from that used in the previous paper and we recall in the notation and conventions below).

In the context of motivic homotopy theory, nilpotent spaces arise more readily than 11-connected spaces. Indeed, the only 11-connected smooth proper variety over a field is Speck\operatorname{Spec}k [AM11, Proposition 5.1.4]. On the other hand, 1{\mathbb{P}}^{1} or, more generally, flag varieties for split simply connected semi-simple algebraic groups are nilpotent [AFH22, Theorem 3.4.8] and have non-abelian fundamental groups! Likewise, affine Grassmannians of split, simply connected, semi-simple group schemes are nilpotent by [Bac19, Theorem 15] in conjunction with [AFH22, Theorem 3.3.17, Example 3.4.1].

In this note, we improve our techniques for dealing with nilpotent motivic spaces from [AFH22]. Before treating nilpotent motivic spaces in general, we further expand the arsenal of tools for dealing with sheaves of groups whose classifying spaces are motivic local: these are F. Morel’s “strongly 𝔸1{\mathbb{A}}^{1}-invariant sheaves of groups. One key structural deficit of the category of strongly 𝔸1{\mathbb{A}}^{1}-invariant sheaves of groups is that given an epimorphism of sheaves of groups whose source is strongly 𝔸1{\mathbb{A}}^{1}-invariant and whose target is 𝔸1{\mathbb{A}}^{1}-invariant, it is unclear whether the target is automatically strongly 𝔸1{\mathbb{A}}^{1}-invariant. We axiomatize this property with the notion of a very strongly 𝔸1{\mathbb{A}}^{1}-invariant sheaf of groups (see Definition 2.6). We then show that the class of very strongly 𝔸1{\mathbb{A}}^{1}-invariant sheaves of groups contains many examples of interest.

We adapt various functorial constructions of classical group theory to the motivic setting. We show that actions of 𝔸1{\mathbb{A}}^{1}-nilpotent sheaves of groups have functorial central series (Proposition 4.16), we improve the relative Hurewicz theorem from [AFH22, Theorem 4.2.1] (Theorem 3.6), we show that nilpotent morphisms of motivic spaces have functorial principal refinements (Theorem 5.1) under mild hypotheses, and we demonstrate the Whitehead theorem for nilpotent motivic spaces (see Theorem 5.2).

In order to state our results, we use the S1S^{1}-stable motivic homotopy category SHS1(k){\mathrm{SH}}^{S^{1}}(k) as described by F. Morel [Mor05, §4]. Recall that there is a stabilization functor ΣS1:Spc(k)SHS1(k)\Sigma^{\infty}_{S^{1}}:\mathrm{Spc}({k})\to{\mathrm{SH}}^{S^{1}}(k) which is Kan extended from the functor sending XSmkX\in\mathrm{Sm}_{k} to its suspension spectrum ΣS1X+\Sigma^{\infty}_{S^{1}}X_{+}, where X+X_{+} is XX with a disjoint base-point attached.

Theorem 1 (See Theorem 5.2).

If kk is a perfect field and f:𝒳𝒴f:\mathscr{X}\to\mathscr{Y} is a morphism of nilpotent motivic kk-spaces such that ΣS1f\Sigma^{\infty}_{S^{1}}f is an equivalence, then ff is an equivalence.

In the body of the text, we actually prove a version of this result with explicit connectivity bounds: loosely, if n0n\geq 0 is any integer, and ΣS1f\Sigma^{\infty}_{S^{1}}f has nn-connected fiber, then ff also has nn-connected fiber. In the classical setting, the corresponding result when n=1n=1 follows from a result of Stallings [Sta65, Theorem 2.1]. Dror’s proof of the Whitehead theorem generalizes Stallings’ approach. Actually, Dror’s result established more than the Whitehead theorem for nilpotent spaces, and an even simpler proof was given slightly later by Gersten [Ger75].

Stallings’ proof uses the Serre spectral sequences, but he suggests a way to avoid their use (though remarks the advantage of such an approach would be “Pyrrhic”). For nilpotent spaces, it is possible to give a proof avoiding the Serre spectral sequence by combining connectivity estimates, the relative Hurewicz theorem and functorial Postnikov resolutions. Unfortunately, the known versions of the Serre spectral sequence in motivic homotopy theory are not useful in our context and so we follow the more hands-on approach suggested in the previous paragraph. We establish and appeal to a mild improvement of the Blakers–Massey theorem in motivic homotopy theory relying on a motivic analog of a result of Ganea [Gan65]. As such, our approach to the Whitehead theorem is undoubtedly related to the treatment in [Too76].

We close with some applications of the Whitehead theorem. In motivic homotopy theory, there are many benefits to being able to check whether a map is a weak equivalence stably. In particular, working stably, one can make use of geometric constructions that allow singularity formation (e.g., by working in the cdhcdh-topology) via a theorem of Voevodsky. As a concrete application of this principle, we establish the following result.

Write HP\mathrm{HP}^{\infty} for the Panin–Walter [PW21] model for BSL2BSL_{2}. The “inclusion of the bottom cell” HP1HP\mathrm{HP}^{1}\hookrightarrow\mathrm{HP}^{\infty} (classifying the Hopf bundle ν\nu) yields a map S4,2HPS^{4,2}\to\mathrm{HP}^{\infty}, which is adjoint to a map S3,1Ω1,1HPS^{3,1}\longrightarrow\Omega^{1,1}\operatorname{HP}^{\infty}. The corresponding map in classical homotopy theory is evidently an equivalence, and the analog of this map in C2C_{2}-equivariant homotopy theory is an equivalence by [HW20, Proposition 3.6]. The next result, which we found surprising because of the “weight shifting” phenomena it displays, asserts that this map is a motivic equivalence; the aforementioned “classical” statements follow immediately by appeal to suitable realizations.

Theorem 2 (See Theorem 6.7).

If kk has characteristic 0, then the map

S3,1Ω1,1HPS^{3,1}\longrightarrow\Omega^{1,1}\operatorname{HP}^{\infty}

deescribed above is an equivalence.

Here, the introduction of singularities arises by our appeal to affine Grassmannians. The affine Grassmannian GrG\operatorname{Gr}_{G} for GG a split, semi-simple, simply connected algebraic group is an increasing union of singular algebraic varieties. Moreover, a motivic variant of a result due to Quillen [Mit88] and Garland-Raghunathan [GR75] due to the second author [Bac19, Theorem 15] implies that GrG\operatorname{Gr}_{G} has the structure of an hh-space. The map from Theorem 2 is obtained from a map of hh-spaces after looping and we use the Whitehead theorem above to check that map is an equivalence.

In another direction, using the full strength of our Whitehead theorem, Proposition 6.8 allows to construct some EHP-sequences for motivic spheres outside the range considered by Wickelgren and Williams [WW19]. We use this exceptional EHP sequence to obtain Corollary 6.10, which provides a motivic analog of F. Cohen’s exceptional 22-local fiber sequence [Coh87, Theorem 3.3] (which he attributes to H. Toda).

Notation, conventions and relation to other work

Assume kk is a field; we will frequently assume kk is perfect. Write Smk\mathrm{Sm}_{k} for the category of smooth kk-schemes. We write P(Smk)\mathrm{P}(\mathrm{Sm}_{k}) for the \infty-category of presheaves of spaces on Smk\mathrm{Sm}_{k}, and ShvNis(Smk)P(Smk){\mathrm{Shv}}_{{\operatorname{Nis}}}(\mathrm{Sm}_{k})\subset\mathrm{P}(\mathrm{Sm}_{k}) for the full subcategory of Nisnevich sheaves of spaces on Smk\mathrm{Sm}_{k}, which is an \infty-topos. There are corresponding pointed versions of all these constructions: we write P(Smk)\mathrm{P}(\mathrm{Sm}_{k})_{*} for the \infty-category of presheaves of pointed spaces and ShvNis(Smk){\mathrm{Shv}}_{{\operatorname{Nis}}}(\mathrm{Sm}_{k})_{*} for the subcategory of Nisnevich sheaves of pointed spaces. We write SiS^{i} for the simplicial ii-circle, and we set Sp,q:=Spq𝔾mqS^{p,q}:=S^{p-q}\wedge{{\mathbb{G}}_{m}^{{\scriptstyle{\wedge}\,}q}}. We then write Σ\Sigma for S1S^{1}\wedge on P(Smk)\mathrm{P}(\mathrm{Sm}_{k})_{*}. We will write 𝝅iNis\bm{\pi}_{i}^{{\operatorname{Nis}}} for the (Nisnevich) homotopy sheaves of any 𝒳P(Smk)\mathscr{X}\in\mathrm{P}(\mathrm{Sm}_{k})_{*}.

Suppose f:𝒳𝒴f:\mathscr{X}\to\mathscr{Y} is a morphism in ShvNis(Smk){\mathrm{Shv}}_{{\operatorname{Nis}}}(\mathrm{Sm}_{k}). We shall say that ff has nn-connected fibers if 𝒳\mathscr{X} is nn-connected as an object of ShvNis(Smk)/𝒴{\mathrm{Shv}}_{\operatorname{Nis}}(\mathrm{Sm}_{k})_{/\mathscr{Y}}. For example, if 𝒴\mathscr{Y} is connected, this is the same as requiring that fib(f)ShvNis(Smk)\mathrm{fib}(f)\in{\mathrm{Shv}}_{\operatorname{Nis}}(\mathrm{Sm}_{k}) be nn-connected.

This paper may be viewed as a prequel to [ABH23] and the results established here will be used there. As is perhaps clear from the discussion above, and in contrast to [AFH22], this paper is written using the language of \infty-categories. As such, terms like limit and colimit or fiber and cofiber will be used in the \infty-categorical sense in contrast to their usage in [AFH22]. The notation and terminology here are chosen to be consistent with [ABH23].

Remark 3.

A previous version of this paper relied on [CH22, Theorem 1.4] asserting that a quotient of a strongly 𝔸1{\mathbb{A}}^{1}-invariant sheaf of groups is 𝔸1{\mathbb{A}}^{1}-invariant if and only if it is strongly 𝔸1{\mathbb{A}}^{1}-invariant. However, at the moment, the proof of [CH22, Lemma 2.9] contains a gap, which renders the proof of [CH22, Theorem 1.4] incomplete. This note circumvents appeal to [CH22] at the expense of some slight additional technical hypotheses which are satisfied in all cases where we will make use of them.

2 Motivic localization and group theory

A presheaf 𝒳P(Smk)\mathscr{X}\in\mathrm{P}(\mathrm{Sm}_{k}) is 𝔸1{\mathbb{A}}^{1}-invariant if for every USmkU\in\mathrm{Sm}_{k} the projection U×𝔸1UU\times{\mathbb{A}}^{1}\to U induces an equivalence 𝒳(U)𝒳(U×𝔸k1)\mathscr{X}(U)\to\mathscr{X}(U\times{\mathbb{A}}^{1}_{k}). Similarly, 𝒳\mathscr{X} is said to be Nisnevich excisive if 𝒳()\mathscr{X}(\emptyset) is contractible and 𝒳\mathscr{X} takes Nisnevich distinguished squares to cartesian squares. The \infty-category of motivic spaces Spc(k)\mathrm{Spc}({k}) is the full subcategory of P(Smk)\mathrm{P}(\mathrm{Sm}_{k}) spanned by 𝔸1{\mathbb{A}}^{1}-invariant and Nisnevich excisive spaces. Since homotopy invariance and Nisnevich excision are defined by a small set of conditions, the inclusion Spc(k)P(Smk)\mathrm{Spc}({k})\subset\mathrm{P}(\mathrm{Sm}_{k}) is an accessible localization. The category Spc(k)\mathrm{Spc}({k}) is a presentable \infty-category and the inclusion Spc(k)P(Smk)\mathrm{Spc}({k})\subset\mathrm{P}(\mathrm{Sm}_{k}) admits a left adjoint

Lmot:P(Smk)Spc(k)\mathrm{L}_{mot}:\mathrm{P}(\mathrm{Sm}_{k})\longrightarrow\mathrm{Spc}({k})

that we call the motivic localization functor.

The functor Lmot\mathrm{L}_{mot} is known to preserve finite products [Hoy14, Proposition C.6]. We write 𝝅i(𝒳)\bm{\pi}_{i}(\mathscr{X}) for the homotopy sheaves of a motivic space (the notation 𝝅i𝔸1()\bm{\pi}_{i}^{{\mathbb{A}}^{1}}(-) is frequently used in previous work). If ff is a map in Spc(k)\mathrm{Spc}({k}), we write fib(f)\mathrm{fib}(f) and cof(f)\mathrm{cof}(f) for its fiber and cofiber. The next result records the key property of the motivic localization functor.

Theorem 2.1 (Morel).

Suppose n2n\geq-2 is an integer and assume kk is a field and 𝒳\mathscr{X} is a space pulled back from a perfect subfield of kk. If 𝒳\mathscr{X} is nn-connected, then Lmot𝒳\mathrm{L}_{mot}\mathscr{X} is also nn-connected.

Proof.

The statement is trivial for n<0n<0, and for n=0n=0 follows essentially from the construction of the motivic localization functor [MV99, §2 Corollary 3.22]. For n>0n>0 it is [Mor12, Theorem 6.38] (though see [AWW17, Theorem 2.2.12] for a detailed proof). ∎

Key to deducing the above results are the notions of strictly and strongly 𝔸1{\mathbb{A}}^{1}-invariant sheaves of groups.

Definition 2.2.

A Nisnevich sheaf of groups 𝐆\mathbf{G} on Smk\mathrm{Sm}_{k} is strongly 𝔸1{\mathbb{A}}^{1}-invariant if the cohomology presheaves Hi(,𝐆)H^{i}(-,\mathbf{G}) are 𝔸1{\mathbb{A}}^{1}-invariant for i=0,1i=0,1. A Nisnevich sheaf of abelian groups is strictly 𝔸1{\mathbb{A}}^{1}-invariant if the (Nisnevich) cohomology presheaves Hi(,𝐀)H^{i}(-,\mathbf{A}) are 𝔸1{\mathbb{A}}^{1}-invariant for all i0i\geq 0.

Write Grpk𝔸1\mathrm{Grp}^{{\mathbb{A}}^{1}}_{k} for the category of strongly 𝔸1{\mathbb{A}}^{1}-invariant sheaves of groups and Abk𝔸1\mathrm{Ab}^{{\mathbb{A}}^{1}}_{k} for the category of strictly 𝔸1{\mathbb{A}}^{1}-invariant sheaves. We use the following result without mention in the sequel.

Proposition 2.3.

Assume kk is a field. If 𝐆\mathbf{G} is a Nisnevich sheaf of groups on Smk\mathrm{Sm}_{k}, then 𝐆\mathbf{G} is strongly 𝔸1{\mathbb{A}}^{1}-invariant if and only if B𝐆B\mathbf{G} is motivic local.

Proof.

This statement is asserted without proof in [Mor12, Remark 1.8] and claims exist (e.g., on MathOverflow) that the two notions are not evidently equivalent because of base-point issues. For any Nisnevich sheaf of groups 𝐆\mathbf{G} we know that B𝐆B\mathbf{G} is 11-truncated and π1i(B𝐆())Hi(,𝐆)\pi_{1-i}(B\mathbf{G}(-))\cong H^{i}(-,\mathbf{G}) (see, for example [MV99, Proposition 4.1.16]). Thus, if B𝐆B\mathbf{G} is motivic local, then 𝐆\mathbf{G} is evidently strongly 𝔸1{\mathbb{A}}^{1}-invariant. Conversely, if 𝐆\mathbf{G} is strongly 𝔸1{\mathbb{A}}^{1}-invariant, then we must show that B𝐆(U)B𝐆(U×𝔸1)B\mathbf{G}(U)\to B\mathbf{G}(U\times{\mathbb{A}}^{1}) is an equivalence for any smooth kk-scheme UU. The remark above shows that the induced map on π0\pi_{0} is a bijection. It remains to prove that for any smooth kk-scheme UU, and any base-point uπ0(B𝐆(U))u\in\pi_{0}(B\mathbf{G}(U)) the induced map π1(B𝐆(U),u)π1(B𝐆(U×𝔸1),u)\pi_{1}(B\mathbf{G}(U),u)\to\pi_{1}(B\mathbf{G}(U\times{\mathbb{A}}^{1}),u) (where we have abused notation for the base-point in B𝐆(U×𝔸1)B\mathbf{G}(U\times{\mathbb{A}}^{1})) is a bijection. This map is once again a bijection when uu corresponds to the trivial torsor by the definition of strong 𝔸1{\mathbb{A}}^{1}-invariance. However, π1\pi_{1} is a sheaf in the Nisnevich topology (it can be described explicitly as the automorphism group scheme of the 𝐆\mathbf{G}-torsor corresponding to uu; see, for example, [Koi22, Corollary 3.6]), isomorphisms of sheaves can be checked locally. Thus, by choosing a Nisnevich cover VUV\to U trivializing uu, we reduce to the case of the base-point corresponding to the trivial torsor, which we have already treated. ∎

If 𝒳\mathscr{X} is a motivic space, F. Morel defined 𝔸1{\mathbb{A}}^{1}-homology sheaves 𝐇i𝔸1(𝒳){{\mathbf{H}}}_{i}^{{\mathbb{A}}^{1}}(\mathscr{X}) [Mor12, Definition 6.29]; these sheaves are strictly 𝔸1{\mathbb{A}}^{1}-invariant. If 𝒳\mathscr{X} is a pointed motivic space, then Morel showed the homotopy sheaves 𝝅i(𝒳)\bm{\pi}_{i}(\mathscr{X}) are strongly 𝔸1{\mathbb{A}}^{1}-invariant for i1i\geq 1. There is a forgetful functor from Abk𝔸1\mathrm{Ab}^{{\mathbb{A}}^{1}}_{k} to the category of strongly 𝔸1{\mathbb{A}}^{1}-invariant sheaves of groups; this functor is fully faithful. If kk is perfect, Morel showed [Mor12, Corollary 5.45] that strongly 𝔸1{\mathbb{A}}^{1}-invariant sheaves of abelian groups are strictly 𝔸1{\mathbb{A}}^{1}-invariant, i.e., the forgetful functor Abk𝔸1\mathrm{Ab}^{{\mathbb{A}}^{1}}_{k} to strongly 𝔸1{\mathbb{A}}^{1}-invariant sheaves of abelian groups is an equivalence.

The category of strictly 𝔸1{\mathbb{A}}^{1}-invariant sheaves is abelian [Mor05, Lemma 6.2.13] as the heart of a tt-structure on a suitable triangulated category. In contrast, while the category Grpk𝔸1\mathrm{Grp}^{{\mathbb{A}}^{1}}_{k} has colimits (e.g. cokernels) for formal reasons, computing them seems tricky. To remedy this defect, we recall some ideas from [BHS15].

Definition 2.4.

For any 𝒳P(Smk)\mathscr{X}\in\mathrm{P}(\mathrm{Sm}_{k}), set 𝐒𝒳:=𝝅0Nis(Sing𝔸1𝒳)\mathbf{S}\mathscr{X}:=\bm{\pi}_{0}^{{\operatorname{Nis}}}(\operatorname{Sing}^{{\mathbb{A}}^{1}}\!\!\mathscr{X}) and define 𝐒=colimn𝐒n\mathbf{S}^{\infty}=\operatorname{colim}_{n}\mathbf{S}^{n} where 𝐒n\mathbf{S}^{n} is the nn-fold iteration of 𝐒\mathbf{S}.

Theorem 2.5.

The following statements hold.

  1. 1.

    The functor 𝐒\mathbf{S}^{\infty} preserves finite products.

  2. 2.

    For any 𝒳P(Smk)\mathscr{X}\in\mathrm{P}(\mathrm{Sm}_{k}), the map 𝒳𝐒𝒳\mathscr{X}\to\mathbf{S}^{\infty}\mathscr{X} factors through an epimorphism 𝝅0Nis(𝒳)𝐒𝒳\bm{\pi}_{0}^{{\operatorname{Nis}}}(\mathscr{X})\to\mathbf{S}^{\infty}\mathscr{X}.

  3. 3.

    The sheaf 𝐒𝒳\mathbf{S}^{\infty}\mathscr{X} is 𝔸1{\mathbb{A}}^{1}-invariant and the map 𝝅0Nis(𝒳)𝐒𝒳\bm{\pi}_{0}^{{\operatorname{Nis}}}(\mathscr{X})\to\mathbf{S}^{\infty}\mathscr{X} is the initial map from the former to an 𝔸1{\mathbb{A}}^{1}-invariant Nisnevich sheaf.

Proof.

The first statement follows because 𝝅0Nis\bm{\pi}_{0}^{{\operatorname{Nis}}} and Sing𝔸1\operatorname{Sing}^{{\mathbb{A}}^{1}}\!\! both preserve finite products. The second statement is immediate because Sing𝔸1\operatorname{Sing}^{{\mathbb{A}}^{1}}\!\! is defined as a colimit and colimits preserve epimorphisms. The final statement is [BHS15, Theorem 2.13]. ∎

Very strongly 𝔸1{\mathbb{A}}^{1}-invariant sheaves

Let 𝐆𝐐\mathbf{G}\to\mathbf{Q} be an epimorphism of Nisnevich sheaves of groups, with 𝐆\mathbf{G} strongly 𝔸1{\mathbb{A}}^{1}-invariant and 𝐐\mathbf{Q} 𝔸1{\mathbb{A}}^{1}-invariant. Ideally, 𝐐\mathbf{Q} would automatically be strongly 𝔸1{\mathbb{A}}^{1}-invariant (see Remark 3 for further discussion of this point). Unfortunately we do not know a proof of this fact in general, and we thus axiomatize the relevant property.

Definition 2.6.

Let 𝐆\mathbf{G} be a Nisnevich sheaf of groups on Smk\mathrm{Sm}_{k}. We call 𝐆\mathbf{G} very strongly 𝔸1{\mathbb{A}}^{1}-invariant if 𝐆\mathbf{G} is 𝔸1{\mathbb{A}}^{1}-invariant and every 𝔸1{\mathbb{A}}^{1}-invariant quotient of 𝐆\mathbf{G} is strongly 𝔸1{\mathbb{A}}^{1}-invariant.

Lemma 2.7.

If 1𝐊𝐆𝐐11\to\mathbf{K}\to\mathbf{G}\to\mathbf{Q}\to 1 is a short exact sequence of Nisnevich sheaves of groups with 𝐆\mathbf{G} strongly 𝔸1{\mathbb{A}}^{1}-invariant, then 𝐐\mathbf{Q} is 𝔸1{\mathbb{A}}^{1}-invariant if and only if 𝐊\mathbf{K} is strongly 𝔸1{\mathbb{A}}^{1}-invariant.

Proof.

Note that B𝐆B\mathbf{G} is connected and motivic local by assumption. Assume 𝐐\mathbf{Q} is 𝔸1{\mathbb{A}}^{1}-invariant and consider the fiber sequence 𝐐B𝐊B𝐆\mathbf{Q}\to B\mathbf{K}\to B\mathbf{G}. In that case [AWW17, Lemma 2.2.10] implies that B𝐊B\mathbf{K} is motivic local as well. Conversely if 𝐊\mathbf{K} is strongly 𝔸1{\mathbb{A}}^{1}-invariant, 𝐐fib(B𝐊B𝐆)\mathbf{Q}\simeq\mathrm{fib}(B\mathbf{K}\to B\mathbf{G}) is 𝔸1{\mathbb{A}}^{1}-invariant. ∎

We first show that the category of very strongly 𝔸1{\mathbb{A}}^{1}-invariant sheaves of groups contains non-trivial objects.

Proposition 2.8.

If kk is a perfect field and 𝐀\mathbf{A} is abelian and strongly 𝔸1{\mathbb{A}}^{1}-invariant, then 𝐀\mathbf{A} is very strongly 𝔸1{\mathbb{A}}^{1}-invariant.

Proof.

Assume 𝐀\mathbf{A} is abelian and strongly 𝔸1{\mathbb{A}}^{1}-invariant; we conclude that 𝐀\mathbf{A} is strictly 𝔸1{\mathbb{A}}^{1}-invariant [Mor12, Corollary 5.45]. Let 𝐀′′\mathbf{A}^{\prime\prime} be an 𝔸1{\mathbb{A}}^{1}-invariant quotient of 𝐀\mathbf{A}. In that case, Lemma 2.7 implies that the kernel 𝐀\mathbf{A}^{\prime} of the homomorphism 𝐀𝐀′′\mathbf{A}\to\mathbf{A}^{\prime\prime} is strongly 𝔸1{\mathbb{A}}^{1}-invariant and abelian, whence once again strictly 𝔸1{\mathbb{A}}^{1}-invariant. In that case, 𝐀𝐀\mathbf{A}^{\prime}\subset\mathbf{A} is a central subsheaf, and the fact that 𝐀′′\mathbf{A}^{\prime\prime} is strongly 𝔸1{\mathbb{A}}^{1}-invariant follows from [AFH22, Lemma 3.1.4(3)]. ∎

Next, we analyze permanence properties of the notion of very strongly 𝔸1{\mathbb{A}}^{1}-invariant sheaves of groups.

Lemma 2.9.

The subcategory of Grpk𝔸1\mathrm{Grp}^{{\mathbb{A}}^{1}}_{k} consisting of very strongly 𝔸1{\mathbb{A}}^{1}-invariant sheaves of groups is closed under formation of

  1. 1.

    quotients,

  2. 2.

    extensions, and

  3. 3.

    filtered colimits.

Proof.

For closure under quotients, observe that if 𝐆\mathbf{G} is very strongly 𝔸1{\mathbb{A}}^{1}-invariant and 𝐐\mathbf{Q} is a quotient of 𝐆\mathbf{G}, then any quotient 𝐐\mathbf{Q}^{\prime} of 𝐐\mathbf{Q} is again a quotient of 𝐆\mathbf{G}. In particular if 𝐐\mathbf{Q}^{\prime} is 𝔸1{\mathbb{A}}^{1}-invariant, it is automatically strongly 𝔸1{\mathbb{A}}^{1}-invariant.

For closure under extensions, suppose we have a short exact sequence of the form

1𝐊𝐆𝐐1.1\longrightarrow\mathbf{K}\longrightarrow\mathbf{G}\longrightarrow\mathbf{Q}\to 1.

If 𝐊\mathbf{K} and 𝐐\mathbf{Q} are strongly 𝔸1{\mathbb{A}}^{1}-invariant, it follows that 𝐆\mathbf{G} is strongly 𝔸1{\mathbb{A}}^{1}-invariant as well by appeal to Lemma 2.7. Suppose 𝐆𝐆\mathbf{G}\to\mathbf{G}^{\prime} is an epimorphism with 𝐆\mathbf{G}^{\prime} an 𝔸1{\mathbb{A}}^{1}-invariant sheaf of groups. Write 𝐊\mathbf{K}^{\prime} for the image of 𝐊\mathbf{K} in 𝐆\mathbf{G}^{\prime}. Since 𝐆\mathbf{G}^{\prime} is 𝔸1{\mathbb{A}}^{1}-invariant, it follows that 𝐊\mathbf{K}^{\prime} is 𝔸1{\mathbb{A}}^{1}-invariant as well, whence strongly 𝔸1{\mathbb{A}}^{1}-invariant by the very strong 𝔸1{\mathbb{A}}^{1}-invariance of 𝐊\mathbf{K}. Since 𝐊\mathbf{K} is a normal subgroup sheaf of 𝐆\mathbf{G}, it also follows that 𝐊\mathbf{K}^{\prime} is a normal subgroup sheaf of 𝐆\mathbf{G}^{\prime}. In that case, write 𝐐\mathbf{Q}^{\prime} for the quotient of 𝐆\mathbf{G}^{\prime} by 𝐊\mathbf{K}^{\prime}. The fiber sequence

𝐆𝐐B𝐊\mathbf{G}^{\prime}\longrightarrow\mathbf{Q}^{\prime}\longrightarrow B\mathbf{K}^{\prime}

in conjunction with the fact that B𝐊B\mathbf{K}^{\prime} is motivic local shows that 𝐐\mathbf{Q}^{\prime} is 𝔸1{\mathbb{A}}^{1}-invariant by appeal to [AWW17, Lemma 2.2.10]. Since 𝐐\mathbf{Q}^{\prime} is also a quotient of 𝐐\mathbf{Q} it is necessarily strongly 𝔸1{\mathbb{A}}^{1}-invariant. It follows that 𝐆\mathbf{G}^{\prime} is an extension of strongly 𝔸1{\mathbb{A}}^{1}-invariant sheaves and therefore itself strongly 𝔸1{\mathbb{A}}^{1}-invariant by appeal to [AFH22, Lemma 3.1.14(2)], as required.

Finally, to establish stability under filtered colimits, suppose 𝐆:𝐈Grpk𝔸1\mathbf{G}:\mathbf{I}\to\mathrm{Grp}^{{\mathbb{A}}^{1}}_{k} is a filtered diagram of strongly 𝔸1{\mathbb{A}}^{1}-invariant sheaves of groups. Consider an 𝔸1{\mathbb{A}}^{1}-invariant quotient colim𝐈𝐆𝐐\operatorname{colim}_{\mathbf{I}}\mathbf{G}\to\mathbf{Q}. Write 𝐐(i)\mathbf{Q}(i) for the image of 𝐆(i)\mathbf{G}(i) in 𝐐\mathbf{Q}. In that case, colimiI𝐐(i)𝐐\operatorname{colim}_{i\in I}\mathbf{Q}(i)\simeq\mathbf{Q}. Each 𝐐(i)\mathbf{Q}(i) is 𝔸1{\mathbb{A}}^{1}-invariant being a subsheaf of 𝐐\mathbf{Q}. Therefore, each 𝐐(i)\mathbf{Q}(i) is strongly 𝔸1{\mathbb{A}}^{1}-invariant as a quotient of the very strongly 𝔸1{\mathbb{A}}^{1}-invariant sheaf 𝐆(i)\mathbf{G}(i). Since motivic local spaces are stable under filtered colimits, we conclude that B𝐐=colimiB𝐐(i)B\mathbf{Q}=\operatorname{colim}_{i}B\mathbf{Q}(i) is motivic local as well and conclude. ∎

Definition 2.10.

A strongly 𝔸1{\mathbb{A}}^{1}-invariant sheaf of groups 𝐆\mathbf{G} is called 𝔸1{\mathbb{A}}^{1}-solvable if there exists a finite increasing filtration:

1=𝐆0𝐆1𝐆n=𝐆1=\mathbf{G}_{0}\subset\mathbf{G}_{1}\subset\cdots\subset\mathbf{G}_{n}=\mathbf{G}

such that, for every ii, 𝐆i\mathbf{G}_{i} is strongly 𝔸1{\mathbb{A}}^{1}-invariant, 𝐆i\mathbf{G}_{i} is a normal subgroup sheaf of 𝐆i+1\mathbf{G}_{i+1}, and the successive subquotients 𝐆i+1/𝐆i\mathbf{G}_{i+1}/\mathbf{G}_{i} are strongly 𝔸1{\mathbb{A}}^{1}-invariant sheaves of abelian groups.

We now give a rather general class of very strongly 𝔸1{\mathbb{A}}^{1}-invariant sheaves of groups. Recall that a Nisnevich sheaf of groups 𝐆\mathbf{G} is locally nilpotent if it admits a descending central series that is stalkwise finite [AFH22, Definition 2.1.3] and nilpotent if it admits a finite descending central series.

Proposition 2.11.

Assume kk is a perfect field.

  1. 1.

    If 𝐆\mathbf{G} is 𝔸1{\mathbb{A}}^{1}-solvable, then 𝐆\mathbf{G} is very strongly 𝔸1{\mathbb{A}}^{1}-invariant.

  2. 2.

    If 𝐆\mathbf{G} is locally nilpotent and strongly 𝔸1{\mathbb{A}}^{1}-invariant, then 𝐆\mathbf{G} is very strongly 𝔸1{\mathbb{A}}^{1}-invariant.

Proof.

The first statement is an immediate consequence of the definition of 𝔸1{\mathbb{A}}^{1}-solvable: use Proposition 2.8 in conjunction with Lemma 2.9 together with a straightforward induction argument on the length of an 𝔸1{\mathbb{A}}^{1}-subnormal series.

For the second statement, write 𝐙i𝐆\mathbf{Z}_{i}\subset\mathbf{G} for the ii-th higher center (see [AFH22, p. 675]) of 𝐆\mathbf{G}, then each 𝐙i\mathbf{Z}_{i} is strongly 𝔸1{\mathbb{A}}^{1}-invariant [AFH22, Proposition 3.1.22]. On the other hand, since 𝐆\mathbf{G} is locally nilpotent, we conclude that 𝐆colimi𝐙i\mathbf{G}\simeq\operatorname{colim}_{i}\mathbf{Z}_{i} since this is true stalkwise by assumption. To establish that 𝐆\mathbf{G} is very strongly 𝔸1{\mathbb{A}}^{1}-invariant, it therefore suffices by appeal to Lemma 2.9(3) to show that each 𝐙i\mathbf{Z}_{i} is very strongly 𝔸1{\mathbb{A}}^{1}-invariant.

By definition there are short exact sequences

1𝐙i𝐙i+1𝐀i+11,1\longrightarrow\mathbf{Z}_{i}\longrightarrow\mathbf{Z}_{i+1}\longrightarrow\mathbf{A}_{i+1}\to 1,

where Ai+1A_{i+1} is strongly 𝔸1{\mathbb{A}}^{1}-invariant and abelian (indeed 𝐀i+1ker(𝐆/𝐙i𝐆/𝐙i+1)\mathbf{A}_{i+1}\simeq\ker(\mathbf{G}/\mathbf{Z}_{i}\to\mathbf{G}/\mathbf{Z}_{i+1}) which is by definition Z(𝐆/𝐙i)Z(\mathbf{G}/\mathbf{Z}_{i}), and thus is strongly 𝔸1{\mathbb{A}}^{1}-invariant by [AFH22, Proposition 3.1.22] again). Since 𝐙1\mathbf{Z}_{1} coincides with the center of 𝐆\mathbf{G}, it and each 𝐀i\mathbf{A}_{i} are strongly 𝔸1{\mathbb{A}}^{1}-invariant and abelian, and thus very strongly 𝔸1{\mathbb{A}}^{1}-invariant by appeal to Proposition 2.8. By induction, it then follows from Lemma 2.9(2) that 𝐙i+1\mathbf{Z}_{i+1} is very strongly 𝔸1{\mathbb{A}}^{1}-invariant for each i2i\geq 2. ∎

Remark 2.12.

It is not clear that a strongly 𝔸1{\mathbb{A}}^{1}-invariant subsheaf of an 𝔸1{\mathbb{A}}^{1}-solvable sheaf of groups is again 𝔸1{\mathbb{A}}^{1}-solvable. In contrast, note that a subsheaf of groups of a locally nilpotent sheaf of groups is again locally nilpotent. It follows that any strongly 𝔸1{\mathbb{A}}^{1}-invariant subsheaf of a locally nilpotent, strongly 𝔸1{\mathbb{A}}^{1}-invariant sheaf of groups is automatically very strongly 𝔸1{\mathbb{A}}^{1}-invariant. We resist the temptation to call such sheaves of groups extremely strongly 𝔸1{\mathbb{A}}^{1}-invariant.

Example 2.13.

In [AFH22, Definition 3.2.1], we also introduced the notion of a (locally) 𝔸1{\mathbb{A}}^{1}-nilpotent sheaf of groups: this is a strongly 𝔸1{\mathbb{A}}^{1}-invariant sheaf of groups that admits a (locally finite) descending central series where all subquotients are strictly 𝔸1{\mathbb{A}}^{1}-invariant. Locally 𝔸1{\mathbb{A}}^{1}-nilpotent sheaves of groups are automatically very strongly 𝔸1{\mathbb{A}}^{1}-invariant. While we showed that nilpotent strongly 𝔸1{\mathbb{A}}^{1}-invariant sheaves of groups are 𝔸1{\mathbb{A}}^{1}-nilpotent [AFH22, Proposition 3.2.3], we do not know whether locally nilpotent strongly 𝔸1{\mathbb{A}}^{1}-invariant sheaves of groups are automatically locally 𝔸1{\mathbb{A}}^{1}-nilpotent because it is unclear whether a locally finite ascending 𝔸1{\mathbb{A}}^{1}-central series gives rise to a locally finite descending 𝔸1{\mathbb{A}}^{1}-central series.

Proposition 2.14.

Suppose kk is a field, and 𝐆Grpk𝔸1\mathbf{G}\in\mathrm{Grp}^{{\mathbb{A}}^{1}}_{k} is a very strongly 𝔸1{\mathbb{A}}^{1}-invariant sheaf of groups. If 𝐇\mathbf{H} is a quotient of 𝐆\mathbf{G}, then 𝐒𝐇\mathbf{S}^{\infty}\mathbf{H} is the initial strongly 𝔸1{\mathbb{A}}^{1}-invariant quotient of 𝐆\mathbf{G} admitting a map from 𝐇\mathbf{H}.

Proof.

The initial 𝔸1{\mathbb{A}}^{1}-invariant and strongly 𝔸1{\mathbb{A}}^{1}-invariant quotient of 𝐇\mathbf{H} coincide, since 𝐆\mathbf{G} is assumed very strongly 𝔸1{\mathbb{A}}^{1}-invariant. The statement thus follows from Theorem 2.5. ∎

3 Connectivity of fibers and cofibers

We now review various facts about comparison of fibers and cofibers, including connectivity estimates.

Comparing fibers and cofibers

We record the following result about comparison of horizontal and vertical fibers in a commutative diagram for lack of a convenient reference. It will be used repeatedly in the sequel; the statement is undoubtedly very old (e.g., see [CMN79, Lemma 2.1] where it stated as “well-known”) and holds in any \infty-category with finite limits. The consequence at the end of the statement about fibers of composites was observed by Quillen [Qui67, 3.10 Remark].

Proposition 3.1.

A commutative square in Spc(k)\mathrm{Spc}({k}) of the form

𝒳00\textstyle{\mathscr{X}_{00}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}g0\scriptstyle{g_{0}}f0\scriptstyle{f_{0}}𝒳10\textstyle{\mathscr{X}_{10}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}f1\scriptstyle{f_{1}}𝒳01\textstyle{\mathscr{X}_{01}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}g1\scriptstyle{g_{1}}𝒳11\textstyle{\mathscr{X}_{11}}

can be embedded in a commutative diagram of the form:

𝒳\textstyle{\mathscr{X}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}fib(f0)\textstyle{\mathrm{fib}(f_{0})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}fib(f1)\textstyle{\mathrm{fib}(f_{1})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}fib(g0)\textstyle{\mathrm{fib}(g_{0})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝒳00\textstyle{\mathscr{X}_{00}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}g0\scriptstyle{g_{0}}f0\scriptstyle{f_{0}}𝒳10\textstyle{\mathscr{X}_{10}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}f1\scriptstyle{f_{1}}fib(g1)\textstyle{\mathrm{fib}(g_{1})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝒳01\textstyle{\mathscr{X}_{01}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}g1\scriptstyle{g_{1}}𝒳11\textstyle{\mathscr{X}_{11}}

where (a) there is an equivalence 𝒳fib(𝒳00𝒳01×𝒳11𝒳10)\mathscr{X}\cong\mathrm{fib}(\mathscr{X}_{00}\to\mathscr{X}_{01}\times_{\mathscr{X}_{11}}\mathscr{X}_{10}), and (b) all rows and columns are fiber sequences. In particular, if f:𝒳𝒴f:\mathscr{X}\to\mathscr{Y} and g:𝒴𝒵g:\mathscr{Y}\to\mathscr{Z} are maps in Spc(k)\mathrm{Spc}({k}), then there is a fiber sequence of the form

fib(f)fib(gf)fib(g).\mathrm{fib}(f)\longrightarrow\mathrm{fib}(g\circ f)\longrightarrow\mathrm{fib}(g).
Proof.

The second statement is a special case of the first arising from comparison of horizontal and vertical homotopy fibers in the commutative diagram:

𝒳\textstyle{\mathscr{X}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}f\scriptstyle{f}gf\scriptstyle{g\circ f}𝒴\textstyle{\mathscr{Y}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}g\scriptstyle{g}𝒵\textstyle{\mathscr{Z}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}id\scriptstyle{id}𝒵.\textstyle{\mathscr{Z}.}

Remark 3.2.

Given a commutative square as in Proposition 3.1, there is, of course, a dual statement comparing vertical and horizontal cofibers. The statement about cofibers of composites in the context of model categories can be found in [Hov99, Proposition 6.3.6].

Connectivity of fibers and cofibers

Proposition 3.3.

Suppose kk is a field. Assume f:𝒳𝒴f:\mathscr{X}\to\mathscr{Y} is a morphism of pointed spaces in Spc(k)\mathrm{Spc}({k}) that is pulled back from a perfect subfield of kk.

  1. 1.

    If ff has (n1)(n-1)-connected fibers, then cof(f)\mathrm{cof}(f) is nn-connected.

  2. 2.

    If 𝒳\mathscr{X} and 𝒴\mathscr{Y} are 11-connected, and cof(f)\mathrm{cof}(f) is nn-connected, then fib(f)\mathrm{fib}(f) is (n1)(n-1)-connected.

Proof.

For the first statement, we proceed as follows. Write 𝒞\mathscr{C} for the cofiber of ff computed in ShvNis(Smk){\mathrm{Shv}}_{{\operatorname{Nis}}}(\mathrm{Sm}_{k}), which fits in a commutative square of the form

𝒳\textstyle{\mathscr{X}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}f\scriptstyle{f}𝒴\textstyle{\mathscr{Y}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\textstyle{\ast\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝒞.\textstyle{\mathscr{C}.}

By definition, cof(f)=Lmot𝒞\mathrm{cof}(f)=\mathrm{L}_{mot}\mathscr{C}. Since the morphism 𝒞\ast\to\mathscr{C} is the cobase change of ff along 𝒳\mathscr{X}\to\ast and since morphisms with (n1)(n-1)-connected fibers are stable under cobase change along arbitrary morphisms [Lur09, Corollary 6.5.1.17], it follows that 𝒞\ast\to\mathscr{C} has (n1)(n-1)-connected fibers, i.e., 𝒞\mathscr{C} is nn-connected. The fact that cof(f)\mathrm{cof}(f) is nn-connected then follows by appeal to Theorem 2.1.

For the second statement, assume that cof(f)\mathrm{cof}(f) is nn-connected. Set :=fib(f)\mathscr{F}:=\mathrm{fib}(f). Since ff is a map of 11-connected spaces, it follows that \mathscr{F} is automatically connected and 𝝅1()\bm{\pi}_{1}(\mathscr{F}) is abelian. Consider the comparison map Ω𝒞\mathscr{F}\to\Omega\mathscr{C} induced by taking fibers in the diagram from the preceding paragraph. The classical Blakers–Massey theorem [GJ09, Theorems 3.10-11] implies that if \mathscr{F} is mm-connected for some integer mm (which we may assume without loss of generality to be 0\geq 0), then Ω𝒞\mathscr{F}\to\Omega\mathscr{C} induces an isomorphism on homotopy sheaves in degrees m+1\leq m+1. We conclude that 𝝅i+1(Ω𝒞)\bm{\pi}_{i+1}(\Omega\mathscr{C}) is strictly 𝔸1{\mathbb{A}}^{1}-invariant for any mi0m\geq i\geq 0. In that case, the map 𝝅i+1(Ω𝒞)𝝅i+1(cof(f))\bm{\pi}_{i+1}(\Omega\mathscr{C})\to\bm{\pi}_{i+1}(\mathrm{cof}(f)) is an isomorphism by [AWW17, Theorem 2.3.8]. Applying this observation iteratively, we deduce that the statement holds as long as mn1m\leq n-1, which is what we wanted to show. ∎

We now recall some consequences of “homotopy distributivity’, i.e., commutativity of certain limits and colimits that will be useful in the sequel. The primordial form of homotopy distributivity we use is that colimits are universal in Spc(k)\mathrm{Spc}({k}); see [Hoy17, Proposition 3.15] for this statement. For topological spaces, the results we describe go back to work of V. Puppe [Pup74] and M. Mather [Mat76]. They were studied in a model-categorical framework in unpublished work of C. Rezk [Rez98]. The results were later analyzed in the context of motivic homotopy theory in the thesis of M. Wendt [Wen11] and the translation to the context of \infty-topoi is straightforward. The first point in the next statement originates from a result of Ganea [Gan65, Theorem 1.1].

Lemma 3.4.

Suppose kk is a field, and fib(f)ιf\mathrm{fib}(f)\stackrel{{\scriptstyle\iota}}{{\to}}\mathscr{E}\stackrel{{\scriptstyle f}}{{\to}}\mathscr{B} is a fiber sequence in Spc(k)\mathrm{Spc}({k}).

  1. 1.

    There is a natural equivalence

    fib(cof(ι))ΣΩfib(f).\mathrm{fib}(\mathrm{cof}(\iota)\to\mathscr{B})\cong\Sigma\Omega\mathscr{B}\wedge\mathrm{fib}(f).
  2. 2.

    There is a cofiber sequence of the form

    Σfib(f)cof(f)cof(cof(ι))\Sigma\mathrm{fib}(f)\longrightarrow\mathrm{cof}(f)\longrightarrow\mathrm{cof}(\mathrm{cof}(\iota)\to\mathscr{B})
Proof.

The first statement may be found in [DH21, Lemma 2.27]. For the second point, consider the commutative diagram

\textstyle{\mathscr{E}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\textstyle{\mathscr{E}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}f\scriptstyle{f}cof(ι)\textstyle{\mathrm{cof}(\iota)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}.\textstyle{\mathscr{B}.}

Comparing vertical and horizontal cofibers via the dual of Proposition 3.1, one obtains a cofiber sequence of the form:

cof(cof(ι))cof(f)cof(cof(ι)).\mathrm{cof}(\mathscr{E}\to\mathrm{cof}(\iota))\longrightarrow\mathrm{cof}(f)\longrightarrow\mathrm{cof}(\mathrm{cof}(\iota)\to\mathscr{B}).

On the other hand, another application of the dual of Proposition 3.1 to the commutative diagram

fib(f)\textstyle{\mathrm{fib}(f)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ι\scriptstyle{\iota}\textstyle{\mathscr{E}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\textstyle{\ast\ignorespaces\ignorespaces\ignorespaces\ignorespaces}cof(ι)\textstyle{\mathrm{cof}(\iota)}

yields a cofiber sequence of the form

Σfib(f)cof(cof(ι)).\Sigma\mathrm{fib}(f)\longrightarrow\mathrm{cof}(\mathscr{E}\to\mathrm{cof}(\iota))\longrightarrow\ast.

Using the two diagrams above, and taking the cofiber of the composite Σfib(f)cof(cof(ι))cof(f)\Sigma\mathrm{fib}(f)\to\mathrm{cof}(\mathscr{E}\to\mathrm{cof}(\iota))\to\mathrm{cof}(f), we obtain an equivalence

cof(Σfib(f)cof(f))cof(cof(ι)),\mathrm{cof}(\Sigma\mathrm{fib}(f)\to\mathrm{cof}(f))\longrightarrow\mathrm{cof}(\mathrm{cof}(\iota)\to\mathscr{B}),

which is precisely what we wanted to show. ∎

The next result improves the motivic analog of the Blakers–Massey theorem of [AF16, Theorem 4.1] or [Str12, Theorem 2.3.8].

Proposition 3.5 (Homotopy excision).

Assume kk is a field, and f:f:\mathscr{E}\to\mathscr{B} is a morphism pulled back from a perfect subfield of kk. If m,n0m,n\geq 0 are integers, the following statements regarding the fiber sequences fib(f)ιf\mathrm{fib}(f)\stackrel{{\scriptstyle\iota}}{{\to}}\mathscr{E}\stackrel{{\scriptstyle f}}{{\to}}\mathscr{B} hold.

  1. 1.

    If fib(f)\mathrm{fib}(f) is mm-connected and \mathscr{B} is nn-connected, then cof(ι)\mathrm{cof}(\iota)\to\mathscr{B} has (m+n+1)(m+n+1)-connected fibers.

  2. 2.

    If fib(f)\mathrm{fib}(f) is mm-connected and \mathscr{B} is nn-connected, then the canonical map Σfib(f)cof(f)\Sigma\mathrm{fib}(f)\to\mathrm{cof}(f) has (m+n+1)(m+n+1)-connected fibers.

Proof.

For the first statement, observe that Ganea’s lemma 3.4 gives an equivalence ΣΩfib(f)fib(cof(ι))\Sigma\Omega\mathscr{B}\wedge\mathrm{fib}(f)\cong\mathrm{fib}(\mathrm{cof}(\iota)\to\mathscr{B}). Since \mathscr{B} is nn-connected, ΣΩ\Sigma\Omega\mathscr{B} is at least nn-connected. Since fib(f)\mathrm{fib}(f) is mm-connected, we conclude that ΣΩfib(f)\Sigma\Omega\mathscr{B}\wedge\mathrm{fib}(f) is at least (m+n+1)(m+n+1)-connected by appeal to [AWW17, Lemma 3.3.1]. Thus, cof(ι)\mathrm{cof}(\iota)\to\mathscr{B} has (m+n+1)(m+n+1)-connected fibers by definition.

For the second point, note that Lemma 3.4(2) yields a cofiber sequence of the form

Σfib(f)cof(ι)cof(cof(ι)).\Sigma\mathrm{fib}(f)\longrightarrow\mathrm{cof}(\iota)\longrightarrow\mathrm{cof}(\mathrm{cof}(\iota)\to\mathscr{B}).

By the conclusion of Point (1), we see that cof(ι)\mathrm{cof}(\iota)\to\mathscr{B} has (m+n+1)(m+n+1)-connected fibers, and then Proposition 3.3(1) implies that cof(cof(ι))\mathrm{cof}(\mathrm{cof}(\iota)\to\mathscr{B}) is (m+n+2)(m+n+2)-connected.

The map ff has mm-connected fibers for some m0m\geq 0 by assumption, so another application of Proposition 3.3(1) implies that cof(f)\mathrm{cof}(f) is at least m+1m+1-connected, in particular simply connected. Likewise, Σfib(f)\Sigma\mathrm{fib}(f) is at least m+1m+1-connected and therefore also at least simply connected. It follows that Σfib(f)cof(f)\Sigma\mathrm{fib}(f)\to\mathrm{cof}(f) is a map of 11-connected spaces whose cofiber is m+n+2m+n+2-connected. Thus, appeal to Proposition 3.3(2) implies that Σfib(f)cof(f)\Sigma\mathrm{fib}(f)\to\mathrm{cof}(f) has (m+n+1)(m+n+1)-connected fibers as well. ∎

We may use the above results to strengthen the relative Hurewicz theorem of [AFH22, Theorem 4.2.1], but we give a self-contained treatment here. Suppose f:f:\mathscr{E}\to\mathscr{B} is a morphism of pointed, connected spaces. There is a morphism of fiber sequences of the form

fib(f)\textstyle{\mathrm{fib}(f)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\textstyle{\mathscr{E}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}f\scriptstyle{f}\textstyle{\mathscr{B}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Ωcof(f)\textstyle{\Omega\mathrm{cof}(f)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\textstyle{\ast\ignorespaces\ignorespaces\ignorespaces\ignorespaces}cof(f).\textstyle{\mathrm{cof}(f).}

The diagram shows that there is an induced action of 𝝅1()\bm{\pi}_{1}(\mathscr{E}) on Ωcof(f)\Omega\mathrm{cof}(f), which is necessarily trivial. The relative Hurewicz theorem analyzes the map fib(f)Ωcof(f)\mathrm{fib}(f)\to\Omega\mathrm{cof}(f) or, rather, the induced map Σfib(f)cof(f)\Sigma\mathrm{fib}(f)\to\mathrm{cof}(f).

Theorem 3.6.

Suppose kk is a field, and assume f:f:\mathscr{E}\to\mathscr{B} is a morphism of pointed connected spaces that is pulled back from a perfect subfield of kk. If ff has (n1)(n-1)-connected fibers, then the following statements hold.

  1. 1.

    The space cof(f)\mathrm{cof}(f) is nn-connected.

  2. 2.

    The relative Hurewicz map 𝝅n(fib(f))𝐇n+1𝔸1(cof(f))\bm{\pi}_{n}(\mathrm{fib}(f))\to{{\mathbf{H}}}_{n+1}^{{\mathbb{A}}^{1}}(\mathrm{cof}(f)) is the initial morphism from 𝝅n(fib(f))\bm{\pi}_{n}(\mathrm{fib}(f)) to a strictly 𝔸1{\mathbb{A}}^{1}-invariant sheaf on which 𝝅1()\bm{\pi}_{1}(\mathscr{E}) acts trivially.

Proof.

The first statement is Proposition 3.3(1); thus we prove the second. We treat only the case n=1n=1, as the case n2n\geq 2 follows the same argument but is easier. The general principle is as follows. If 𝒳\mathscr{X} is an (n1)(n-1)-connected Nisnevich sheaf of spaces, then 𝝅n(Lmot𝒳)\bm{\pi}_{n}(\mathrm{L}_{mot}\mathscr{X}) is the initial strongly 𝔸1{\mathbb{A}}^{1}-invariant sheaf under 𝝅n(𝒳)\bm{\pi}_{n}(\mathscr{X}). By the classical relative Hurewicz theorem (in ShvNis(Smk){\mathrm{Shv}}_{{\operatorname{Nis}}}(\mathrm{Sm}_{k})) we know that 𝝅1(fib(f))𝐇2(𝒞)\bm{\pi}_{1}(\mathrm{fib}(f))\to{{\mathbf{H}}}_{2}(\mathscr{C}) is the initial map from 𝝅1(fib(f))\bm{\pi}_{1}(\mathrm{fib}(f)) to a sheaf on which 𝝅1()\bm{\pi}_{1}(\mathscr{E}) acts trivially. Since 𝐇2𝔸1(cof(f))𝝅2(cof(f)){{\mathbf{H}}}_{2}^{{\mathbb{A}}^{1}}(\mathrm{cof}(f))\simeq\bm{\pi}_{2}(\mathrm{cof}(f)) is the initial strictly 𝔸1{\mathbb{A}}^{1}-invariant sheaf under 𝐇2(𝒞){{\mathbf{H}}}_{2}(\mathscr{C}), the result follows.

For later use, we record the following version of Morel’s suspension theorem [Mor12, Theorem 6.61].

Theorem 3.7.

Assume n0n\geq 0 is an integer and 𝒳\mathscr{X} is a pointed, nn-connected motivic space.

  1. 1.

    For any integer i0i\geq 0, the canonical map 𝒳ΩiΣi𝒳\mathscr{X}\to\Omega^{i}\Sigma^{i}\mathscr{X} has 2n2n-connected fibers.

  2. 2.

    The map 𝒳ΩS1ΣS1𝒳\mathscr{X}\to\Omega^{\infty}_{S^{1}}\Sigma^{\infty}_{S^{1}}\mathscr{X} has 2n2n-connected fibers.

Proof.

For i=0i=0 there is nothing to show, so we assume i1i\geq 1. In that case, Theorem 2.1 implies that Σi𝒳\Sigma^{i}\mathscr{X} is at least n+in+i-connected and thus that ΩiΣi𝒳\Omega^{i}\Sigma^{i}\mathscr{X} is at least nn-connected as well. Theorem 3.6 implies that 𝝅n+1(ΩiΣi𝒳)\bm{\pi}_{n+1}(\Omega^{i}\Sigma^{i}\mathscr{X}) coincides with 𝐇n+1𝔸1(𝒳){{\mathbf{H}}}_{n+1}^{{\mathbb{A}}^{1}}(\mathscr{X}). With that in mind, when n=0n=0, the conclusion of the theorem is that the map 𝝅1(𝒳)𝐇1𝔸1(𝒳)\bm{\pi}_{1}(\mathscr{X})\to{{\mathbf{H}}}_{1}^{{\mathbb{A}}^{1}}(\mathscr{X}) is an epimorphism, which follows immediately from Theorem 3.6 (this particular case also is contained in [CH22, Theorem 1.4]).

We now assume that n>0n>0. In that case, write 𝒳\mathscr{X^{\prime}} for kk-fold suspension of 𝒳\mathscr{X} in ShvNis(Smk){\mathrm{Shv}}_{{\operatorname{Nis}}}(\mathrm{Sm}_{k}). The map 𝒳Ωi𝒳\mathscr{X}\to\Omega^{i}\mathscr{X}^{\prime} in ShvNis(Smk){\mathrm{Shv}}_{{\operatorname{Nis}}}(\mathrm{Sm}_{k}) has 2n2n-connected fibers by the classical Freudenthal suspension theorem applied stalkwise. In other words, there is a fiber sequence in ShvNis(Smk){\mathrm{Shv}}_{{\operatorname{Nis}}}(\mathrm{Sm}_{k}) of the form

𝒳Ωi𝒳\mathscr{F}\longrightarrow\mathscr{X}\longrightarrow\Omega^{i}\mathscr{X}^{\prime}

with \mathscr{F} 2n2n-connected. In particular, we conclude that 𝝅j(Ωi𝒳)\bm{\pi}_{j}(\Omega^{i}\mathscr{X}^{\prime}) is strongly 𝔸1{\mathbb{A}}^{1}-invariant for j2nj\leq 2n. Since n1n\geq 1, we conclude from [AWW17, Theorem 2.3.3] that the canonical map LmotLmot(fib(𝒳Ωi𝒳))\mathrm{L}_{mot}\mathscr{F}\to\mathrm{L}_{mot}(\mathrm{fib}(\mathscr{X}\to\Omega^{i}\mathscr{X}^{\prime})) is an equivalence. On the other hand, repeated application of [AWW17, Theorem 2.4.1] implies that LmotΩi𝒳ΩiLmot𝒳\mathrm{L}_{mot}\Omega^{i}\mathscr{X}^{\prime}\to\Omega^{i}\mathrm{L}_{mot}\mathscr{X}^{\prime} is an equivalence as well. Combining the observations above, there is thus a fiber sequence of the form

Lmot𝒳ΩiΣi𝒳.\mathrm{L}_{mot}\mathscr{F}\longrightarrow\mathscr{X}\longrightarrow\Omega^{i}\Sigma^{i}\mathscr{X}.

The unstable connectivity theorem implies Lmot\mathrm{L}_{mot}\mathscr{F} is 2n2n-connected, which concludes the verification of Point(1). Point (2) follows immediately from Point (1) by passage to the limit. ∎

Example 3.8.

Assume 𝐆\mathbf{G} is a connected group (e.g., SL2SL_{2}). In that case, the identity map on 𝐆\mathbf{G} factors through a map 𝐆ΩΣ𝐆𝐆\mathbf{G}\to\Omega\Sigma\mathbf{G}\to\mathbf{G}, which shows that 𝐆\mathbf{G} is a retract of ΩΣ𝐆\Omega\Sigma\mathbf{G}. The other factor can be described explicitly as follows. Applying Lemma 3.4 to the fiber sequence 𝐆B𝐆\mathbf{G}\to\ast\to B\mathbf{G}, one concludes the existence of a fiber sequence of the form

Σ𝐆𝐆Σ𝐆B𝐆,\Sigma\mathbf{G}\wedge\mathbf{G}\longrightarrow\Sigma\mathbf{G}\longrightarrow B\mathbf{G},

where the map Σ𝐆𝐆Σ𝐆\Sigma\mathbf{G}\wedge\mathbf{G}\to\Sigma\mathbf{G} is the Hopf construction of the multiplication on 𝐆\mathbf{G} [Mor12, p. 191]. The retraction map described in the first paragraph provides a section of the corresponding fibration after looping and we conclude that

ΩΣ𝐆ΩΣ(𝐆𝐆)×𝐆\Omega\Sigma\mathbf{G}\cong\Omega\Sigma(\mathbf{G}\wedge\mathbf{G})\times\mathbf{G}

It follows that the fiber of 𝐆ΩΣ𝐆\mathbf{G}\to\Omega\Sigma\mathbf{G} can be identified with Ω2Σ𝐆𝐆\Omega^{2}\Sigma\mathbf{G}\wedge\mathbf{G}. Since 𝐆\mathbf{G} is a connected group, the map 𝝅1(𝐆)𝝅1(ΩΣ𝐆)\bm{\pi}_{1}(\mathbf{G})\to\bm{\pi}_{1}(\Omega\Sigma\mathbf{G}) is an isomorphism by the Hurewicz theorem 3.6. On the other hand 𝝅1(Ω2Σ𝐆𝐆)\bm{\pi}_{1}(\Omega^{2}\Sigma\mathbf{G}\wedge\mathbf{G}) is non-trivial in general (e.g., for SL2SL_{2}). One concludes that connectivity assertion in Theorem 3.7 is optimal for n=0n=0. For n1n\geq 1, we refer the reader to [AWW17, Theorem 3.2.1].

4 Abelianization and 𝔸1{\mathbb{A}}^{1}-lower central series

Theorem 2.5 and Proposition 2.14 allow us to build appropriate analogs of functorial constructions from classical group theory. To illustrate this, we discuss abelianization and lower central series for strongly 𝔸1{\mathbb{A}}^{1}-invariant sheaves of groups. Unfortunately, all of these notions are only provably well-behaved for very strongly 𝔸1{\mathbb{A}}^{1}-invariant sheaves of groups in the sense of Definition 2.6.

Lemma 4.1.

Assuming kk a field, the following statements hold.

  1. 1.

    The inclusion functor Abk𝔸1Grpk𝔸1\mathrm{Ab}^{{\mathbb{A}}^{1}}_{k}\hookrightarrow\mathrm{Grp}^{{\mathbb{A}}^{1}}_{k} admits a left adjoint ()𝔸1ab(-)^{ab}_{{\mathbb{A}}^{1}} (which is thus right exact).

  2. 2.

    If kk is furthermore perfect and 𝐆\mathbf{G} is a very strongly 𝔸1{\mathbb{A}}^{1}-invariant sheaf, then the unit of the adjunction 𝐆𝐆𝔸1ab\mathbf{G}\to\mathbf{G}^{ab}_{{\mathbb{A}}^{1}} is an epimorphism. In fact 𝐆𝔸1ab𝐒𝐆ab\mathbf{G}^{ab}_{{\mathbb{A}}^{1}}\simeq\mathbf{S}^{\infty}\mathbf{G}^{ab}.

Proof.

The existence of an adjoint is a consequence of Morel’s stable connectivity theorem. We establish the second point by analyzing universal properties; to this end, write Grpk\mathrm{Grp}_{k} for the category of Nisnevich sheaves of groups on Smk\mathrm{Sm}_{k}, and Abk\mathrm{Ab}_{k} for the category of Nisnevich sheaves of abelian groups on Smk\mathrm{Sm}_{k}. Suppose 𝐆\mathbf{G} is a very strongly 𝔸1{\mathbb{A}}^{1}-invariant sheaf of groups.

If 𝐀\mathbf{A} is a strongly 𝔸1{\mathbb{A}}^{1}-invariant abelian sheaf of groups, then writing 𝐆ab\mathbf{G}^{ab} for abelianization as a Nisnevich sheaf of groups, there is a bijection

HomGrpk𝔸1(𝐆,𝐀)HomAbk(𝐆ab,𝐀),\operatorname{Hom}_{\mathrm{Grp}^{{\mathbb{A}}^{1}}_{k}}(\mathbf{G},\mathbf{A})\cong\operatorname{Hom}_{\mathrm{Ab}_{k}}(\mathbf{G}^{ab},\mathbf{A}),

by the universal property of abelianization. On the other hand, since 𝐀\mathbf{A} is 𝔸1{\mathbb{A}}^{1}-invariant, we deduce from Theorem 2.5 that there is a bijection:

HomAbk(𝐆ab,𝐀)HomAbk(𝐒𝐆ab,𝐀).\operatorname{Hom}_{\mathrm{Ab}_{k}}(\mathbf{G}^{ab},\mathbf{A})\cong\operatorname{Hom}_{\mathrm{Ab}_{k}}(\mathbf{S}^{\infty}\mathbf{G}^{ab},\mathbf{A}).

Proposition 2.14 implies that 𝐒𝐆ab\mathbf{S}^{\infty}\mathbf{G}^{ab} is strongly 𝔸1{\mathbb{A}}^{1}-invariant, and since it is abelian it is thus strictly 𝔸1{\mathbb{A}}^{1}-invariant. We then conclude that

HomGrpk𝔸1(𝐆,𝐀)HomAbk𝔸1(𝐒𝐆ab,𝐀),\operatorname{Hom}_{\mathrm{Grp}^{{\mathbb{A}}^{1}}_{k}}(\mathbf{G},\mathbf{A})\cong\operatorname{Hom}_{\mathrm{Ab}^{{\mathbb{A}}^{1}}_{k}}(\mathbf{S}^{\infty}\mathbf{G}^{ab},\mathbf{A}),

i.e.., that 𝐒𝐆ab\mathbf{S}^{\infty}\mathbf{G}^{ab} and 𝐆𝔸1ab\mathbf{G}^{ab}_{{\mathbb{A}}^{1}} satisfy the same universal property, which is what we wanted to show. ∎

Remark 4.2.

If kk is a field and 𝐆\mathbf{G} is a strongly 𝔸1{\mathbb{A}}^{1}-invariant sheaf of groups, then 𝐆𝔸1ab𝐇1𝔸1(B𝐆)\mathbf{G}^{ab}_{{\mathbb{A}}^{1}}\cong{{\mathbf{H}}}_{1}^{{\mathbb{A}}^{1}}(B\mathbf{G}) by Morel’s Hurewicz Theorem  [Mor12, Theorem 6.35]. If kk is perfect field and 𝐆\mathbf{G} is very strongly 𝔸1{\mathbb{A}}^{1}-invariant, then 𝐇1𝔸1(B𝐆)𝐒𝐆ab{{\mathbf{H}}}_{1}^{{\mathbb{A}}^{1}}(B\mathbf{G})\cong\mathbf{S}^{\infty}\mathbf{G}^{ab} by the preceding result.

Remark 4.3.

Analyzing the proof of Lemma 4.1 we observe that for any very strongly 𝔸1{\mathbb{A}}^{1}-invariant sheaf 𝐆\mathbf{G} there is always an epimorphism 𝐆ab𝐆𝔸1ab\mathbf{G}^{ab}\to\mathbf{G}^{ab}_{{\mathbb{A}}^{1}} and that this map is an isomorphism if 𝐆ab\mathbf{G}^{ab} happens to be 𝔸1{\mathbb{A}}^{1}-invariant. We do not know whether 𝐆ab\mathbf{G}^{ab} is 𝔸1{\mathbb{A}}^{1}-invariant in general, though this does happen: see Example 4.12.

Example 4.4 (Contraction does not commute with abelianization).

In general, abelianization does not commute with contraction for an 𝔸1{\mathbb{A}}^{1}-nilpotent sheaf of groups. For example, one knows that 𝝅:=𝝅1(1)\bm{\pi}:=\bm{\pi}_{1}({\mathbb{P}}^{1}) is a central extension of 𝔾m{{\mathbb{G}}_{m}} by 𝐊2MW{{\mathbf{K}}}^{MW}_{2} by [Mor12, Theorem 7.29]. In that case, 𝐇1𝔸1(1)𝐊1MW{{\mathbf{H}}}_{1}^{{\mathbb{A}}^{1}}({\mathbb{P}}^{1})\cong{{\mathbf{K}}}^{MW}_{1} by the suspension isomorphism, i.e., 𝝅ab𝐊1MW\bm{\pi}^{ab}\cong{{\mathbf{K}}}^{MW}_{1}. We know that (𝐊1MW)1𝐊0MW({{\mathbf{K}}}^{MW}_{1})_{-1}\cong{{\mathbf{K}}}^{MW}_{0}. On the other hand, the contraction of 𝝅1(1)\bm{\pi}_{1}({\mathbb{P}}^{1}) is already abelian and isomorphic to 𝐊1MW{{\mathbf{K}}}^{MW}_{1}\oplus{\mathbb{Z}} by appeal to [Mor12, Corollary 7.34] and therefore coincides with its abelianization.

Definition 4.5.

If 𝐆\mathbf{G} is a very strongly 𝔸1{\mathbb{A}}^{1}-invariant sheaf of groups, then we define [𝐆,𝐆]𝔸1[\mathbf{G},\mathbf{G}]_{{\mathbb{A}}^{1}} to be the kernel of the epimorphism 𝐆𝐆𝔸1ab\mathbf{G}\to\mathbf{G}^{ab}_{{\mathbb{A}}^{1}}.

Remark 4.6.

Assume kk is a perfect field. If 𝐆\mathbf{G} is very strongly 𝔸1{\mathbb{A}}^{1}-invariant, it is not clear that one can iterate the construction of the 𝔸1{\mathbb{A}}^{1}-commutator subgroup sheaf to define an 𝔸1{\mathbb{A}}^{1}-derived series because [𝐆,𝐆]𝔸1[\mathbf{G},\mathbf{G}]_{{\mathbb{A}}^{1}} is not evidently very strongly 𝔸1{\mathbb{A}}^{1}-invariant. In particular, it is not clear that analogs of the various equivalent characterizations of solvable groups hold for their motivic variants.

The next result is an analog for strictly 𝔸1{\mathbb{A}}^{1}-invariant sheaves of a result of Stallings [Sta65, Theorem 2.1].

Lemma 4.7.

Suppose 1𝐍𝐆𝐐11\to\mathbf{N}\to\mathbf{G}\to\mathbf{Q}\to 1 is a short exact sequence of strongly 𝔸1{\mathbb{A}}^{1}-invariant sheaves of groups.

  1. 1.

    There exists a strictly 𝔸1{\mathbb{A}}^{1}-invariant sheaf 𝐍/[𝐍,𝐆]𝔸1\mathbf{N}/[\mathbf{N},\mathbf{G}]_{{\mathbb{A}}^{1}} under 𝐍\mathbf{N}, which is initial among strictly 𝔸1{\mathbb{A}}^{1}-invariant sheaves under 𝐍ab\mathbf{N}^{ab} on which 𝐆\mathbf{G} acts trivially.

  2. 2.

    There is an exact sequence of strictly 𝔸1{\mathbb{A}}^{1}-invariant sheaves of the form

    𝐇2𝔸1(B𝐆)𝐇2𝔸1(B𝐐)𝐍/[𝐍,𝐆]𝔸1𝐆𝔸1ab𝐐𝔸1ab0.{{\mathbf{H}}}_{2}^{{\mathbb{A}}^{1}}(B\mathbf{G})\longrightarrow{{\mathbf{H}}}_{2}^{{\mathbb{A}}^{1}}(B\mathbf{Q})\longrightarrow\mathbf{N}/[\mathbf{N},\mathbf{G}]_{{\mathbb{A}}^{1}}\longrightarrow\mathbf{G}^{ab}_{{\mathbb{A}}^{1}}\longrightarrow\mathbf{Q}^{ab}_{{\mathbb{A}}^{1}}\longrightarrow 0.
  3. 3.

    The exact sequence of the previous point is functorial in morphisms of short exact sequences.

  4. 4.

    If 𝐍\mathbf{N} is very strongly 𝔸1{\mathbb{A}}^{1}-invariant then 𝐍/[𝐍,𝐆]𝔸1𝐒𝐍/[𝐍,𝐆]\mathbf{N}/[\mathbf{N},\mathbf{G}]_{{\mathbb{A}}^{1}}\simeq\mathbf{S}^{\infty}\mathbf{N}/[\mathbf{N},\mathbf{G}].

Remark 4.8.

If 𝐍\mathbf{N} is not very strongly 𝔸1{\mathbb{A}}^{1}-invariant then it is not clear to us if 𝐍𝐍/[𝐍,𝐆]𝔸1\mathbf{N}\to\mathbf{N}/[\mathbf{N},\mathbf{G}]_{{\mathbb{A}}^{1}} needs to be an epimorphism; the notation may be somewhat misleading in this case.

Proof.

(1)-(3). We establish the three points simultaneously. Set 𝒞=cof(B𝐆B𝐐)\mathcal{C}=\mathrm{cof}(B\mathbf{G}\to B\mathbf{Q}). Since B𝐍B\mathbf{N} is 0-connected and B𝐐B\mathbf{Q} is 0-connected, the canonical map ΣB𝐍𝒞\Sigma B\mathbf{N}\to\mathscr{C} has 11-connected fibers by Proposition 3.5(2). Since ΣB𝐍\Sigma B\mathbf{N} is 11-connected, 𝒞\mathscr{C} is 11-connected as well. The long exact sequence in homology takes the form (use Remark 4.2)

𝐇2𝔸1(B𝐆)𝐇2𝔸1(B𝐐)𝐇2𝔸1(𝒞)𝐆𝔸1ab𝐐𝔸1ab0,{{\mathbf{H}}}_{2}^{{\mathbb{A}}^{1}}(B\mathbf{G})\longrightarrow{{\mathbf{H}}}_{2}^{{\mathbb{A}}^{1}}(B\mathbf{Q})\longrightarrow{{\mathbf{H}}}_{2}^{{\mathbb{A}}^{1}}(\mathscr{C})\longrightarrow\mathbf{G}^{ab}_{{\mathbb{A}}^{1}}\longrightarrow\mathbf{Q}^{ab}_{{\mathbb{A}}^{1}}\longrightarrow 0,

and moreover the map 𝐇2𝔸1(ΣB𝐍)𝐇2𝔸1(𝒞){{\mathbf{H}}}_{2}^{{\mathbb{A}}^{1}}(\Sigma B\mathbf{N})\to{{\mathbf{H}}}_{2}^{{\mathbb{A}}^{1}}(\mathscr{C}) is an epimorphism.

Morel’s Hurewicz theorem [Mor12, Theorem 6.37] implies that there is an isomorphism 𝐇2𝔸1(ΣB𝐍)𝐇1𝔸1(B𝐍)𝐍𝔸1ab{{\mathbf{H}}}_{2}^{{\mathbb{A}}^{1}}(\Sigma B\mathbf{N})\cong{{\mathbf{H}}}_{1}^{{\mathbb{A}}^{1}}(B\mathbf{N})\cong\mathbf{N}^{ab}_{{\mathbb{A}}^{1}}. The relative Hurewicz theorem 3.6 implies that 𝐇2𝔸1(𝒞){{\mathbf{H}}}_{2}^{{\mathbb{A}}^{1}}(\mathscr{C}) is the initial strictly 𝔸1{\mathbb{A}}^{1}-invariant sheaf under 𝐍\mathbf{N} on which 𝐆\mathbf{G} acts trivially. Combining the points above, we write 𝐍/[𝐍,𝐆]𝔸1\mathbf{N}/[\mathbf{N},\mathbf{G}]_{{\mathbb{A}}^{1}} for this sheaf.

(4). As in the proof of Lemma 4.1, it follows from Theorem 2.5 that 𝐒𝐍/[𝐍,𝐆]\mathbf{S}^{\infty}\mathbf{N}/[\mathbf{N},\mathbf{G}] is the initial 𝔸1{\mathbb{A}}^{1}-invariant sheaf under 𝐍\mathbf{N} with a trivial action by 𝐆\mathbf{G}, and this sheaf is strongly (hence strictly) 𝔸1{\mathbb{A}}^{1}-invariant by Proposition 2.14. It thus coincides with 𝐍/[𝐍,𝐆]𝔸1\mathbf{N}/[\mathbf{N},\mathbf{G}]_{{\mathbb{A}}^{1}} by comparison of universal properties. ∎

Construction 4.9.

If 𝐍\mathbf{N} is a very strongly 𝔸1{\mathbb{A}}^{1}-invariant normal subgroup sheaf of 𝐆\mathbf{G} fitting into a short exact sequence 1𝐍𝐆𝐐11\to\mathbf{N}\to\mathbf{G}\to\mathbf{Q}\to 1, we define [𝐍,𝐆]𝔸1[\mathbf{N},\mathbf{G}]_{{\mathbb{A}}^{1}} as the kernel of the epimorphism

𝐍𝐍/[𝐍,𝐆]𝔸1.\mathbf{N}\longrightarrow\mathbf{N}/[\mathbf{N},\mathbf{G}]_{{\mathbb{A}}^{1}}.

Note that [𝐍,𝐆]𝔸1[\mathbf{N},\mathbf{G}]_{{\mathbb{A}}^{1}} is strongly 𝔸1{\mathbb{A}}^{1}-invariant by [AFH22, Lemma 3.1.14](1).

To iterate this construction and define an 𝔸1{\mathbb{A}}^{1}-lower central series, we need to know that [𝐍,𝐆]𝔸1[\mathbf{N},\mathbf{G}]_{{\mathbb{A}}^{1}} is again very strongly 𝔸1{\mathbb{A}}^{1}-invariant. This is automatic if 𝐆\mathbf{G} is locally nilpotent, by Remark 2.12. We restrict our definition to this case.

Definition 4.10.

Let 𝐆\mathbf{G} be a strongly 𝔸1{\mathbb{A}}^{1}-invariant, locally nilpotent sheaf of groups. The 𝔸1{\mathbb{A}}^{1}-lower central series Γ𝔸1i𝐆\Gamma^{i}_{{\mathbb{A}}^{1}}\mathbf{G}, i1i\geq 1 is defined inductively by setting Γ𝔸11𝐆=𝐆\Gamma^{1}_{{\mathbb{A}}^{1}}\mathbf{G}=\mathbf{G} and Γ𝔸1i𝐆:=[Γ𝔸1i𝐆,𝐆]𝔸1\Gamma^{i}_{{\mathbb{A}}^{1}}\mathbf{G}:=[\Gamma_{{\mathbb{A}}^{1}}^{i}\mathbf{G},\mathbf{G}]_{{\mathbb{A}}^{1}}.

Proposition 4.11.

The following statements hold.

  1. 1.

    The 𝔸1{\mathbb{A}}^{1}-lower central series is functorial, i.e., if f:𝐆𝐆f:\mathbf{G}\to\mathbf{G}^{\prime} is a morphism of strongly 𝔸1{\mathbb{A}}^{1}-invariant sheaves of groups, there are induced morphisms Γ𝔸1i(f):Γ𝔸1i𝐆Γ𝔸1i𝐆\Gamma^{i}_{{\mathbb{A}}^{1}}(f):\Gamma^{i}_{{\mathbb{A}}^{1}}\mathbf{G}\to\Gamma^{i}_{{\mathbb{A}}^{1}}\mathbf{G}^{\prime} for all i0i\geq 0.

  2. 2.

    A strongly 𝔸1{\mathbb{A}}^{1}-invariant sheaf of groups 𝐆\mathbf{G} is 𝔸1{\mathbb{A}}^{1}-nilpotent of nilpotence class c\leq c if and only if the 𝔸1{\mathbb{A}}^{1}-lower central series has length cc.

Proof.

For the first statement, observe that abelianization is functorial by Lemma 4.1 while the lower central series is built inductively by appeal to the exact sequence of Lemma 4.7(2), which is functorial by Lemma 4.7(3). The second statement is established exactly as in the classical setting: an 𝔸1{\mathbb{A}}^{1}-lower central series is an 𝔸1{\mathbb{A}}^{1}-central series in the sense of [AFH22, Definition 3.2.1], and the 𝔸1{\mathbb{A}}^{1}-lower central series is the 𝔸1{\mathbb{A}}^{1}-central series of minimal length. ∎

Example 4.12.

The 𝔸1{\mathbb{A}}^{1}-lower central series of 𝝅1𝔸1(1)\bm{\pi}_{1}^{{\mathbb{A}}^{1}}({\mathbb{P}}^{1}) differs from the upper central series (see [AFH22, Proposition 3.1.22] for discussion of the latter). Indeed, 𝝅1(1)\bm{\pi}_{1}({\mathbb{P}}^{1}) has center 𝐊2MW{{\mathbf{K}}}^{MW}_{2} and 𝔸1{\mathbb{A}}^{1}-nilpotence class 22. In contrast, 𝐇1𝔸1(1)𝐊1MW{{\mathbf{H}}}_{1}^{{\mathbb{A}}^{1}}({\mathbb{P}}^{1})\cong{{\mathbf{K}}}^{MW}_{1} and the epimorphism 𝝅1(1)𝐇1𝔸1(1)\bm{\pi}_{1}({\mathbb{P}}^{1})\to{{\mathbf{H}}}_{1}^{{\mathbb{A}}^{1}}({\mathbb{P}}^{1}) is studied in [CH22, §4]: the kernel is the actual commutator subgroup of 𝝅1(1)\bm{\pi}_{1}({\mathbb{P}}^{1}).

The following result is a motivic analog of a presumably well-known fact about nilpotent groups, though we did not manage to locate a reference.

Proposition 4.13.

Let 𝐆𝐆\mathbf{G}^{\prime}\to\mathbf{G} be a morphism of strongly 𝔸1{\mathbb{A}}^{1}-invariant sheaves of groups, with 𝐆\mathbf{G} locally 𝔸1{\mathbb{A}}^{1}-nilpotent and 𝐆\mathbf{G}^{\prime} very strongly 𝔸1{\mathbb{A}}^{1}-invariant. If 𝐆𝔸1ab𝐆𝔸1ab\mathbf{G}^{\prime ab}_{{\mathbb{A}}^{1}}\to\mathbf{G}^{ab}_{{\mathbb{A}}^{1}} is surjective, then so is 𝐆𝐆\mathbf{G}^{\prime}\to\mathbf{G}.

Proof.

By Lemma 2.9(1), we we may replace 𝐆\mathbf{G}^{\prime} by its image 𝐇\mathbf{H} in 𝐆\mathbf{G}.

We begin with a preparatory observation. Assume 𝐊𝐆\mathbf{K}\subset\mathbf{G} is a strongly 𝔸1{\mathbb{A}}^{1}-invariant subsheaf of groups. For i1i\geq 1 let 𝐊i𝐆\mathbf{K}_{i}\subset\mathbf{G} be the subsheaf generated by Γ𝔸1i𝐆\Gamma^{i}_{{\mathbb{A}}^{1}}\mathbf{G} and 𝐊\mathbf{K}. Since Γ𝔸1i𝐆\Gamma^{i}_{{\mathbb{A}}^{1}}\mathbf{G} is normal in 𝐆\mathbf{G}, there are exact sequences of the form

1Γ𝔸1i𝐆𝐊i𝐊/(𝐊Γ𝔸1i𝐆)1.1\longrightarrow\Gamma^{i}_{{\mathbb{A}}^{1}}\mathbf{G}\longrightarrow\mathbf{K}_{i}\longrightarrow\mathbf{K}/(\mathbf{K}\cap\Gamma^{i}_{{\mathbb{A}}^{1}}\mathbf{G})\longrightarrow 1.

Since 𝐊/(𝐊Γ𝔸1i𝐆)\mathbf{K}/(\mathbf{K}\cap\Gamma^{i}_{{\mathbb{A}}^{1}}\mathbf{G}) is the image of 𝐊\mathbf{K} under the quotient morphism 𝐆𝐆/Γ𝔸1i\mathbf{G}\to\mathbf{G}/\Gamma^{i}_{{\mathbb{A}}^{1}}, it is strongly 𝔸1{\mathbb{A}}^{1}-invariant (𝐊\mathbf{K} being very strongly 𝔸1{\mathbb{A}}^{1}-invariant by Remark 2.12). Thus 𝐊i\mathbf{K}_{i} is strongly 𝔸1{\mathbb{A}}^{1}-invariant by appeal to [AFH22, Lemma 3.1.14(2)].

Next, we claim that 𝐊i+1𝐊i\mathbf{K}_{i+1}\subset\mathbf{K}_{i} is again normal. Since Γ𝔸1i+1𝐆𝐆\Gamma^{i+1}_{{\mathbb{A}}^{1}}\mathbf{G}\subset\mathbf{G} is normal, it suffices to prove (abusing notation slightly) that if k𝐊k\in\mathbf{K} and gΓ𝔸1i𝐆g\in\Gamma^{i}_{{\mathbb{A}}^{1}}\mathbf{G} then kg𝐊i+1k^{g}\in\mathbf{K}_{i+1}. This statement holds since

kg=kk1g1kg=k[k1,g1]𝐊Γ𝔸1i+1𝐆,k^{g}=kk^{-1}g^{-1}kg=k[k^{-1},g^{-1}]\in\mathbf{K}\cdot\Gamma^{i+1}_{{\mathbb{A}}^{1}}\mathbf{G},

the filtration being central.

Now we apply this construction with 𝐊:=𝐇\mathbf{K}:=\mathbf{H}; we shall prove by induction that 𝐇i=𝐆\mathbf{H}_{i}=\mathbf{G} for all ii. Since also

𝐇=i𝐇i\mathbf{H}=\bigcap_{i}\mathbf{H}_{i}

by local nilpotency (if ss^{*} is any stalk functor, then s𝐇s(i𝐇i)s𝐇is^{*}\mathbf{H}\subset s^{*}(\bigcap_{i}\mathbf{H}_{i})\subset s^{*}\mathbf{H}_{i} for any ii, but by assumption s𝐇i=s𝐇s^{*}\mathbf{H}_{i}=s^{*}\mathbf{H} for ii large enough), this will conclude the proof. Noting that 𝐇1=𝐆\mathbf{H}_{1}=\mathbf{G} by definition, we assume that 𝐇n=𝐆\mathbf{H}_{n}=\mathbf{G}. The discussion above shows that 𝐇n+1𝐇n=𝐆\mathbf{H}_{n+1}\subset\mathbf{H}_{n}=\mathbf{G} is normal; set 𝐐:=𝐇n/𝐇n+1\mathbf{Q}:=\mathbf{H}_{n}/\mathbf{H}_{n+1}. The sheaf 𝐇n+1\mathbf{H}_{n+1} is very strongly 𝔸1{\mathbb{A}}^{1}-invariant (Remark 2.12) and the sheaf 𝐐\mathbf{Q} is strongly 𝔸1{\mathbb{A}}^{1}-invariant by appeal to Lemma 2.7.

Since 𝐇𝔸1ab\mathbf{H}^{ab}_{{\mathbb{A}}^{1}} surjects onto 𝐆𝔸1ab(𝐇n)𝔸1ab\mathbf{G}^{ab}_{{\mathbb{A}}^{1}}\simeq(\mathbf{H}_{n})^{ab}_{{\mathbb{A}}^{1}}, also (𝐇n+1)𝔸1ab(𝐇n)𝔸1ab(\mathbf{H}_{n+1})^{ab}_{{\mathbb{A}}^{1}}\to(\mathbf{H}_{n})^{ab}_{{\mathbb{A}}^{1}} is surjective. The functor ()𝔸1ab(-)^{ab}_{{\mathbb{A}}^{1}} is a left adjoint and hence preserves quotients; thus 𝐐𝔸1ab=0\mathbf{Q}^{ab}_{{\mathbb{A}}^{1}}=0. Note also that 𝐐\mathbf{Q} is necessarily locally 𝔸1{\mathbb{A}}^{1}-nilpotent as a quotient of the locally 𝔸1{\mathbb{A}}^{1}-nilpotent sheaf of groups 𝐆\mathbf{G} [AFH22, Lemma 3.2.7(2)]. Since 𝐐𝔸1ab\mathbf{Q}^{ab}_{{\mathbb{A}}^{1}} is trivial and 𝐐\mathbf{Q} is locally 𝔸1{\mathbb{A}}^{1}-nilpotent, one may inductively deduce that 𝐐=0\mathbf{Q}=0. Thus, we conclude that 𝐇n+1=𝐆\mathbf{H}_{n+1}=\mathbf{G} as well. ∎

The next result is an analog of [Sta65, Lemma 3.1].

Proposition 4.14.

If f:𝐆1𝐆2f:\mathbf{G}_{1}\to\mathbf{G}_{2} is a morphism of strongly 𝔸1{\mathbb{A}}^{1}-invariant, locally nilpotent sheaves of groups such that 𝐇1𝔸1(Bf){{\mathbf{H}}}_{1}^{{\mathbb{A}}^{1}}(Bf) is an isomorphism and 𝐇2𝔸1(Bf){{\mathbf{H}}}_{2}^{{\mathbb{A}}^{1}}(Bf) is an epimorphism, then

𝐆1/Γ𝔸1i+1𝐆1𝐆2/Γ𝔸1i+1𝐆2\mathbf{G}_{1}/\Gamma^{i+1}_{{\mathbb{A}}^{1}}\mathbf{G}_{1}\longrightarrow\mathbf{G}_{2}/\Gamma^{i+1}_{{\mathbb{A}}^{1}}\mathbf{G}_{2}

is an isomorphism for all i1i\geq 1. In particular, if 𝐆1\mathbf{G}_{1} and 𝐆2\mathbf{G}_{2} are locally 𝔸1{\mathbb{A}}^{1}-nilpotent, then ff is an isomorphism.

Proof.

Taking 𝐆=𝐆i\mathbf{G}=\mathbf{G}_{i}, 𝐍=Γ𝔸1r𝐆i\mathbf{N}=\Gamma^{r}_{{\mathbb{A}}^{1}}\mathbf{G}_{i}, Lemma 4.7 in conjunction with the definition of the 𝔸1{\mathbb{A}}^{1}-lower central series yields a morphism of commutative diagrams of the form

𝐇2𝔸1(B𝐆1)\textstyle{{{\mathbf{H}}}_{2}^{{\mathbb{A}}^{1}}(B\mathbf{G}_{1})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝐇2𝔸1(B𝐆1/Γ𝔸1r𝐆1)\textstyle{{{\mathbf{H}}}_{2}^{{\mathbb{A}}^{1}}(B\mathbf{G}_{1}/\Gamma^{r}_{{\mathbb{A}}^{1}}\mathbf{G}_{1})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Γ𝔸1r𝐆1/Γ𝔸1r+1𝐆1\textstyle{\Gamma^{r}_{{\mathbb{A}}^{1}}\mathbf{G}_{1}/\Gamma^{r+1}_{{\mathbb{A}}^{1}}\mathbf{G}_{1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝐇1𝔸1(B𝐆1)\textstyle{{{\mathbf{H}}}_{1}^{{\mathbb{A}}^{1}}(B\mathbf{G}_{1})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝐇1𝔸1(B𝐆1/Γ𝔸1r𝐆1)\textstyle{{{\mathbf{H}}}_{1}^{{\mathbb{A}}^{1}}(B\mathbf{G}_{1}/\Gamma^{r}_{{\mathbb{A}}^{1}}\mathbf{G}_{1})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝐇2𝔸1(B𝐆2)\textstyle{{{\mathbf{H}}}_{2}^{{\mathbb{A}}^{1}}(B\mathbf{G}_{2})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝐇2𝔸1(B𝐆2/Γ𝔸1r𝐆2)\textstyle{{{\mathbf{H}}}_{2}^{{\mathbb{A}}^{1}}(B\mathbf{G}_{2}/\Gamma^{r}_{{\mathbb{A}}^{1}}\mathbf{G}_{2})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Γ𝔸1r𝐆2/Γ𝔸1r+1𝐆2\textstyle{\Gamma^{r}_{{\mathbb{A}}^{1}}\mathbf{G}_{2}/\Gamma^{r+1}_{{\mathbb{A}}^{1}}\mathbf{G}_{2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝐇1𝔸1(B𝐆2)\textstyle{{{\mathbf{H}}}_{1}^{{\mathbb{A}}^{1}}(B\mathbf{G}_{2})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝐇1𝔸1(B𝐆2/Γ𝔸1r𝐆2)\textstyle{{{\mathbf{H}}}_{1}^{{\mathbb{A}}^{1}}(B\mathbf{G}_{2}/\Gamma^{r}_{{\mathbb{A}}^{1}}\mathbf{G}_{2})}

Assume inductively that 𝐆1/Γ𝔸1r𝐆1𝐆2/Γ𝔸1r𝐆2\mathbf{G}_{1}/\Gamma^{r}_{{\mathbb{A}}^{1}}\mathbf{G}_{1}\to\mathbf{G}_{2}/\Gamma^{r}_{{\mathbb{A}}^{1}}\mathbf{G}_{2} is an isomorphism; the base case r=2r=2 follows from the assumption that 𝐇1(Bf){{\mathbf{H}}}_{1}(Bf) is an isomorphism. In that case, the second and fifth vertical arrows from the left are isomorphisms. By hypothesis, the first vertical arrow is an epimorphism and the fourth vertical arrow is an isomorphism by assumption. The five lemma implies that the third vertical arrow is an isomorphism as well. Then, functoriality of the lower central series in conjunction with the short exact sequences

1\textstyle{1\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Γ𝔸1r𝐆i/Γ𝔸1r+1𝐆i\textstyle{\Gamma^{r}_{{\mathbb{A}}^{1}}\mathbf{G}_{i}/\Gamma^{r+1}_{{\mathbb{A}}^{1}}\mathbf{G}_{i}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝐆i/Γ𝔸1r+1𝐆i\textstyle{\mathbf{G}_{i}/\Gamma^{r+1}_{{\mathbb{A}}^{1}}\mathbf{G}_{i}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝐆i/Γ𝔸1r𝐆i\textstyle{\mathbf{G}_{i}/\Gamma^{r}_{{\mathbb{A}}^{1}}\mathbf{G}_{i}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}1\textstyle{1}

together with the five lemma allow us to conclude that 𝐆1/Γ𝔸1r+1𝐆1𝐆2/Γ𝔸1r+1𝐆2\mathbf{G}_{1}/\Gamma^{r+1}_{{\mathbb{A}}^{1}}\mathbf{G}_{1}\to\mathbf{G}_{2}/\Gamma^{r+1}_{{\mathbb{A}}^{1}}\mathbf{G}_{2} is an isomorphism. ∎

More generally, we can refine the 𝔸1{\mathbb{A}}^{1}-lower central series of a strongly 𝔸1{\mathbb{A}}^{1}-invariant sheaf 𝐆\mathbf{G} to an 𝔸1{\mathbb{A}}^{1}-lower central series for an action.

Construction 4.15.

Suppose 𝝅\bm{\pi} is a strongly 𝔸1{\mathbb{A}}^{1}-invariant sheaf of groups and 𝐆\mathbf{G} is a very strongly 𝔸1{\mathbb{A}}^{1}-invariant sheaf with 𝝅\bm{\pi}-action. In that case, there is a largest quotient sheaf of groups 𝐆𝝅\mathbf{G}_{\bm{\pi}} on which 𝝅\bm{\pi} acts trivially. The sheaf 𝐆𝝅\mathbf{G}_{\bm{\pi}} need not be 𝔸1{\mathbb{A}}^{1}-invariant. However, Theorem 2.5 implies that

𝐆𝐆𝝅𝐒𝐆𝝅\mathbf{G}\longrightarrow\mathbf{G}_{\bm{\pi}}\longrightarrow\mathbf{S}^{\infty}\mathbf{G}_{\bm{\pi}}

is an epimorphism with 𝐒𝐆𝝅\mathbf{S}^{\infty}\mathbf{G}_{\bm{\pi}} 𝔸1{\mathbb{A}}^{1}-invariant. Very strong invariance of 𝐆\mathbf{G} implies that 𝐒𝐆𝝅\mathbf{S}^{\infty}\mathbf{G}_{\bm{\pi}} is strongly 𝔸1{\mathbb{A}}^{1}-invariant. It is thus necessarily the initial strongly 𝔸1{\mathbb{A}}^{1}-invariant quotient of 𝐆\mathbf{G} with trivial 𝝅\bm{\pi}-action. In that case, we define

Γ𝝅2𝐆:=ker(𝐆𝐒𝐆𝝅),\Gamma^{2}_{\bm{\pi}}\mathbf{G}:=\ker(\mathbf{G}\to\mathbf{S}^{\infty}\mathbf{G}_{\bm{\pi}}),

which is strongly 𝔸1{\mathbb{A}}^{1}-invariant by appeal to [AFH22, Lemma 3.1.14(1)]. If 𝐆\mathbf{G} is locally nilpotent, then for n3n\geq 3, we inductively define

Γ𝝅n𝐆:=Γ𝝅2Γ𝝅n1𝐆.\Gamma^{n}_{\bm{\pi}}\mathbf{G}:=\Gamma^{2}_{\bm{\pi}}\Gamma^{n-1}_{\bm{\pi}}\mathbf{G}.

By definition, Γ𝝅i𝐆\Gamma^{i}_{\bm{\pi}}\mathbf{G} is a 𝝅\bm{\pi}-central series for 𝐆\mathbf{G} in the sense of [AFH22, Definition 3.2.1].

Proposition 4.16.

Suppose 𝐆\mathbf{G} is a strongly 𝔸1{\mathbb{A}}^{1}-invariant, locally nilpotent sheaf with action of 𝛑\bm{\pi}.

  1. 1.

    The series Γ𝝅i𝐆\Gamma^{i}_{\bm{\pi}}\mathbf{G} is functorial in 𝝅\bm{\pi} and 𝐆\mathbf{G}.

  2. 2.

    When 𝝅=𝐆\bm{\pi}=\mathbf{G} acting by conjugation on 𝐆\mathbf{G}, Γ𝐆i𝐆\Gamma^{i}_{\mathbf{G}}\mathbf{G} coincides with the 𝔸1{\mathbb{A}}^{1}-lower central series of 𝐆\mathbf{G} of Definition 4.10.

  3. 3.

    The action of 𝝅\bm{\pi} on 𝐆\mathbf{G} is 𝔸1{\mathbb{A}}^{1}-nilpotent if and only if Γ𝝅i𝐆\Gamma^{i}_{\bm{\pi}}\mathbf{G} terminates after finitely many steps.

Proof.

Under the assumptions, 𝐆\mathbf{G} is very strongly 𝔸1{\mathbb{A}}^{1}-invariant by appeal to Proposition 2.11. Granted that observation, the first statement is immediate from the fact that 𝐒\mathbf{S}^{\infty} is functorial (Theorem 2.5(1)). The second statement follows from the fact that the corresponding statement holds for the classical central series. The final statement is immediate from the definitions. ∎

4.17.

Suppose 𝐀\mathbf{A} is a strictly 𝔸1{\mathbb{A}}^{1}-invariant sheaf with an action of a strongly 𝔸1{\mathbb{A}}^{1}-invariant sheaf of groups 𝝅\bm{\pi}. We may consider the twisted Eilenberg-Mac Lane space K𝝅(𝐀,n)K^{\bm{\pi}}(\mathbf{A},n) for any integer n2n\geq 2. In that case, there is a fiber sequence

K(𝐀,n)K𝝅(𝐀,n)fB𝝅.K(\mathbf{A},n)\longrightarrow K^{\bm{\pi}}(\mathbf{A},n)\stackrel{{\scriptstyle f}}{{\longrightarrow}}B\bm{\pi}.

The relative Hurewicz theorem 3.6 tells us that 𝐇n𝔸1(cof(f)){{\mathbf{H}}}_{n}^{{\mathbb{A}}^{1}}(\mathrm{cof}(f)) is the largest strictly 𝔸1{\mathbb{A}}^{1}-invariant quotient of 𝐀\mathbf{A} on which 𝝅\bm{\pi} acts trivially. We write 𝐇0𝔸1(𝝅,𝐀){{\mathbf{H}}}_{0}^{{\mathbb{A}}^{1}}(\bm{\pi},\mathbf{A}) for this strictly 𝔸1{\mathbb{A}}^{1}-invariant sheaf, which we call the sheaf of 𝔸1{\mathbb{A}}^{1}-coinvariants of the 𝝅\bm{\pi}-action on 𝐀\mathbf{A}. In particular if the action of 𝝅\bm{\pi} on 𝐀\mathbf{A} is trivial, then 𝐇0𝔸1(𝝅;𝐀)=𝐀{{\mathbf{H}}}_{0}^{{\mathbb{A}}^{1}}(\bm{\pi};\mathbf{A})=\mathbf{A}.

Proposition 4.18.

Assume 𝐀\mathbf{A} is a 𝛑\bm{\pi}-module.

  1. 1.

    The sheaf 𝐇0𝔸1(𝝅;𝐀){{\mathbf{H}}}_{0}^{{\mathbb{A}}^{1}}(\bm{\pi};\mathbf{A}) is functorial in 𝝅\bm{\pi} and 𝐀\mathbf{A}.

  2. 2.

    The sheaf 𝐇0𝔸1(𝝅;𝐀){{\mathbf{H}}}_{0}^{{\mathbb{A}}^{1}}(\bm{\pi};\mathbf{A}) coincides with 𝐒𝐀𝝅\mathbf{S}^{\infty}\mathbf{A}_{\bm{\pi}} from Construction 4.15.

Proof.

The first statement is immediate from the definitions. The second statement follows because 𝐇0𝔸1(𝝅;𝐀){{\mathbf{H}}}_{0}^{{\mathbb{A}}^{1}}(\bm{\pi};\mathbf{A}) and 𝐒𝐀𝝅\mathbf{S}^{\infty}\mathbf{A}_{\bm{\pi}} have the same universal property: they are the largest strictly 𝔸1{\mathbb{A}}^{1}-invariant quotients of 𝐀\mathbf{A} with trivial action of 𝝅\bm{\pi}. ∎

Remark 4.19.

Another example arises from the mod pp lower central series (or Stallings filtration) of a group [Sta65]. Recall that if pp is a fixed integer, GG is a group and UU is a subgroup, then GpUG\sharp_{p}U is the subgroup generated by [g,u]vp[g,u]v^{p} for gG,u,vUg\in G,u,v\in U. If UU is normal in GG, then GpUG\sharp_{p}U is normal in GG. One then defines the mod pp lower central series inductively by setting Γp1G=G\Gamma^{1}_{p}G=G, Γpi+1G:=GpΓpiG\Gamma^{i+1}_{p}G:=G\sharp_{p}\Gamma^{i}_{p}G. The construction above yields a central series which for p=0p=0 coincides with the lower central series, and for pp a prime is the largest descending central series with subquotients that are 𝔽p\mathbb{F}_{p}-vector spaces. Since the above construction is evidently functorial it makes sense for presheaves of groups and then for sheaves of groups by sheafication. In that case, we can define the mod pp-𝔸1{\mathbb{A}}^{1}-lower central series following the procedure above. If 𝐆\mathbf{G} is a strongly 𝔸1{\mathbb{A}}^{1}-invariant, locally nilpotent sheaf of groups, we define Γ𝔸1,pi𝐆\Gamma^{i}_{{\mathbb{A}}^{1},p}\mathbf{G} by setting Γ𝔸1,p1𝐆=𝐆\Gamma^{1}_{{\mathbb{A}}^{1},p}\mathbf{G}=\mathbf{G} and then inductively defining

Γ𝔸1,pi+1𝐆:=ker(𝐆𝐒𝐆/𝐆pΓ𝔸1,pi𝐆).\Gamma^{i+1}_{{\mathbb{A}}^{1},p}\mathbf{G}:=\ker(\mathbf{G}\longrightarrow\mathbf{S}^{\infty}\mathbf{G}/\mathbf{G}\sharp_{p}\Gamma^{i}_{{\mathbb{A}}^{1},p}\mathbf{G}).

Theorem 2.5 and Proposition 2.14 then imply that Γ𝔸1,pi+1𝐆\Gamma^{i+1}_{{\mathbb{A}}^{1},p}\mathbf{G} is a functorial decreasing central series with successive subquotients that are strictly 𝔸1{\mathbb{A}}^{1}-invariant sheaves of 𝔽p{\mathbb{F}}_{p}-vector spaces.

5 Principal refinements of Moore–Postnikov factorizations

Recall that a pointed, connected motivic space 𝒳\mathscr{X} is called nilpotent if 𝝅1(𝒳)\bm{\pi}_{1}(\mathscr{X}) is 𝔸1{\mathbb{A}}^{1}-nilpotent and 𝝅1(𝒳)\bm{\pi}_{1}(\mathscr{X}) acts 𝔸1{\mathbb{A}}^{1}-nilpotently on higher homotopy sheaves [AFH22, Definition 3.3.1]. More generally if f:f:\mathscr{E}\to\mathscr{B} is a morphism of pointed connected motivic spaces, then ff is nilpotent if fib(f)\mathrm{fib}(f) is connected and the action of 𝝅1()\bm{\pi}_{1}(\mathscr{E}) on 𝝅i(fib(f))\bm{\pi}_{i}(\mathrm{fib}(f)) is 𝔸1{\mathbb{A}}^{1}-nilpotent for all i1i\geq 1. The relative Hurewicz theorem 3.6 in conjunction with the functorial central series for nilpotent actions from Proposition 4.16 allows us to construct a functorial principal refinement of the Moore-Postnikov tower improving [AFH22, Theorem 3.3.13].

Theorem 5.1.

Assume kk is a perfect field. Suppose f:f:\mathscr{E}\to\mathscr{B} is a morphism of pointed, connected motivic spaces. If ff is nilpotent, and 𝛑1(fib(f))\bm{\pi}_{1}(\mathrm{fib}(f)) is very strongly 𝔸1{\mathbb{A}}^{1}-invariant, then ff admits a functorial principal refinement. More precisely, if f:f^{\prime}:\mathscr{E}^{\prime}\to\mathscr{B} is another nilpotent morphism of pointed, connected motivic spaces with 𝛑1(fib(f))\bm{\pi}_{1}(\mathrm{fib}(f^{\prime})) very strongly 𝔸1{\mathbb{A}}^{1}-invariant, and g:g:\mathscr{E}\to\mathscr{E}^{\prime} is a morphism of spaces over \mathscr{B}, then gg induces homomorphisms

Γi(g):Γ𝝅1()i𝝅j(fib(f))Γ𝝅1()i𝝅j(fib(f))\Gamma^{i}(g):\Gamma^{i}_{\bm{\pi}_{1}(\mathscr{E})}\bm{\pi}_{j}(\mathrm{fib}(f))\longrightarrow\Gamma^{i}_{\bm{\pi}_{1}(\mathscr{E}^{\prime})}\bm{\pi}_{j}(\mathrm{fib}(f^{\prime}))

for all i,ji,j which induce morphisms of layers of the principal refinements of ff and ff^{\prime}.

Proof.

Repeat the proof of [AFH22, Corollary 4.2.4] using the 𝔸1{\mathbb{A}}^{1}-lower central series for the action of 𝝅1()\bm{\pi}_{1}(\mathscr{E}) on fib(f)\mathrm{fib}(f) of Construction 4.15. Functoriality follows from Proposition 4.16 upon replacing appeals to [AFH22, Theorem 4.2.1] with appeals to Theorem 3.6 (in view of Proposition 4.18). ∎

With the above preparations at hand, we can now establish the analog of the Whitehead theorem for nilpotent motivic spaces extending [AFH22, Theorem 4.2.2]. In order to formulate the next theorem, write SHS1(k){\mathrm{SH}}^{S^{1}}(k) for the stabilization of Spc(k)\mathrm{Spc}({k}) in the sense of [Lur17, §1.4]. We write ΣS1:Spc(k)SHS1(k):Ω\Sigma^{\infty}_{S^{1}}:\mathrm{Spc}({k})_{*}\leftrightarrows{\mathrm{SH}}^{S^{1}}(k):\Omega^{\infty} for the canonical adjunction. The functor ΣS1\Sigma^{\infty}_{S^{1}} preserves cofiber sequences, being a left adjoint.

Theorem 5.2.

Assume kk is a perfect field. Suppose f:𝒳𝒴f:\mathscr{X}\to\mathscr{Y} is a (pointed) morphism of nilpotent motivic spaces. If n0n\geq 0 is an integer, the following statements are equivalent.

  1. 1.

    The map ff has (n1)(n-1)-connected fibers.

  2. 2.

    The map ΣS1f\Sigma^{\infty}_{S^{1}}f has (n1)(n-1)-connected fiber.

  3. 3.

    The maps 𝐇i𝔸1(f){{\mathbf{H}}}_{i}^{{\mathbb{A}}^{1}}(f) are isomorphisms for i<ni<n and 𝐇n𝔸1(f){{\mathbf{H}}}_{n}^{{\mathbb{A}}^{1}}(f) is an epimorphism.

Proof.

The implications (1) \Longrightarrow (2) \Longrightarrow (3) follow by appeal to Proposition 3.3(1). For the reverse implications we proceed as follows. Assume (3); we shall prove (1).

The statement is immediate if n=0n=0 since nilpotent spaces are connected by assumption. If n=1n=1 we must prove that 𝝅1(f)\bm{\pi}_{1}(f) is surjective knowing only that 𝝅1(f)𝔸1ab\bm{\pi}_{1}(f)^{ab}_{{\mathbb{A}}^{1}} is surjective; this is precisely Proposition 4.13.

We therefore assume n2n\geq 2, in which case we know that 𝝅1(f)\bm{\pi}_{1}(f) is an epimorphism by what we just observed. Then, by appeal to Theorem 5.1 we know that the Moore–Postnikov factorization of ff admits a principal refinement. In other words, there are fiber sequences

𝒴j+1𝒴jK(𝐀j,nj),\mathscr{Y}_{j+1}\longrightarrow\mathscr{Y}_{j}\longrightarrow K(\mathbf{A}_{j},n_{j}),

with njn_{j} a weakly increasing sequence of integers, nj2n_{j}\geq 2, 𝒴0=𝒴\mathscr{Y}_{0}=\mathscr{Y} and such that the connectivity of 𝒳𝒴N\mathscr{X}\to\mathscr{Y}_{N} is tends to infinity as NN tends to \infty. Without loss of generality, we may also assume that 𝐀j0\mathbf{A}_{j}\neq 0.

Essentially by construction, n0=t+1n_{0}=t+1 if and only if 𝝅tfib(f)\bm{\pi}_{t}\mathrm{fib}(f) is the first non-vanishing homotopy group. Our goal is thus to prove that n0n+1n_{0}\geq n+1. Suppose, in the interest of obtaining a contradiction that n0nn_{0}\leq n. The long exact homology sequence for the cofiber of 𝒴j+1𝒴j\mathscr{Y}_{j+1}\to\mathscr{Y}_{j} together with the relative Hurewicz Theorem 3.6 yield (using that the action of 𝝅1𝒳\bm{\pi}_{1}\mathscr{X} on 𝐀j\mathbf{A}_{j} is trivial)

𝐇n0𝔸1(𝒴j+1)𝐇n0𝔸1(𝒴j)𝐀j𝐇n01𝔸1(𝒴j+1)𝐇n01𝔸1(𝒴j)0;{{\mathbf{H}}}_{n_{0}}^{{\mathbb{A}}^{1}}(\mathscr{Y}_{j+1})\longrightarrow{{\mathbf{H}}}_{n_{0}}^{{\mathbb{A}}^{1}}(\mathscr{Y}_{j})\longrightarrow\mathbf{A}_{j}^{\prime}\longrightarrow{{\mathbf{H}}}_{n_{0}-1}^{{\mathbb{A}}^{1}}(\mathscr{Y}_{j+1})\longrightarrow{{\mathbf{H}}}_{n_{0}-1}^{{\mathbb{A}}^{1}}(\mathscr{Y}_{j})\longrightarrow 0;

where 𝐀j=𝐀j\mathbf{A}_{j}^{\prime}=\mathbf{A}_{j} if nj=n0n_{j}=n_{0} and 𝐀j=0\mathbf{A}_{j}^{\prime}=0 otherwise. We deduce that 𝐇n01𝔸1(𝒴j+1)𝐇n01𝔸1(𝒴j){{\mathbf{H}}}_{n_{0}-1}^{{\mathbb{A}}^{1}}(\mathscr{Y}_{j+1})\to{{\mathbf{H}}}_{n_{0}-1}^{{\mathbb{A}}^{1}}(\mathscr{Y}_{j}) is surjective for every jj. Since 𝒳𝒴N\mathscr{X}\to\mathscr{Y}_{N} is highly connected and we have already established that (1) implies (3), we find that 𝐇n01𝔸1(𝒳)𝐇n01𝔸1(𝒴N)𝐇n01𝔸1(𝒴1){{\mathbf{H}}}_{n_{0}-1}^{{\mathbb{A}}^{1}}(\mathscr{X})\to{{\mathbf{H}}}_{n_{0}-1}^{{\mathbb{A}}^{1}}(\mathscr{Y}_{N})\to{{\mathbf{H}}}_{n_{0}-1}^{{\mathbb{A}}^{1}}(\mathscr{Y}_{1}) is surjective. On the other hand since n0nn_{0}\leq n, 𝐇n01𝔸1(𝒳)𝐇n01𝔸1(𝒴){{\mathbf{H}}}_{n_{0}-1}^{{\mathbb{A}}^{1}}(\mathscr{X})\to{{\mathbf{H}}}_{n_{0}-1}^{{\mathbb{A}}^{1}}(\mathscr{Y}) is an isomorphism, whence also 𝐇n01𝔸1(𝒴1)𝐇n01𝔸1(𝒴){{\mathbf{H}}}_{n_{0}-1}^{{\mathbb{A}}^{1}}(\mathscr{Y}_{1})\simeq{{\mathbf{H}}}_{n_{0}-1}^{{\mathbb{A}}^{1}}(\mathscr{Y}). Finally since 𝐇n0𝔸1(𝒳)𝐇n0𝔸1(𝒴){{\mathbf{H}}}_{n_{0}}^{{\mathbb{A}}^{1}}(\mathscr{X})\to{{\mathbf{H}}}_{n_{0}}^{{\mathbb{A}}^{1}}(\mathscr{Y}) is surjective so is 𝐇n0𝔸1(𝒴1)𝐇n0𝔸1(𝒴){{\mathbf{H}}}_{n_{0}}^{{\mathbb{A}}^{1}}(\mathscr{Y}_{1})\to{{\mathbf{H}}}_{n_{0}}^{{\mathbb{A}}^{1}}(\mathscr{Y}), whence we get (from the above six term exact sequence)

0𝐀0𝐇n01𝔸1(𝒴1)𝐇n01𝔸1(𝒴)0.0\longrightarrow\mathbf{A}_{0}\longrightarrow{{\mathbf{H}}}_{n_{0}-1}^{{\mathbb{A}}^{1}}(\mathscr{Y}_{1})\longrightarrow{{\mathbf{H}}}_{n_{0}-1}^{{\mathbb{A}}^{1}}(\mathscr{Y})\longrightarrow 0.

Since the second map is an isomorphism as shown above, we learn that 𝐀0=0\mathbf{A}_{0}=0, which is the desired contradiction. ∎

Remark 5.3.

Assume 𝒳\mathscr{X} is a pointed connected motivic space with 𝝅1(𝒳)=:𝝅1\bm{\pi}_{1}(\mathscr{X})=:\bm{\pi}_{1}. The map 𝒳B𝝅1\mathscr{X}\to B\bm{\pi}_{1} has 11-connected fibers, and the argument for Theorem 5.2 implies the motivic analog of a result of Hopf [Hop42, Satz II] that 𝐇2𝔸1(𝒳)𝐇2𝔸1(B𝝅1){{\mathbf{H}}}_{2}^{{\mathbb{A}}^{1}}(\mathscr{X})\to{{\mathbf{H}}}_{2}^{{\mathbb{A}}^{1}}(B\bm{\pi}_{1}) is an epimorphism.

Remark 5.4.

Of course, the Whitehead theorem for nilpotent motivic spaces could also be deduced following Bousfield [Bou75, Lemma 8.9]. Consider Morel’s 𝔸1{\mathbb{A}}^{1}-homology theory 𝐇𝔸1{{\mathbf{H}}}^{{\mathbb{A}}^{1}}_{*}. We can localize Spc(k)\mathrm{Spc}({k}) with respect to this homology theory; following [AFH22, Definition 4.3.1] this localization yields the {\mathbb{Z}}-𝔸1{\mathbb{A}}^{1}-local homotopy category. A morphism f:𝒳𝒴f:\mathscr{X}\to\mathscr{Y} is an 𝐇𝔸1{{\mathbf{H}}}^{{\mathbb{A}}^{1}}_{*}-local equivalence if it induces an isomorphism after applying 𝐇𝔸1{{\mathbf{H}}}^{{\mathbb{A}}^{1}}_{*}. Theorem 5.2 implies that nilpotent motivic spaces are 𝐇𝔸1{{\mathbf{H}}}^{{\mathbb{A}}^{1}}_{*}-local.

Ignoring base-point issues, one could check the converse as follows. First, observe that K(𝐀,n)K(\mathbf{A},n) is an 𝐇𝔸1{{\mathbf{H}}}^{{\mathbb{A}}^{1}}_{*}-local space for any n0n\geq 0. Indeed, this follows from the existence of the universal coefficient spectral sequence (see [AFH22, Proof of Proposition 4.1.2]). In that case, any limit of 𝐇𝔸1{{\mathbf{H}}}^{{\mathbb{A}}^{1}}_{*}-local spaces is 𝐇𝔸1{{\mathbf{H}}}^{{\mathbb{A}}^{1}}_{*}-local. The existence of a principal refinement of the Postnikov tower then implies that nilpotent motivic spaces are 𝐇𝔸1{{\mathbf{H}}}^{{\mathbb{A}}^{1}}_{*}-local.

Aside: Moore–Postnikov towers for locally 𝔸1{\mathbb{A}}^{1}-nilpotent morphisms

In [AFH22, Definition 3.3.1], the notion of a locally 𝔸1{\mathbb{A}}^{1}-nilpotent morphism of motivic spaces was also introduced: these are morphisms f:f:\mathscr{E}\to\mathscr{B} such that fib(f)\mathrm{fib}(f) is connected, and the action of 𝝅1()\bm{\pi}_{1}(\mathscr{E}) on 𝝅1(fib(f))\bm{\pi}_{1}(\mathrm{fib}(f)) is locally 𝔸1{\mathbb{A}}^{1}-nilpotent for all i1i\geq 1. The functorial principal refinements of Moore–Postnikov factorizations described above can be extended to locally 𝔸1{\mathbb{A}}^{1}-nilpotent morphisms of motivic spaces by limiting processes; we briefly explain this here.

Let us revisit the construction of principal refinements of nilpotent morphisms. Suppose f:f:\mathscr{E}\to\mathscr{B} is a morphism of pointed, connected motivic spaces such that 𝝅1(f)\bm{\pi}_{1}(f) is an epimorphism. Set =fib(f)\mathscr{F}=\mathrm{fib}(f). Set i=τi\mathscr{E}_{i}=\tau_{\leq i}\mathscr{E}, where the Postnikov truncation is taken in the topos ShvNis(Smk)/{\mathrm{Shv}}_{\operatorname{Nis}}(\mathrm{Sm}_{k})_{/\mathscr{B}}. Viewed as an object of ShvNis(Smk){\mathrm{Shv}}_{\operatorname{Nis}}(\mathrm{Sm}_{k}) (i.e., ignoring the map to \mathscr{B}), i\mathscr{E}_{i} is just τi\tau_{\leq i}\mathscr{F}. The object i+1ShvNis(Smk)/i\mathscr{E}_{i+1}\in{\mathrm{Shv}}_{\operatorname{Nis}}(\mathrm{Sm}_{k})_{/\mathscr{E}_{i}} is an (i+1)(i+1)-gerbe and hence (at least if i>1i>1) classified by a map Ki(𝐀i+1,i+2)ShvNis(Smk)/i*\to K_{\mathscr{E}_{i}}(\mathbf{A}_{i+1},i+2)\in{\mathrm{Shv}}_{\operatorname{Nis}}(\mathrm{Sm}_{k})_{/\mathscr{E}_{i}} [Lur09, Theorem 7.2.2.26], for some 𝐀i+1Ab(ShvNis(Smk)/i)\mathbf{A}_{i+1}\in\mathrm{Ab}({\mathrm{Shv}}_{\operatorname{Nis}}(\mathrm{Sm}_{k})_{/\mathscr{E}_{i}}). Concretely, 𝐀i+1=𝝅i+1\mathbf{A}_{i+1}=\bm{\pi}_{i+1}\mathscr{F}, with its action by 𝝅1i𝝅1\bm{\pi}_{1}\mathscr{E}_{i}\simeq\bm{\pi}_{1}\mathscr{E}.

Suppose given a (strongly 𝔸1{\mathbb{A}}^{1}-invariant) quotient 𝐀i+1,0:=𝐀i+1𝐐i+1,1\mathbf{A}_{i+1,0}:=\mathbf{A}_{i+1}\to\mathbf{Q}_{i+1,1}. Let i,1\mathscr{E}_{i,1} be the fiber of the composite i,0:=iKi(𝐐i+1,1,i+2)\mathscr{E}_{i,0}:=\mathscr{E}_{i}\to K_{\mathscr{E}_{i}}(\mathbf{Q}_{i+1,1},i+2). Then there is a canonical map i+1i,1\mathscr{E}_{i+1}\to\mathscr{E}_{i,1} exhibiting i+1ShvNis(Smk)/i,1\mathscr{E}_{i+1}\in{\mathrm{Shv}}_{\operatorname{Nis}}(\mathrm{Sm}_{k})_{/\mathscr{E}_{i,1}}. This is another (i+1)(i+1)-gerbe, this time banded by 𝐀i+1,1:=ker(𝐀i+1,0𝐐i+1,1)\mathbf{A}_{i+1,1}:=\ker(\mathbf{A}_{i+1,0}\to\mathbf{Q}_{i+1,1}), and hence classified by a map Ki,1(𝐀i+1,1,i+2)ShvNis(Smk)/i,1*\to K_{\mathscr{E}_{i,1}}(\mathbf{A}_{i+1,1},i+2)\in{\mathrm{Shv}}_{\operatorname{Nis}}(\mathrm{Sm}_{k})_{/\mathscr{E}_{i,1}}. Provided we are given a decreasing filtration of 𝐀i+1\mathbf{A}_{i+1}, we can keep repeating this process. Assuming furthermore that the action of 𝝅1()\bm{\pi}_{1}(\mathscr{E}) on 𝐐i+1,j\mathbf{Q}_{i+1,j} is trivial, we obtain principalized fiber sequences

i,j+1i,jK(𝐐i+1,j,i+2)Spc(k).\mathscr{E}_{i,j+1}\longrightarrow\mathscr{E}_{i,j}\longrightarrow K(\mathbf{Q}_{i+1,j},i+2)\in\mathrm{Spc}(k)_{*}.

If ff is nilpotent, then the filtrations can be chosen to be finite, and so i,ji+1\mathscr{E}_{i,j}\simeq\mathscr{E}_{i+1} for j0j\gg 0. This way \mathscr{E} is built from \mathscr{B} by a sequence of principal fibrations.

Now suppose that ff is only locally 𝔸1{\mathbb{A}}^{1}-nilpotent. Then all steps in this process can still be performed. The only issue is that we may not have i,ji+1\mathscr{E}_{i,j}\simeq\mathscr{E}_{i+1} for any finite jj. However, we still have the following.

Proposition 5.5.

Let f:f:\mathscr{E}\to\mathscr{B} be a locally 𝔸1{\mathbb{A}}^{1}-nilpotent morphism of motivic spaces. Using notation as above, we have

ilimji,j.\mathscr{E}_{i}\simeq\lim_{j}\mathscr{E}_{i,j}.
Proof.

Let XSmkX\in\mathrm{Sm}_{k} be connected. It will suffice to show the equivalence after restricting to the small Zariski site XZarX_{\operatorname{Zar}} (note that this restriction preserves limits). By assumption the filtration on the generic stalk of XX is of finite length. Since the 𝐀i,j\mathbf{A}_{i,j} are unramified [Mor12, Corollary 6.9] we deduce that 𝐀i,j|XZar=0\mathbf{A}_{i,j}|_{X_{\operatorname{Zar}}}=0 for j0j\gg 0. Hence i+1|XZari,j|XZar\mathscr{E}_{i+1}|_{X_{\operatorname{Zar}}}\simeq\mathscr{E}_{i,j}|_{X_{\operatorname{Zar}}} for such jj. This proves the claim. ∎

Corollary 5.6.

Locally 𝔸1{\mathbb{A}}^{1}-nilpotent motivic spaces lie in the subcategory of Spc(k)\mathrm{Spc}(k)_{*} generated under (iterated) limits by S1S^{1}-infinite loop spaces. In particular, if f:𝒳𝒴Spc(k)f:\mathscr{X}\to\mathscr{Y}\in\mathrm{Spc}(k)_{*} is a morphism of locally 𝔸1{\mathbb{A}}^{1}-nilpotent motivic spaces such that ΣS1f\Sigma^{\infty}_{S^{1}}f is an equivalence, then ff is an equivalence.

Proof.

The first statement is clear by considering the construction of Proposition 5.5 for 𝒳\mathscr{X}\to*. For the second statement, by Yoneda, it suffices to show that for any locally 𝔸1{\mathbb{A}}^{1}-nilpotent motivic space 𝒵\mathscr{Z} we have Map(𝒳,𝒵)Map(𝒴,𝒵)\operatorname{Map}_{*}(\mathscr{X},\mathscr{Z})\simeq\operatorname{Map}_{*}(\mathscr{Y},\mathscr{Z}). The class of spaces 𝒵\mathscr{Z} such that this holds is closed under (iterated) limits and contains all S1S^{1}-infinite loop spaces. Hence we conclude by the first statement. ∎

6 Applications

In this section, we describe some applications of the Whitehead theorem building on a comment made in the introduction: in checking that a morphism between spaces is an S1S^{1}-stable equivalence, singular spaces may intercede. In order to make sense of this principle, we begin by recalling how to define S1S^{1}-stable homotopy types of singular (ind-schemes) in characteristic 0. Then, we analyze the main example: the Schubert filtration on the affine Grassmannian for SL2SL_{2}. Using this filtration we are able to deduce Theorem 2, which provides an exotic equivalence in unstable motivic homotopy theory.

In what follows, if 𝒳\mathscr{X} is a pointed space, then we write Ωp,q𝒳\Omega^{p,q}\mathscr{X} for the internal (pointed) mapping space Map¯(Sp,q,𝒳)\underline{\operatorname{Map}}_{*}(S^{p,q},\mathscr{X}). Before getting to the more complicated case mentioned above, we describe a simple motivating example.

Example 6.1.

Consider the space B𝔾m{\mathbb{P}}^{\infty}\cong B{{\mathbb{G}}_{m}} [MV99, §4 Proposition 3.8]. In fact, =colimnn{\mathbb{P}}^{\infty}=\operatorname{colim}_{n}{\mathbb{P}}^{n} and in particular there is an inclusion 1{\mathbb{P}}^{1}\hookrightarrow{\mathbb{P}}^{\infty} of the “bottom cell”. Under the equivalence 1S2,1{\mathbb{P}}^{1}\cong S^{2,1}, this map is adjoint to a map

S1Ω1,1.S^{1}\longrightarrow\Omega^{1,1}{\mathbb{P}}^{\infty}.

The space {\mathbb{P}}^{\infty} has a commutative hh-space structure that is inherited by Ω1,1\Omega^{1,1}{\mathbb{P}}^{\infty} and the displayed map is an hh-map for this structure. In fact, the above map is an equivalence by explicit computation. Indeed, Ω1,1\Omega^{1,1}{\mathbb{P}}^{\infty} is connected, 11-truncated, and an explicit computation shows that 𝝅1(Ω1,1)(𝐊1M)1\bm{\pi}_{1}(\Omega^{1,1}{\mathbb{P}}^{\infty})\cong({{\mathbf{K}}}^{M}_{1})_{-1}\cong{\mathbb{Z}} and the displayed map induces an isomorphism on 𝝅1\bm{\pi}_{1} and is thus an equivalence.

S1S^{1}-stable homotopy types of singular varieties

Write Schkft\mathrm{Sch}^{ft}_{k} for the category of finite type kk-schemes and consider the inclusion functor SmkSchkft\mathrm{Sm}_{k}\hookrightarrow\mathrm{Sch}^{ft}_{k} (which is fully-faithful). There is an induced restriction functor R:P(Schkft)P(Smk)R:\mathrm{P}(\mathrm{Sch}^{ft}_{k})\to\mathrm{P}(\mathrm{Sm}_{k}), where P(Schkft)\mathrm{P}(\mathrm{Sch}^{ft}_{k}) is the \infty-category of presheaves of spaces on Schkft\mathrm{Sch}^{ft}_{k}. We would like to show 𝒳P(Schkft)\mathscr{X}\in\mathrm{P}(\mathrm{Sch}^{ft}_{k}) has a well-defined homotopy type. To make this precise, recall that one may equip Schkft\mathrm{Sch}^{ft}_{k} with the structure of a site as follows. Consider the cd-structure [AHW17, Definition 2.1.1] on Schkft\mathrm{Sch}^{ft}_{k} generated by Nisnevich distinguished squares and squares arising from diagrams of the form ZiXpYZ\stackrel{{\scriptstyle i}}{{\to}}X\stackrel{{\scriptstyle p}}{{\leftarrow}}Y where ii is a closed immersion, pp is an isomorphism on the complement of ii. The cd-structure just described defines a topology on Schkft\mathrm{Sch}^{ft}_{k} called the cdh-topology.

Recall that 𝒳P(Schkft)\mathscr{X}\in\mathrm{P}(\mathrm{Sch}^{ft}_{k}) is cdh-excisive if 𝒳()\mathscr{X}(\emptyset) is contractible and if 𝒳\mathscr{X} takes squares in the cd-structure described above to cartesian squares. We define Hft(k)\mathrm{H}^{ft}(k) to be the subcategory of P(Schkft)\mathrm{P}(\mathrm{Sch}^{ft}_{k}) spanned by presheaves of spaces that are cdh-excisive and 𝔸1{\mathbb{A}}^{1}-invariant (see the beginning of Section 2). As before there is a motivic localization functor

Lmotcdh:P(Schkft)Hft(k);\mathrm{L}_{mot}^{cdh}:\mathrm{P}(\mathrm{Sch}^{ft}_{k})\longrightarrow\mathrm{H}^{ft}(k);

that preserves finite products.

The restriction functor sends cdhcdh-excisive and 𝔸1{\mathbb{A}}^{1}-invariant sheaves to Nisnevich and 𝔸1{\mathbb{A}}^{1}-invariant sheaves on Smk\mathrm{Sm}_{k} and one thus has a diagram of functors, which need not commute:

P(Schkft)\textstyle{\mathrm{P}(\mathrm{Sch}^{ft}_{k})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}R\scriptstyle{R}Lmotcdh\scriptstyle{\mathrm{L}_{mot}^{cdh}}P(Smk)\textstyle{\mathrm{P}(\mathrm{Sm}_{k})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Lmot\scriptstyle{\mathrm{L}_{mot}}Hft(k)\textstyle{\mathrm{H}^{ft}(k)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Spc(k).\textstyle{\mathrm{Spc}({k}).}

One can stabilize both sides with respect to the simplicial circle. We write SHftS1(k){\mathrm{SH}}^{S^{1}}_{ft}(k) for the category obtained by inverting S1S^{1} on Hft(k)\mathrm{H}^{ft}(k)_{*} (as described in [Rob15, §4.1]). Abusing terminology slightly, we write Lmotcdh𝒳+\mathrm{L}_{mot}^{cdh}\mathscr{X}_{+} for the functor P(Schkft)SHftS1(k)\mathrm{P}(\mathrm{Sch}^{ft}_{k})\to{\mathrm{SH}}^{S^{1}}_{ft}(k) and Lmot𝒳+\mathrm{L}_{mot}\mathscr{X}_{+} for the functor P(Smk)SHftS1(k)\mathrm{P}(\mathrm{Sm}_{k})\to{\mathrm{SH}}^{S^{1}}_{ft}(k).

This functor RR has a left adjoint E:P(Smk)P(Schkft)E:\mathrm{P}(\mathrm{Sm}_{k})\to\mathrm{P}(\mathrm{Sch}^{ft}_{k}). The above inclusion functor corresponds to a morphism of sites

π:(Schkft)cdh(Smk)Nis.\pi:(\mathrm{Sch}^{ft}_{k})_{cdh}\longrightarrow(\mathrm{Sm}_{k})_{{\operatorname{Nis}}}.

Voevodsky established that the functor π:SHcdhS1(k)SHS1(k)\pi^{*}:{\mathrm{SH}}^{S^{1}}_{cdh}(k)\to{\mathrm{SH}}^{S^{1}}(k) is an equivalence [Voe10, Theorem 4.2]. The stable motivic homotopy type of 𝒳P(Schkft)\mathscr{X}\in\mathrm{P}(\mathrm{Sch}^{ft}_{k}) is then defined to be πLmotcdh𝒳+\pi^{*}\mathrm{L}_{mot}^{cdh}\mathscr{X}_{+}. The following result establishes that singular schemes have well-defined stable motivic homotopy types (cf. [Bac19, Proposition 17]).

Proposition 6.2.

Assume kk has characteristic 0.

  1. 1.

    For any 𝒳P(Schkft)\mathscr{X}\in\mathrm{P}(\mathrm{Sch}^{ft}_{k}), there is an equivalence

    LmotR𝒳+πLmotcdh𝒳+.\mathrm{L}_{mot}R\mathscr{X}_{+}\cong\pi^{*}\mathrm{L}_{mot}^{cdh}\mathscr{X}_{+}.
  2. 2.

    If X=colimnXnX=\operatorname{colim}_{n}X_{n} is a filtered colimit of finite type schemes, then LmotRX+colimnLmotcdh(Xn)+\mathrm{L}_{mot}RX_{+}\cong\operatorname{colim}_{n}\mathrm{L}_{mot}^{cdh}(X_{n})_{+}.

Proof.

Since the functor SmkSchkft\mathrm{Sm}_{k}\hookrightarrow\mathrm{Sch}^{ft}_{k} is fully-faithful, the composite P(Smk)EP(Schkft)RP(Smk)\mathrm{P}(\mathrm{Sm}_{k})\stackrel{{\scriptstyle E}}{{\to}}\mathrm{P}(\mathrm{Sch}^{ft}_{k})\stackrel{{\scriptstyle R}}{{\to}}\mathrm{P}(\mathrm{Sm}_{k}) is the identity functor on P(Smk)\mathrm{P}(\mathrm{Sm}_{k}). It follows that RER𝒳+R𝒳+RER\mathscr{X}_{+}\cong R\mathscr{X}_{+}, i.e., if η:ERid\eta:ER\to id is the counit of adjunction, then R(η)R(\eta) is an equivalence. On the other hand, since kk has characteristic 0, resolution of singularities implies that all schemes are cdhcdh-locally smooth [Voe10, Lemma 4.3]. Thus, the assertion that R(η)R(\eta) is an equivalence is sufficient to guarantee that η\eta becomes an equivalence after applying Lmotcdh()+\mathrm{L}_{mot}^{cdh}(-)_{+}, which is what we wanted to show.

The second statement follows from the first since all functors under consideration commute with filtered colimits. ∎

Remark 6.3.

Assume that kk is a field with characteristic pp and assume \ell is a prime different from pp. If strong resolution of singularities was known in positive characteristic, then Proposition 6.2 would extend to that situation as well. Unfortunately, we are not aware of a version of Proposition 6.2 that holds, even assuming we invert the characteristic exponent of pp, say working with the dh\ell dh-topology of [Kel17, Definition 2.1.11], even though finite-type singular kk-schemes are dh\ell dh-locally smooth [Kel17, Corollary 2.1.15], this is not sufficient to guarantee dh\ell dh-descent for suspension spectra.

Affine Grassmannian models of Tate loop spaces

Suppose GG is a split reductive group scheme (which we will always assume to be SL2SL_{2} below). One model of the affine grassmannian GrG\operatorname{Gr}_{G} is an ind-scheme representing the fppf sheafification of the functor on Schk\mathrm{Sch}_{k} defined by

UG(Γ(U,𝒪U)((t)))/G(Γ(U,𝒪U)[[t]]);U\mapsto G(\Gamma(U,\mathcal{O}_{U})((t)))/G(\Gamma(U,\mathcal{O}_{U})[[t]]);

see [BL94, §2] for more details in the case of interest for us (i.e., G=SL2G=SL_{2}). Quillen [Mit88] and Garland–Raghunathan [GR75] observed that GrG()\operatorname{Gr}_{G}({\mathbb{C}}) had the structure of an associative hh-space: if KK is the maximal compact subgroup of G()G({\mathbb{C}}), then there is a homotopy equivalence ΩKGrG\Omega K\cong\operatorname{Gr}_{G}. The following result can be viewed as a lift of these results to the motivic homotopy category and equips GrG\operatorname{Gr}_{G} with an associative hh-space structure in Spc(k)\mathrm{Spc}({k}).

Theorem 6.4 ([Bac19, Theorem 15]).

If kk is a field, and GG is a split reductive group scheme, then there is an equivalence

Ω1,1GGrG.\Omega^{1,1}G\cong\operatorname{Gr}_{G}.

As mentioned above, the affine Grassmannian of GrG\operatorname{Gr}_{G} has the structure of an ind-scheme: the Schubert cell decomposition explicitly allows us to see GrG\operatorname{Gr}_{G} as a filtered colimit of projective kk-schemes that are frequently singular. The Schubert cell description in the case of SL2SL_{2} is particularly simple and was elucidated and made extremely explicit in the work of Mitchell [Mit86], who also analyzed the interaction of this filtration (which he called the Bott filtration) with the multiplicative structure induced by the weak equivalence GrG()ΩK\operatorname{Gr}_{G}({\mathbb{C}})\cong\Omega K mentioned above.

Theorem 6.5 (Mitchell).

The ind-scheme GrSL2\operatorname{Gr}_{SL_{2}} admits a Schubert filtration, i.e., there exists a directed sequence of projective schemes F2,rF_{2,r} such that GrSL2=colimrF2,r\operatorname{Gr}_{SL_{2}}=\operatorname{colim}_{r}F_{2,r} and such that the schemes F2,rF_{2,r} have the following properties.

  1. 1.

    The scheme F2,1F_{2,1} is isomorphic to 1{\mathbb{P}}^{1}.

  2. 2.

    The complement F2,rF2,r1F_{2,r}\smallsetminus F_{2,r-1} is isomorphic to an affine space of dimension rr and F2,r/F2,r1F_{2,r}/F_{2,r-1} is equivalent to S2r,rS^{2r,r}.

  3. 3.

    The multiplication for the hh-space structure induces a morphism F2,1×rF2,rF_{2,1}^{\times r}\to F_{2,r} that is a resolution of singularities.

  4. 4.

    The inclusion F2,1×r1F2,1×rF2,rS2r,rF_{2,1}^{\times r-1}\hookrightarrow F_{2,1}^{\times r}\to F_{2,r}\to S^{2r,r} (as the first r1r-1-factors) is the constant map to the base-point.

Proof.

Mitchell states these results only in the case k=k={\mathbb{C}}, but the proofs work over any characteristic 0 field. For a construction of the relevant varieties as schemes see [BL94, Proposition 2.6]. With those observations in mind, these results are a summary of [Mit86, 2.7-8, 2.11-12] in the case n=2n=2. ∎

Remark 6.6.

The map in Theorem 6.5(3) is an affine Bott–Samelson–Demazure resolution (see [BS58, p. 920] and [Dem74, §3] for a scheme-theoretic treatment). The maximal compact subgroup of SL2()SL_{2}({\mathbb{C}}) is SU(2)S3SU(2)\cong S^{3}. Under the identification ΩS3=ΩΣS2J(S2)\Omega S^{3}=\Omega\Sigma S^{2}\cong J(S^{2}), the space ΩΣS2\Omega\Sigma S^{2} carries an increasing James filtration. Under the weak equivalence ΩSU(2)GrSL2()\Omega SU(2)\cong\operatorname{Gr}_{SL_{2}}({\mathbb{C}}), the James filtration is carried to the Schubert filtration. Moreover, the inclusion F2,1GrSL2F_{2,1}\hookrightarrow\operatorname{Gr}_{SL_{2}} realizes Bott’s generating complex [Bot58, §5].

An unstable exotic periodicity

After Theorem 6.4 and Theorem 6.5(1) there is a morphism of (ind-)schemes

1GrSL2,{\mathbb{P}}^{1}\longrightarrow\operatorname{Gr}_{SL_{2}},

corresponding to the inclusion F2,1GrSL2F_{2,1}\hookrightarrow\operatorname{Gr}_{SL_{2}}. The right hand side has the structure has the structure of an hh-group by means of Theorem 6.4, in particular, it is an associative monoid.

Given any connected space 𝒳\mathscr{X}, the James construction J(𝒳)J(\mathscr{X}) [AWW17, Theorem 2.4.2] comes equipped with a morphism 𝒳J(𝒳)\mathscr{X}\to J(\mathscr{X}) that is universal among maps from 𝒳\mathscr{X} to associative monoids: if \mathscr{M} is an associative monoid in Spc(k)\mathrm{Spc}({k}), then given a map f:𝒳f:\mathscr{X}\to\mathscr{M}, ff factors uniquely through the map 𝒳J(𝒳)\mathscr{X}\to J(\mathscr{X}). The universal property of the James construction applied to the map 1GrSL2{\mathbb{P}}^{1}\to\operatorname{Gr}_{SL_{2}} of the previous point factors through a morphism:

ψ:J(1)GrSL2.\psi:J({\mathbb{P}}^{1})\longrightarrow\operatorname{Gr}_{SL_{2}}.

Moreover, the left hand side has an increasing filtration by spaces of the form Jn(1)J_{n}({\mathbb{P}}^{1}) with successive subquotients equivalent to S2n,nS^{2n,n}; we refer to this filtration as the James filtration. After a single simplicial suspension, the left hand side splits as n0S2n+1,n\bigvee_{n\geq 0}S^{2n+1,n} by [WW19, Proposition 5.6]. The next result, which implies Theorem 2 essentially follows from assertion of Theorem 6.5 that the James filtration coincides with the Schubert filtration; in the motivic context, we view the “weight shifting” displayed by ψ\psi as an exotic periodicity.

Theorem 6.7.

Assume kk is a field having characteristic 0.

  1. 1.

    The morphism ΣS1ψ\Sigma^{\infty}_{S^{1}}\psi is an equivalence.

  2. 2.

    The morphism ψ:Ω1,0Σ1,0S2,1Ω1,1Σ1,1S2,1\psi:\Omega^{1,0}\Sigma^{1,0}S^{2,1}\to\Omega^{1,1}\Sigma^{1,1}S^{2,1} is an equivalence.

  3. 3.

    The morphism S3,1Ω1,1HPS^{3,1}\to\Omega^{1,1}\operatorname{HP}^{\infty} of the introduction is a weak equivalence.

Proof.

Consider the Schubert filtration on GrSL2\operatorname{Gr}_{SL_{2}} from Theorem 6.5. By definition of the James filtration on J(1)J({\mathbb{P}}^{1}), the layers Jr(1)J_{r}({\mathbb{P}}^{1}) are precisely the images of 1×r{{\mathbb{P}}^{1}}^{\times r} under the monoid structure. Thus, Theorem 6.5(3) implies that the product map factors through a map

Jr(1)F2,r.J_{r}({\mathbb{P}}^{1})\longrightarrow F_{2,r}.

By definition, the morphism in question is an equivalence when r=1r=1. Theorem 6.5(4) implies that the induced maps Jr(1)/Jr1(1)F2,r/F2,r1J_{r}({\mathbb{P}}^{1})/J_{r-1}({\mathbb{P}}^{1})\to F_{2,r}/F_{2,r-1} are equivalences. An induction argument using Proposition 6.2(1) implies that the maps Jr(1)F2,rJ_{r}({\mathbb{P}}^{1})\to F_{2,r} are S1S^{1}-stable equivalences. In that case, Proposition 6.2(2) again in conjunction with Theorem 6.5 implies that the map J(1)GrSL2J({\mathbb{P}}^{1})\to\operatorname{Gr}_{SL_{2}} is an S1S^{1}-stable equivalence. The second point follows from the first by appeal to Theorem 5.2 since both spaces are connected hh-spaces (in particular, nilpotent).

The map ψ\psi is an hh-map by construction and therefore B(ψ)B(\psi) is an equivalence. In that case, we have

S3,1BΩS3,1B(J(1))B(GrSL2)B(Ω1,1SL2)Ω1,1BSL2.S^{3,1}\cong B\Omega S^{3,1}\cong B(J({\mathbb{P}}^{1}))\cong B(\operatorname{Gr}_{SL_{2}})\cong B(\Omega^{1,1}SL_{2})\cong\Omega^{1,1}BSL_{2}.

The inclusion of the bottom cell HP1HP\operatorname{HP}^{1}\to\operatorname{HP}^{\infty} is classified by the Hopf bundle ν\nu, which is a GW(k)GW(k)-module generator of 𝝅4,2(BSL2)GW(k)\bm{\pi}_{4,2}(BSL_{2})\cong GW(k). Up to multiplication by a unit, the adjoint of this map coincides with the map inducing the equivalence in the previous point. Replacing the equivalence of the previous point by its product with inverse of the unit in GW(k)GW(k) just described if necessary, the result from the introduction follows. ∎

An exceptional EHP sequence

The combinatorial James–Hopf invariant (in this context, [AWW17, Definition 3.1.3]) defines a morphism:

Ω1,0Σ1,0S2,1Ω1,0Σ1,0S4,2.\Omega^{1,0}\Sigma^{1,0}S^{2,1}\longrightarrow\Omega^{1,0}\Sigma^{1,0}S^{4,2}.

We can identify the 22-local fiber of this map in a fashion that extends the EHP fiber sequence of Wickelgren–Williams [WW19]. The following result uses the full strength of Theorem 5.2.

Proposition 6.8.

Over any field kk, there is 22-local homotopy fiber sequence of motivic spaces of the form

S2,1Ω1,0S3,1Ω1,0S5,2.S^{2,1}\longrightarrow\Omega^{1,0}S^{3,1}\longrightarrow\Omega^{1,0}S^{5,2}.
Proof.

The James–Hopf invariant described above is a map of pointed connected and simple spaces so its fiber is nilpotent by appeal to [AFH22, Theorem 3.3.6]. The composite map S2,1Ω1,0S5,2S^{2,1}\to\Omega^{1,0}S^{5,2} is null because the target is simply connected. A choice of null-homotopy then determines a map f:1fib(Ω1,0S3,1Ω1,0S5,2)f:{\mathbb{P}}^{1}\to\mathrm{fib}(\Omega^{1,0}S^{3,1}\to\Omega^{1,0}S^{5,2}). The space S2,1S^{2,1} is a nilpotent motivic space by [AFH22, Theorem 3.4.8] (take G=SL2G=SL_{2} in that statement). Now, let us localize all spaces and maps at 22; via [AFH22, Theorem 4.3.9] there is a well-behaved 22-localization for nilpotent motivic spaces.

To check that ff is an equivalence after 22-localization, by the Whitehead theorem for nilpotent motivic spaces 5.2, it suffices to check this is the case after S1S^{1}-suspension. In that case, the proof of [WW19, Theorem 8.3], in conjunction with [WW19, Proposition 4.7] shows that Σf\Sigma^{\infty}f is a 22-local equivalence, as claimed. ∎

Remark 6.9.

The connected motivic spheres Sq+1,qS^{q+1,q} are only known to be hh-spaces when q=0q=0 and q=2q=2, so the above argument cannot be generalized much further.

Consider the map S4,2K((2),4)S^{4,2}\to K({\mathbb{Z}}(2),4) classifying a generator. This map is adjoint to maps S3,1K((1),3)S^{3,1}\to K({\mathbb{Z}}(1),3), S3,2K((2),3)S^{3,2}\to K({\mathbb{Z}}(2),3) and S2,1K((1),2)S^{2,1}\to K({\mathbb{Z}}(1),2) and moreover these adjoint maps are also the relevant fundamental class maps. We define

S3,i3:=fib(S3,iK((i),3)).S^{3,i}\langle 3\rangle:=\mathrm{fib}(S^{3,i}\to K({\mathbb{Z}}(i),3)).

We obtain analogs of Cohen’s exceptional 22-local fiber sequence [Coh87, Theorem 3.3].

Corollary 6.10.

Assume kk is a field. There is a 22-local fiber sequence of the form

S3,2Ω1,0S3,13Ω1,0S5,2S^{3,2}\longrightarrow\Omega^{1,0}S^{3,1}\langle 3\rangle\longrightarrow\Omega^{1,0}S^{5,2}

and, if kk has characteristic 0, then there is a fiber sequence of the form

S3,2Ω1,1S3,23Ω1,0S5,2.S^{3,2}\longrightarrow\Omega^{1,1}S^{3,2}\langle 3\rangle\longrightarrow\Omega^{1,0}S^{5,2}.

In both cases, the left hand map is induced by a suitable generator of 𝛑3,2(Ω1,iS3,ii)\bm{\pi}_{3,2}(\Omega^{1,i}S^{3,i}\langle i\rangle).

Proof.

There is a commutative square of the form:

Ω1,0S3,1\textstyle{\Omega^{1,0}S^{3,1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Ω1,0S5,3\textstyle{\Omega^{1,0}S^{5,3}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}K((1),2)\textstyle{K({\mathbb{Z}}(1),2)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\textstyle{\ast}

where the left vertical map is obtained by looping the fundamental class map S3,1K((1),3)S^{3,1}\to K({\mathbb{Z}}(1),3). Proposition 6.8 describes the (22-local) fiber of the top row, and the unwinding the definition of the map Ω1,0S3,1K((1),2)\Omega^{1,0}S^{3,1}\to K({\mathbb{Z}}(1),2) we conclude that the induced map of fibers is the fundamental class map S2,1K((1),2)S^{2,1}\to K({\mathbb{Z}}(1),2). By definition of η\eta, one identifies fib(S2,1K((1),2))S3,2\mathrm{fib}(S^{2,1}\to K({\mathbb{Z}}(1),2))\cong S^{3,2} and the induced map as η\eta. Taking vertical fibers yields the first statement.

For the second statement, assuming kk has characteristic 0, consider the exotic equivalence Ω1,0Σ1,0S2,1Ω1,1Σ1,1S2,1\Omega^{1,0}\Sigma^{1,0}S^{2,1}\to\Omega^{1,1}\Sigma^{1,1}S^{2,1} of Theorem 6.7. Under that equivalence, there is an induced map Ω1,1S3,2K((1),2)\Omega^{1,1}S^{3,2}\to K({\mathbb{Z}}(1),2) that coincides with the map obtained by applying Ω1,1\Omega^{1,1} to the fundamental class S3,2K((2),3)S^{3,2}\to K({\mathbb{Z}}(2),3). Repeating the argument of the preceding paragraph with this new map yields the second fiber sequence. ∎

References

  • [ABH23] A. Asok, T. Bachmann, and M.J. Hopkins. On 1{\mathbb{P}}^{1}-stabilization in unstable motivic homotopy theory. 2023. Available at https://arxiv.org/abs/2306.04631.
  • [AF16] A. Asok and J. Fasel. Comparing Euler classes. Q. J. Math., 67(4):603–635, 2016.
  • [AFH22] A. Asok, J. Fasel, and M. J. Hopkins. Localization and nilpotent spaces in 𝔸1\mathbb{A}^{1}-homotopy theory. Compos. Math., 158(3):654–720, 2022.
  • [AHW17] A. Asok, M. Hoyois, and M. Wendt. Affine representability results in 𝔸1\mathbb{A}^{1}-homotopy theory, I: vector bundles. Duke Math. J., 166(10):1923–1953, 2017.
  • [AM11] A. Asok and F. Morel. Smooth varieties up to 𝔸1\mathbb{A}^{1}-homotopy and algebraic hh-cobordisms. Adv. Math., 227(5):1990–2058, 2011.
  • [AWW17] A. Asok, K. Wickelgren, and T.B. Williams. The simplicial suspension sequence in 𝔸1\mathbb{A}^{1}-homotopy. Geom. Topol., 21(4):2093–2160, 2017.
  • [Bac19] T. Bachmann. Affine Grassmannians in 𝔸1\mathbb{A}^{1}-homotopy theory. Selecta Math. (N.S.), 25(2):Paper No. 25, 14, 2019.
  • [BHS15] C. Balwe, A. Hogadi, and A. Sawant. 𝔸1\mathbb{A}^{1}-connected components of schemes. Adv. Math., 282:335–361, 2015.
  • [BL94] A. Beauville and Y. Laszlo. Conformal blocks and generalized theta functions. Comm. Math. Phys., 164(2):385–419, 1994.
  • [Bot58] R. Bott. The space of loops on a Lie group. Michigan Math. J., 5:35–61, 1958.
  • [Bou75] A. K. Bousfield. The localization of spaces with respect to homology. Topology, 14:133–150, 1975.
  • [BS58] R. Bott and H. Samelson. Applications of the theory of Morse to symmetric spaces. Amer. J. Math., 80:964–1029, 1958.
  • [CH22] U. Choudhury and A. Hogadi. The Hurewicz map in motivic homotopy theory. Annals of K-Theory, 7(1):179–190, jun 2022.
  • [CMN79] F. R. Cohen, J. C. Moore, and J. A. Neisendorfer. The double suspension and exponents of the homotopy groups of spheres. Ann. of Math. (2), 110(3):549–565, 1979.
  • [Coh87] F. R. Cohen. A course in some aspects of classical homotopy theory. In Algebraic topology (Seattle, Wash., 1985), volume 1286 of Lecture Notes in Math., pages 1–92. Springer, Berlin, 1987.
  • [Dem74] M. Demazure. Désingularisation des variétés de Schubert généralisées. Ann. Sci. École Norm. Sup. (4), 7:53–88, 1974.
  • [DF71] E. Dror-Farjoun. A generalization of the Whitehead theorem. In Symposium on Algebraic Topology (Battelle Seattle Res. Center, Seattle, Wash., 1971), Lecture Notes in Math., Vol. 249, pages 13–22. Springer, Berlin, 1971.
  • [DH21] S. Devalapurkar and P. Haine. On the James and Hilton-Milnor Splittings, and the metastable EHP sequence. Doc. Math., 26:1423–1464, 2021.
  • [Gan65] T. Ganea. A generalization of the homology and homotopy suspension. Comment. Math. Helv., 39:295–322, 1965.
  • [Ger75] S. M. Gersten. The Whitehead theorem for nilpotent spaces. Proc. Amer. Math. Soc., 47:259–260, 1975.
  • [GJ09] P. G. Goerss and J. F. Jardine. Simplicial homotopy theory. Modern Birkhäuser Classics. Birkhäuser Verlag, Basel, 2009.
  • [GR75] H. Garland and M. S. Raghunathan. A Bruhat decomposition for the loop space of a compact group: a new approach to results of Bott. Proc. Nat. Acad. Sci. U.S.A., 72(12):4716–4717, 1975.
  • [Hop42] H. Hopf. Fundamentalgruppe und zweite Bettische Gruppe. Comment. Math. Helv., 14:257–309, 1942.
  • [Hov99] M. Hovey. Model categories, volume 63 of Mathematical Surveys and Monographs. American Mathematical Society, Providence, RI, 1999.
  • [Hoy14] M. Hoyois. A quadratic refinement of the Grothendieck-Lefschetz-Verdier trace formula. Algebr. Geom. Topol., 14(6):3603–3658, 2014.
  • [Hoy17] M. Hoyois. The six operations in equivariant motivic homotopy theory. Adv. Math., 305:197–279, 2017.
  • [HW20] J. Hahn and D. Wilson. Eilenberg–Mac Lane spectra as equivariant Thom spectra. Geom. Topol., 24(6):2709–2748, 2020.
  • [Kel17] S. Kelly. Voevodsky motives and lldh descent. Astérisque, 391, 2017.
  • [Koi22] J. Koizumi. Zeroth 𝔸1\mathbb{A}^{1}-homology of smooth proper varieties. New York J. Math., 28:824–834, 2022.
  • [Lur09] J. Lurie. Higher topos theory, volume 170 of Annals of Mathematics Studies. Princeton University Press, Princeton, NJ, 2009.
  • [Lur17] J. Lurie. Higher Algebra. 2017. Available at https://www.math.ias.edu/~lurie/papers/HA.pdf.
  • [Mat76] M. Mather. Pull-backs in homotopy theory. Canadian J. Math., 28(2):225–263, 1976.
  • [Mit86] S. A. Mitchell. A filtration of the loops on SU(n){\rm SU}(n) by Schubert varieties. Math. Z., 193(3):347–362, 1986.
  • [Mit88] S. A. Mitchell. Quillen’s theorem on buildings and the loops on a symmetric space. Enseign. Math. (2), 34(1-2):123–166, 1988.
  • [Mor05] F. Morel. The stable 𝔸1{\mathbb{A}}^{1}-connectivity theorems. KK-Theory, 35(1-2):1–68, 2005.
  • [Mor12] F. Morel. 𝔸1\mathbb{A}^{1}-algebraic topology over a field, volume 2052 of Lecture Notes in Mathematics. Springer, Heidelberg, 2012.
  • [MV99] F. Morel and V. Voevodsky. 𝔸1\mathbb{A}^{1}-homotopy theory of schemes. Publications Mathématiques de l’Institut des Hautes Études Scientifiques, 90(1):45–143, 1999.
  • [Pup74] V. Puppe. A remark on “homotopy fibrations”. Manuscripta Math., 12:113–120, 1974.
  • [PW21] I. Panin and C. Walter. Quaternionic Grassmannians and Borel classes in algebraic geometry. Algebra i Analiz, 33(1):136–193, 2021.
  • [Qui67] D. G. Quillen. Homotopical algebra. Lecture Notes in Mathematics, No. 43. Springer-Verlag, Berlin-New York, 1967.
  • [Rez98] C. Rezk. Fibrations and homotopy colimits of simplicial sheaves, 1998. Available at https://arxiv.org/abs/math/9811038.
  • [Rob15] M. Robalo. KK-theory and the bridge from motives to noncommutative motives. Adv. Math., 269:399–550, 2015.
  • [Sta65] J. Stallings. Homology and central series of groups. J. Algebra, 2:170–181, 1965.
  • [Str12] F. Strunk. On motivic spherical bundles. 2012. Thesis Institut für Mathematik Universität Osnabrück, available at https://repositorium.uni-osnabrueck.de/bitstream/urn:nbn:de:gbv:700-2013052710851/3/thesis_strunk.pdf.
  • [Too76] G. H. Toomer. Homology equivalences and a technique of Ganea. Math. Z., 150(3):273–279, 1976.
  • [Voe10] V. Voevodsky. Unstable motivic homotopy categories in Nisnevich and cdh-topologies. J. Pure Appl. Algebra, 214(8):1399–1406, 2010.
  • [Wen11] M. Wendt. Classifying spaces and fibrations of simplicial sheaves. J. Homotopy Relat. Struct., 6(1):1–38, 2011.
  • [WW19] K. Wickelgren and T.B. Williams. The simplicial EHP sequence in 𝔸1\mathbb{A}^{1}-algebraic topology. Geom. Topol., 23(4):1691–1777, 2019.

A. Asok, Department of Mathematics, University of Southern California, 3620 S. Vermont Ave., Los Angeles, CA 90089-2532, United States; E-mail address: asok@usc.edu

T. Bachmann, Department of Mathematics, JGU Mainz, Staudingerweg 9, 55128 Mainz, Germany; E-mail address: tom.bachmann@zoho.com

M.J. Hopkins, Department of Mathematics, Harvard University, One Oxford Street, Cambridge, MA 02138, United States E-mail address: mjh@math.harvard.edu