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.eduwww.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.desamhughesmaths.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 of manifolds with abelian fundamental group. We give sufficient conditions for 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 , the -th higher topological complexity is the sectional category of the fibration , that is
where denotes the space of paths in and
is the usual -th evaluation map. That is, in the reduced version used here, is one less than the minimal number of open sets covering , over each of which the fibration admits a section.
Topological complexity 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 as the space of configurations of a mechanical system, the integer provides a topological measure of the complexity of planning motion in from an initial configuration to a terminal configuration, passing through specified intermediate configurations.
Despite a huge body of research into these invariants, there are very few complete computations of . 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 is not simply connected, this maximal value is , where is the homotopy dimension of , see [2]. Work of Cohen–Vandembroucq [6] explored the non-maximality of when is a manifold with abelian fundamental group. In this paper we extend these ideas to for .
Espinosa Baro, Farber, Mescher, and Oprea [12] have recently characterized the maximality of of a finite-dimensional CW-complex in terms of a canonical cohomology class generalizing the ‘Costa–Farber class’ introduced in [7] (see Section 1). Restricting our attention to a manifold with an abelian fundamental group and following the strategy of [6], we first
express
this characterization in terms of a homology class of the group (see Proposition 2.3 and Corollary 2.4). This permits us to establish the non-maximality of in some cases. For example, when is orientable, we obtain the following result (see Section 3):
Theorem A.
Let be an orientable -dimensional connected closed manifold. In each of the following cases, we have :
(1)
with ;
(2)
;
(3)
with .
Computations of cohomological lower bounds of the -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 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 be an orientable -dimensional connected closed manifold with maximal Lusternik–Schnirelmann category, that is, .
(1)
If mod and , then for sufficiently large.
(2)
If mod and where is a prime, then for sufficiently large.
See Corollaries 4.4 and 4.9 for a more explicit description of the condition “ 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 for . For instance, for the real projective plane , we have , see [16].
Furthermore,
it has been shown in [5] and [9] that, when is even, the real projective space satisfies for sufficiently large. For a fixed even integer , the sequence forms an increasing sequence starting at , equal to the immersion dimension of ([15]), and stabilizing to when 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 , we have:
As before, this result can be extended to a larger family of manifolds:
Theorem C.
Let be a non-orientable -dimensional connected closed manifold with and where . Then, for any even , we have .
If moreover , then and for .
Notation and conventions
For a topological space ,
we use to denote the homotopy dimension of . The integral homology of is denoted by , and the reduced homology by . If , we denote the cohomology of with coefficients in the local system
determined by the -module by .
For an element of a group , we often denote the inverse of by .
We use the reduced version of sectional category throughout, so that for a fibration , when finite, is one less than the minimal number of open sets covering , over each of which the fibration admits a section.
1. A 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 be a CW-complex.
The standard dimensional upper bound for higher topological complexity is
(1.1)
Although (1.1) can be improved in terms of the connectivity of , we are interested in the improvements coming from obstruction-theory techniques in cases where 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 is .
In [12], the homological obstruction for sectioning over the 1-dimensional skeleton of is identified with a canonical twisted class,
(1.2)
where and denotes the augmentation ideal of , viewed as a -submodule of . Here the action of on , which corresponds to the
monodromy associated with the fibration , is given by
The class can also be described as the cohomology class induced by the crossed homomorphism
given by
where is the unit element of . Obstruction-theoretic arguments lead then to the following result:
Let be a CW-complex of dimension . Then if and only if the -th cup-power .
Here,
lies in the cohomology of with coefficients in
the -th tensor power of endowed with the diagonal action of , denoted by .
The construction of the class generalizes the canonical class of [7] and provides for an analogue of the classical Berstein-Schwarz class . Note that, in this case, is the augmentation ideal of endowed with the left -module structure induced by the multiplication of . As is well-known, the Lusternik–Schnirelmann category of , , satisfies if and only if (see [3], [19] and [11] for a proof including the case ).
Remark 1.2.
We conclude this section with a brief remark regarding the functoriality of these classes. For , applying the above result to the classifying space yields a crossed homomorphism and associated cohomology class, which we denote by and respectively.
Recall that if is a map, and is a -module, then denotes the -module whose underlying abelian group is and the action of on is given by .
Taking to be a classifying map, the isomorphism yields
Similar considerations apply to the Berstein-Schwarz class (resp., ), induced by the crossed homomorphism , . Namely, the isomorphism yields .
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 is abelian.
We consider the group homomorphism given by
Note that the -module is exactly the -module . With the notation regarding canonical classes, Berstein-Schwarz classes, and crossed homomorphisms of the previous section, we also have, for any ,
We then have in , and, for any ,
where is a classifying map.
In order to establish our results, it is useful to consider the cofiber of the diagonal map . We denote it by .
We will more generally use the notation to denote the -diagonal of a set and suppress the superscript when the context is clear.
Proposition 2.1.
Let be an -dimensional CW-complex with . Suppose that is abelian and let be a classifying map. Then for any we have
(1)
in where is the identification map.
(2)
if and only if .
(3)
If then, for any -module and for any homology class , the class satisfies .
Proof.
First observe that the homomorphism is trivial. Consequently, the map obtained after applying the functor is also trivial. By identifying with and with , we have a commutative diagram of the following form
where is induced by the quotient property.
Since is abelian, we have an exact sequence and, using the Van Kampen theorem, we can see that and that is an isomorphism. Consequently is a classifying map and the Berstein-Schwarz class of is given by
. By the commutativity of the diagram we then get and as claimed in the first item.
The equality established above implies that .
For dimensional reasons, the map is an isomorphism. We therefore have if and only if , which implies the second item.
We now prove the last item. Let be a nonzero class and let . We have . Note that is a homology class of degree . Since we have . Therefore the classifying map factors up to homotopy through an -dimensional space. Consequently, and the result follows. ∎
Remark 2.2.
In the situation of Proposition 2.1, if is a trivial -module and is an element such that the class satisfies then .
Item (3) of Proposition 2.1 is sharp under reasonably general conditions. Let be an -dimensional connected closed manifold with fundamental group and let be the homomorphism determined by the first Stiefel-Whitney class of . Recall that the orientation module of , denoted by , is the abelian group given with a structure of -module determined by for , . Note that , which additively is with action given by .
Proposition 2.3.
Let be an -dimensional connected closed manifold with and abelian. Assume there is a -module such that the -modules and are isomorphic. Then the following two conditions are equivalent:
(1)
The class satisfies in .
(2)
.
Here we denote by the image of the fundamental class of under the homomorphism induced by . Note that we also denote by the local system over arising from the isomorphism induced by the classifying map .
Proof.
From the naturality of the cap-product and the assumption that is a -module satisfying we get the following diagram.
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
in degree . It is induced by the obvious isomophism between the underlying -modules and and is an isomorphism on the coinvariants because is surjective.
Let be the fundamental class. Since the third column of the diagram is comprised of isomorphisms, we have
This is equivalent to saying
The hypothesis yields . By Poincaré duality, we can then conclude that and consequently .
∎
Corollary 2.4.
Let be an orientable -dimensional manifold with and abelian fundamental group . The class satisfies in if and only if .
Proof.
Since is orientable, the orientation module is just with trivial action. Taking also with trivial action, we have
and the result follows from Proposition 2.3.
∎
Remark 2.5.
When is non-orientable, the -module satisfies the assumptions of Proposition 2.3 for ([6]) but fails to do so for all .
For instance, set and suppose that is an element for which the orientation character satisfies .
Then, for and we have
while
This shows that the map does not induce a homomorphism from to .
Note that, in Proposition 2.3, must be, as an abelian group, isomorphic to . Furthermore, since is surjective, the -module structure on is forced by the hypothesis and this condition is impossible when is odd.
For instance, again set , choose as above and assume . The equalities then lead to the impossible
Nonetheless, when , , the -module does satisfy , and we explore its usage in Section 5.
3. Some calculations in the orientable case
Let be an orientable connected closed manifold with abelian. In this section we will use Corollary 2.4 to establish the non-maximality for some families of manifolds with abelian fundamental groups.
Let be a classifying map and let . Since is orientable, we will suppress the -coefficients from the notation. In all cases, we will see that in .
In our first result, we suppose that 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 be an orientable -dimensional connected closed manifold with and let . If then .
Proof.
Let , let be a classifying map and let . For degree reasons, we can see that in . Indeed and for . Consequently if .
∎
In general, observe that the homomorphism given by
can be decomposed as
(3.1)
where is the diagonal map and . Denote by the inversion.
Since can be seen as the multiplication of , , precomposed with , we have, for classes ,
where is the Pontryagin product, that is, the product induced by in homology, see [4, V.5].
In the results below, we consider the cyclic group and work at the chain level. Recall the classical resolution of as a trivial -module given by
(3.2)
where .
In the following lemma, we recall the morphisms induced by the diagonal , the multiplication and the inversion on the level of resolutions (see [4, page 108] and [6, §3.2]). Let denote the generator of degree in (3.2), and write for the binomial coefficient .
We denote by the -chain complex obtained by tensoring the resolution (3.2) with over .
(3.3)
Recall that the homology of this chain complex gives . In positive degrees, is concentrated in odd degrees.
As in [6], we denote by the Pontryagin product, which is given by the formulæ of Lemma 3.2:
(3.4)
We denote by the morphism induced by the inversion, which is from Lemma 3.2 given by
(3.5)
In these terms, the chain map induced by can be described as the composite
We will also use the diagonal approximation of , obtained from Lemma 3.2 :
(3.6)
Here if is even and if is odd.
Proposition 3.3.
Let be an orientable -dimensional connected closed manifold with . Then, for any , we have .
Proof.
Let be a classifying map, where , and let . We will see that in . If is even, this is immediate since is concentrated in odd degrees, which implies . We then suppose that . A cycle representing the class is of the form for some . In order to compute we use the decomposition (3.1) and analyze the element which is given by
The element is given by a -linear combination of elements of the form
where for any . Setting , there will be necessarily some such that and are both odd. Applying to the element above yields
If and are both odd, the corresponding factor vanishes since is a multiple of and the Pontryagin product of two odd degree elements is zero. Consequently, we obtain and .
∎
Proposition 3.4.
Let be an orientable -dimensional connected closed manifold with such that . Then, for any , we have .
Proof.
Let be a classifying map, where , and let . By the Künneth formula, we have . Since , we can write where and with each of degree . Since is
concentrated in odd degrees, a cycle representing can be described as a sum of terms of the form where , and is a class regarded as a cycle.
The element is therefore given by a -linear combination of elements of the form
(3.7)
where , for , and are all odd and the elements belong to . The calculation of on (say) is made componentwise and gives rise to factors of the form
As in the proof of Proposition 3.3, there will be necessarily, in the expression (3.7), some such that and are both odd. After applying , the corresponding factor will be . Consequently, we obtain and . We can hence conclude that .
∎
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 where is a prime.
Example 3.5.
A manifold with and .
Set and consider . We will write instead of . We first consider the cycle and denote by its homology class. We will see that . By the Universal Coefficient Theorem, it is actually sufficient to see that where corresponds to in . As and for all , we will continue to write .
Using the diagonal approximation associated to the resolution (3.2) described in Lemma 3.2 (or, tensoring the diagonal (3.6) by ) we can check that the homology diagonal of satisfies
Consequently, the homology diagonal of satisfies:
We have to compute:
A term of the form is given in by a componentwise calculation:
Taking into account the formulas for the inversion and for the Pontryagin product (induced in -homology by the formulas (3.5) and (3.4) given above) we have
which vanishes since we are working with coefficients in . We can thus check that
and that this is the only term belonging to in the expansion of . Since does not vanish in , we can conclude that
. Consequently .
We can next follow the same strategy as in [6] to show that there exists a manifold with fundamental group and maximal . More precisely, considering the lens spaces and , we can realize the class as the image of the fundamental class of under the map induced by
We can then use surgery to replace by a manifold with and by a classifying map . In this way, and, from and Proposition 2.1 (3), we can deduce that .
In [6], it has been shown that the regular topological complexity of a non-orientable manifold with abelian fundamental group is never maximal. This is not longer true for with . For instance, for the real projective plane , -zero-divisor cuplength considerations imply that , 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 given by the -zero-divisor cup length or -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 be a field. Recall that, for a space , the (-coefficients) -zero-divisor cup length, , is the maximum of the set
In some cases, a better lower bound can be obtained through the notion of -weight. Recall (see [14, §2]) that if is a fibration and is a nontrivial class, the weight of associated to , , is the largest integer such that for any map satisfying . If , then if and only if and . Moreover, if satisfy then
For a space , the -weight, denoted by , is the weight associated to the fibration . Taking coefficients in , the morphism can be identified with the -fold cup-product and we can define the (-coefficients) weighted -zero divisor cup length, , to be the maximum of the set
We have . We also note that if is a map and satisfies then .
4.1. Manifolds with and
The -coefficient -zero-divisor cuplength of the real projective space 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 when and . We first recall some results from [9].
For an integer with binary expansion , i.e., , with digits , let
•
denote the exponent in the maximal 2-power dividing , i.e., is the minimal with ;
•
, the set of binary positions starting (from left to right) a block of consecutive 1’s of length at least 2;
•
, the complement of the binary expansion of mod .
Building on [5], Davis [9] proves that, for , the -coefficient -zero-divisors cuplength of the -dimensional real projective space is given by
where . In particular, for even (so that is non-orientable), has maximal possible , that is, , whenever
When is even and its binary expansion has no blocks of two or more consecutive 1’s, we have and the inequality (4.1) reduces to . Note that this condition for the maximality of is sharp, since ([15], [7, Theorem 1]).
(b)
For , we have , and (4.1) becomes . Note that, when , we have . We will see in Section 5 that so that the condition for the maximality of is sharp again.
Thanks to the following result, Davis’ computations of have impact on more general manifolds.
Proposition 4.2.
Let be an -dimensional connected closed manifold with and . Then, for any , .
Proof.
Let be a classifying map and let be the generator. For dimensional reasons, factors as . Let .
The hypothesis that implies that , see [3].
Consequently as well as are monomorphisms. As the image by of a -zero-divisor of over gives rise to a -zero-divisor of over , we can conclude that and the result follows.
∎
From Example 4.1 and the discussion above, we directly obtain:
Corollary 4.3.
Let be an -dimensional connected closed manifold with and . If is even and satisfies the inequality (4.1), then . In particular:
(a)
If is even and its binary expansion of contains no consecutive digits equal to , then for any .
(b)
If , then for any .
By using Davis’ computations in combination with Proposition 3.3, we can also state:
Corollary 4.4.
Let be an orientable connected closed manifold satisfying
the conditions
and . If , then for satisfying the inequality (4.1).
Proof.
In this case so that for 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 , implies that is odd. Indeed, if were even, the image of the fundamental class of by would vanish. But this fact would force, by Poincaré duality, the th power of the Berstein-Schwarz class of to vanish, contradicting the equality .
4.2. Odd dimensional manifolds with and
Throughout this section we consider a prime . Recall that the classifying space can be identified with the infinite dimensional lens space .
In order to have an analogue of Proposition 4.2, we first note the following result:
Lemma 4.6.
Let be an orientable connected closed -manifold
satisfying the conditions
and . If is a classifying map, then is non-zero.
Proof.
Considering the Berstein–Schwarz class , we have if and only if . By Poincaré duality, the second statement is equivalent to . Taking cap products we obtain
Since induces an isomorphism at the level of fundamental groups, naturality of the cap-products yields
Hence, in . Since we can conclude that in .
∎
Theorem 4.7.
Let be a closed orientable -manifold with and where is a prime. Then, .
Proof.
Let be a classifying map.
We have a commutative triangle
(4.2)
where the inclusion is simply the -skeleton of the infinite dimensional lens space . Since , we know by the previous lemma that in . Consequently in .
Recall that
where , . In particular, .
We first check that is a monomorphism.
It suffices to show that and are non-trivial in . Consider the following cap-product diagram:
Both horizontal morphisms are isomorphisms by Poincaré duality. As in , is not divisible by in . Consequently, and using the diagram we have that . We thus have in and hence and for are non-zero. Therefore is a monomorphism. By Künneth formula, we obtain that is also a monomorphism. Let . Then and . Consequently and the results follows by definition of .
∎
There are extensive computations of , 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 and let be a closed orientable -manifold with and where is prime. Then,
where are any integers such that does not divide .
Proof.
Here we apply the computation [8, Theorem 5.2] and Theorem 4.7.
∎
In many situations, for example if is much larger than the dimension of , we obtain an exact computation.
Corollary 4.9.
Let and let be a closed orientable -manifold with and where is a prime. If does not divide , then
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 for non-orientable manifolds having . The case is well understood ([7, Theorem 1]), so we assume from now on. For such cases the non-maximality of demands further restrictions on . The aim of this section is to establish the following result.
Theorem 5.1.
Let be a non-orientable -dimensional manifold with and . Then, for any even no greater than , we have .
By Corollary 4.3(b), we know that, for , for , so that in this case the upper limiting restriction on in Theorem 5.1 is in fact sharp and we have:
Corollary 5.2.
If the manifold in Theorem 5.1 has , then for and for .
Proof.
The equality for follows from (see Example 4.1(b)), Proposition 4.2 and Theorem 5.1.
∎
Corollary 5.2 should be compared to the fact that is maximal for , but ([15]). Worth noting is the fact that the case in Corollary 5.2 (with ) upgrades the observation in [5, (7.4)] that to an equality, giving evidence for what would be regular behavior of the higher topological complexity of projective spaces with .
Suitable analogues of Theorem 5.1 should hold for more general values of , 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 and with . Set , the -module of Remark 2.5. By Proposition 2.3, it suffices to establish the triviality of
(5.1)
where
(5.2)
As in §3,
our starting point is the free -resolution (3.2) of with . Recall that denotes the generator of degree . In addition to the chain complex of (3.3), we will also need the complex ,
(5.3)
obtained by tensoring (3.2) with over . Abusing notation, we continue using for the generators of both and .
The homology groups in (5.2) can be computed from the complexes and
(5.4)
In both cases, we will use the shorthand for a tensor product . The Künneth formula and the fact that the homology of is 2-torsion (in all degrees) gives:
Lemma 5.3.
The element in (5.1) is torsion. Indeed, both groups in (5.2) are torsion.
Let denote the quotient of resulting from killing all boundaries, and consider the obvious monomorphism . 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 is an element of the torsion-free (graded) subgroup of .
In computing the homology groups in (5.2) using the complexes and , we will use for a factor where is meant to be taken, reserving the notation for factors where is meant to be taken. For instance, the diagonal morphism and the group-multiplication morphism extend to morphisms , and that are compatible with the implied module structures. In these terms, since
in the present case,
the inversion morphism plays no role and the map factors as
(5.5)
Recalling that denotes the binomial coefficient , the formulæ of Lemma 3.2 written with the shorthand in use in this section read
(5.6)
We note also that, since , the formula of Lemma 3.2(a) giving on generators at the level of resolutions can be written
(5.7)
where generates and if is odd and otherwise.
The class is either trivial or, else, represented by the cycle in (5.3) —recall is even. For the purposes of proving Theorem 5.1, we may safely assume the latter possibility. Then, is represented in by the corresponding tensor product . We chase the latter element under the first map of the composite (5.5) to get, in view of (5.7),
Note that in the latter expression we are extending in the obvious way the convention above regarding the use of and . Then, after tensoring with the needed coefficients (thus dropping the indicators), this becomes
Since is even, the parity of each agrees with the one of the corresponding , so the last expression can be rewritten as
where the sum now runs over , and . Using the formulae (5.6), we finally obtain the image of under the entire composition in (5.5). This image may be expressed as
(5.8)
where
and the sum runs over the same indices as above, except now that no two consecutive and can simultaneously equal 1, in view of (5.6). In what follows we set .
Having described a cycle representing the obstruction in (5.1), we next spell out the complex (5.4) where it lies. Degreewise, is -free with basis given by elements for non-negative integers and , and with differential
(5.9)
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 , let denote the set of binary positions where the binary expansion of has digit 1. For instance, and .
A standard well known fact is:
Lemma 5.6.
A binomial coefficient is even if and only if .
The following result is the only place where the special assumptions in Theorem 5.1 (i.e., with ) are needed.
We use without further notice the fact coming from Lemma 5.6 that any binomial coefficient is even whenever . Recall and
(5.11)
with . Assume for a contradiction that some coefficient (5.10) is odd (i.e., that all of its binomial-coefficient factors are odd) and has . Recall the forced conditions
(1)
with and ;
(2)
and implies ,
for . Let be all the indices (if any) with . Note that
The coefficient is odd by hypothesis, so and —the latter equality holds in view of (1) since . Actually, the same argument can be used iteratively for (so ) with the binomial coefficients (e.g. ) to show that
and .
Next, since , is odd and , we get
and ,
and now the process repeats with a slight adjustment. For starters, , is odd and is forced by (2), so that and . We then iterate the latter argument: For , the assumption and the fact that is odd with yield
and .
Of course, the last two inequalities now hold for all . The next round of iterations start with the fact that is odd with and , to get
and ,
and the process has a corresponding new obvious adjustment to yield
and ,
for , whereas
and .
Just before the last adjustment we get
and ,
for , whereas
and .
However, after this point the conditions and are kept for all . In particular, , in view of (5.11) and (5.12). But then the final factor of (5.10) is even, a contradiction.
∎
The right hand-side of (5.9) yields the defining relations in . Namely, for each tuple of non-negative integers there is a defining relation
(5.13)
where
The tuple and the basis element of are said to be even (respectively, odd) when is even (respectively, odd). In the odd case, (5.13) gives a way to write the double of the class in 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 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 for the boundary morphism 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 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 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 Algebra222 (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 . 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.