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

On the higher topological complexity of manifolds with abelian fundamental group

N. Cadavid-Aguilar Departamento de Matemáticas, Centro de Investigación y de Estudios Avanzados del IPN, Av. Instituto Politécnico Nacional 2508, San Pedro Zacatenco, Ciudad de México 07000 ncadavia@gmail.com D. Cohen Department of Mathematics, Louisiana State University, Baton Rouge, Louisiana 70803 cohen@math.lsu.edu www.math.lsu.edu/˜cohen J. González Departamento de Matemáticas, Centro de Investigación y de Estudios Avanzados del IPN, Av. Instituto Politécnico Nacional 2508, San Pedro Zacatenco, Ciudad de México 07000 jesus.glz-espino@cinvestav.mx S. Hughes Rheinische Friedrich-Wilhelms-Universität Bonn, Mathematical Institute, Endenicher Allee 60, 53115 Bonn, Germany sam.hughes.maths@gmail.com; hughes@math.uni-bonn.de samhughesmaths.github.io  and  L. Vandembroucq Centro de Matemática, Universidade do Minho, Braga, Portugal lucile@math.uminho.pt
Abstract.

We study the higher (or sequential) topological complexity TCs\operatorname{TC}_{s} of manifolds with abelian fundamental group. We give sufficient conditions for TCs\operatorname{TC}_{s} to be non-maximal in both the orientable and non-orientable cases. In combination with cohomological lower bounds, we also obtain some exact values for certain families of manifolds.

Introduction

For a path-connected space XX, the ss-th higher topological complexity TCs(X)\operatorname{TC}_{s}(X) is the sectional category of the fibration es:PXXse_{s}\colon PX\to X^{s}, that is

TCs(X)=secat(es:PXXs),\operatorname{TC}_{s}(X)=\operatorname{secat}(e_{s}\colon PX\to X^{s}),

where PXPX denotes the space of paths in XX and

es(γ)=(γ(0),γ(1s1),,γ(s2s1),γ(1))e_{s}(\gamma)=\Bigl{(}\gamma(0),\gamma\bigl{(}\frac{1}{s-1}\bigr{)},\ldots,\gamma\bigl{(}\frac{s-2}{s-1}\bigr{)},\gamma(1)\Bigr{)}

is the usual ss-th evaluation map. That is, in the reduced version used here, TCs(X)\operatorname{TC}_{s}(X) is one less than the minimal number of open sets covering XsX^{s}, over each of which the fibration ese_{s} admits a section.

Topological complexity TC(X)=TC2(X)\operatorname{TC}(X)=\operatorname{TC}_{2}(X) was introduced by Farber in [13] and the ‘higher’ invariants were introduced by Rudyak in [18]. The invariants were developed and motivated by applications for motion planning problems in robotics. More precisely, viewing XX as the space of configurations of a mechanical system, the integer TCs(X)\operatorname{TC}_{s}(X) provides a topological measure of the complexity of planning motion in XX from an initial configuration to a terminal configuration, passing through s2s-2 specified intermediate configurations.

Despite a huge body of research into these invariants, there are very few complete computations of TCs(X)\operatorname{TC}_{s}(X). Examples for which the full spectrum of invariants is known include products of spheres, surfaces, path-connected topological groups whose Lusternik-Schnirelmann category is known, closed simply-connected symplectic manifolds, classifying spaces of hyperbolic groups and some (additional) polyhedral product type spaces, see [2, 16, 17, 1]. In a number of these examples, the higher topological complexities attain the maximal values possible.

If XX is not simply connected, this maximal value is TCs(X)sdim(X)\operatorname{TC}_{s}(X)\leq s\dim(X), where dim(X)\dim(X) is the homotopy dimension of XX, see [2]. Work of Cohen–Vandembroucq [6] explored the non-maximality of TC2(M)\operatorname{TC}_{2}(M) when MM is a manifold with abelian fundamental group. In this paper we extend these ideas to TCs(M)\operatorname{TC}_{s}(M) for s2s\geq 2.

Espinosa Baro, Farber, Mescher, and Oprea [12] have recently characterized the maximality of TCs\operatorname{TC}_{s} of a finite-dimensional CW-complex XX in terms of a canonical cohomology class generalizing the ‘Costa–Farber class’ introduced in [7] (see Section 1). Restricting our attention to a manifold MM with an abelian fundamental group π\pi and following the strategy of [6], we first express this characterization in terms of a homology class of the group πs1\pi^{s-1} (see Proposition 2.3 and Corollary 2.4). This permits us to establish the non-maximality of TCs(M)\operatorname{TC}_{s}(M) in some cases. For example, when MM is orientable, we obtain the following result (see Section 3):

Theorem A.

Let MM be an orientable nn-dimensional connected closed manifold. In each of the following cases, we have TCs(M)<sn\operatorname{TC}_{s}(M)<sn:

  1. (1)

    π1(M)=r\pi_{1}(M)={\mathbb{Z}}^{r} with (s1)r<sdim(M)(s-1)r<s\dim(M);

  2. (2)

    π1(M)=q\pi_{1}(M)={\mathbb{Z}}_{q};

  3. (3)

    π1(M)=r×q\pi_{1}(M)={\mathbb{Z}}^{r}\times{\mathbb{Z}}_{q} with r<dim(M)r<\dim(M).

Computations of cohomological lower bounds of the ss-th topological complexity of the real projective spaces and lens spaces have attracted much interest [5, 9, 14, 8]. In Section 4, we show how these results provide lower bounds of TCs\operatorname{TC}_{s} for larger families of manifolds (see Proposition 4.2 and Theorem 4.7). Then Theorem A enables us to obtain the following exact values:

Theorem B.

Let MM be an orientable nn-dimensional connected closed manifold with maximal Lusternik–Schnirelmann category, that is, cat(M)=n\operatorname{\mathrm{cat}}(M)=n.

  1. (1)

    If n1n\equiv 1 mod 44 and π1(M)=2\pi_{1}(M)={\mathbb{Z}}_{2}, then TCs(M)=sn1\operatorname{TC}_{s}(M)=sn-1 for ss sufficiently large.

  2. (2)

    If n1n\equiv 1 mod 22 and π1(M)=p\pi_{1}(M)={\mathbb{Z}}_{p} where p3p\geq 3 is a prime, then TCs(M)=sn1\operatorname{TC}_{s}(M)=sn-1 for ss sufficiently large.

See Corollaries 4.4 and 4.9 for a more explicit description of the condition “ss sufficiently large”.

The case of non-orientable manifolds is much more complicated. By [6, Theorem 1.2(1)], the topological complexity of a non-orientable manifold with abelian fundamental group is always non-maximal. However, it is well-known that there exist such non-orientable manifolds with maximal TCs\operatorname{TC}_{s} for s3s\geq 3. For instance, for the real projective plane P2P^{2}, we have TC3(P2)=6=3dim(P2)\operatorname{TC}_{3}(P^{2})=6=3\dim(P^{2}), see [16]. Furthermore, it has been shown in [5] and [9] that, when nn is even, the real projective space PnP^{n} satisfies TCs(Pn)=sn\operatorname{TC}_{s}(P^{n})=sn for ss sufficiently large. For a fixed even integer nn, the sequence (TCs(Pn))s2(\operatorname{TC}_{s}(P^{n}))_{s\geq 2} forms an increasing sequence starting at TC2(Pn)\operatorname{TC}_{2}(P^{n}), equal to the immersion dimension of PnP^{n} ([15]), and stabilizing to snsn when ss is sufficiently large. As explained in [5], it would be very interesting to better understand this sequence. Our methods, developed in Section 5, permit us to obtain new information in this direction. In particular, in combination with Davis’ results [9], when n=2r2n=2^{r}-2, we have:

TCs(Pn)sn1for even sn,TCn(Pn)=n21,andTCs(Pn)=snfor s>n.\operatorname{TC}_{s}(P^{n})\leq sn-1\ \mbox{for even }s\leq n,\quad\operatorname{TC}_{n}(P^{n})=n^{2}-1,\quad\mbox{and}\quad\operatorname{TC}_{s}(P^{n})=sn\ \mbox{for }s>n.

As before, this result can be extended to a larger family of manifolds:

Theorem C.

Let MM be a non-orientable nn-dimensional connected closed manifold with π1(M)=2\pi_{1}(M)=\mathbb{Z}_{2} and n=2r2n=2^{r}-2 where r3r\geq 3. Then, for any even sns\leq n, we have TCs(M)<sn\operatorname{TC}_{s}(M)<sn. If moreover catM=n\operatorname{\mathrm{cat}}{M}=n, then TCn(M)=n21\operatorname{TC}_{n}(M)=n^{2}-1 and TCs(M)=sn\operatorname{TC}_{s}(M)=sn for s>ns>n.

Notation and conventions

For a topological space YY, we use dim(Y)\dim(Y) to denote the homotopy dimension of YY. The integral homology of YY is denoted by H(Y)H_{*}(Y), and the reduced homology by H~(Y)\widetilde{H}_{*}(Y). If π=π1(Y)\pi=\pi_{1}(Y), we denote the cohomology of YY with coefficients in the local system determined by the [π]{\mathbb{Z}}[\pi]-module VV by H(Y;V)H^{*}(Y;V).

For an element aa of a group π\pi, we often denote the inverse of aa by a¯\overline{a}.

We use the reduced version of sectional category throughout, so that for a fibration p:EBp\colon E\to B, when finite, secat(p:EB)\operatorname{secat}(p\colon E\to B) is one less than the minimal number of open sets covering BB, over each of which the fibration admits a section.

1. A TCs\operatorname{TC}_{s} canonical class

Canonical cohomology classes for higher topological complexity were recently introduced and studied by Espinosa Baro, Farber, Mescher, and Oprea, see [12]. In this brief preliminary section, with this work as a general reference ([12, §§5–6] in particular), we recall and discuss aspects of these classes which will be of subsequent use.

Let XX be a CW-complex. The standard dimensional upper bound for higher topological complexity is

(1.1) TCs(X)sdim(X).\operatorname{TC}_{s}(X)\leq s\dim(X).

Although (1.1) can be improved in terms of the connectivity of XX, we are interested in the improvements coming from obstruction-theory techniques in cases where XX is not simply connected. A fundamental concept in this context is the notion of homological obstruction as considered in Schwarz’ monograph [19]. Recall that the fiber of es:PXXse_{s}\colon PX\to X^{s} is ΩXs1=(ΩX)s1\Omega X^{s-1}=(\Omega X)^{s-1}. In [12], the homological obstruction for sectioning ese_{s} over the 1-dimensional skeleton of XsX^{s} is identified with a canonical twisted class,

(1.2) 𝔳X,sH1(Xs;Is(πs1))=H1(Xs;H~0(ΩXs1)),\mathfrak{v}_{X,s}\in H^{1}(X^{s};I_{s}(\pi^{s-1}))=H^{1}(X^{s};\widetilde{H}_{0}(\Omega X^{s-1})),

where π:=π1(X)\pi:=\pi_{1}(X) and Is(πs1)I_{s}(\pi^{s-1}) denotes the augmentation ideal of πs1\pi^{s-1}, viewed as a [πs]{\mathbb{Z}}[\pi^{s}]-submodule of [πs1]{\mathbb{Z}}[\pi^{s-1}]. Here the action of πs\pi^{s} on Is(πs1)I_{s}(\pi^{s-1}), which corresponds to the monodromy associated with the fibration ese_{s}, is given by

(a1,,as)(b1,,bs1)=(a1b1a2¯,a2b2a3¯,,as1bs1as¯).(a_{1},\ldots,a_{s})\cdot(b_{1},\ldots,b_{s-1})=(a_{1}b_{1}\overline{a_{2}},a_{2}b_{2}\overline{a_{3}},\ldots,a_{s-1}b_{s-1}\overline{a_{s}}).

The class 𝔳X,s\mathfrak{v}_{X,s} can also be described as the cohomology class induced by the crossed homomorphism νX,s:πsIs(πs1)\nu_{X,s}\colon\pi^{s}\to I_{s}(\pi^{s-1}) given by

νX,s(a1,,as)=(a1a2¯,a2a3¯,,as1as¯)1s1,\nu_{X,s}(a_{1},\ldots,a_{s})=\left(a_{1}\overline{a_{2}},a_{2}\overline{a_{3}},\ldots,a_{s-1}\overline{a_{s}}\right)-1_{s-1},

where 1s11_{s-1} is the unit element of πs1\pi^{s-1}. Obstruction-theoretic arguments lead then to the following result:

Theorem 1.1 ([12]).

Let XX be a CW-complex of dimension n2n\geq 2. Then TCs(X)<sn\operatorname{TC}_{s}(X)<sn if and only if the snsn-th cup-power 𝔳X,ssn=0\mathfrak{v}_{X,s}^{sn}=0.

Here, 𝔳X,ssn\mathfrak{v}_{X,s}^{sn} lies in the cohomology of XsX^{s} with coefficients in the snsn-th tensor power of Is(πs1)I_{s}(\pi^{s-1}) endowed with the diagonal action of πs\pi_{s}, denoted by Issn(πs1)I_{s}^{sn}(\pi^{s-1}).

The construction of the class 𝔳X,s\mathfrak{v}_{X,s} generalizes the TC\operatorname{TC} canonical class of [7] and provides for TCs\operatorname{TC}_{s} an analogue of the classical Berstein-Schwarz class 𝔟XH1(X;I(π))\mathfrak{b}_{X}\in H^{1}(X;I(\pi)). Note that, in this case, I(π)I(\pi) is the augmentation ideal of π\pi endowed with the left [π]{\mathbb{Z}}[\pi]-module structure induced by the multiplication of π\pi. As is well-known, the Lusternik–Schnirelmann category of XX, cat(X)\operatorname{\mathrm{cat}}(X), satisfies cat(X)=dimX\operatorname{\mathrm{cat}}(X)=\dim X if and only if 𝔟Xdim(X)0\mathfrak{b}_{X}^{\dim(X)}\neq 0 (see [3], [19] and [11] for a proof including the case dim(X)=2\dim(X)=2).

Remark 1.2.

We conclude this section with a brief remark regarding the functoriality of these classes. For π=π1(X)\pi=\pi_{1}(X), applying the above result to the classifying space BπB\pi yields a crossed homomorphism and associated cohomology class, which we denote by νπ,s\nu_{\pi,s} and 𝔳π,s\mathfrak{v}_{\pi,s} respectively.

Recall that if f:XYf:X\to Y is a map, and AA is a [π1(Y)]{\mathbb{Z}}[\pi_{1}(Y)]-module, then f(A)f^{*}(A) denotes the [π1(X)]{\mathbb{Z}}[\pi_{1}(X)]-module whose underlying abelian group is AA and the action of gπ1(X)g\in\pi_{1}(X) on aAa\in A is given by ga:=π1(f)(g)ag\cdot a:=\pi_{1}(f)(g)\cdot a. Taking f:XY=Bπf\colon X\to Y=B\pi to be a classifying map, the isomorphism Is(πs1)(fs)Is(πs1){I}_{s}(\pi^{s-1})\cong(f^{s})^{*}{I}_{s}(\pi^{s-1}) yields

𝔳X,s=(fs)𝔳π,s.{\mathfrak{v}}_{X,s}=(f^{s})^{*}{\mathfrak{v}}_{\pi,s}.

Similar considerations apply to the Berstein-Schwarz class 𝔟πH1(π;I(π))\mathfrak{b}_{\pi}\in H^{1}(\pi;I(\pi)) (resp., 𝔟XH1(X;I(π))\mathfrak{b}_{X}\in H^{1}(X;I(\pi))), induced by the crossed homomorphism βπ:πI(π)\beta_{\pi}\colon\pi\to I(\pi), αα1\alpha\mapsto\alpha-1. Namely, the isomorphism I(π)fI(π)I(\pi)\cong f^{*}I(\pi) yields 𝔟X=f𝔟π\mathfrak{b}_{X}=f^{*}\mathfrak{b}_{\pi}.

2. Abelian fundamental group

In this section we extend to higher topological complexity some results of [6] which will be useful for our computations. The arguments are therefore similar to those of [6] as well as some of [10].

Assume from now on that π=π1(X)\pi=\pi_{1}(X) is abelian. We consider the group homomorphism χs:πsπs1{}^{s}\chi:\pi^{s}\to\pi^{s-1} given by

χs(a1,,as)=(a1a2¯,a2a3¯,,as1as¯).{}^{s}\chi(a_{1},\dots,a_{s})=(a_{1}\overline{a_{2}},a_{2}\overline{a_{3}},\dots,a_{s-1}\overline{a_{s}}).

Note that the [πs]{\mathbb{Z}}[\pi^{s}]-module χs(I(πs1)){}^{s}\chi^{*}(I(\pi^{s-1})) is exactly the [πs]{\mathbb{Z}}[\pi^{s}]-module Is(πs1){I}_{s}(\pi^{s-1}). With the notation regarding canonical classes, Berstein-Schwarz classes, and crossed homomorphisms of the previous section, we also have, for any (a1,,as)πs(a_{1},\dots,a_{s})\in\pi^{s},

νπ,s(a1,,as)=βπs1(a1a2¯,a2a3¯,,as1as¯)=βπs1(χs(a1,,as)).\nu_{\pi,s}(a_{1},\dots,a_{s})=\beta_{\pi^{s-1}}(a_{1}\overline{a_{2}},a_{2}\overline{a_{3}},\dots,a_{s-1}\overline{a_{s}})=\beta_{\pi^{s-1}}({{}^{s}\chi}(a_{1},\dots,a_{s})).

We then have 𝔳π,s=χs𝔟πs1{\mathfrak{v}}_{\pi,s}={{}^{s}\chi}^{*}{\mathfrak{b}}_{\pi^{s-1}} in H1(πs;Is(πs1))=H1(Bπs;Is(πs1))H^{1}(\pi^{s};{I}_{s}(\pi^{s-1}))=H^{1}(B\pi^{s};{I}_{s}(\pi^{s-1})), and, for any kk,

𝔳X,sk=(γs)𝔳π,sk=(γs)(sχ)𝔟πs1k in Hk(Xs;Isk(πs1)){\mathfrak{v}}^{k}_{X,s}=(\gamma^{s})^{*}{\mathfrak{v}}^{k}_{\pi,s}=(\gamma^{s})^{*}{(^{s}\chi)}^{*}{\mathfrak{b}}^{k}_{\pi^{s-1}}\,\,\mbox{ in }H^{k}(X^{s};{I}_{s}^{k}(\pi^{s-1}))

where γ:XBπ\gamma:X\to B\pi is a classifying map.

In order to establish our results, it is useful to consider the cofiber of the diagonal map Δs=ΔsX:XXs\Delta_{s}=\Delta_{s}^{X}:X\to X^{s}. We denote it by CΔs(X)C_{\Delta_{s}}(X). We will more generally use the notation ΔsZ:ZZs\Delta_{s}^{Z}:Z\to Z^{s} to denote the ss-diagonal of a set ZZ and suppress the superscript when the context is clear.

Proposition 2.1.

Let XX be an nn-dimensional CW-complex with n2n\geq 2. Suppose that π=π1(X)\pi=\pi_{1}(X) is abelian and let γ:XBπ\gamma:X\to B\pi be a classifying map. Then for any s2s\geq 2 we have

  1. (1)

    𝔳X,s=q𝔟CΔs(X)\mathfrak{v}_{X,s}=q^{*}\mathfrak{b}_{C_{\Delta_{s}}(X)} in H1(Xs;Is(πs1))H^{1}(X^{s};I_{s}(\pi^{s-1})) where q:XsCΔs(X)q:X^{s}\to C_{\Delta_{s}}(X) is the identification map.

  2. (2)

    TCs(X)<sn\operatorname{TC}_{s}(X)<sn if and only if cat(CΔs(X))<sn\operatorname{\mathrm{cat}}(C_{\Delta_{s}}(X))<sn.

  3. (3)

    If TCs(X)<sn\operatorname{TC}_{s}(X)<sn then, for any [πs1]{\mathbb{Z}}[\pi^{s-1}]-module AA and for any homology class 𝖼Hsn(Xs;(sχγs)A)\mathsf{c}\in H_{sn}(X^{s};(^{s}\chi\gamma^{s})^{*}A), the class 𝔠=γs(𝖼)Hsn(πs;(sχ)A)\mathfrak{c}=\gamma^{s}_{*}(\mathsf{c})\in H_{sn}(\pi^{s};(^{s}\chi)^{*}A) satisfies χs(𝔠)=0{}^{s}\chi_{*}(\mathfrak{c})=0.

Proof.

First observe that the homomorphism χsΔsπ{}^{s}\chi{\circ\Delta_{s}^{\pi}} is trivial. Consequently, the map BsχBΔsπB^{s}\chi\circ B{{\Delta_{s}^{\pi}}} obtained after applying the functor BB is also trivial. By identifying BπsB\pi^{s} with (Bπ)s(B\pi)^{s} and BΔsπB{\Delta_{s}^{\pi}} with ΔsBπ\Delta_{s}^{B\pi}, we have a commutative diagram of the following form

X\textstyle{X\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ΔsX\scriptstyle{\Delta_{s}^{X}}γ\scriptstyle{\gamma}Xs\textstyle{X^{s}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}γs\scriptstyle{\gamma^{s}}q\scriptstyle{q}CΔs(X)\textstyle{C_{\Delta_{s}}(X)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ξ\scriptstyle{\xi}Bπ\textstyle{B\pi\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ΔsBπ\scriptstyle{\Delta_{s}^{B\pi}}(Bπ)s\textstyle{(B\pi)^{s}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Bsχ\scriptstyle{{B^{s}\chi}}Bπs1\textstyle{B\pi^{s-1}}

where ξ\xi is induced by the quotient property. Since π\pi is abelian, we have an exact sequence 1πΔsππsχsπs111\to\pi\stackrel{{\scriptstyle\Delta_{s}^{\pi}}}{{\longrightarrow}}{\pi^{s}}\stackrel{{{}^{s}\chi}}{{\longrightarrow}}{\pi^{s-1}}\to 1 and, using the Van Kampen theorem, we can see that π1(CΔs(X))=πs1\pi_{1}(C_{\Delta_{s}}(X))=\pi^{s-1} and that π1(ξ)\pi_{1}(\xi) is an isomorphism. Consequently ξ\xi is a classifying map and the Berstein-Schwarz class of CΔs(X)C_{\Delta_{s}}(X) is given by 𝔟CΔ=ξ𝔟πs1{\mathfrak{b}}_{C_{\Delta}}=\xi^{*}{\mathfrak{b}}_{\pi^{s-1}}. By the commutativity of the diagram we then get q(I(πs1))χs(I(πs1))Is(πs1)q^{*}(I(\pi^{s-1}))\cong{{}^{s}\chi}^{*}(I(\pi^{s-1}))\cong I_{s}(\pi^{s-1}) and q𝔟CΔs(X)=𝔳X,sq^{*}{\mathfrak{b}}_{C_{\Delta_{s}}(X)}=\mathfrak{v}_{X,s} as claimed in the first item.

The equality established above implies that q𝔟CΔs(X)sn=𝔳X,ssnq^{*}{\mathfrak{b}}^{sn}_{C_{\Delta_{s}}(X)}=\mathfrak{v}^{sn}_{X,s}. For dimensional reasons, the map q:Hsn(CΔs(X);I(πs1))Hsn(Xs;Is(πs1))q^{*}:H^{sn}(C_{\Delta_{s}}(X);I(\pi^{s-1}))\to H^{sn}(X^{s};I_{s}(\pi^{s-1})) is an isomorphism. We therefore have 𝔟CΔs(X)sn=0{\mathfrak{b}}^{sn}_{C_{\Delta_{s}}(X)}=0 if and only if 𝔳X,ssn=0\mathfrak{v}^{sn}_{X,s}=0, which implies the second item.

We now prove the last item. Let 𝖼Hsn(Xs;(sχγs)A)\mathsf{c}\in H_{sn}(X^{s};(^{s}\chi\gamma^{s})^{*}A) be a nonzero class and let 𝔠=γs(𝖼)\mathfrak{c}=\gamma^{s}_{*}(\mathsf{c}). We have χs(𝔠)=ξq(𝖼){}^{s}\chi_{*}(\mathfrak{c})=\xi_{*}q_{*}(\mathsf{c}). Note that q(𝖼)q_{*}(\mathsf{c}) is a homology class of degree snsn. Since TCs(X)<sn\operatorname{TC}_{s}(X)<sn we have cat(CΔs(X))<sn\operatorname{\mathrm{cat}}(C_{\Delta_{s}}(X))<sn. Therefore the classifying map ξ\xi factors up to homotopy through an (sn1)(sn-1)-dimensional space. Consequently, ξq(𝖼)=0\xi_{*}q_{*}(\mathsf{c})=0 and the result follows. ∎

Remark 2.2.

In the situation of Proposition 2.1, if AA is a trivial [πs1]{\mathbb{Z}}[\pi^{s-1}]-module and 𝖼Hsn(Xs;A)\mathsf{c}\in H_{sn}(X^{s};A) is an element such that the class 𝔠=γs(𝖼)Hsn(πs;A)\mathfrak{c}=\gamma^{s}_{*}(\mathsf{c})\in H_{sn}(\pi^{s};A) satisfies χs(𝔠)0{}^{s}\chi_{*}(\mathfrak{c})\neq 0 then TCs(X)=sn\operatorname{TC}_{s}(X)=sn.

Item (3) of Proposition 2.1 is sharp under reasonably general conditions. Let MM be an nn-dimensional connected closed manifold with fundamental group π=π1(M)\pi=\pi_{1}(M) and let ω=ωM:π{±1}\omega=\omega_{M}:\pi\to\{\pm 1\} be the homomorphism determined by the first Stiefel-Whitney class of MM. Recall that the orientation module of MM, denoted by ~=~M\widetilde{{\mathbb{Z}}}=\widetilde{{\mathbb{Z}}}_{M}, is the abelian group {\mathbb{Z}} given with a structure of [π]{\mathbb{Z}}[\pi]-module determined by at=ω(a)ta\cdot t=\omega(a)t for aπa\in\pi, tt\in{\mathbb{Z}}. Note that ~Ms=~Ms\widetilde{{\mathbb{Z}}}_{M^{s}}=\widetilde{{\mathbb{Z}}}_{M}^{\otimes s}, which additively is {\mathbb{Z}} with πs\pi^{s} action given by (a1,a2,,as)t=ω(a1)ω(a2)ω(as)t(a_{1},a_{2},\dots,a_{s})t=\omega(a_{1})\omega(a_{2})\cdots\omega(a_{s})t.

Proposition 2.3.

Let MM be an nn-dimensional connected closed manifold with n2n\geq 2 and π=π1(M)\pi=\pi_{1}(M) abelian. Assume there is a [πs1]{\mathbb{Z}}[\pi^{s-1}]-module AA such that the [πs]{\mathbb{Z}}[\pi^{s}]-modules χs(A){}^{s}\chi^{*}(A) and ~s\widetilde{{\mathbb{Z}}}^{\otimes s} are isomorphic. Then the following two conditions are equivalent:

  1. (1)

    The class 𝔪:=γ([M])Hn(π;~)\mathfrak{m}:=\gamma_{*}([M])\in H_{n}(\pi;\widetilde{{\mathbb{Z}}}) satisfies χs(𝔪×s)=0{}^{s}\chi_{*}(\mathfrak{m}^{\times s})=0 in Hsn(πs1;A)H_{sn}(\pi^{s-1};A).

  2. (2)

    TCs(X)<sn\operatorname{TC}_{s}(X)<sn.

Here we denote by 𝔪×sHsn(πs;~s)\mathfrak{m}^{\times s}\in H_{sn}(\pi^{s};\widetilde{{\mathbb{Z}}}^{\otimes s}) the image of the fundamental class of MsM^{s} under the homomorphism induced by γs:MsBπs\gamma^{s}:M^{s}\to B\pi^{s}. Note that we also denote by ~\widetilde{{\mathbb{Z}}} the local system over BπB\pi arising from the isomorphism π1(γ)\pi_{1}(\gamma) induced by the classifying map γ:MBπ\gamma\colon M\to B\pi.

Proof.

From the naturality of the cap-product and the assumption that AA is a [πs1]{\mathbb{Z}}[\pi^{s-1}]-module satisfying χs(A)~s{}^{s}\chi^{*}(A)\cong\widetilde{{\mathbb{Z}}}^{\otimes s} we get the following diagram.

Hsn(Ms;~s)Hsn(Ms;Issn(πs1))\textstyle{H_{sn}(M^{s};\widetilde{{\mathbb{Z}}}^{\otimes s})\otimes H^{sn}(M^{s};{I}_{s}^{sn}(\pi^{s-1}))\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}(γs)\scriptstyle{(\gamma^{s})_{*}}\scriptstyle{\cap}\scriptstyle{\cong}Issn(πs1)πs~s\textstyle{{I}_{s}^{sn}(\pi^{s-1})\otimes_{\pi^{s}}\widetilde{{\mathbb{Z}}}^{\otimes s}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}=\scriptstyle{=}Hsn(Bπs;~s)Hsn(Bπs;Issn(πs1))\textstyle{H_{sn}(B\pi^{s};\widetilde{{\mathbb{Z}}}^{\otimes s})\otimes H^{sn}(B\pi^{s};{I}_{s}^{sn}(\pi^{s-1}))\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}(γs)\scriptstyle{(\gamma^{s})^{*}}(sχ)\scriptstyle{(^{s}\chi)_{*}}\scriptstyle{\cap}Issn(πs1)πs~s\textstyle{{I}_{s}^{sn}(\pi^{s-1})\otimes_{\pi^{s}}\widetilde{{\mathbb{Z}}}^{\otimes s}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}(sχ)\scriptstyle{(^{s}\chi)_{*}}\scriptstyle{\cong}Hsn(Bπs1;A)Hsn(Bπs1;Isn(πs1))\textstyle{H_{sn}(B\pi^{s-1};A)\otimes H^{sn}(B\pi^{s-1};{I}^{sn}(\pi^{s-1}))\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}(sχ)\scriptstyle{(^{s}\chi)^{*}}\scriptstyle{\cap}Isn(πs1)πs1A\textstyle{{I}^{sn}(\pi^{s-1})\otimes_{\pi^{s-1}}A}

The cap-product on the first line is an isomorphism by Poincaré duality. The bottom vertical map in the third column corresponds to the morphism

χ:H(Bπs;Issn(πs1)~s)H(Bπs1;Isn(πs1)A)\chi_{*}:H_{*}(B\pi^{s};{I}_{s}^{sn}(\pi^{s-1})\otimes\widetilde{{\mathbb{Z}}}^{\otimes s})\to H_{*}(B\pi^{s-1};{I}^{sn}(\pi^{s-1})\otimes A)

in degree 0. It is induced by the obvious isomophism between the underlying {\mathbb{Z}}-modules Isn(πs1)~sI^{sn}(\pi^{s-1})\otimes\widetilde{{\mathbb{Z}}}^{\otimes s} and Isn(πs1)AI^{sn}(\pi^{s-1})\otimes A and is an isomorphism on the coinvariants because χs{}^{s}\chi is surjective.

Let [M]Hn(M;~)[M]\in H_{n}(M;\widetilde{{\mathbb{Z}}}) be the fundamental class. Since the third column of the diagram is comprised of isomorphisms, we have

(sχ)(γs)([Ms])𝔟πs1sn=0 if and only if [Ms](γs)(sχ)𝔟πs1sn=0.(^{s}\chi)_{*}(\gamma^{s})_{*}([M^{s}])\cap\mathfrak{b}^{sn}_{\pi^{s-1}}=0\mbox{ if and only if }[M^{s}]\cap(\gamma^{s})^{*}(^{s}\chi)^{*}\mathfrak{b}^{sn}_{\pi^{s-1}}=0.

This is equivalent to saying

(sχ)(𝔪×s)𝔟πs1sn=0 if and only if [Ms]𝔳M,ssn=0.(^{s}\chi)_{*}(\mathfrak{m}^{\times s})\cap\mathfrak{b}^{sn}_{\pi^{s-1}}=0\mbox{ if and only if }[M^{s}]\cap\mathfrak{v}^{sn}_{M,s}=0.

The hypothesis χs(𝔪×s)=0{}^{s}\chi_{*}(\mathfrak{m}^{\times s})=0 yields [Ms]𝔳M,ssn=0[M^{s}]\cap\mathfrak{v}^{sn}_{M,s}=0. By Poincaré duality, we can then conclude that 𝔳M,ssn=0\mathfrak{v}^{sn}_{M,s}=0 and consequently TCs(M)<sn\operatorname{TC}_{s}(M)<sn. ∎

Corollary 2.4.

Let MM be an orientable nn-dimensional manifold with n2n\geq 2 and abelian fundamental group π=π1(M)\pi=\pi_{1}(M). The class 𝔪=γ([M])Hn(π;)\mathfrak{m}=\gamma_{*}([M])\in H_{n}(\pi;{\mathbb{Z}}) satisfies χs(𝔪×s)=0{}^{s}\chi_{*}(\mathfrak{m}^{\times s})=0 in Hsn(πs1;)H_{sn}(\pi^{s-1};{\mathbb{Z}}) if and only if TCs(X)<sn\operatorname{TC}_{s}(X)<sn.

Proof.

Since MM is orientable, the orientation module ~\widetilde{{\mathbb{Z}}} is just {\mathbb{Z}} with trivial action. Taking A=A={\mathbb{Z}} also with trivial action, we have χs(A)s{}^{s}\chi^{*}(A)\cong{{\mathbb{Z}}}^{\otimes s} and the result follows from Proposition 2.3. ∎

Remark 2.5.

When MM is non-orientable, the [πs1]{\mathbb{Z}}[\pi^{s-1}]-module A=~s1A=\widetilde{{\mathbb{Z}}}^{\otimes s-1} satisfies the assumptions of Proposition 2.3 for s=2s=2 ([6]) but fails to do so for all s>2s>2. For instance, set s=3s=3 and suppose that bπb\in\pi is an element for which the orientation character ω:π{±1}\omega:\pi\to\{\pm 1\} satisfies ω(b)=1\omega(b)=-1. Then, for a,cπa,c\in\pi and tt\in{\mathbb{Z}} we have

(a,b,c)t=ω(a)ω(b)ω(c)t=ω(a)ω(c)t(a,b,c)\cdot t=\omega(a)\omega(b)\omega(c)t=-\omega(a)\omega(c)t

while

χ3(a,b,c)t=(ab¯,bc¯)t=ω(ab¯)ω(bc¯)t=ω(a)ω(c)t.{}^{3}\chi(a,b,c)\cdot t=(a\overline{b},b\overline{c})\cdot t=\omega(a\overline{b})\omega(b\overline{c})t=\omega(a)\omega(c)t.

This shows that the map χ3{}^{3}\chi does not induce a homomorphism from H(π3;~3)H_{*}(\pi^{3};\widetilde{{\mathbb{Z}}}^{\otimes 3}) to H(π2;~2)H_{*}(\pi^{2};\widetilde{{\mathbb{Z}}}^{\otimes 2}). Note that, in Proposition 2.3, AA must be, as an abelian group, isomorphic to {\mathbb{Z}}. Furthermore, since χs{}^{s}\chi is surjective, the [πs1]{\mathbb{Z}}[\pi^{s-1}]-module structure on AA is forced by the hypothesis χs(A)~s{}^{s}\chi^{*}(A)\cong\widetilde{{\mathbb{Z}}}^{\otimes s} and this condition is impossible when ss is odd. For instance, again set s=3s=3, choose bπb\in\pi as above and assume χ3(A)~3{}^{3}\chi^{*}(A)\cong\widetilde{{\mathbb{Z}}}^{\otimes 3}. The equalities χ3(b,1,1)=χ3(1,b¯,b¯)=(b,1){}^{3}\chi(b,1,1)={}^{3}\chi(1,\overline{b},\overline{b})=(b,1) then lead to the impossible

t=ω(b¯)ω(b¯)t=χ3(1,b¯,b¯)t=(b,1)t=χ3(b,1,1)t=ω(b)t=t.t=\omega(\overline{b})\omega(\overline{b})t={}^{3}\chi(1,\overline{b},\overline{b})\cdot t=(b,1)\cdot t={}^{3}\chi(b,1,1)\cdot t=\omega(b)t=-t.

Nonetheless, when s=2σs=2\sigma, σ1\sigma\geq 1, the [πs1]{\mathbb{Z}}[\pi^{s-1}]-module A=~(~)σ1A=\widetilde{{\mathbb{Z}}}\otimes({\mathbb{Z}}\otimes\widetilde{{\mathbb{Z}}})^{\sigma-1} does satisfy χs(A)~s{}^{s}\chi^{*}(A)\cong\widetilde{{\mathbb{Z}}}^{\otimes s}, and we explore its usage in Section 5.

3. Some calculations in the orientable case

Let MM be an orientable connected closed manifold with π=π1(M)\pi=\pi_{1}(M) abelian. In this section we will use Corollary 2.4 to establish the non-maximality TCs(M)<sdim(M)\operatorname{TC}_{s}(M)<s\dim(M) for some families of manifolds with abelian fundamental groups.

Let γ:MBπ\gamma:M\to B\pi be a classifying map and let 𝔪=γ([M])Hn(π;)\mathfrak{m}=\gamma_{*}([M])\in H_{n}(\pi;{\mathbb{Z}}). Since MM is orientable, we will suppress the {\mathbb{Z}}-coefficients from the notation. In all cases, we will see that χs(𝔪s)=0{}^{s}\chi_{*}(\mathfrak{m}^{\otimes s})=0 in Hsn(πs1)H_{sn}(\pi^{s-1}).

In our first result, we suppose that π\pi is a free abelian group. This case has already been considered in [12] in the more general context of finite CW-complexes. Here, restricting to closed manifolds, we obtain a slightly stronger statement than [12, Corollary 6.14].

Proposition 3.1.

Let MM be an orientable nn-dimensional connected closed manifold with π1(M)=r\pi_{1}(M)={\mathbb{Z}}^{r} and let s2s\geq 2. If sn>(s1)rsn>(s-1)r then TCs(M)<sn\operatorname{TC}_{s}(M)<sn.

Proof.

Let π=r\pi={\mathbb{Z}}^{r}, let γ:MBπ\gamma:M\to B\pi be a classifying map and let 𝔪=γ([M])Hn(π))\mathfrak{m}=\gamma_{*}([M])\in H_{n}(\pi)). For degree reasons, we can see that χs(𝔪×s)=0{}^{s}\chi_{*}(\mathfrak{m}^{\times s})=0 in Hsn(πs1)H_{sn}(\pi^{s-1}). Indeed Bπ=(S1)rB\pi=(S^{1})^{r} and Hk(π)=0H_{k}(\pi)=0 for k>rk>r. Consequently Hsn(πs1)=0H_{sn}(\pi^{s-1})=0 if sn>(s1)rsn>(s-1)r. ∎

In general, observe that the homomorphism χs:πsπs1{}^{s}\chi:\pi^{s}\to\pi^{s-1} given by

χs(a1,,as)=(a1a2¯,a2a3¯,,as1as¯){}^{s}\chi(a_{1},\dots,a_{s})=(a_{1}\overline{a_{2}},a_{2}\overline{a_{3}},\dots,a_{s-1}\overline{a_{s}})

can be decomposed as

(3.1) sχ=(χ××χs1)(Id×Δ××Δs2×Id)^{s}\chi=(\underbrace{\chi\times\cdots\times\chi}_{s-1})\circ({\rm Id}\times\underbrace{\Delta\times\cdots\times\Delta}_{s-2}\times{\rm Id})

where Δ=Δ2π:ππ×π\Delta=\Delta_{2}^{\pi}:\pi\to\pi\times\pi is the diagonal map and χ=χ2\chi={{}^{2}\chi}. Denote by j:ππj:\pi\to\pi the inversion. Since χ\chi can be seen as the multiplication of π\pi, μ:π×ππ\mu:\pi\times\pi\to\pi, precomposed with Id×j{\rm Id}\times j, we have, for classes 𝔞,𝔟H(π)\mathfrak{a},\mathfrak{b}\in H_{*}(\pi),

χ(𝔞×𝔟)=𝔞j(𝔟)\chi_{*}(\mathfrak{a}\times\mathfrak{b})=\mathfrak{a}\wedge j_{*}(\mathfrak{b})

where \wedge is the Pontryagin product, that is, the product induced by μ\mu in homology, see [4, V.5].

In the results below, we consider the cyclic group q=v|vq=1{\mathbb{Z}}_{q}=\langle v~|~v^{q}=1\rangle and work at the chain level. Recall the classical resolution of {\mathbb{Z}} as a trivial [q]{\mathbb{Z}}[{\mathbb{Z}}_{q}]-module given by

(3.2) \textstyle{\cdots\ignorespaces\ignorespaces\ignorespaces\ignorespaces}[q]\textstyle{{\mathbb{Z}}[{\mathbb{Z}}_{q}]\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Nq(v)\scriptstyle{N_{q}(v)}[q]\textstyle{{\mathbb{Z}}[{\mathbb{Z}}_{q}]\ignorespaces\ignorespaces\ignorespaces\ignorespaces}v1\scriptstyle{v-1}[q]\textstyle{{\mathbb{Z}}[{\mathbb{Z}}_{q}]\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Nq(v)\scriptstyle{N_{q}(v)}[q]\textstyle{{\mathbb{Z}}[{\mathbb{Z}}_{q}]\ignorespaces\ignorespaces\ignorespaces\ignorespaces}v1\scriptstyle{v-1}[q]\textstyle{{\mathbb{Z}}[{\mathbb{Z}}_{q}]\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ε\scriptstyle{\varepsilon}\textstyle{\mathbb{Z}}

where Nq(v)=1+v++vq1N_{q}(v)=1+v+\cdots+v^{q-1}.

In the following lemma, we recall the morphisms induced by the diagonal Δ\Delta, the multiplication μ\mu and the inversion jj on the level of resolutions (see [4, page 108] and [6, §3.2]). Let [k][k] denote the generator of degree kk in (3.2), and write Bi,jB_{i,j} for the binomial coefficient (i+ji)\binom{\,i+j\,}{i}.

Lemma 3.2.

At the level of the resolution (3.2),

  1. (a)(a)

    Δ\Delta is given on generators by [p]k+l=pΔkl[p][p]\mapsto\sum_{k+l=p}\Delta_{kl}[p] where

    Δkl[p]={[k][l]k even;[k]v[l]k odd, l even;0i<jq1vi[k]vj[l]k odd, l odd.\Delta_{kl}[p]=\left\{\begin{array}[]{ll}[k]\otimes[l]&k\mbox{ even;}\\ \mathopen{[}k\mathclose{]}\otimes v[l]&k\mbox{ odd, }l\mbox{ even;}\\ \sum_{0\leq i<j\leq q-1}v^{i}[k]\otimes v^{j}[l]&k\mbox{ odd, }l\mbox{ odd.}\end{array}\right.
  2. (b)(b)

    μ\mu is given on generators by the formulæ

    [2i][2j]\displaystyle[2i]\otimes[2j] Bi,j[2(i+j)];\displaystyle\mapsto B_{i,j}\,[2(i+j)];
    [2i][2j+1]\displaystyle[2i]\otimes[2j+1] Bi,j[2(i+j)+1];\displaystyle\mapsto B_{i,j}\,[2(i+j)+1];
    [2i+1][2j]\displaystyle[2i+1]\otimes[2j] Bi,j[2(i+j)+1];\displaystyle\mapsto B_{i,j}\,[2(i+j)+1];
    [2i+1][2j+1]\displaystyle[2i+1]\otimes[2j+1] 0.\displaystyle\mapsto 0.
  3. (c)(c)

    jj is given on generators by

    [i]Nq1k(v)[i]if i{2k,2k1}.[i]\to N_{q-1}^{k}(v)[i]\qquad\mbox{if }i\in\{2k,2k-1\}.

We denote by C(q)C_{\bullet}({\mathbb{Z}}_{q}) the {\mathbb{Z}}-chain complex obtained by tensoring the resolution (3.2) with {\mathbb{Z}} over q{\mathbb{Z}}_{q}.

(3.3) C(q):\textstyle{C_{\bullet}({\mathbb{Z}}_{q}):}\textstyle{\cdots\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0\scriptstyle{0}[2k]\textstyle{{\mathbb{Z}}[2k]\ignorespaces\ignorespaces\ignorespaces\ignorespaces}q\scriptstyle{q}[2k1]\textstyle{{\mathbb{Z}}[2k-1]\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0\scriptstyle{0}\textstyle{\cdots\ignorespaces\ignorespaces\ignorespaces\ignorespaces}q\scriptstyle{q}[1]\textstyle{{\mathbb{Z}}[1]\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0\scriptstyle{0}[0]\textstyle{{\mathbb{Z}}[0]}

Recall that the homology of this chain complex gives H(q)=H(q;)H_{*}({\mathbb{Z}}_{q})=H_{*}({\mathbb{Z}}_{q};{\mathbb{Z}}). In positive degrees, H+(q)H_{+}({\mathbb{Z}}_{q}) is concentrated in odd degrees.

As in [6], we denote by :C(q)C(q)C(q)\wedge:C_{\bullet}({\mathbb{Z}}_{q})\otimes C_{\bullet}({\mathbb{Z}}_{q})\to C_{\bullet}({\mathbb{Z}}_{q}) the Pontryagin product, which is given by the formulæ (b)(b) of Lemma 3.2:

(3.4) {[2i][2k]=Bi,k[2i+2k],[2i+1][2k+1]=0,[2i][2k+1]=[2k+1][2i]=Bi,k[2i+2k+1].\left\{\begin{array}[]{l}~[2i]\wedge[2k]=B_{i,k}[2i+2k],\quad\quad[2i+1]\wedge[2k+1]=0,\\ \\ ~[2i]\wedge[2k+1]=[2k+1]\wedge[2i]=B_{i,k}[2i+2k+1].\end{array}\right.

We denote by 𝐣:C(q)C(q)\mathbf{j}:C_{\bullet}({\mathbb{Z}}_{q})\to C_{\bullet}({\mathbb{Z}}_{q}) the morphism induced by the inversion, which is from Lemma 3.2 (c)(c) given by

(3.5) 𝐣([i])=(q1)k[i]if i{2k,2k1}.\mathbf{j}([i])=(q-1)^{k}[i]\qquad\mbox{if }i\in\{2k,2k-1\}.

In these terms, the chain map χ=χ2\chi_{\bullet}={{}^{2}\chi}_{\bullet} induced by χ=χ2\chi={{}^{2}\chi} can be described as the composite

C(q)C(q)\textstyle{C_{\bullet}({\mathbb{Z}}_{q})\otimes C_{\bullet}({\mathbb{Z}}_{q})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Id𝐣\scriptstyle{\mathrm{Id}\otimes\mathbf{j}}C(q)C(q)\textstyle{C_{\bullet}({\mathbb{Z}}_{q})\otimes C_{\bullet}({\mathbb{Z}}_{q})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\wedge}C(q).\textstyle{C_{\bullet}({\mathbb{Z}}_{q}).}

We will also use the diagonal approximation of C(q)C_{\bullet}({\mathbb{Z}}_{q}), obtained from Lemma 3.2 (a)(a):

(3.6) Δ:C(q)C(q)C(q)[p]k+l=pαkl[k][l].\begin{array}[]{rcl}\Delta_{\bullet}:C_{\bullet}({\mathbb{Z}}_{q})&\to&C_{\bullet}({\mathbb{Z}}_{q})\otimes C_{\bullet}({\mathbb{Z}}_{q})\\[2.0pt] {[p]}&\mapsto&\sum\limits_{k+l=p}\alpha_{kl}[k]\otimes[l].\end{array}

Here αkl=1\alpha_{kl}=1 if klkl is even and αkl=(q1)q/2\alpha_{kl}=(q-1)q/2 if klkl is odd.

Proposition 3.3.

Let MM be an orientable nn-dimensional connected closed manifold with π1(M)=q\pi_{1}(M)={\mathbb{Z}}_{q}. Then, for any s2s\geq 2, we have TCs(M)<sn\operatorname{TC}_{s}(M)<sn.

Proof.

Let γ:MBπ\gamma:M\to B\pi be a classifying map, where π=q\pi={\mathbb{Z}}_{q}, and let 𝔪=γ([M])Hn(π;))\mathfrak{m}=\gamma_{*}([M])\in H_{n}(\pi;{\mathbb{Z}})). We will see that χs(𝔪×s)=0{}^{s}\chi_{*}(\mathfrak{m}^{\times s})=0 in Hsn(πs1;)H_{sn}(\pi^{s-1};{\mathbb{Z}}). If dimM\dim M is even, this is immediate since H+(π)H_{+}(\pi) is concentrated in odd degrees, which implies 𝔪=0\mathfrak{m}=0. We then suppose that dimM=2p+1\dim M=2p+1. A cycle 𝐦C2p+1(q)\mathbf{m}\in C_{2p+1}({\mathbb{Z}}_{q}) representing the class 𝔪\mathfrak{m} is of the form 𝐦=λ[2p+1]\mathbf{m}=\lambda[2p+1] for some λ\lambda\in{\mathbb{Z}}. In order to compute χs(𝔪×s){}^{s}\chi_{*}(\mathfrak{m}^{\times s}) we use the decomposition (3.1) and analyze the element χs(𝐦s){}^{s}\chi_{\bullet}(\mathbf{m}^{\otimes s}) which is given by

(χ)s1(𝐦Δ𝐦Δ𝐦s2𝐦).(\chi_{\bullet})^{\otimes s-1}(\mathbf{m}\otimes\underbrace{\Delta_{\bullet}\mathbf{m}\otimes\cdots\otimes\Delta_{\bullet}\mathbf{m}}_{s-2}\otimes\mathbf{m}).

The element 𝐦Δ𝐦Δ𝐦𝐦\mathbf{m}\otimes\Delta_{\bullet}\mathbf{m}\otimes\cdots\otimes\Delta_{\bullet}\mathbf{m}\otimes\mathbf{m} is given by a {\mathbb{Z}}-linear combination of elements of the form

[2p+1][k1][l1][ks2][ls2][2p+1][2p+1]\otimes[k_{1}]\otimes[l_{1}]\otimes\cdots\otimes[k_{s-2}]\otimes[l_{s-2}]\otimes[2p+1]

where ki+li=2p+1k_{i}+l_{i}=2p+1 for any 1is21\leq i\leq s-2. Setting l0=ks1=2p+1l_{0}=k_{s-1}=2p+1, there will be necessarily some i{0,,s2}i\in\{0,\dots,s-2\} such that lil_{i} and ki+1k_{i+1} are both odd. Applying (χ)s1(\chi_{\bullet})^{\otimes s-1} to the element above yields

([2p+1]𝐣[k1])([l1]𝐣[k2])([ls2]𝐣[ks1]).([2p+1]\wedge\mathbf{j}[k_{1}])\otimes([l_{1}]\wedge\mathbf{j}[k_{2}])\otimes\cdots\otimes([l_{s-2}]\wedge\mathbf{j}[k_{s-1}]).

If lil_{i} and ki+1k_{i+1} are both odd, the corresponding factor ([li]𝐣[ki+1])([l_{i}]\wedge\mathbf{j}[k_{i+1}]) vanishes since 𝐣[ki+1])\mathbf{j}[k_{i+1}]) is a multiple of [ki+1][k_{i+1}] and the Pontryagin product of two odd degree elements is zero. Consequently, we obtain χs(𝐦s)=0{}^{s}\chi_{\bullet}(\mathbf{m}^{\otimes s})=0 and χs(𝔪×s)=0{}^{s}\chi_{*}(\mathfrak{m}^{\times s})=0. ∎

Proposition 3.4.

Let MM be an orientable nn-dimensional connected closed manifold with π1(M)=r×q\pi_{1}(M)={\mathbb{Z}}^{r}\times{\mathbb{Z}}_{q} such that r<nr<n. Then, for any s2s\geq 2, we have TCs(M)<sn\operatorname{TC}_{s}(M)<sn.

Proof.

Let γ:MBπ\gamma:M\to B\pi be a classifying map, where π=r×q\pi={\mathbb{Z}}^{r}\times{\mathbb{Z}}_{q}, and let 𝔪=γ([M])Hn(π;))\mathfrak{m}=\gamma_{*}([M])\in H_{n}(\pi;{\mathbb{Z}})). By the Künneth formula, we have H(r×q)=H(r)H(q)H_{*}({\mathbb{Z}}^{r}\times{\mathbb{Z}}_{q})=H_{*}({\mathbb{Z}}^{r})\otimes H_{*}({\mathbb{Z}}_{q}). Since n>rn>r, we can write 𝔪=σiαi\mathfrak{m}=\sum\sigma_{i}\otimes\alpha_{i} where αiH+(q)\alpha_{i}\in H_{+}({\mathbb{Z}}_{q}) and σiH(r)=(x1,,xr)\sigma_{i}\in H_{*}({\mathbb{Z}}^{r})=\bigwedge(x_{1},\dots,x_{r}) with each xjx_{j} of degree 11. Since H+(q)H_{+}({\mathbb{Z}}_{q}) is concentrated in odd degrees, a cycle 𝐦\mathbf{m} representing 𝔪\mathfrak{m} can be described as a sum of terms of the form λσ[2p+1]\lambda\sigma\otimes[2p+1] where λ\lambda\in{\mathbb{Z}}, p0p\geq 0 and σ(x1,,xr)\sigma\in\bigwedge(x_{1},\cdots,x_{r}) is a class regarded as a cycle. The element 𝐦Δ𝐦Δ𝐦𝐦\mathbf{m}\otimes\Delta_{\bullet}\mathbf{m}\otimes\cdots\otimes\Delta_{\bullet}\mathbf{m}\otimes\mathbf{m} is therefore given by a {\mathbb{Z}}-linear combination of elements of the form

(3.7) σ0[l0]σ1[k1]σ~1[l1]σs2[ks2]σ~s2[ls2]σs1[ks1]\sigma_{0}\otimes[l_{0}]\otimes\sigma_{1}\otimes[k_{1}]\otimes\tilde{\sigma}_{1}\otimes[l_{1}]\otimes\cdots\otimes\sigma_{s-2}\otimes\otimes[k_{s-2}]\otimes\tilde{\sigma}_{s-2}\otimes[l_{s-2}]\otimes\sigma_{s-1}[k_{s-1}]

where l0l_{0}, ki+lik_{i}+l_{i} for 1is21\leq i\leq s-2, and ks1k_{s-1} are all odd and the elements σi,σ~i\sigma_{i},\tilde{\sigma}_{i} belong to (x1,,xr)\bigwedge(x_{1},\cdots,x_{r}). The calculation of χ\chi_{\bullet} on (say) σ0[l0]σ1[k1]\sigma_{0}\otimes[l_{0}]\otimes\sigma_{1}\otimes[k_{1}] is made componentwise and gives rise to factors of the form

±(σ0σ1)([l0]𝐣[k1]).\pm(\sigma_{0}\wedge\sigma_{1})\otimes([l_{0}]\wedge{\mathbf{j}}[k_{1}]).

As in the proof of Proposition 3.3, there will be necessarily, in the expression (3.7), some i{0,,s2}i\in\{0,\dots,s-2\} such that lil_{i} and ki+1k_{i+1} are both odd. After applying χ\chi_{\bullet}, the corresponding factor will be 0. Consequently, we obtain χs(𝐦s)=0{}^{s}\chi_{\bullet}(\mathbf{m}^{\otimes s})=0 and χs(𝔪×s)=0{}^{s}\chi_{*}(\mathfrak{m}^{\times s})=0. We can hence conclude that TCs(M)<sn\operatorname{TC}_{s}(M)<sn. ∎

Limiting examples

Examples 4.1 and 4.2 from [6] show that the conditions in Propositions 3.1 and 3.4 are sharp. We now show that Proposition 3.3 cannot be extended to manifolds whose fundamental group is of the form p×p{\mathbb{Z}}_{p}\times{\mathbb{Z}}_{p} where pp is a prime.

Example 3.5.

A manifold NN with π1(N)=3×3\pi_{1}(N)={\mathbb{Z}}_{3}\times{\mathbb{Z}}_{3} and TC3(N)=3dim(N)\operatorname{TC}_{3}(N)=3\dim(N).

Set π=3×3\pi={\mathbb{Z}}_{3}\times{\mathbb{Z}}_{3} and consider C(π)=C(3)C(3)C_{\bullet}(\pi)=C_{\bullet}({\mathbb{Z}}_{3})\otimes C_{\bullet}({\mathbb{Z}}_{3}). We will write [ik][ik] instead of [i][k][i]\otimes[k]. We first consider the cycle 𝐦=[05]+[50]\mathbf{m}=[05]+[50] and denote by 𝔪\mathfrak{m} its homology class. We will see that χ3(𝔪×3)0{}^{3}\chi_{*}(\mathfrak{m}^{\times 3})\neq 0. By the Universal Coefficient Theorem, it is actually sufficient to see that χ3(𝔪3×3)0{}^{3}\chi_{*}(\mathfrak{m}^{\times 3}_{{\mathbb{Z}}_{3}})\neq 0 where 𝔪3\mathfrak{m}_{{\mathbb{Z}}_{3}} corresponds to 𝔪\mathfrak{m} in H5(π;3)H_{5}(\pi;{\mathbb{Z}}_{3}). As H(π;3)H(3;3)H(3;3)H_{*}(\pi;{\mathbb{Z}}_{3})\cong H_{*}({\mathbb{Z}}_{3};{\mathbb{Z}}_{3})\otimes H_{*}({\mathbb{Z}}_{3};{\mathbb{Z}}_{3}) and H(3;3)=3[k]H_{*}({\mathbb{Z}}_{3};{\mathbb{Z}}_{3})={\mathbb{Z}}_{3}[k] for all k0k\geq 0, we will continue to write 𝔪3=[05]+[50]\mathfrak{m}_{{\mathbb{Z}}_{3}}=[05]+[50].

Using the diagonal approximation associated to the resolution (3.2) described in Lemma 3.2 (or, tensoring the diagonal (3.6) by q{\mathbb{Z}}_{q}) we can check that the homology diagonal of H(3;3)H_{*}({\mathbb{Z}}_{3};{\mathbb{Z}}_{3}) satisfies

Δ[0]=[0][0]Δ[5]=k+l=5[k][l].\Delta_{*}[0]=[0]\otimes[0]\qquad\Delta_{*}[5]=\sum\limits_{k+l=5}[k]\otimes[l].

Consequently, the homology diagonal of H(π;3)H_{*}(\pi;{\mathbb{Z}}_{3}) satisfies:

Δ[05]=k+l=5[0k][0l]Δ[50]=k+l=5[k0][l0].\Delta_{*}[05]=\sum\limits_{k+l=5}[0k]\otimes[0l]\qquad\Delta_{*}[50]=\sum\limits_{k+l=5}[k0]\otimes[l0].

We have to compute:

(χχ)(([05]+[50])(Δ[05]+Δ[50])([05]+[50])).(\chi_{*}\otimes\chi_{*})\left(([05]+[50])\otimes(\Delta_{*}[05]+\Delta_{*}[50])\otimes([05]+[50])\right).

A term of the form χ([kl][kl])\chi_{*}([kl]\otimes[k^{\prime}l^{\prime}]) is given in H(π;3)=H(3;3)H(3;3)H_{*}(\pi;{\mathbb{Z}}_{3})=H_{*}({\mathbb{Z}}_{3};{\mathbb{Z}}_{3})\otimes H_{*}({\mathbb{Z}}_{3};{\mathbb{Z}}_{3}) by a componentwise calculation:

χ([kl][kl])=(1)lk([k]j[k])([l]j[l]).\chi_{*}([kl]\otimes[k^{\prime}l^{\prime}])=(-1)^{lk^{\prime}}([k]\wedge j_{*}[k^{\prime}])\otimes([l]\wedge j_{*}[l^{\prime}]).

Taking into account the formulas for the inversion and for the Pontryagin product (induced in 3{\mathbb{Z}}_{3}-homology by the formulas (3.5) and (3.4) given above) we have

χ([04][05]])=([0]j[0])([4]j[5])=[0]([4]([5]))=[0](6[9])=6[09]\chi_{*}([04]\otimes[05]])=([0]\wedge j_{*}[0])\otimes([4]\wedge j_{*}[5])=[0]\otimes([4]\wedge(-[5]))=-[0]\otimes(6[9])=-6[09]

which vanishes since we are working with coefficients in 3{\mathbb{Z}}_{3}. We can thus check that

(χχ)([50]([01][04])([05]+[50]))=[51][54](\chi_{*}\otimes\chi_{*})\left([50]\otimes([01]\otimes[04])\otimes([05]+[50])\right)=[51]\otimes[54]

and that this is the only term belonging to 3[51]H(π;3){\mathbb{Z}}_{3}[51]\otimes H_{*}(\pi;{\mathbb{Z}}_{3}) in the expansion of χ3(𝔪3×3){}^{3}\chi_{*}(\mathfrak{m}^{\times 3}_{{\mathbb{Z}}_{3}}). Since [51][54][51]\otimes[54] does not vanish in 3[51]H(π;3){\mathbb{Z}}_{3}[51]\otimes H_{*}(\pi;{\mathbb{Z}}_{3}), we can conclude that χ3(𝔪3×3)0{}^{3}\chi_{*}(\mathfrak{m}^{\times 3}_{{\mathbb{Z}}_{3}})\neq 0. Consequently χ3(𝔪×3)0{}^{3}\chi_{*}(\mathfrak{m}^{\times 3})\neq 0.

We can next follow the same strategy as in [6] to show that there exists a manifold NN with fundamental group π=3×3\pi={\mathbb{Z}}_{3}\times{\mathbb{Z}}_{3} and maximal TC3\operatorname{TC}_{3}. More precisely, considering the lens spaces L35=S5/3L_{3}^{5}=S^{5}/{\mathbb{Z}}_{3} and L3=S/3=B3L_{3}^{\infty}=S^{\infty}/{\mathbb{Z}}_{3}=B{\mathbb{Z}}_{3}, we can realize the class [05]+[50]H(π)[05]+[50]\in H_{*}(\pi) as the image of the fundamental class of M=L35#L35M=L_{3}^{5}\#L_{3}^{5} under the map induced by

f:MpinchL35L35⸦⟶L3L3=B(π).f\colon M\xrightarrow{\ \text{pinch}\ }L_{3}^{5}\vee L_{3}^{5}{\lhook\joinrel\longrightarrow}L_{3}^{\infty}\vee L_{3}^{\infty}=B(\pi).

We can then use surgery to replace MM by a manifold NN with π1(N)=π\pi_{1}(N)=\pi and ff by a classifying map γ:NBπ\gamma:N\to B\pi. In this way, 𝔪=γ([N])\mathfrak{m}=\gamma_{*}([N]) and, from χ3(𝔪×3)0{}^{3}\chi_{*}(\mathfrak{m}^{\times 3})\neq 0 and Proposition 2.1 (3), we can deduce that TC3(N)=3n\operatorname{TC}_{3}(N)=3n.

In [6], it has been shown that the regular topological complexity TC=TC2\operatorname{TC}=\operatorname{TC}_{2} of a non-orientable manifold with abelian fundamental group is never maximal. This is not longer true for TCs\operatorname{TC}_{s} with s3s\geq 3. For instance, for the real projective plane P2P^{2}, ss-zero-divisor cuplength considerations imply that TC3(P2)=6\operatorname{TC}_{3}(P^{2})=6, see [9] and the discussion in §4 below. With the approach of this paper, we pursue more general maximality results of this nature next.

4. Cohomological lower bounds

In this section, we use cohomological lower bounds on TCs\operatorname{TC}_{s} given by the ss-zero-divisor cup length or TCs\operatorname{TC}_{s}-weights as well as specific calculations from [9, 8] to obtain lower bounds on the higher topological complexity of families of manifolds with finite cyclic fundamental group and maximal LS-category. In some cases, exact values are given by using our results from Section 3.

Let 𝕜\Bbbk be a field. Recall that, for a space XX, the (𝕜\Bbbk-coefficients) ss-zero-divisor cup length, zcls(X)=zcls(X;𝕜)\operatorname{\mathrm{zcl}}_{s}(X)=\operatorname{\mathrm{zcl}}_{s}(X;\Bbbk), is the maximum of the set

{|u1u0,ui𝐙s(X;𝕜)}\left\{\ell\ |\ u_{1}\dots u_{\ell}\neq 0,u_{i}\in\mathbf{Z}_{s}(X;\Bbbk)\right\}

where

𝐙s(X;𝕜)=ker(i=1sH(X;𝕜)H(X;𝕜)).\mathbf{Z}_{s}(X;\Bbbk)=\ker\left(\bigotimes_{i=1}^{s}H^{\ast}(X;\Bbbk)\stackrel{{\scriptstyle\cup}}{{\to}}H^{\ast}(X;\Bbbk)\right).

We have zcls(X)TCs(X)\operatorname{\mathrm{zcl}}_{s}(X)\leq\operatorname{TC}_{s}(X), see [2].

In some cases, a better lower bound can be obtained through the notion of TCs\operatorname{TC}_{s}-weight. Recall (see [14, §2]) that if p:EBp:E\to B is a fibration and uH~(B;𝕜)u\in\widetilde{H}^{*}(B;\Bbbk) is a nontrivial class, the weight of uu associated to pp, wgtp(u)\operatorname{\mathrm{wgt}}_{p}(u), is the largest integer kk such that f(u)=0f^{*}(u)=0 for any map f:YBf:Y\to B satisfying secat(f(p))<k\operatorname{secat}(f^{*}(p))<k. If u0u\neq 0, then wgtp(u)>0\operatorname{\mathrm{wgt}}_{p}(u)>0 if and only if p(u)=0p^{*}(u)=0 and secat(p)wgtp(u)\operatorname{secat}(p)\geq\operatorname{\mathrm{wgt}}_{p}(u). Moreover, if u1,,ulH~(B;𝕜)u_{1},\dots,u_{l}\in\widetilde{H}^{*}(B;\Bbbk) satisfy u1ul0u_{1}\cup\cdots\cup u_{l}\neq 0 then

wgtp(u1ul)wgtp(u1)++wgtp(ul).\operatorname{\mathrm{wgt}}_{p}(u_{1}\cup\cdots\cup u_{l})\geq\operatorname{\mathrm{wgt}}_{p}(u_{1})+\cdots+\operatorname{\mathrm{wgt}}_{p}(u_{l}).

For a space XX, the TCs\operatorname{TC}_{s}-weight, denoted by wgts\operatorname{\mathrm{wgt}}_{s}, is the weight associated to the fibration es:PXXse_{s}:PX\to X^{s}. Taking coefficients in 𝕜\Bbbk, the morphism ese_{s}^{*} can be identified with the ss-fold cup-product and we can define the (𝕜\Bbbk-coefficients) weighted ss-zero divisor cup length, zclsw¯(X)=zclsw¯(X;𝕜)\operatorname{\operatorname{\mathrm{zcl}}^{\underline{w}}_{s}}(X)=\operatorname{\operatorname{\mathrm{zcl}}^{\underline{w}}_{s}}(X;\Bbbk), to be the maximum of the set

{i=1wgts(ui)|u1u0,ui𝐙s(X;𝕜)}.\left\{\sum_{i=1}^{\ell}\operatorname{\mathrm{wgt}}_{s}(u_{i})\ |\ u_{1}\dots u_{\ell}{\neq 0,u_{i}}\in\mathbf{Z}_{s}(X;\Bbbk)\ \right\}.

We have TCs(X)zclsw¯(X)\operatorname{TC}_{s}(X)\geq\operatorname{\operatorname{\mathrm{zcl}}^{\underline{w}}_{s}}(X). We also note that if f:YXf:Y\to X is a map and uH(Xs;𝕜)u\in H^{*}(X^{s};\Bbbk) satisfies (fs)(u)0(f^{s})^{*}(u)\neq 0 then wgts(f(u))wgts(u)\operatorname{\mathrm{wgt}}_{s}(f^{*}(u))\geq\operatorname{\mathrm{wgt}}_{s}(u).

4.1. Manifolds with π1(M)=2\pi_{1}(M)={\mathbb{Z}}_{2} and cat(M)=dimM\operatorname{\mathrm{cat}}(M)=\dim M

The 2{\mathbb{Z}}_{2}-coefficient ss-zero-divisor cuplength of the real projective space PnP^{n} has been studied extensively, see [5] and [9]. We will see in Proposition 4.2 below how to use these results to obtain information on TCs(M)\operatorname{TC}_{s}(M) when π1(M)=2\pi_{1}(M)={\mathbb{Z}}_{2} and cat(M)=dimM\operatorname{\mathrm{cat}}(M)=\dim M. We first recall some results from [9].

For an integer n>0n>0 with binary expansion d2d1d0\cdots d_{2}d_{1}d_{0}, i.e., n=i0di2in=\sum_{i\geq 0}d_{i}2^{i}, with digits di{0,1}d_{i}\in\{0,1\}, let

  • ν(n)\nu(n) denote the exponent in the maximal 2-power dividing nn, i.e., ν(n)\nu(n) is the minimal ii with di=1d_{i}=1;

  • S(n):={i>0:di+1didi1=011}S(n):=\{i>0:\cdots d_{i+1}d_{i}d_{i-1}\cdots=\cdots 011\cdots\}, the set of binary positions ii starting (from left to right) a block of consecutive 1’s of length at least 2;

  • Zi(n):=j=0i(1dj)2jZ_{i}(n):=\sum_{j=0}^{i}(1-d_{j})2^{j}, the complement of the binary expansion of nn mod 2i+12^{i+1}.

Building on [5], Davis [9] proves that, for s3s\geq 3, the 2{\mathbb{Z}}_{2}-coefficient ss-zero-divisors cuplength of the nn-dimensional real projective space PnP^{n} is given by

zcls(Pn)=snmn,s\operatorname{\mathrm{zcl}}_{s}(P^{n})=sn-m_{n,s}

where mn,s=max{2ν(n+1)1,2i+11sZi(n):iS(n)}m_{n,s}=\max\{2^{\nu(n+1)}-1,2^{i+1}-1-sZ_{i}(n):i\in S(n)\}. In particular, for even nn (so that PnP^{n} is non-orientable), PnP^{n} has maximal possible TCs(Pn)\operatorname{TC}_{s}(P^{n}), that is, TCs(Pn)=sn\operatorname{TC}_{s}(P^{n})=sn, whenever

(4.1) smax{3,2i+11Zi(n):iS(n)}.s\geq\max\left\{3,\left\lceil\frac{2^{i+1}-1}{Z_{i}(n)}\right\rceil:i\in S(n)\right\}.

We specialize two cases of the condition (4.1):

Example 4.1.
  • (a)

    When nn is even and its binary expansion has no blocks of two or more consecutive 1’s, we have S(n)=S(n)=\emptyset and the inequality (4.1) reduces to s3s\geq 3. Note that this condition for the maximality of TCs(Pn)\operatorname{TC}_{s}(P^{n}) is sharp, since TC2(Pn)<2n\operatorname{TC}_{2}(P^{n})<2n ([15], [7, Theorem 1]).

  • (b)

    For n=2r+12n=2^{r+1}-2, we have S(n)={r}S(n)=\{r\}, mn,s=2r+11sm_{n,s}=2^{r+1}-1-s and (4.1) becomes s2r+11s\geq 2^{r+1}-1. Note that, when s=n=2r+12s=n=2^{r+1}-2, we have TCn(Pn){sn,sn1}\operatorname{TC}_{n}(P^{n})\in\{sn,sn-1\}. We will see in Section 5 that TCn(Pn)=sn1\operatorname{TC}_{n}(P^{n})=sn-1 so that the condition s2r+11s\geq 2^{r+1}-1 for the maximality of TCs(Pn)\operatorname{TC}_{s}(P^{n}) is sharp again.

Thanks to the following result, Davis’ computations of zcls(Pn)\operatorname{\mathrm{zcl}}_{s}(P^{n}) have impact on more general manifolds.

Proposition 4.2.

Let MM be an nn-dimensional connected closed manifold with cat(M)=n\operatorname{\mathrm{cat}}(M)=n and π1(M)=2\pi_{1}(M)={\mathbb{Z}}_{2}. Then, for any s2s\geq 2, TCs(M)zcls(Pn;2)\operatorname{TC}_{s}(M)\geq\operatorname{\mathrm{zcl}}_{s}(P^{n};{\mathbb{Z}}_{2}).

Proof.

Let γ:MP=B2\gamma:M\to P^{\infty}=B{\mathbb{Z}}_{2} be a classifying map and let xH1(B2;2)=2x\in H_{1}(B{\mathbb{Z}}_{2};{\mathbb{Z}}_{2})={\mathbb{Z}}_{2} be the generator. For dimensional reasons, γ\gamma factors as McPnPM\stackrel{{\scriptstyle c}}{{\to}}P^{n}\hookrightarrow P^{\infty}. Let xM=γ(x)=c(x)H1(M;2)x_{M}=\gamma^{*}(x)=c^{*}(x)\in H^{1}(M;{\mathbb{Z}}_{2}). The hypothesis that cat(M)=n\operatorname{\mathrm{cat}}(M)=n implies that xM0x_{M}\neq 0, see [3]. Consequently c:H(Pn;2)H(M;2)c^{*}:H^{*}(P^{n};{\mathbb{Z}}_{2})\to H^{*}(M;{\mathbb{Z}}_{2}) as well as (cs):H((Pn)s;2)H(Ms;2)(c^{s})^{*}:H^{*}((P^{n})^{s};{\mathbb{Z}}_{2})\to H^{*}(M^{s};{\mathbb{Z}}_{2}) are monomorphisms. As the image by (cs)(c^{s})^{*} of a ss-zero-divisor of PnP^{n} over 2{\mathbb{Z}}_{2} gives rise to a ss-zero-divisor of MM over 2{\mathbb{Z}}_{2}, we can conclude that zcls(M;2)zcls(Pn;2)\operatorname{\mathrm{zcl}}_{s}(M;{\mathbb{Z}}_{2})\geq\operatorname{\mathrm{zcl}}_{s}(P^{n};{\mathbb{Z}}_{2}) and the result follows. ∎

From Example 4.1 and the discussion above, we directly obtain:

Corollary 4.3.

Let MM be an nn-dimensional connected closed manifold with cat(M)=n\operatorname{\mathrm{cat}}(M)=n and π1(M)=2\pi_{1}(M)={\mathbb{Z}}_{2}. If nn is even and ss satisfies the inequality (4.1), then TCs(M)=sn\operatorname{TC}_{s}(M)=sn. In particular:

  • (a)

    If nn is even and its binary expansion of nn contains no consecutive digits equal to 11, then TCs(M)=sn\operatorname{TC}_{s}(M)=sn for any s3s\geq 3.

  • (b)

    If n=2r+12n=2^{r+1}-2, then TCs(M)=sn\operatorname{TC}_{s}(M)=sn for any s2r+11s\geq 2^{r+1}-1.

By using Davis’ computations in combination with Proposition 3.3, we can also state:

Corollary 4.4.

Let MM be an orientable connected closed manifold satisfying the conditions n=dim(M)=cat(M)n=\dim(M)=\operatorname{\mathrm{cat}}(M) and π1(M)=2\pi_{1}(M)={\mathbb{Z}}_{2}. If n1mod4n\equiv 1\bmod 4, then TCs(M)=sn1\operatorname{TC}_{s}(M)=sn-1 for ss satisfying the inequality (4.1).

Proof.

In this case mn,s=1m_{n,s}=1 so that TCs(M)sn1\operatorname{TC}_{s}(M)\geq sn-1 for ss satisfying the inequality (4.1). The other direction follows from Proposition 3.3. ∎

Remark 4.5.

Observe that the orientability hypothesis, together with the condition dim(M)=cat(M)\dim(M)=\operatorname{\mathrm{cat}}(M), implies that nn is odd. Indeed, if nn were even, the image of the fundamental class of MM by γ:Hn(M)Hn(B2)=0\gamma_{*}:H_{n}(M)\to H_{n}(B{\mathbb{Z}}_{2})=0 would vanish. But this fact would force, by Poincaré duality, the nnth power of the Berstein-Schwarz class of MM to vanish, contradicting the equality dim(M)=cat(M)\dim(M)=\operatorname{\mathrm{cat}}(M).

4.2. Odd dimensional manifolds with π1(M)=p\pi_{1}(M)={\mathbb{Z}}_{p} and cat(M)=dimM\operatorname{\mathrm{cat}}(M)=\dim M

Throughout this section we consider a prime p3p\geq 3. Recall that the classifying space BpB{\mathbb{Z}}_{p} can be identified with the infinite dimensional lens space LpL_{p}^{\infty}. In order to have an analogue of Proposition 4.2, we first note the following result:

Lemma 4.6.

Let MM be an orientable connected closed (2n+1)(2n+1)-manifold satisfying the conditions cat(M)=2n+1\operatorname{\mathrm{cat}}(M)=2n+1 and π1(M)=p\pi_{1}(M)={\mathbb{Z}}_{p}. If γ:MBp\gamma\colon M\to B{\mathbb{Z}}_{p} is a classifying map, then γ([M])H2n+1(p;p)\gamma_{\ast}([M])\in H_{2n+1}({\mathbb{Z}}_{p};{\mathbb{Z}}_{p}) is non-zero.

Proof.

Considering the Berstein–Schwarz class 𝔟MH1(M;I(π))\mathfrak{b}_{M}\in H^{1}(M;I(\pi)), we have cat(M)=dim(M)=2n+1\operatorname{\mathrm{cat}}(M)=\dim(M)=2n+1 if and only if 𝔟M2n+10\mathfrak{b}_{M}^{2n+1}\neq 0. By Poincaré duality, the second statement is equivalent to 𝔟M2n+1([M])0\mathfrak{b}_{M}^{2n+1}([M])\neq 0. Taking cap products we obtain

[M]𝔟M2n+1=[M]γ(𝔟p2n+1)0.[M]\cap\mathfrak{b}_{M}^{2n+1}=[M]\cap\gamma^{\ast}(\mathfrak{b}^{2n+1}_{{\mathbb{Z}}_{p}})\neq 0.

Since γ\gamma induces an isomorphism at the level of fundamental groups, naturality of the cap-products yields

γ([M])𝔟p2n+10.\gamma_{\ast}([M])\cap\mathfrak{b}^{2n+1}_{{\mathbb{Z}}_{p}}\neq 0.

Hence, γ([M])0\gamma_{\ast}([M])\neq 0 in H2n+1(p;)=pH_{2n+1}({\mathbb{Z}}_{p};{\mathbb{Z}})={\mathbb{Z}}_{p}. Since H2n+1(p;)=H2n+1(p;p)=pH_{2n+1}({\mathbb{Z}}_{p};{\mathbb{Z}})=H_{2n+1}({\mathbb{Z}}_{p};{\mathbb{Z}}_{p})={\mathbb{Z}}_{p} we can conclude that γ([M])0\gamma_{\ast}([M])\neq 0 in H2n+1(p;p)H_{2n+1}({\mathbb{Z}}_{p};{\mathbb{Z}}_{p}). ∎

Theorem 4.7.

Let MM be a closed orientable (2n+1)(2n+1)-manifold with cat(M)=2n+1\operatorname{\mathrm{cat}}(M)=2n+1 and π1(M)=p\pi_{1}(M)={\mathbb{Z}}_{p} where p3p\geq 3 is a prime. Then, TCs(M)zclsw¯(M;p)zclsw¯(Lp2n+1;p)\operatorname{TC}_{s}(M)\geq\operatorname{\operatorname{\mathrm{zcl}}^{\underline{w}}_{s}}(M;{\mathbb{Z}}_{p})\geq\operatorname{\operatorname{\mathrm{zcl}}^{\underline{w}}_{s}}(L_{p}^{2n+1};{\mathbb{Z}}_{p}).

Proof.

Let γ:MBp\gamma\colon M\to B{\mathbb{Z}}_{p} be a classifying map. We have a commutative triangle

(4.2) M{M}L2n+1{L^{2n+1}}Lp{L^{\infty}_{p}}ϕ\scriptstyle{\phi}γ\scriptstyle{\gamma}

where the inclusion is simply the (2n+1)(2n+1)-skeleton of the infinite dimensional lens space LpBpL^{\infty}_{p}\simeq B{\mathbb{Z}}_{p}. Since cat(M)=2n+1\operatorname{\mathrm{cat}}(M)=2n+1, we know by the previous lemma that γ([M])0\gamma_{\ast}([M])\neq 0 in H2n+1(p;p)H_{2n+1}({\mathbb{Z}}_{p};{\mathbb{Z}}_{p}). Consequently ϕ([M])0\phi_{\ast}([M])\neq 0 in H2n+1(L2n+1;p)H_{2n+1}(L^{2n+1};{\mathbb{Z}}_{p}). Recall that

H(Lp2n+1;p)=p[x,y]/(yn+1,x2)H^{\ast}(L^{2n+1}_{p};{\mathbb{Z}}_{p})={\mathbb{Z}}_{p}[x,y]/(y^{n+1},x^{2})

where |x|=1|x|=1, |y|=2|y|=2. In particular, H2n+1(Lp2n+1;p)=pxynH^{2n+1}(L^{2n+1}_{p};{\mathbb{Z}}_{p})={\mathbb{Z}}_{p}xy^{n}. We first check that ϕ:H(Lp2n+1;p)H(M;p)\phi^{*}:H^{*}(L_{p}^{2n+1};{\mathbb{Z}}_{p})\to H^{*}(M;{\mathbb{Z}}_{p}) is a monomorphism. It suffices to show that ϕ(x)\phi^{\ast}(x) and ϕ(y),,ϕ(yn)\phi^{\ast}(y),\dots,\phi^{\ast}(y^{n}) are non-trivial in H(M;p)H^{\ast}(M;{\mathbb{Z}}_{p}). Consider the following cap-product diagram:

H2n+1(M;)H2n+1(M;p)\textstyle{H_{2n+1}(M;{\mathbb{Z}})\otimes H^{2n+1}(M;{\mathbb{Z}}_{p})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ϕ\scriptstyle{\phi_{*}}\scriptstyle{\cap}\scriptstyle{\cong}H0(M;p)=p\textstyle{H_{0}(M;{\mathbb{Z}}_{p})={\mathbb{Z}}_{p}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}=\scriptstyle{=}H2n+1(Lp2n+1;p)H2n+1(Lp2n+1;p)\textstyle{H_{2n+1}(L_{p}^{2n+1};{\mathbb{Z}}_{p})\otimes H^{2n+1}(L_{p}^{2n+1};{\mathbb{Z}}_{p})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ϕ\scriptstyle{\phi^{*}}\scriptstyle{\cap}\scriptstyle{\cong}H0(Lp2n+1;p)=p\textstyle{H_{0}(L_{p}^{2n+1};{\mathbb{Z}}_{p})={\mathbb{Z}}_{p}}

Both horizontal morphisms are isomorphisms by Poincaré duality. As ϕ([M])0\phi_{\ast}([M])\neq 0 in H2n+1(L2n+1;p)H_{2n+1}(L^{2n+1};{\mathbb{Z}}_{p}), ϕ([M]\phi_{*}([M] is not divisible by pp in H2n+1(L2n+1;)=H_{2n+1}(L^{2n+1};{\mathbb{Z}})={\mathbb{Z}}. Consequently, ϕ([M])xyn0\phi_{*}([M])\cap xy^{n}\neq 0 and using the diagram we have that [M]ϕ(xyn)0[M]\cap\phi^{\ast}(xy^{n})\neq 0. We thus have ϕ(xyn)0\phi^{*}(xy^{n})\neq 0 in H(M;p)H^{*}(M;{\mathbb{Z}}_{p}) and hence ϕ(x)\phi^{\ast}(x) and ϕ(yi)\phi^{\ast}(y^{i}) for i=1,,ni=1,\dots,n are non-zero. Therefore ϕ\phi^{*} is a monomorphism. By Künneth formula, we obtain that (ϕs)(\phi^{s})^{*} is also a monomorphism. Let u0𝐙s(Lp2n+1;p)u\neq 0\in\mathbf{Z}_{s}(L_{p}^{2n+1};{\mathbb{Z}}_{p}). Then (ϕs)(u)𝐙s(M;p)(\phi^{s})^{*}(u)\in\mathbf{Z}_{s}(M;{\mathbb{Z}}_{p}) and (ϕs)(u)0(\phi^{s})^{*}(u)\neq 0. Consequently wgts((ϕs)(u))wgts(u)\operatorname{\mathrm{wgt}}_{s}((\phi^{s})^{*}(u))\geq\operatorname{\mathrm{wgt}}_{s}(u) and the results follows by definition of zclsw¯\operatorname{\operatorname{\mathrm{zcl}}^{\underline{w}}_{s}}. ∎

There are extensive computations of zclsw¯(Lpn;p)\operatorname{\operatorname{\mathrm{zcl}}^{\underline{w}}_{s}}(L^{n}_{p};{\mathbb{Z}}_{p}), see [14, §5] and [8, Section 5]. By using the previous theorem, we can use the information coming from these computations for a larger class of manifolds.

Corollary 4.8.

Let s2s\geq 2 and let MM be a closed orientable (2n+1)(2n+1)-manifold with cat(M)=2n+1\operatorname{\mathrm{cat}}(M)=2n+1 and π1(M)=p\pi_{1}(M)={\mathbb{Z}}_{p} where p3p\geq 3 is prime. Then,

TCs(M){s(++1)1if s is even;(s1)(+)+s+2n1if s is odd\operatorname{TC}_{s}(M)\geq\begin{cases}s\cdot(\ell+\ell^{\prime}+1)-1&\text{if $s$ is even};\\ (s-1)\cdot(\ell+\ell^{\prime})+s+2n-1&\text{if $s$ is odd}\end{cases}

where 0,n0\leq\ell,\ell^{\prime}\leq n are any integers such that mm does not divide (+)s/2\binom{\ell+\ell^{\prime}}{\ell}^{\lfloor s/2\rfloor}.

Proof.

Here we apply the computation [8, Theorem 5.2] and Theorem 4.7. ∎

In many situations, for example if ss is much larger than the dimension of MM, we obtain an exact computation.

Corollary 4.9.

Let s2s\geq 2 and let MM be a closed orientable (2n+1)(2n+1)-manifold with cat(M)=2n+1\operatorname{\mathrm{cat}}(M)=2n+1 and π1(M)=p\pi_{1}(M)={\mathbb{Z}}_{p} where p3p\geq 3 is a prime. If ss does not divide (2nn)s/2\binom{2n}{n}^{\lfloor s/2\rfloor}, then

TCs(M)=s(2n+1)1.\operatorname{TC}_{s}(M)=s(2n+1)-1.
Proof.

The lower bound follows from Corollary 4.8 (see also [8, Theorem 5.3]) and the upper bound is Proposition 3.3. ∎

5. Some calculations in the non-orientable case

We now address the (non-)maximality of TCs(M)\operatorname{TC}_{s}(M) for non-orientable manifolds having π1(M)=2\pi_{1}(M)={\mathbb{Z}}_{2}. The case s=2s=2 is well understood ([7, Theorem 1]), so we assume s3s\geq 3 from now on. For such cases the non-maximality of TCs(M)\operatorname{TC}_{s}(M) demands further restrictions on ss. The aim of this section is to establish the following result.

Theorem 5.1.

Let MM be a non-orientable nn-dimensional manifold with π1(M)=2\pi_{1}(M)=\mathbb{Z}_{2} and n=2r+12n=2^{r+1}-2. Then, for any even ss no greater than 2r+122^{r+1}-2, we have TCs(M)<sn\operatorname{TC}_{s}(M)<sn.

By Corollary 4.3(b), we know that, for n=2r+12n=2^{r+1}-2, TCs(M)=sn\operatorname{TC}_{s}(M)=sn for s2r+11s\geq 2^{r+1}-1, so that in this case the upper limiting restriction on ss in Theorem 5.1 is in fact sharp and we have:

Corollary 5.2.

If the manifold MM in Theorem 5.1 has catM=n\operatorname{\mathrm{cat}}{M}=n, then TCs(M)=sn1\operatorname{TC}_{s}(M)=sn-1 for s=2r+12s=2^{r+1}-2 and TCs(M)=sn\operatorname{TC}_{s}(M)=sn for s2r+11s\geq 2^{r+1}-1.

Proof.

The equality TCs(M)=sn1\operatorname{TC}_{s}(M)=sn-1 for s=2r+12s=2^{r+1}-2 follows from m2r+12,2r+12=1m_{2^{r+1}-2,2^{r+1}-2}=1 (see Example 4.1(b)), Proposition 4.2 and Theorem 5.1. ∎

Corollary 5.2 should be compared to the fact that TCs(P2r)\operatorname{TC}_{s}(P^{2^{r}}) is maximal for s3s\geq 3, but TC2(P2r)=Imm(P2r)=2r+11\operatorname{TC}_{2}(P^{2^{r}})=\operatorname{Imm}(P^{2^{r}})=2^{r+1}-1 ([15]). Worth noting is the fact that the case M=P6M=P^{6} in Corollary 5.2 (with r=2r=2) upgrades the observation in [5, (7.4)] that δ6(6)1\delta_{6}(6)\leq 1 to an equality, giving evidence for what would be regular behavior of the higher topological complexity of projective spaces PmP^{m} with m=2a+2a+1m=2^{a}+2^{a+1}.

Suitable analogues of Theorem 5.1 should hold for more general values of nn, but the complexity of calculations seems to be a major obstacle towards obtaining corresponding proofs.

We now start working towards the proof of Theorem 5.1. From now on π:=2\pi:={\mathbb{Z}}_{2} and s=2σs=2\sigma with 1σ2r1=n/21\leq\sigma\leq 2^{r}-1=n/2. Set ^:=~(~)σ1\widehat{{\mathbb{Z}}}:=\widetilde{{\mathbb{Z}}}\otimes({\mathbb{Z}}\otimes\widetilde{{\mathbb{Z}}})^{\sigma-1}, the [πs1]{\mathbb{Z}}[\pi^{s-1}]-module of Remark 2.5. By Proposition 2.3, it suffices to establish the triviality of

(5.1) sχ(𝔪×s)Hsn(πs1;^)^{s}\chi_{*}(\mathfrak{m}^{\times s})\in H_{sn}(\pi^{s-1};\widehat{\mathbb{Z}})

where

(5.2) sχ:H(πs;~s)H(πs1;^).^{s}\chi_{*}\colon H_{*}(\pi^{s};\widetilde{\mathbb{Z}}^{\otimes s})\to H_{*}(\pi^{s-1};\widehat{\mathbb{Z}}).

As in §3, our starting point is the free [q]{\mathbb{Z}}[{\mathbb{Z}}_{q}]-resolution (3.2) of {\mathbb{Z}} with q=2q=2. Recall that [k][k] denotes the generator of degree kk. In addition to the chain complex C(π)C_{\bullet}(\pi) of (3.3), we will also need the complex C~(π)\widetilde{C}_{\bullet}(\pi),

(5.3) 2[2k]0[2k1]20[1]2[0],\lx@xy@svg{\hbox{\raise 2.5pt\hbox{\kern 3.0pt\hbox{\ignorespaces\ignorespaces\ignorespaces\hbox{\vtop{\kern 0.0pt\offinterlineskip\halign{\entry@#!@&&\entry@@#!@\cr&&&&&\crcr}}}\ignorespaces{\hbox{\kern-3.0pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise-2.5pt\hbox{$\textstyle{\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces\ignorespaces\ignorespaces{\hbox{\kern 14.20659pt\raise 5.25555pt\hbox{{}\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\hbox{\hbox{\kern 0.0pt\raise-2.25555pt\hbox{$\scriptstyle{-2}$}}}\kern 3.0pt}}}}}}\ignorespaces{\hbox{\kern 27.0pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}{\hbox{\kern 27.0pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise-2.5pt\hbox{$\textstyle{{\mathbb{Z}}[2k]\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces\ignorespaces\ignorespaces{\hbox{\kern 65.38889pt\raise 5.25555pt\hbox{{}\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\hbox{\hbox{\kern 0.0pt\raise-2.25555pt\hbox{$\scriptstyle{0}$}}}\kern 3.0pt}}}}}}\ignorespaces{\hbox{\kern 79.18753pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}{\hbox{\kern 79.18753pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise-2.5pt\hbox{$\textstyle{{\mathbb{Z}}[2k-1]\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces\ignorespaces\ignorespaces{\hbox{\kern 118.94615pt\raise 5.25555pt\hbox{{}\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\hbox{\hbox{\kern 0.0pt\raise-2.25555pt\hbox{$\scriptstyle{-2}$}}}\kern 3.0pt}}}}}}\ignorespaces{\hbox{\kern 144.15274pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}{\hbox{\kern 144.15274pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise-2.5pt\hbox{$\textstyle{\cdots\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces\ignorespaces\ignorespaces{\hbox{\kern 166.95135pt\raise 5.25555pt\hbox{{}\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\hbox{\hbox{\kern 0.0pt\raise-2.25555pt\hbox{$\scriptstyle{0}$}}}\kern 3.0pt}}}}}}\ignorespaces{\hbox{\kern 181.65274pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}{\hbox{\kern 181.65274pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise-2.5pt\hbox{$\textstyle{{\mathbb{Z}}[1]\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces\ignorespaces\ignorespaces{\hbox{\kern 209.97914pt\raise 5.25555pt\hbox{{}\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\hbox{\hbox{\kern 0.0pt\raise-2.25555pt\hbox{$\scriptstyle{-2}$}}}\kern 3.0pt}}}}}}\ignorespaces{\hbox{\kern 228.31943pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}{\hbox{\kern 228.31943pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise-2.5pt\hbox{$\textstyle{{\mathbb{Z}}[0]}$}}}}}}}\ignorespaces}}}}\ignorespaces,

obtained by tensoring (3.2) with ~\widetilde{{\mathbb{Z}}} over π\pi. Abusing notation, we continue using [k][k] for the generators of both C(π)C_{\bullet}(\pi) and C~(π)\widetilde{C}_{\bullet}(\pi).

The homology groups in (5.2) can be computed from the complexes C~(π)s\widetilde{C}_{\bullet}(\pi)^{\otimes s} and

(5.4) 𝒟:=C~(π)(C(π)C~(π))σ1.\mathcal{D}_{\bullet}:=\widetilde{C}_{\bullet}(\pi)\otimes({C}_{\bullet}(\pi)\otimes\widetilde{C}_{\bullet}(\pi))^{\sigma-1}.

In both cases, we will use the shorthand [i1,,i][i_{1},\ldots,i_{\ell}] for a tensor product [i1][i][i_{1}]\otimes\cdots\otimes[i_{\ell}]. The Künneth formula and the fact that the homology of C~(π)\widetilde{C}_{\bullet}(\pi) is 2-torsion (in all degrees) gives:

Lemma 5.3.

The element χs(𝔪×s){}^{s}\chi_{*}(\mathfrak{m}^{\times s}) in (5.1) is torsion. Indeed, both groups in (5.2) are torsion.

Let \mathcal{H}_{\bullet} denote the quotient of 𝒟\mathcal{D}_{\bullet} resulting from killing all boundaries, and consider the obvious monomorphism ι:H(πs1;^)\iota:H_{\bullet}(\pi^{s-1};\widehat{\mathbb{Z}})\hookrightarrow\mathcal{H}_{\bullet}. The triviality of the element in (5.1) follows from Lemma 5.3 and the following key result, whose proof is addressed in the rest of the section through a direct analysis of (5.1) and (5.2).

Proposition 5.4.

The class ι(sχ(𝔪×s))\iota(^{s}\chi_{*}(\mathfrak{m}^{\times s})) is an element of the torsion-free (graded) subgroup of \mathcal{H}_{\bullet}.

In computing the homology groups in (5.2) using the complexes C~(π)s\widetilde{C}_{\bullet}(\pi)^{\otimes s} and 𝒟\mathcal{D}_{\bullet}, we will use π~\widetilde{\pi} for a factor where C~(π)\widetilde{C}_{\bullet}(\pi) is meant to be taken, reserving the notation π\pi for factors where C(π)C_{\bullet}(\pi) is meant to be taken. For instance, the diagonal morphism Δ:ππ×π\Delta:\pi\to\pi\times\pi and the group-multiplication morphism μ:π×ππ\mu:\pi\times\pi\to\pi extend to morphisms Δ:π~π~×π\Delta:\widetilde{\pi}\to\widetilde{\pi}\times\pi, Δ:π~π×π~\Delta:\widetilde{\pi}\to\pi\times\widetilde{\pi} and μ:π~×π~π~\mu:\widetilde{\pi}\times\widetilde{\pi}\to\widetilde{\pi} that are compatible with the implied module structures. In these terms, since π=2=vv2=1\pi={\mathbb{Z}}_{2}=\langle v\mid v^{2}=1\rangle in the present case, the inversion morphism plays no role and the map χs{}^{s}\chi factors as

(5.5) π~s1×(Δ×Δ)σ1×1\displaystyle\widetilde{\pi}^{s}\stackrel{{\scriptstyle 1\times(\Delta\times\Delta)^{\sigma-1}\times 1}}{{-\!\!\!-\!\!\!-\!\!\!-\!\!\!-\!\!\!-\!\!\!-\!\!\!-\!\!\!-\!\!\!-\!\!\!\longrightarrow}}{} π~×(π~×π×π×π~)σ1×π~\displaystyle\widetilde{\pi}\times(\widetilde{\pi}\times\pi\times\pi\times\widetilde{\pi})^{\sigma-1}\times\widetilde{\pi}
=\displaystyle= (π~×π~)×(π×π×π~×π~)σ1μ×(μ×μ)σ1π~×(π×π~)σ1.\displaystyle(\widetilde{\pi}\times\widetilde{\pi})\times(\pi\times\pi\times\widetilde{\pi}\times\widetilde{\pi})^{\sigma-1}\stackrel{{\scriptstyle\mu\times(\mu\times\mu)^{\sigma-1}}}{{-\!\!\!-\!\!\!-\!\!\!-\!\!\!-\!\!\!-\!\!\!-\!\!\!\longrightarrow}}\widetilde{\pi}\times(\pi\times\widetilde{\pi})^{\sigma-1}.

Recalling that Bi,jB_{i,j} denotes the binomial coefficient (i+ji)\binom{\,i+j\,}{i}, the formulæ of Lemma 3.2 (b)(b) written with the shorthand in use in this section read

(5.6) [2i,2j]\displaystyle{}[2i,2j] Bi,j[2(i+j)];\displaystyle\mapsto B_{i,j}\,[2(i+j)];
[2i,2j+1]\displaystyle{}[2i,2j+1] Bi,j[2(i+j)+1];\displaystyle\mapsto B_{i,j}\,[2(i+j)+1];
[2i+1,2j]\displaystyle{}[2i+1,2j] Bi,j[2(i+j)+1];\displaystyle\mapsto B_{i,j}\,[2(i+j)+1];
[2i+1,2j+1]\displaystyle{}[2i+1,2j+1] 0.\displaystyle\mapsto 0.

We note also that, since π=2\pi={\mathbb{Z}}_{2}, the formula of Lemma 3.2(a) giving Δ\Delta on generators at the level of resolutions can be written

(5.7) [k]p+q=k[p]vodd(p)[q],[k]\to\sum_{p+q=k}[p]\otimes v^{\mathrm{odd}(p)}\hskip-1.70717pt\cdot[q],

where vv generates π\pi and odd(p)=1{\mathrm{odd}(p)}=1 if pp is odd and 0 otherwise.

The class 𝔪H(π;~)\mathfrak{m}\in H_{*}(\pi;\widetilde{\mathbb{Z}}) is either trivial or, else, represented by the cycle [n][n] in (5.3) —recall nn is even. For the purposes of proving Theorem 5.1, we may safely assume the latter possibility. Then, 𝔪×s\mathfrak{m}^{\times s} is represented in C~(π)s\widetilde{C}_{\bullet}(\pi)^{\otimes s} by the corresponding tensor product [n,n,,n][n,n,\ldots,n]. We chase the latter element under the first map of the composite (5.5) to get, in view of (5.7),

[n~,n~,,n~][n~]((p+q=n[p~]vodd(p)[q])(p+q=n[p]vodd(p)[q~]))σ1[n~].\displaystyle[\widetilde{n},\widetilde{n},\ldots,\widetilde{n}]\mapsto[\widetilde{n}]\otimes\left(\hskip-2.84526pt\left(\,\sum_{p+q=n}[\widetilde{p}]\otimes v^{\text{odd}(p)}\cdot[q]\right)\otimes\left(\,\sum_{p+q=n}[p]\otimes v^{\text{odd}(p)}\cdot[\widetilde{q}]\right)\hskip-2.84526pt\right)^{\hskip-2.84526pt\otimes\sigma-1}\hskip-5.69054pt\otimes[\widetilde{n}].

Note that in the latter expression we are extending in the obvious way the convention above regarding the use of π\pi and π~\widetilde{\pi}. Then, after tensoring with the needed coefficients (thus dropping the \sim indicators), this becomes

[n,n,,n]\displaystyle[{n},{n},\ldots,{n}]\mapsto [n]((p+q=n[p,q])(p+q=n(1)p[p,q]))σ1[n]\displaystyle[{n}]\otimes\left(\hskip-2.84526pt\left(\,\sum_{p+q=n}[p,q]\right)\otimes\left(\,\sum_{p+q=n}(-1)^{p}[p,q]\right)\hskip-2.84526pt\right)^{\hskip-2.84526pt\otimes\sigma-1}\hskip-5.69054pt\otimes[{n}]
=\displaystyle= pi+qi=n1is2(1)j=1σ1p2j[n,p1,q1,p2,q2,,ps2,qs2,n].\displaystyle\sum_{\genfrac{}{}{0.0pt}{}{p_{i}+q_{i}=n}{1\leq i\leq s-2}}(-1)^{{}^{\sum_{j=1}^{\sigma-1}p_{2j}}}[n,p_{1},q_{1},p_{2},q_{2},\cdots,p_{s-2},q_{s-2},n].

Since nn is even, the parity of each pip_{i} agrees with the one of the corresponding qiq_{i}, so the last expression can be rewritten as

(1)j=1σ1δ2j[n,2p1+δ1,2q1+δ1,2p2+δ2,2q2+δ2,,2ps2+δs2,2qs2+δs2,n],\displaystyle\sum(-1)^{{}^{\sum_{j=1}^{\sigma-1}\delta_{2j}}}[n,2p_{1}+\delta_{1},2q_{1}+\delta_{1},2p_{2}+\delta_{2},2q_{2}+\delta_{2},\cdots\hskip-0.85358pt,2p_{s-2}+\delta_{s-2},2q_{s-2}+\delta_{s-2},n],

where the sum now runs over 1is21\leq i\leq s-2, δi{0,1}\delta_{i}\in\{0,1\} and pi+qi=n/2δip_{i}+q_{i}=n/2-\delta_{i}. Using the formulae (5.6), we finally obtain the image of [n,,n][n,\ldots,n] under the entire composition in (5.5). This image may be expressed as

(5.8) [n,,n](1)j=1σ1δ2jBn2,p1Bq1,p2Bqs3,ps2Bqs2,n2𝐆,[n,\ldots,n]\mapsto\sum(-1)^{{}^{\sum_{j=1}^{\sigma-1}\delta_{2j}}}B_{\frac{n}{2},p_{1}}B_{q_{1},p_{2}}\cdots B_{q_{s-3},p_{s-2}}B_{q_{s-2},\frac{n}{2}}\cdot\mathbf{G},

where

𝐆=[n+2p1+δ1,2(q1+p2)+δ1+δ2,,2(qs3+ps2)+δs3+δs2,n+2qs2+δs2]\mathbf{G}=[n+2p_{1}+\delta_{1},2(q_{1}+p_{2})+\delta_{1}+\delta_{2},\ldots,2(q_{s-3}+p_{s-2})+\delta_{s-3}+\delta_{s-2},n+2q_{s-2}+\delta_{s-2}]

and the sum runs over the same indices as above, except now that no two consecutive δj\delta_{j} and δj+1\delta_{j+1} can simultaneously equal 1, in view of (5.6). In what follows we set m=n/2m={n}/{2}.

Having described a cycle representing the obstruction in (5.1), we next spell out the complex (5.4) where it lies. Degreewise, 𝒟\mathcal{D}_{\bullet} is \mathbb{Z}-free with basis given by elements [u1,v1,,uσ1,vσ1,uσ][u_{1},v_{1},\ldots,u_{\sigma-1},v_{\sigma-1},u_{\sigma}] for non-negative integers uiu_{i} and viv_{i}, and with differential

[u1,\displaystyle\partial[u_{1}, v1,,uσ1,vσ1,uσ]=\displaystyle v_{1},\ldots,u_{\sigma-1},v_{\sigma-1},u_{\sigma}]=
=\displaystyle= 2odd(u1)[u11,v1,u2,v2,,uσ1,vσ1,uσ]\displaystyle-2\hskip 2.27621pt\text{odd}\hskip 1.13809pt(u_{1})[u_{1}-1,v_{1},u_{2},v_{2},\ldots,u_{\sigma-1},v_{\sigma-1},u_{\sigma}]
+(1)u12even(v1)[u1,v11,u2,v2,,uσ1,vσ1,uσ]\displaystyle\quad+(-1)^{u_{1}}\hskip 0.85358pt2\hskip 2.27621pt\text{even}\hskip 1.13809pt(v_{1})[u_{1},v_{1}-1,u_{2},v_{2},\ldots,u_{\sigma-1},v_{\sigma-1},u_{\sigma}]
(1)u1+v12odd(u2)[u1,v1,u21,v2,,uσ1,vσ1,uσ]\displaystyle\quad-(-1)^{u_{1}+v_{1}}\hskip 0.85358pt2\hskip 2.27621pt\text{odd}\hskip 1.13809pt(u_{2})[u_{1},v_{1},u_{2}-1,v_{2},\ldots,u_{\sigma-1},v_{\sigma-1},u_{\sigma}]
(5.9) +(1)u1+v1+u22even(v2)[u1,v1,u2,v21,,uσ1,vσ1,uσ]\displaystyle\quad+(-1)^{u_{1}+v_{1}+u_{2}}\hskip 0.85358pt2\hskip 2.27621pt\text{even}\hskip 1.13809pt(v_{2})[u_{1},v_{1},u_{2},v_{2}-1,\ldots,u_{\sigma-1},v_{\sigma-1},u_{\sigma}]
±\displaystyle\quad\pm\cdots
(1)u1+v1++uσ2+vσ22odd(uσ1)[u1,v1,u2,v2,,uσ11,vσ1,uσ]\displaystyle\quad-(-1)^{u_{1}+v_{1}+\cdots+u_{\sigma-2}+v_{\sigma-2}}\hskip 0.85358pt2\hskip 2.27621pt\text{odd}\hskip 1.13809pt(u_{\sigma-1})[u_{1},v_{1},u_{2},v_{2},\ldots,u_{\sigma-1}-1,v_{\sigma-1},u_{\sigma}]
+(1)u1+v1++uσ2+vσ2+uσ12even(vσ1)[u1,v1,u2,v2,,uσ1,vσ11,uσ]\displaystyle\quad+(-1)^{u_{1}+v_{1}+\cdots+u_{\sigma-2}+v_{\sigma-2}+u_{\sigma-1}}\hskip 0.85358pt2\hskip 2.27621pt\text{even}\hskip 1.13809pt(v_{\sigma-1})[u_{1},v_{1},u_{2},v_{2},\ldots,u_{\sigma-1},v_{\sigma-1}-1,u_{\sigma}]
(1)u1+v1++uσ1+vσ12odd(uσ)[u1,v1,u2,v2,,uσ1,vσ1,uσ1],\displaystyle\quad-(-1)^{u_{1}+v_{1}+\cdots+u_{\sigma-1}+v_{\sigma-1}}\hskip 0.85358pt2\hskip 2.27621pt\text{odd}\hskip 1.13809pt(u_{\sigma})[u_{1},v_{1},u_{2},v_{2},\ldots,u_{\sigma-1},v_{\sigma-1},u_{\sigma}-1],

where a basis element with a negative entry is meant to be interpreted as zero.

All the information needed to prove Proposition 5.4 is of course contained in (5.8) and (5.9). The following constructions are meant to organize a proof argument.

Definition 5.5.

For a positive integer kk, let p(k)p(k) denote the set of binary positions where the binary expansion of kk has digit 1. For instance, p(5=4+1)={0,2}p(5=4+1)=\{0,2\} and p(42=32+8+2)={1,3,5}p(42=32+8+2)=\{1,3,5\}.

A standard well known fact is:

Lemma 5.6.

A binomial coefficient Ba,bB_{a,b} is even if and only if p(a)p(b)p(a)\cap p(b)\neq\varnothing.

The following result is the only place where the special assumptions in Theorem 5.1 (i.e., s=2σn=2m=2r+12s=2\sigma\leq n=2m=2^{r+1}-2 with r1r\geq 1) are needed.

Proposition 5.7.

Any coefficient

(5.10) Bm,p1Bq1,p2Bqs3,ps2Bqs2,mB_{m,p_{1}}B_{q_{1},p_{2}}\cdots B_{q_{s-3},p_{s-2}}B_{q_{s-2},m}

in (5.8) with p1+q1=mp_{1}+q_{1}=m (i.e., with δ1=0\delta_{1}=0) is even.

Proof.

We use without further notice the fact coming from Lemma 5.6 that any binomial coefficient Bmj,iB_{m-j,i} is even whenever 0j<im0\leq j<i\leq m. Recall m=2r1m=2^{r}-1 and

(5.11) s=2σ2r+12s=2\sigma\leq 2^{r+1}-2

with r1r\geq 1. Assume for a contradiction that some coefficient (5.10) is odd (i.e., that all of its binomial-coefficient factors are odd) and has δ1=0\delta_{1}=0. Recall the forced conditions

  1. (1)

    pi+qi=mδip_{i}+q_{i}=m-\delta_{i} with pi,qi0p_{i},q_{i}\geq 0 and δi{0,1}\delta_{i}\in\{0,1\};

  2. (2)

    1i<s21\leq i<s-2 and δi=1\delta_{i}=1 implies δi+1=0\delta_{i+1}=0,

for 1is21\leq i\leq s-2. Let 2i1<i2<<iks22\leq i_{1}<i_{2}<\cdots<i_{k}\leq s-2 be all the indices jj (if any) with δj=1\delta_{j}=1. Note that

(5.12) 0kσ1,0\leq k\leq\sigma-1,

in view of (2).

The coefficient Bm,p1B_{m,p_{1}} is odd by hypothesis, so p1=0p_{1}=0 and q1=mq_{1}=m —the latter equality holds in view of (1) since δ1=0\delta_{1}=0. Actually, the same argument can be used iteratively for 1j<i11\leq j<i_{1} (so δj=0\delta_{j}=0) with the binomial coefficients Bqj1,pjB_{q_{j-1},p_{j}} (e.g. q0:=mq_{0}:=m) to show that

pj=0p_{j}=0   and   qj=mq_{j}=m.

Next, since m=qi11m=q_{i_{1}-1}, Bqi11,pi1B_{q_{i_{1}-1},p_{i_{1}}} is odd and δi1=1\delta_{i_{1}}=1, we get

pi1=0p_{i_{1}}=0   and   qi1=m1q_{i_{1}}=m-1,

and now the process repeats with a slight adjustment. For starters, qi1=m1q_{i_{1}}=m-1, Bqi1,pi1+1B_{q_{i_{1}},p_{i_{1}+1}} is odd and δi1+1=0\delta_{i_{1}+1}=0 is forced by (2), so that pi1+11p_{i_{1}+1}\leq 1 and qi1+1m1q_{i_{1}+1}\geq m-1. We then iterate the latter argument: For i1<j<i2i_{1}<j<i_{2}, the assumption δj=0\delta_{j}=0 and the fact that Bqj1,pjB_{q_{j-1},p_{j}} is odd with qj1m1q_{j-1}\geq m-1 yield

pj1p_{j}\leq 1   and   qjm1q_{j}\geq m-1.

Of course, the last two inequalities now hold for all 1j<i21\leq j<i_{2}. The next round of iterations start with the fact that Bqi21,pi2B_{q_{i_{2}-1},p_{i_{2}}} is odd with qi21m1q_{i_{2}-1}\geq m-1 and δi2=1\delta_{i_{2}}=1, to get

pi21p_{i_{2}}\leq 1   and   qi2m2q_{i_{2}}\geq m-2,

and the process has a corresponding new obvious adjustment to yield

pj2p_{j}\leq 2   and   qjm2q_{j}\geq m-2,

for 1j<i31\leq j<i_{3}, whereas

pi32p_{i_{3}}\leq 2   and   qi3m3q_{i_{3}}\geq m-3.

Just before the last adjustment we get

pjk1p_{j}\leq k-1   and   qjmk+1q_{j}\geq m-k+1,

for 1j<ik1\leq j<i_{k}, whereas

pikk1p_{i_{k}}\leq k-1   and   qikmkq_{i_{k}}\geq m-k.

However, after this point the conditions pjkp_{j}\leq k and qjmkq_{j}\geq m-k are kept for all js2j\leq s-2. In particular, qs2mk=2r1k1q_{s-2}\geq m-k=2^{r}-1-k\geq 1, in view of (5.11) and (5.12). But then the final factor Bqs2,mB_{q_{s-2},m} of (5.10) is even, a contradiction. ∎

Proof of Proposition 5.4.

The right hand-side of (5.9) yields the defining relations in \mathcal{H}_{\bullet}. Namely, for each tuple (u1,v1,,uσ1,vσ1,uσ)(u_{1},v_{1},\ldots,u_{\sigma-1},v_{\sigma-1},u_{\sigma}) of non-negative integers there is a defining relation

(5.13) 0=U1+V1+U2+V2++Uσ1+Vσ1+Uσ0=U_{1}+V_{1}+U_{2}+V_{2}+\cdots+U_{\sigma-1}+V_{\sigma-1}+U_{\sigma}

where

Ui:=\displaystyle U_{i}:= (1)pi 2odd(ui)[u1,v1,,ui1,vi1,ui1,vi,ui+1,vi+1,,uσ1,vσ1,uσ],\displaystyle(-1)^{p_{i}}\,2\,\text{odd}(u_{i})\,[u_{1},v_{1},\ldots,u_{i-1},v_{i-1},u_{i}-1,v_{i},u_{i+1},v_{i+1},\ldots,u_{\sigma-1},v_{\sigma-1},u_{\sigma}],
Vi:=\displaystyle V_{i}:= (1)qi 2even(vi)[u1,v1,,ui1,vi1,ui,vi1,ui+1,vi+1,,uσ1,vσ1,uσ],\displaystyle(-1)^{q_{i}}\,2\,\text{even}(v_{i})\,[u_{1},v_{1},\ldots,u_{i-1},v_{i-1},u_{i},v_{i}-1,u_{i+1},v_{i+1},\ldots,u_{\sigma-1},v_{\sigma-1},u_{\sigma}],
pi:=\displaystyle p_{i}:= 1+1j<i(uj+vj) and qi:=1j<i(uj+vj)+ui.\displaystyle 1+\sum_{1\leq j<i}(u_{j}+v_{j})\text{ \ \ and \ \ }q_{i}:=\sum_{1\leq j<i}(u_{j}+v_{j})+u_{i}.

The tuple (u1,v1,,uσ1,vσ1,uσ)(u_{1},v_{1},\ldots,u_{\sigma-1},v_{\sigma-1},u_{\sigma}) and the basis element [u1,v1,,uσ1,vσ1,uσ][u_{1},v_{1},\ldots,u_{\sigma-1},v_{\sigma-1},u_{\sigma}] of 𝒟\mathcal{D}_{\bullet} are said to be even (respectively, odd) when u1u_{1} is even (respectively, odd). In the odd case, (5.13) gives a way to write the double of the class in \mathcal{H}_{\bullet} represented by an even basis element as a linear combination of the doubles of classes represented by odd basis elements. On the other hand, in the even case U1=0U_{1}=0 and the right hand-side of (5.13) is a linear combination of doubles of classes represented by even basis elements. A straightforward computation111The calculation is formally identical to the standard verification that 2=0\partial^{2}=0 for the boundary morphism \partial in the singular complex of a given space. Details are left as an exercise for the reader. shows that the latter linear combination vanishes directly in 𝒟\mathcal{D}_{\bullet} when (the double of) each even summand is replaced by the corresponding linear combination of odd basis elements. Thus, relations (5.13) coming from even tuples are irrelevant. Since each even basis element appears in a single relation (5.13) coming from an odd tuple, we see that the subgroup of \mathcal{H}_{\bullet} spanned by the classes represented by odd basis elements is torsion free. The proof is complete in view of Proposition 5.7. ∎

Acknowledgments

The research of L. Vandembroucq was partially financed by Portuguese Funds through FCT (Fundação para a Ciência e a Tecnologia) within the Project UID/00013: Centro de Matemática da Universidade do Minho (CMAT/UM). Sam Hughes was supported by a Humboldt Research Fellowship at Universität Bonn.

References

  • [1] J. Aguilar, J. González, J. Oprea, Right-angled Artin groups, polyhedral products and the TC-generating function, Proceedings of the Royal Society of Edinburgh Section A, 152 (2022), no. 3, 649–673.
  • [2] I. Basabe, J. González, Y.B. Rudyak, D. Tamaki, Higher topological complexity and its symmetrization, Algebr. Geom. Topol. 14 (2014), 2103–2124.
  • [3] I. Berstein, On the Lusternik–Schnirelmann category of Grassmannians, Math. Proc. Cambridge Philos. Soc. 79 (1976), 129–134.
  • [4] K. S. Brown, Cohomology of Groups, Graduate Texts in Mathematics, Vol. 87, Springer-Verlag, New York-Berlin, 1982.
  • [5] N. Cadavid-Aguilar, J. González, D. Gutiérrez, A. Guzmán-Sáenz, Aldo, A. Lara, Sequential motion planning algorithms in real projective spaces: an approach to their immersion dimension. Forum Math. 30 (2018), no. 2, 397–417.
  • [6] D. Cohen, L. Vandembroucq, On the topological complexity of manifolds with abelian fundamental group, Forum Math. 33 (2021), 1403-1421.
  • [7] A. Costa and M. Farber. Motion planning in spaces with small fundamental groups. Commun. Contemp. Math. 12 (2010), 107–119.
  • [8] N. Daundkar, Group actions and higher topological complexity of lens spaces, J. Appl. Comput. Topol. 8, No. 7, 2051–2067 (2024).
  • [9] D. Davis, Alower bound for higher topological complexity of real projective space, Journal of Pure and Applied Algebra 222 (2018), 2881–2887.
  • [10] A. Dranishnikov, The topological complexity and the homotopy cofiber of the diagonal map for non-orientable surfaces, Proc. Amer. Math. Soc. 144 (2016), 4999–5014.
  • [11] A. Dranishnikov, Y. Rudyak, On the Berstein-Svarc theorem in dimension 22. Math. Proc. Cambridge Philos. Soc. 146 (2009), 407–413.
  • [12] A. Espinosa Baro, M. Farber, S. Mescher, J. Oprea, Sequential topological complexity of aspherical spaces and sectional categories of subgroup inclusions, Math. Ann. 391(3), 4555–4605 (2025).
  • [13] M. Farber, Topological complexity of motion planning, Discrete Comput. Geom. 29 (2003), 211–221.
  • [14] M. Farber and M. Grant, Robot motion planning, weights of cohomology classes, and cohomology operations, Proc. Amer. Math. Soc. 136 (2008), no.9, 3339–3349
  • [15] M. Farber, S. Tabachnikov and, S. Yuzvinsky, Topological robotics: Motion planning in projective spaces, Int. Math. Res. Not. 30 (2003), 1853–1870.
  • [16] J. González, B. Gutiérrez, D. Gutiérrez and A. Lara, Motion planning in real flag manifolds, Homology Homotopy Appl. 18, 359–375 (2016).
  • [17] S. Hughes and K. Li, Higher topological complexity of hyperbolic groups, J. Appl. Comput. Topol. 6 (2022), no.3, 323–329.
  • [18] Y.B. Rudyak, On higher analogs of topological complexity, Topol. Appl. 157(5) (2010), 916–920.
  • [19] A. S. Schwarz. The genus of a fiber space. Amer. Math. Soc. Transl. (2) 55 (1966), 49–140.