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

Higher topological complexity of planar polygon spaces having small genetic codes

Sutirtha Datta Department of Mathematics, Indian Institute of Science Education and Research Pune, India sutirtha2702@gmail.com Navnath Daundkar Department of Mathematics, Indian Institute of Science Education and Research Pune, India navnath.daundkar@acads.iiserpune.ac.in  and  Abhishek Sarkar Department of Mathematics, Indian Institute of Science Education and Research Pune, India abhisheksarkar49@gmail.com
Abstract.

We study the higher (sequential) topological complexity, a numerical homotopy invariant for the planar polygon spaces. For these spaces with a small genetic codes and dimension mm, Davis showed that their topological complexity is either 2m2m or 2m+12m+1. We extend these bounds to the setting of higher topological complexity. In particular, when mm is power of 22, we show that the kk-th higher topological complexity of these spaces is either kmkm or km+1.km+1.

Key words and phrases:
LS category, higher topological complexity, planar polygon spaces
2020 Mathematics Subject Classification:
55M30, 58D29, 55R80

1. Introduction

A motion planning algorithm in a path-connected topological space XX is defined as a section of the free path space fibration π:XIX×X,\pi:X^{I}\to X\times X, where π(γ)=(γ(0),γ(1)),\pi(\gamma)=(\gamma(0),\gamma(1)), and XIX^{I} denotes the free path space of XX, equipped with the compact open topology. To analyze the complexity of designing a motion planning algorithm for the configuration space XX of a mechanical system, Farber [11] introduced the notion of topological complexity. The topological complexity of a space XX, denoted by TC(X)\mathrm{TC}(X), is defined as the smallest positive integer rr for which X×XX\times X can be covered by open sets {U1,,Ur}\{U_{1},\dots,U_{r}\}, with each UiU_{i} admitting a continuous local section of π\pi. The number TC(X)\mathrm{TC}(X) represents the minimal number of continuous rules required to implement a motion planning algorithm in the space XX. Farber [11, Theorem 3] showed that TC(X)\mathrm{TC}(X) is a numerical homotopy invariant of a space XX.

In [16], Rudyak introduced the higher analogue of topological complexity. For a path-connected space XX, consider the fibration πk:XIXk\pi_{k}:X^{I}\to X^{k} defined by

πk(γ)=(γ(0),γ(1k1),,γ(k2k1),γ(1)).\pi_{k}(\gamma)=\bigg{(}\gamma(0),\gamma\bigg{(}\frac{1}{k-1}\bigg{)},\dots,\gamma\bigg{(}\frac{k-2}{k-1}\bigg{)},\gamma(1)\bigg{)}. (1)

The higher topological complexity of XX, denoted by TCk(X)\mathrm{TC}_{k}(X), is the smallest positive integer rr for which XkX^{k} can be covered by open sets {U1,,Ur}\{U_{1},\dots,U_{r}\}, such that each UiU_{i} admits a continuous local section of πk\pi_{k} . Note that when k=2k=2, the higher topological complexity TCk(X)\mathrm{TC}_{k}(X) coincides with TC(X)\mathrm{TC}(X). The above definition makes also sense for k=1k=1, but TC1(X)\mathrm{TC}_{1}(X) is always equal to 11 (see [16]). Similar to TC(X)\mathrm{TC}(X), the invariant TCk(X)\mathrm{TC}_{k}(X) is associated with motion planning problems, where the input includes not only an initial and final point but also an additional k2k-2 intermediate points.

An older homotopy invariant of topological spaces, known as the Lusternik-Schnirelmann (LS) category, was introduced by Lusternik and Schnirelmann in [15]. The LS-category of a space XX, denoted by cat(X)\mathrm{cat}(X), is the smallest positive integer rr such that XX can be covered by open subsets V1,,VrV_{1},\dots,V_{r} such that each inclusion ViXV_{i}\hookrightarrow{}X is null-homotopic. The following inequalities were established in [1] shows how the LS category and higher topological complexity are related:

cat(Xk1)TCk(X)cat(Xk).\mathrm{cat}(X^{k-1})\leq\mathrm{TC}_{k}(X)\leq\mathrm{cat}(X^{k}). (2)

Determining the precise values of these invariants is usually a challenging task. Over the past two decades, several mathematicians have significantly contributed to approximating these invariants with bounds. To be more specific, Farber [11, Theorem 7] gave a cohomological lower bound on the topological complexity, and this concept was extended to the higher topological complexity by Rudyak in [16, Proposition 3.4]. Let Δk:XXk\Delta_{k}:X\to X^{k} be the diagonal map and zclk(X)\mathrm{zcl}_{k}(X) denote the cup-length of ker(Δk)\mathrm{ker}(\Delta_{k}^{*}), where cup-length is the length of the largest nontrivial product of cohomology classes. Rudyak in [16, Proposition 3.4] proved the following inequality, generalizing Farber’s cohomological lower bound on the topological complexity.

TCk(X)zclk(X)+1.\mathrm{TC}_{k}(X)\geq\mathrm{zcl}_{k}(X)+1. (3)

We refer to the non-negative integer zclk(X)\mathrm{zcl}_{k}(X) as the higher zero-divisors-cup-length of XX. On the other hand, if cl(X)\mathrm{cl}(X) denotes the cup-length of a cohomology ring of XX. Then there is an inequality

cat(X)cl(X)+1\mathrm{cat}(X)\geq\mathrm{cl}(X)+1 (4)

(see [2, Proposition 1.5]). For a paracompact space, there is a usual dimensional upper bound on the higher topological complexity given as follows:

TCk(X)kdim(X)+1.\mathrm{TC}_{k}(X)\leq k\cdot\mathrm{dim}(X)+1. (5)

Moreover, an additional upper bound for TCk(X)\mathrm{TC}_{k}(X) is formulated in terms of the homotopy dimension of the space (see [1, Theorem 3.9]).

In this paper, we are interested in studying the (higher) motion planning problem for planar polygon spaces. These spaces can be viewed as equivalence classes of oriented planar nn-gons with consecutive side lengths α1,,αn(0,)\alpha_{1},\dots,\alpha_{n}\in(0,\infty) for some nn\in\mathbb{N}, identified under translation, rotation, and reflection. Practically, one can regard the sides of a polygon as the linked arms of a robot. Then the higher topological complexity of a polygon space describes the minimum number of motion-planning rules required to maneuver the robot from an initial configuration to a final configuration, ensuring the motion passes through a fixed sequence of intermediate configurations.

We now briefly recall what planar polygon spaces are and some basic information. A length vector is a tuple of positive real numbers. The planar polygon space associated with a length vector α=(α1,,αn)\alpha=(\alpha_{1},\dots,\alpha_{n}), denoted by Mα\mathrm{M}_{\alpha}, is the collection of all closed piecewise linear paths in the plane with side lengths α1,α2,,αn\alpha_{1},\alpha_{2},\dots,\alpha_{n}, considered up to orientation-preserving isometries. Equivalently, we can describe Mα\mathrm{M}_{\alpha} as

Mα:={(v1,v2,,vn)(S1)ni=1nαivi=0}/SO2,\mathrm{M}_{\alpha}:=\{(v_{1},v_{2},\dots,v_{n})\in(S^{1})^{n}\mid\sum_{i=1}^{n}\alpha_{i}v_{i}=0\}/\mathrm{SO}_{2},

where S1S^{1} is the unit circle and the group of orientation-preserving isometries SO2\mathrm{SO}_{2} acts diagonally. The planar polygon space (associated with α\alpha) viewed up to isometries is defined as

M¯α{(v1,v2,,vn)(S1)ni=1nαivi=0}/O2,\overline{\mathrm{M}}_{\alpha}\coloneqq\{(v_{1},v_{2},\dots,v_{n})\in(S^{1})^{n}\mid\displaystyle\sum_{i=1}^{n}\alpha_{i}v_{i}=0\}/\mathrm{O}_{2},

where O2\mathrm{O}_{2} is the orthogonal group acting on 2\mathbb{R}^{2}, including all linear isometries.

The spaces Mα\mathrm{M}_{\alpha} and M¯α\overline{\mathrm{M}}_{\alpha} are also called moduli spaces of planar polygons. It is clear that Mα\mathrm{M}_{\alpha} is a double cover of M¯α\overline{\mathrm{M}}_{\alpha}. A length vector α\alpha is called generic if i=1n±αi0\sum_{i=1}^{n}\pm\alpha_{i}\neq 0. For such a length vector α\alpha, the moduli spaces Mα\mathrm{M}_{\alpha} and M¯α\overline{\mathrm{M}}_{\alpha} are closed, smooth manifolds of dimension m:=n3m:=n-3 (see [12]). In the rest of this paper, the length vectors are assumed to be generic unless stated otherwise.

The planar polygon spaces have been studied extensively. For example, Farber and Schütz [10] proved that the integral homology groups of Mα\mathrm{M}_{\alpha} are torsion-free. They also described the Betti numbers in terms of the combinatorial data associated with the length vector. The mod-22 cohomology ring of M¯α\overline{\mathrm{M}}_{\alpha} was computed by Hausmann and Knutson in [13].

The program of studying the motion planning problem for planar polygon spaces was initiated by Donald Davis. In his several works [6, 7, 8, 9], he has shown that in most cases TC(M¯α)\mathrm{TC}(\overline{\mathrm{M}}_{\alpha}) is either 2m2m or 2m+12m+1. The estimates for TC(M¯α)\mathrm{TC}(\overline{\mathrm{M}}_{\alpha}), are primarily determined by the usual dimensional upper bound and cohomological lower bound, respectively, with only a few exceptions. Davis utilized the genetic code description to investigate these bounds, noting that the homeomorphism type of M¯α\overline{\mathrm{M}}_{\alpha} is determined by its genetic code α\alpha. In this paper, we initiate the study of the higher topological complexity of planar polygon spaces, focusing on spaces with small genetic codes. We establish higher analogues of Davis’s results [7] through a systematic investigation, leveraging a novel small genetic code description for the mod-22 cohomology ring of M¯α\overline{\mathrm{M}}_{\alpha}.

1.1. Structure of the article:

In Section˜2, we begin with the digression of genetic codes and the description of the mod-22 cohomology ring of the planar polygon spaces. Then we compute the exact value of the LS category of these spaces in Proposition˜2.7 and briefly explain the key strategy for our proofs.

In Section˜3, we compute bounds on TCk(M¯α)\mathrm{TC}_{k}(\overline{\mathrm{M}}_{\alpha}) when the genetic code of α\alpha contains a gene of size 22. Our results Theorem˜3.1 and Theorem˜3.3 address the cases where α\alpha has a monogenic code of size 22 and where the genetic code includes a gene of size 22, respectively. In particular, in Theorem˜3.1 we identify cases for which TCk(Mα)\mathrm{TC}_{k}(\mathrm{M}_{\alpha}) is either kmkm or km+1km+1. In Proposition˜3.4, we explicitly consider the case of a genetic code having a gene of size 22 and under some minor conditions we show that TCk(M¯α)\mathrm{TC}_{k}(\overline{\mathrm{M}}_{\alpha}) is either kmkm or km+1km+1.

Section˜4 consists of the cases of genetic code containing a gene of size 33. In Theorem˜4.2, we establish that TCk(M¯α)\mathrm{TC}_{k}(\overline{\mathrm{M}}_{\alpha}) is either kmkm or km+1km+1 under mild assumptions. In general, we provide a novel bound in Theorem˜4.3, which extends [7, Theorem 2.4].

In Section˜5, we deal with genetic codes of Type 22. Specifically, our result Theorem˜5.1 extends [7, Theorem 3.1] to the sequential case.

In Section˜6, we obtain the bounds for TCk(M¯α)\mathrm{TC}_{k}(\overline{\mathrm{M}}_{\alpha}) where α\alpha having monogenic codes of size 44. We begin this section by classifying (see Lemma˜6.1) such genetic codes for which the Rm0R^{m}\neq 0 (see Theorem˜2.4 for the cohomology class RR). This is crucial in identifying the genetic codes of α\alpha for which we can have TCk(M¯α)\mathrm{TC}_{k}(\overline{\mathrm{M}}_{\alpha}) to be either kmkm or km+1km+1. We have achieved this in Theorem˜6.2. We end this section by proving our general result Theorem˜6.3, which is a higher analogue of Davis’s results [7, Proposition 4.3 and Proposition 4.4].

Finally, in Section˜7, we take care of two sorts of genetic codes. Theorem˜7.2 and Theorem˜7.3 deal with the higher topological complexity of the genetic codes consisting of two genes, each of size 33, with the former one being a strong bound and the latter one generalizing the result from section 55 of [7]. Our final result Theorem˜7.4 involves genetic codes of Type 1, a generalization of [7, Proposition 6.2].

2. Planar polygon spaces

In this section, we provide the necessary background on planar polygon spaces, which is essential for stating and proving our upcoming results.

We denote the set {1,,r}\{1,\dots,r\} by [r][r], for rr\in\mathbb{N}. There are two important combinatorial objects associated with the length vector α=(α1,α2,,αn)\alpha=(\alpha_{1},\alpha_{2},\dots,\alpha_{n}).

Definition 2.1.

A subset I[n]I\subset[n] is called α\alpha-short if

iIαi<jIαj\sum_{i\in I}\alpha_{i}<\sum_{j\not\in I}\alpha_{j}

and α\alpha-long otherwise.

We may simply use "short" as a shorthand for α\alpha-short when the context is clear. The collection of short subsets can be quite large. However, there is another combinatorial object associated with length vectors that further compacts the short subset data. It is important to note that the diffeomorphism type of a planar polygon space does not depend on the ordering of the side lengths. Therefore, we assume that the length vector satisfies α1α2αn\alpha_{1}\leq\alpha_{2}\leq\dots\leq\alpha_{n}.

For a length vector α\alpha, consider the collection of subsets of [n][n] :

Sn(α):={J[n]: nJ and J is short}S_{n}(\alpha):=\{J\subset[n]:\text{ $n\in J$ and $J$ is short}\}

along with the partial order \leq defined as follows:

IJifI={i1,,it}and{ji,,jt}Jwithisjsfor1st.I\leq J~\text{if}~I=\{i_{1},\dots,i_{t}\}~\text{and}~\{j_{i},\dots,j_{t}\}\subseteq J~\text{with}~i_{s}\leq j_{s}~\text{for}~1\leq s\leq t.
Definition 2.2.

The genetic code of α\alpha is the set of maximal elements of Sn(α)S_{n}(\alpha) with respect to the partial order defined above.

If A1,A2,,AkA_{1},A_{2},\dots,A_{k} are the maximal elements of Sn(α)S_{n}(\alpha) with respect to \leq then the genetic code of α\alpha is denoted by A1,,Ak\langle A_{1},\dots,A_{k}\rangle. Each AiA_{i}’s is called a gene and a subset Ai{n}A_{i}\setminus\{n\} is called gee for each 1ik1\leq i\leq k (see [7] for more details).

Example 2.3.

Let α=(1,,1,n2)\alpha=(1,\dots,1,n-2) (nn-tuple) be a length vector. Then the genetic code of α\alpha is {n}\langle\{n\}\rangle. Moreover, MαSn3\mathrm{M}_{\alpha}\cong S^{n-3} and M¯αPn3\overline{\mathrm{M}}_{\alpha}\cong\mathbb{R}P^{n-3}.

Let GG be the genetic code of α\alpha. Then it is easy to see that GG uniquely determines the collection Sn(α)S_{n}(\alpha). Consequently, it follows from [14, Lemma 1.2] that if the genetic codes of length vectors α\alpha and β\beta are the same, then the corresponding planar polygon spaces are diffeomorphic.

We now recall the mod-22 cohomology algebra H(M¯α;2)H^{*}(\overline{\mathrm{M}}_{\alpha};\mathbb{Z}_{2}) of M¯α\overline{\mathrm{M}}_{\alpha} in terms of the genetic code information. This result was originally established by Hausmann and Knutson in [13, Corollary 9.2], and later reinterpreted by Davis (see [7, Theorem 2.1]) as follows.

Theorem 2.4.

The algebra H(M¯α;2)H^{*}(\overline{\mathrm{M}}_{\alpha};\mathbb{Z}_{2}) is generated by classes R,V1,,Vn1H1(M¯α;2)R,V_{1},\dots,V_{n-1}\in H^{1}(\overline{\mathrm{M}}_{\alpha};\mathbb{Z}_{2}) subject to the following relations:

  1. (1)

    RVi+Vi2RV_{i}+V_{i}^{2}, for i[n1]i\in[n-1].

  2. (2)

    VS:=iSViV_{S}:=\prod\limits_{i\in S}V_{i}, unless SS is a subgee.

  3. (3)

    For every subgee SS with |S|nd2|S|\geq n-d-2,

    TS=Rd|T|VT,\sum\limits_{T\cap S=\emptyset}R^{d-|T|}V_{T}, (6)

    where TT is a subgee.

The following result is an immediate application of the previous theorem, which is crucial in proving some of the results in the next section.

Corollary 2.5 ( [7]).

For the genetic code {a,n}\langle\{a,n\}\rangle of α\alpha, the following set

{Rd,Rd1V1,,Rd1Va}\{R^{d},R^{d-1}V_{1},\dots,R^{d-1}V_{a}\}

forms a basis for Hd(M¯α;2)H^{d}(\overline{\mathrm{M}}_{\alpha};\mathbb{Z}_{2}), for 1dn41\leq d\leq n-4.

We now set up some notations that will be used throughout the paper. Let pj:XkXp_{j}:X^{k}\rightarrow X be the projection onto the kk-th factor. There is an induced map on the cohomology ring

pj:H(X;𝕂)H(X;𝕂)k,p_{j}^{*}:H^{*}(X;\mathbb{K})\rightarrow H^{*}(X;\mathbb{K})^{\otimes k},

where 𝕂\mathbb{K} is a field. Now considering X=M¯αX=\overline{\mathrm{M}}_{\alpha} and 𝕂=2\mathbb{K}=\mathbb{Z}_{2} we set the following notions

  1. (1)

    Rj:=pj(R)R_{j}:=p_{j}^{*}(R)

  2. (2)

    wj:=pj(V1)w_{j}:=p_{j}^{*}(V_{1})

  3. (3)

    R¯j:=Rj+Rj1\bar{R}_{j}:=R_{j}+R_{j-1}

  4. (4)

    w¯j:=wj+wj1\bar{w}_{j}:=w_{j}+w_{j-1}

Observe that R¯jker(Δk)\bar{R}_{j}\in\mathrm{ker}(\Delta_{k}^{*}) and w¯jker(Δk)\bar{w}_{j}\in\mathrm{ker}(\Delta_{k}^{*}), where Δk:XXk\Delta_{k}:X\to X^{k} is the diagonal map.

Remark 2.6.
  1. (1)

    From Corollary 2.5, it follows that {Rd,Rd1V1}\{R^{d},R^{d-1}V_{1}\} is linearly independent. Thus,

    {Rjd,Rjd1wj}\{R_{j}^{d},R_{j}^{d-1}w_{j}\}

    forms a linearly independent set in the cohomology algebra Hd(M¯α;2)kH^{d}(\overline{\mathrm{M}}_{\alpha};\mathbb{Z}_{2})^{\otimes k}.

  2. (2)

    As part-(1) extends Corollary˜2.5, we repeatedly utilize similar canonical approaches to explore the relations among cohomology classes in Hd(M¯α;2)kH^{d}(\overline{\mathrm{M}}_{\alpha};\mathbb{Z}_{2})^{\otimes k} for different kinds of small genetic codes, using these as key tools to extend the results of Davis.

We now compute the LS category of planar polygon spaces and obtain bounds on the higher topological complexity of M¯α\overline{\mathrm{M}}_{\alpha} whose dimension is m=n3m=n-3.

Proposition 2.7.

Let α\alpha be a length vector. Then cat(M¯α)=m+1\mathrm{cat}(\overline{\mathrm{M}}_{\alpha})=m+1. Moreover, we have

(k1)m+1TCk(M¯α)km+1.(k-1)m+1\leq\mathrm{TC}_{k}(\overline{\mathrm{M}}_{\alpha})\leq km+1.
Proof.

Observe that for a gee JJ of maximal size, we have Rm|J|iJViR^{m-|J|}\prod_{i\in J}V_{i} generates Hm(M¯α;2)H^{m}(\overline{\mathrm{M}}_{\alpha};\mathbb{Z}_{2}). Therefore, for some J[n1]J\subseteq[n-1] we have Rm|J|iJVi0R^{m-|J|}\prod_{i\in J}V_{i}\neq 0 This gives us cl(M¯α)=m\mathrm{cl}(\overline{\mathrm{M}}_{\alpha})=m. Then using [2, Proposition 1.7] we have cat(M¯α)=m+1\mathrm{cat}(\overline{\mathrm{M}}_{\alpha})=m+1. Now it is easy to see that the product j=1r(Rjm|J|iJwi)\prod_{j=1}^{r}(R_{j}^{m-|J|}\prod_{i\in J}w_{i}) is nonzero, giving us cl(M¯αr)=rm\mathrm{cl}(\overline{\mathrm{M}}_{\alpha}^{r})=rm for any postive integer rr. Consequently, cat(M¯αr)=rm+1\mathrm{cat}(\overline{\mathrm{M}}_{\alpha}^{r})=rm+1. The desired inequality then follows from (2). ∎

In most cases, to establish lower bounds for TC(M¯α)\mathrm{TC}(\overline{\mathrm{M}}_{\alpha}), Davis utilized two key group homomorphisms, ϕ\phi and ψ\psi. Specifically, ϕ:Hm(M¯α;2)2\phi:H^{m}(\overline{\mathrm{M}}_{\alpha};\mathbb{Z}_{2})\rightarrow\mathbb{Z}_{2} is the Poincaré duality isomorphism, and by choosing an uniform homomorphism ψ:Hm1(M¯α;2)2\psi:H^{m-1}(\overline{\mathrm{M}}_{\alpha};\mathbb{Z}_{2})\rightarrow\mathbb{Z}_{2} such that (ϕψ)(z)0(\phi\otimes\psi)(z)\neq 0, where zH2m1(M¯α×M¯α;2)z\in H^{2m-1}(\overline{\mathrm{M}}_{\alpha}\times\overline{\mathrm{M}}_{\alpha};\mathbb{Z}_{2}) a product of (2m1)(2m-1)-many cohomology classes from the ker(Δ)\mathrm{ker}(\Delta^{*}). Thus, one can conclude that zcl(X)2m1\mathrm{zcl}(X)\geq 2m-1.

For example, if the genetic code of a length vector α\alpha is {a,a+b,n}\langle\{a,a+b,n\}\rangle, then the uniform homomorphism in the sense of Davis is a homomorphism ψ:Hm1(M¯α)2\psi:H^{m-1}(\overline{\mathrm{M}}_{\alpha})\to\mathbb{Z}_{2} satisfying

  • ψ(Vi)=ψ(Vj)\psi(V_{i})=\psi(V_{j}) if i,jai,j\leq a or if a<i,ja+ba<i,j\leq a+b, and

  • ψ(ViVj)=ψ(ViVk)\psi(V_{i}V_{j})=\psi(V_{i}V_{k}) if j,kaj,k\leq a or if a<j,ka+ba<j,k\leq a+b.

(Above, we have omitted writing powers of RR accompanying VV’s as done by Davis in [7]).

Our main strategy is to determine lower bounds for TCk(M¯α)\mathrm{TC}_{k}(\overline{\mathrm{M}}_{\alpha}) as follows. We will construct classes ξH2m1(M¯α×M¯α;2)l\xi\in H^{2m-1}(\overline{\mathrm{M}}_{\alpha}\times\overline{\mathrm{M}}_{\alpha};\mathbb{Z}_{2})^{\otimes l} when k=2lk=2l and ξH2m1(M¯α×M¯α;2)lHm(M¯α;2)\xi^{\prime}\in H^{2m-1}(\overline{\mathrm{M}}_{\alpha}\times\overline{\mathrm{M}}_{\alpha};\mathbb{Z}_{2})^{\otimes l}\otimes H^{m}(\overline{\mathrm{M}}_{\alpha};\mathbb{Z}_{2}) when k=2l+1k=2l+1 such that (ϕψ)l(ξ)0(\phi\otimes\psi)^{\otimes l}(\xi)\neq 0 and ((ϕψ)lϕ)(ξ)0((\phi\otimes\psi)^{\otimes l}\otimes\phi)(\xi^{\prime})\neq 0. Here the map (ϕψ)l(\phi\otimes\psi)^{\otimes l} is given by the usual tensor product of maps as

(ϕψ)l:(Hm(M¯α;2)Hm1(M¯α;2))l2l2,(\phi\otimes\psi)^{\otimes l}:(H^{m}(\overline{\mathrm{M}}_{\alpha};\mathbb{Z}_{2})\otimes H^{m-1}(\overline{\mathrm{M}}_{\alpha};\mathbb{Z}_{2}))^{\otimes l}\rightarrow\mathbb{Z}_{2}^{\otimes l}\rightarrow\mathbb{Z}_{2},

defined by (ϕψ)l(w1wl)=(ϕψ)(w1)(ϕψ)(wl)(\phi\otimes\psi)^{\otimes l}(w_{1}\otimes\cdots\otimes w_{l})=(\phi\otimes\psi)(w_{1})\cdots(\phi\otimes\psi)(w_{l}) for w1,,wlHm(M¯α;2)Hm1(M¯α;2)w_{1},\dots,w_{l}\in H^{m}(\overline{\mathrm{M}}_{\alpha};\mathbb{Z}_{2})\otimes H^{m-1}(\overline{\mathrm{M}}_{\alpha};\mathbb{Z}_{2}), and (ϕψ)(ww′′)=ϕ(w)ψ(w′′)(\phi\otimes\psi)(w^{\prime}\otimes w^{\prime\prime})=\phi(w^{\prime})\psi(w^{\prime\prime}) for wHm(M¯α;2)w^{\prime}\in H^{m}(\overline{\mathrm{M}}_{\alpha};\mathbb{Z}_{2}), w′′Hm1(M¯α;2)w^{\prime\prime}\in H^{m-1}(\overline{\mathrm{M}}_{\alpha};\mathbb{Z}_{2}). Also, we will use the following homomorphisms

(ϕψ)jl:=11ϕψ11,(\phi\otimes\psi)_{j}^{\otimes l}:=1\otimes\cdots\otimes 1\otimes\phi\otimes\psi\otimes 1\otimes\cdots\otimes 1, (7)

where the 2j12j-1 th and 2j2j th entries are the maps ϕ\phi and ψ\psi respectively, for 1jl1\leq j\leq l.

In [7], Davis has explicitly shown that TC(M¯α){2m,2m+1}\mathrm{TC}(\overline{\mathrm{M}}_{\alpha})\in\{2m,2m+1\}, for the genetic code α\alpha having gene sizes from the table below. For genes of size up to 44, the following table depicts the number of possible genetic codes for the first few values of nn.

Gene Size n=5n=5 n=6n=6 n=7n=7 n=8n=8
2 4 5 6 7
3 5 15 21
4 4 21
3,3 15 35
4,3 Type 1 8 20
4,3 Type 2 10 10
3,3,3 Type 2 1 1
4,3,3 Type 2 14 14
4,3,3,3 Type 2 2 2
anything, 2 8 55 559
Table 1. Number of occurrences.

A genetic code S,S\langle S,S^{\prime}\rangle is said to be of Type 1 if 1SS1\in S\cap S^{\prime}. For the case of n=7n=7 and onward, we need the notion of Type 2. These are the genetic codes of specified size and which are not of Type 1 and their gees are those of a genetic code with n=7n=7. Note that there are 2727 genetic codes of Type 22 as highlighted in Table˜1. We refer the reader to [7, Table 2] for the list of all possible gees of the genetic codes of Type 22 cases. For example, {1,2,4,n},{3,4,n},{2,5,n},{1,6,n}\langle\{1,2,4,n\},\{3,4,n\},\{2,5,n\},\{1,6,n\}\rangle is the genetic code of Type 22 having genes of size 44, 3,3,33,3,3 (see [7] for more details).

3. The case of genetic code having gene of size 2

In this section, we obtain sharp lower bounds on TCk(M¯α)\mathrm{TC}_{k}(\overline{\mathrm{M}}_{\alpha}) when the genetic code of α\alpha has a gene of size 22.

Since M¯α\overline{\mathrm{M}}_{\alpha} is a connected smooth manifold of dimension m=n3m=n-3, we know that TCk(M¯α)km+1\mathrm{TC}_{k}(\overline{\mathrm{M}}_{\alpha})\leq km+1. In [7, Theorem 2.2], Davis has shown that the TC(M¯α){2m,2m+1}\mathrm{TC}(\overline{\mathrm{M}}_{\alpha})\in\{2m,2m+1\} if the genetic code of α\alpha is {a,n}\langle\{a,n\}\rangle. We will now generalize this result in the higher setting.

Theorem 3.1.

Let {a,n}\langle\{a,n\}\rangle be the genetic code of α\alpha. Then

kmk2+1TCk(M¯α)km+1.km-\bigg{\lfloor}\frac{k}{2}\bigg{\rfloor}+1\leq\mathrm{TC}_{k}(\overline{\mathrm{M}}_{\alpha})\leq km+1. (8)

Moreover, if m=2tm=2^{t} for some non-negative integer tt, then

kmTCk(M¯α)km+1.km\leq\mathrm{TC}_{k}(\overline{\mathrm{M}}_{\alpha})\leq km+1. (9)
Proof.

For k=2k=2, to obtain the lower bound on TC(M¯α)\mathrm{TC}(\overline{\mathrm{M}}_{\alpha}), Davis proved that w¯2mR¯2m10\bar{w}_{2}^{m}\bar{R}_{2}^{m-1}\neq 0 by expanding it in bidegree (m1,m)(m-1,m) as follows:

w¯2mR¯2m1=βR1m2w1R2m1w2+R1m1R2m1w2,\bar{w}_{2}^{m}\bar{R}_{2}^{m-1}=\beta R_{1}^{m-2}w_{1}R_{2}^{m-1}w_{2}+R_{1}^{m-1}R_{2}^{m-1}w_{2},

where β=(2m1m1)+1\beta=\binom{2m-1}{m-1}+1, and using the linear independence of {R1m1,R1m2w1}\{R_{1}^{m-1},R_{1}^{m-2}w_{1}\} together with the fact that R2m1w20R_{2}^{m-1}w_{2}\neq 0.

We now build on this idea to argue in the general case k2k\geq 2. We first prove the left inequality of (8) when k=2lk=2l. Consider the following product of cohomology classes:

A¯l:=j=1l(w¯2j)m(R¯2j)m1.\bar{A}_{l}:=\prod\limits_{j=1}^{l}(\bar{w}_{2j})^{m}(\bar{R}_{2j})^{m-1}~.

We now expand (w¯2j)m(R¯2j)m1(\bar{w}_{2j})^{m}(\bar{R}_{2j})^{m-1} in the multidegree (0,,0,m1,m,0,,0)(0,\dots,0,m-1,m,0,\dots,0), where m1m-1 and mm appear at 2j12j-1-th and 2j2j-th position, respectively. We then have

(w¯2j)m(R¯2j)m1=βR2j1m2w2j1R2jm1w2j+R2j1m1R2jm1w2j.(\bar{w}_{2j})^{m}(\bar{R}_{2j})^{m-1}=\beta R_{2j-1}^{m-2}w_{2j-1}R_{2j}^{m-1}w_{2j}+R_{2j-1}^{m-1}R_{2j}^{m-1}w_{2j}. (10)

By comparing the 2j12j-1 th position in (10) and applying Remark 2.6, the above product is non-zero. Now, A¯l\bar{A}_{l} has the following expression:

A¯l=j=1l(βR2j1m2w2j1R2jm1w2j+R2j1m1R2jm1w2j).\bar{A}_{l}=\prod_{j=1}^{l}\left(\beta R_{2j-1}^{m-2}w_{2j-1}R_{2j}^{m-1}w_{2j}+R_{2j-1}^{m-1}R_{2j}^{m-1}w_{2j}\right).

Observe that the above expansion contains the following (m1,m,m1,m,,m1,m)(m-1,m,m-1,m,\dots,m-1,m)-multidegree term

j=1lR2j1m1w2jR2jm1.\prod_{j=1}^{l}R_{2j-1}^{m-1}w_{2j}R_{2j}^{m-1}.

Every element in the above product is non-zero, which follows from the Remark˜2.6 and the fact V1Rm10V_{1}R^{m-1}\neq 0. Moreover, this expression can not be annihilated by any other term, since no other term in the expansion of A¯l\bar{A}_{l} is devoid of V1V_{1} in the odd positions.

Thus,

zclk(M¯α)l(m+m1)=kmk2\mathrm{zcl}_{k}(\overline{\mathrm{M}}_{\alpha})\geq l(m+m-1)=km-\bigg{\lfloor}\frac{k}{2}\bigg{\rfloor}

holds, and using the relation (3) we get the required result.

Next, we obtain the desired result for k=2l+1k=2l+1. Consider the following product of cohomology classes:

A¯l(w¯2l+1R¯2l+1m1).\bar{A}_{l}\left(\overline{w}_{2l+1}\overline{R}_{2l+1}^{m-1}\right).

Observe that the (m1,m,m1,m,,m1,m)(m-1,m,m-1,m,\dots,m-1,m)-multidegree term of the above product is A¯lw2l+1R2l+1m1\bar{A}_{l}w_{2l+1}R^{m-1}_{2l+1}, which is nonzero. This gives us the inequality

zclk(M¯α)l(m+m1)+m=kmk2\mathrm{zcl}_{k}(\overline{\mathrm{M}}_{\alpha})\geq l(m+m-1)+m=km-\bigg{\lfloor}\frac{k}{2}\bigg{\rfloor}

giving us the desired lower bound of (8).

Now, we give the sharp lower bound of TCk(M¯α)\mathrm{TC}_{k}(\overline{\mathrm{M}}_{\alpha}), when m=2tm=2^{t} for some non-negative integer tt. We will now prove the left inequality of (9) for the case k=2lk=2l.

We consider another product of cohomology classes:

C¯l=j=1l1(w2j1+w2j+1),\bar{C}_{l}=\prod\limits_{j=1}^{l-1}(w_{2j-1}+w_{2j+1}),

for l2l\geq 2, and C¯1\bar{C}_{1} is the empty product. Thus for l=1l=1, we have A¯1C¯1=A¯10\bar{A}_{1}\bar{C}_{1}=\bar{A}_{1}\neq 0.

Note that multiplying the class A¯l\bar{A}_{l} by the class (w2j1+w2j+1)(w_{2j-1}+w_{2j+1}) brings V1V_{1} into the (2j+1)(2j+1)-th position, keeping it non-zero. Our claim is

A¯lC¯l=(β+1)j=1,j odd 2li=1,ij2lwii=12lRim1.\bar{A}_{l}\bar{C}_{l}=(\beta+1)\sum\limits_{j=1,j\text{~odd~}}^{2l}\prod\limits_{i=1,i\neq j}^{2l}w_{i}\prod\limits_{i=1}^{2l}R_{i}^{m-1}. (11)

We induct on ll. Suppose the statement is true for l=l=\ell. Then A¯+1C¯+1\bar{A}_{\ell+1}\bar{C}_{\ell+1} is

j=1+1(βw2jR2jm1w2j1R2j1m2+w2jR2jm1R2j1m1)j=1(w2j1+w2j+1)=A¯C¯(βw2+2R2+2m1w2+1R2+1m2+w2+2R2+2m1R2+1m1)(w21+w2+1)=(β+1)(j=1,j odd2i=1,ij2wii=12Rim1)(w21w2+2R2+1m1R2+2m1+w2+1w2+2R2+1m1R2+2m1)=(β+1)j=1,j odd 2+2i=1,ij2+2wii=12+2Rim1.\displaystyle\begin{split}&\prod\limits_{j=1}^{\ell+1}\left(\beta w_{2j}R_{2j}^{m-1}w_{2j-1}R_{2j-1}^{m-2}+w_{2j}R_{2j}^{m-1}R_{2j-1}^{m-1}\right)\prod\limits_{j=1}^{\ell}\left(w_{2j-1}+w_{2j+1}\right)\\ &=\bar{A}_{\ell}\bar{C}_{\ell}\left(\beta w_{2\ell+2}R_{2\ell+2}^{m-1}w_{2\ell+1}R_{2\ell+1}^{m-2}+w_{2\ell+2}R_{2\ell+2}^{m-1}R_{2\ell+1}^{m-1}\right)\left(w_{2\ell-1}+w_{2\ell+1}\right)\\ &=(\beta+1)\left(\sum_{j=1,j\text{~odd}}^{2\ell}\prod\limits_{i=1,i\neq j}^{2\ell}w_{i}\prod\limits_{i=1}^{2\ell}R_{i}^{m-1}\right)\left(w_{2\ell-1}w_{2\ell+2}R_{2\ell+1}^{m-1}R_{2\ell+2}^{m-1}+w_{2\ell+1}w_{2\ell+2}R_{2\ell+1}^{m-1}R_{2\ell+2}^{m-1}\right)\\ &=(\beta+1)\sum\limits_{j=1,j\text{~odd~}}^{2\ell+2}\prod\limits_{i=1,i\neq j}^{2\ell+2}w_{i}\prod\limits_{i=1}^{2\ell+2}R_{i}^{m-1}.\end{split}

In the penultimate step, we have used the induction hypothesis and the fact that β(β+1)\beta(\beta+1) is even. Therefore, whenever (β+1)0(\beta+1)\neq 0, it follows that A¯lC¯l0\bar{A}_{l}\bar{C}_{l}\neq 0, providing us :

zclk(M¯α)l(m+m1)+(l1)=2lm1=km1.\displaystyle\mathrm{zcl}_{k}(\overline{\mathrm{M}}_{\alpha})\geq l(m+m-1)+(l-1)=2lm-1=km-1.

We can impose conditions on mm to deduce when (β+1)0(\beta+1)\neq 0. For m=2tm=2^{t}, Lucas’s theorem implies that the binomial coefficient β\beta is even. Thus, in this situation, we get

TCk(M¯α)zclk(M¯α)+1km.\mathrm{TC}_{k}(\overline{\mathrm{M}}_{\alpha})\geq\mathrm{zcl}_{k}(\overline{\mathrm{M}}_{\alpha})+1\geq km.

We now address the case when k=2l+1k=2l+1. Consider the element

D:=A¯lC¯l(w¯2l+1R¯2l+1m1).D:=\bar{A}_{l}\bar{C}_{l}\left(\overline{w}_{2l+1}\overline{R}_{2l+1}^{m-1}\right).

Using (11) we have the expression:

(β+1)j=1,j odd 2li=1,ij2lwii=12lRim1(w¯2l+1R¯2l+1m1).\displaystyle(\beta+1)\sum\limits_{j=1,j\text{~odd~}}^{2l}\prod\limits_{i=1,i\neq j}^{2l}w_{i}\prod\limits_{i=1}^{2l}R_{i}^{m-1}\left(\overline{w}_{2l+1}\overline{R}_{2l+1}^{m-1}\right).

The rightmost term expands as (w2l+1+w2l)(r=0m1(m1r)R2l+1rR2lmr1)(w_{2l+1}+w_{2l})\left(\sum_{r=0}^{m-1}\binom{m-1}{r}R_{2l+1}^{r}R_{2l}^{m-r-1}\right). We consider only the term w2l+1R2l+1m1w_{2l+1}R_{2l+1}^{m-1} from the expansion. Observe that A¯lC¯l(w2l+1R2l+1m1)0\bar{A}_{l}\bar{C}_{l}\left(w_{2l+1}R_{2l+1}^{m-1}\right)\neq 0 and therefore D0D\neq 0, as the aforementioned term cannot be annihilated by any other term from the expansion of DD, due to positional differences. Hence,

zclk(M¯α)l(2m1)+l1+1+m1=m(2l+1)1=km1\displaystyle\mathrm{zcl}_{k}(\overline{\mathrm{M}}_{\alpha})\geq l(2m-1)+l-1+1+m-1=m(2l+1)-1=km-1

Consequently, TCk(M¯α)zclk(M¯α)+1km,\mathrm{TC}_{k}(\overline{\mathrm{M}}_{\alpha})\geq\mathrm{zcl}_{k}(\overline{\mathrm{M}}_{\alpha})+1\geq km, when m=2tm=2^{t} for some t{0}t\in\mathbb{N}\cup\{0\}. ∎

Remark 3.2.

For β:=(2m1m1)+1\beta:=\binom{2m-1}{m-1}+1, by Lucas’s theorem β+10\beta+1\neq 0 ((mod 2)2) when m=2tm=2^{t} for some t{0}t\in\mathbb{N}\cup\{0\}.

For genetic codes having a gene of size two of the length vector α\alpha, Davis in [7, Theorem 2.3] has shown that TC(M¯α){2m,2m+1}\mathrm{TC}(\overline{\mathrm{M}}_{\alpha})\in\{2m,2m+1\}. We will now prove a higher analogue of this result. In particular, we have:

Theorem 3.3.

Suppose the length vector α\alpha has a genetic code with genes {a,n}\{a,n\}. Then

kmk2+1TCk(M¯α)km+1.km-\bigg{\lfloor}\frac{k}{2}\bigg{\rfloor}+1\leq\mathrm{TC}_{k}(\overline{\mathrm{M}}_{\alpha})\leq km+1.
Proof.

First, we prove our assertion when k=2lk=2l. Note that, there exists a non-empty subset {i1,,it}[b1]\{i_{1},\dots,i_{t}\}\subset[b-1] such that ϕ(RmtVi1Vit)=1\phi(R^{m-t}V_{i_{1}}\cdots V_{i_{t}})=1, as done in the proof of [7, Theorem 2.3], where ϕ:Hm(M¯α;2)2\phi:H^{m}(\overline{\mathrm{M}}_{\alpha};\mathbb{Z}_{2})\to\mathbb{Z}_{2} is the Poincaré duality isomorphism. Define

Aj:=w¯2j,i1m+1tw¯2j,i2w¯2j,itR¯2jm1,wherewj,is:=pj(Vis)andw¯j,is=wj,is+wj1,is.A_{j}:=\bar{w}_{2j,i_{1}}^{m+1-t}\bar{w}_{2j,i_{2}}\cdots\bar{w}_{2j,i_{t}}\bar{R}_{2j}^{m-1},~\text{where}~{w}_{j,i_{s}}:=p_{j}^{*}(V_{i_{s}})~\text{and}~\bar{w}_{j,i_{s}}={w}_{j,i_{s}}+{w}_{j-1,i_{s}}.

Now using similar arguments as presented in the proof of [7, Theorem 2.3], we show that the product A¯l:=j=1lAj\bar{A}_{l}:=\prod_{j=1}^{l}A_{j} is nonzero. Recall from the proof of [7, Theorem 2.3] that there is a homomorphism ψ:Hm1(M¯α;2)2\psi:H^{m-1}(\overline{\mathrm{M}}_{\alpha};\mathbb{Z}_{2})\to\mathbb{Z}_{2} which sends Rm1R^{m-1} and Rm2VbR^{m-2}V_{b} to 11 and all other monomials to zero.

In the following we apply (ϕψ)l(\phi\otimes\psi)^{\otimes l} on A¯l\bar{A}_{l}, using the homomorphisms (7).

(ϕψ)l(A¯l)=j=1l(ϕψ)jl(Aj)=j=1l(ϕψ)jl(w¯2j,i1m+1tw¯2j,i2w¯2j,itR¯2jm1)=j=1lϕ(w2j1,i1m+1tw2j1,i2w2j1,it)ψ(R2jm1)=1.\displaystyle\begin{split}(\phi\otimes\psi)^{\otimes l}(\bar{A}_{l})&=\prod_{j=1}^{l}(\phi\otimes\psi)^{\otimes l}_{j}(A_{j})\\ &=\prod_{j=1}^{l}(\phi\otimes\psi)^{\otimes l}_{j}(\bar{w}_{2j,i_{1}}^{m+1-t}\bar{w}_{2j,i_{2}}\cdots\bar{w}_{2j,i_{t}}\bar{R}_{2j}^{m-1})\\ &=\prod_{j=1}^{l}\phi({w}_{2j-1,i_{1}}^{m+1-t}{w}_{2j-1,i_{2}}\cdots{w}_{2j-1,i_{t}})\psi({R}_{2j}^{m-1})=1.\end{split}

This implies A¯l0\bar{A}_{l}\neq 0, thereby completing the proof of our assertion.

Now we consider the case k=2l+1k=2l+1. Consider the product:

A¯lw¯2l+1,i1w¯2l+1,i2w¯2l+1,itR¯2l+1mt.\bar{A}_{l}\bar{w}_{2l+1,i_{1}}\bar{w}_{2l+1,i_{2}}\dots\bar{w}_{2l+1,i_{t}}\bar{R}_{2l+1}^{m-t}.

Observe that the (m,m1,m,m1,m)(m,m-1,\dots m,m-1,m)-multidegree term of the above product is A¯lw2l+1,i1w2l+1,i2w2l+1,itR2l+1mt\bar{A}_{l}{w}_{2l+1,i_{1}}{w}_{2l+1,i_{2}}\dots w_{2l+1,i_{t}}{R}_{2l+1}^{m-t}. Applying (ϕψ)lϕ(\phi\otimes\psi)^{\otimes l}\otimes\phi on the previous product obtains:

((ϕψ)lϕ)(A¯lw2l+1,i1w2l+1,i2w2l+1,itR2l+1mt)=(ϕψ)l(A¯l)ϕ(w2l+1,i1w2l+1,i2w2l+1,itR2l+1mt)=1.\displaystyle\begin{split}&\left((\phi\otimes\psi)^{\otimes l}\otimes\phi\right)\left(\bar{A}_{l}{w}_{2l+1,i_{1}}{w}_{2l+1,i_{2}}\dots w_{2l+1,i_{t}}{R}_{2l+1}^{m-t}\right)\\ &=(\phi\otimes\psi)^{\otimes l}\left(\bar{A}_{l}\right)\phi\left({w}_{2l+1,i_{1}}{w}_{2l+1,i_{2}}\dots w_{2l+1,i_{t}}{R}_{2l+1}^{m-t}\right)\\ &=1.\end{split}

This completes the proof. ∎

Next we compute sharp bounds on TCk(M¯α)\mathrm{TC}_{k}(\overline{\mathrm{M}}_{\alpha}), when the genetic code of α\alpha contains a gene {a,n}\{a,n\}. In particular, using the result [3, Example 4.15] for the genetic code {2,4,n},{a,n}\langle\{2,4,n\},\{a,n\}\rangle we get the following.

Proposition 3.4.

Let a5a\geq 5 be odd and {2,4,n},{a,n}\langle\{2,4,n\},\{a,n\}\rangle be the genetic code of the length vector α\alpha. If mm is a 22-power, then

kmTCk(M¯α)km+1.km\leq\mathrm{TC}_{k}(\overline{\mathrm{M}}_{\alpha})\leq km+1.
Proof.

Using [3, Example 4.15] we get that RmR^{m} is non-zero if aa is odd. Thus,

R¯2m1=i=02m1(2m1i)RiR2mi1=(β+1)(Rm1Rm+RmRm1),\displaystyle\begin{split}\bar{R}^{2m-1}=&~\sum_{i=0}^{2m-1}\binom{2m-1}{i}R^{i}\otimes R^{2m-i-1}\\ =&~(\beta+1)(R^{m-1}\otimes R^{m}+R^{m}\otimes R^{m-1}),\end{split} (12)

is a non-zero cohomology class (using Remark 3.2), where the class R¯:=R1+1RH(M¯α×M¯α,2)\bar{R}:=R\otimes 1+1\otimes R\in H^{*}(\overline{\mathrm{M}}_{\alpha}\times\overline{\mathrm{M}}_{\alpha},\mathbb{Z}_{2}).

Assume kk to be even, say k=2lk=2l. Consider the following product of cohomology classes:

  • A¯l:=j=1l(R¯2j)2m1=(β+1)lj=1l(R2j1m1R2jm+R2j1mR2jm1)\bar{A}_{l}:=\prod\limits_{j=1}^{l}(\bar{R}_{2j})^{2m-1}=(\beta+1)^{l}\prod\limits_{j=1}^{l}(R_{2j-1}^{m-1}R_{2j}^{m}+R_{2j-1}^{m}R_{2j}^{m-1}), for ll\in\mathbb{N} (using (12)).

  • C¯l=j=1l1(R2j1+R2j+1)\bar{C}_{l}=\prod\limits_{j=1}^{l-1}(R_{2j-1}+R_{2j+1}) for l2l\geq 2, and C¯1\bar{C}_{1} is the empty product.

We inductively show that for each ll\in\mathbb{N}, the following identity holds.

A¯lC¯l=(β+1)l(j=12(Rjm1i=1,ij2Rim)).\bar{A}_{l}\bar{C}_{l}=(\beta+1)^{l}\left(\sum_{j=1}^{2\ell}(R_{j}^{m-1}\prod\limits_{i=1,i\neq j}^{2\ell}R_{i}^{m})\right)~. (13)

Consider the case for l=1l=1, the identity A¯1C¯1=R¯22m1=R1m1R2m+R2m1R1m\bar{A}_{1}\bar{C}_{1}=\bar{R}_{2}^{2m-1}=R_{1}^{m-1}R_{2}^{m}+R_{2}^{m-1}R_{1}^{m} holds.

Assuming the required identity holds for some l=l=\ell\in\mathbb{N}, we show that it also holds for the case of +1\ell+1. Note that A¯+1C¯+1\bar{A}_{\ell+1}\bar{C}_{\ell+1} expands as

A¯C¯R¯2+22m1(R21+R2+1)=(β+1)(j=12(Rjm1i=1,ij2Rim))(β+1)(R2+1m1R2+2m+R2+1mR2+2m1)(R21+R2+1)=(β+1)+1(j=12(Rjm1i=1,ij2Rim))(R21R2+1m1R2+2m+R21R2+1mR2+2m1+R2+1mR2+2m)=(β+1)+1(R2+1m1i=1,i2+12+2Rim+R2+2m1i=1,i2+22+2Rim+j=12(Rjm1i=1,ij2+2Rim))=(β+1)+1(j=12+2(Rjm1i=1,ij2+2Rim)).\displaystyle\begin{split}&\bar{A}_{\ell}\bar{C}_{\ell}\bar{R}_{2\ell+2}^{2m-1}\left(R_{2\ell-1}+R_{2\ell+1}\right)\\ &=(\beta+1)^{\ell}\left(\sum_{j=1}^{2\ell}(R_{j}^{m-1}\prod\limits_{i=1,i\neq j}^{2\ell}R_{i}^{m})\right)(\beta+1)(R_{2\ell+1}^{m-1}R_{2\ell+2}^{m}+R_{2\ell+1}^{m}R_{2\ell+2}^{m-1})\left(R_{2\ell-1}+R_{2\ell+1}\right)\\ &=(\beta+1)^{\ell+1}\left(\sum_{j=1}^{2\ell}(R_{j}^{m-1}\prod\limits_{i=1,i\neq j}^{2\ell}R_{i}^{m})\right)(R_{2\ell-1}R_{2\ell+1}^{m-1}R_{2\ell+2}^{m}+R_{2\ell-1}R_{2\ell+1}^{m}R_{2\ell+2}^{m-1}+R_{2\ell+1}^{m}R_{2\ell+2}^{m})\\ &=(\beta+1)^{\ell+1}\left(R_{2\ell+1}^{m-1}\prod\limits_{i=1,i\neq 2\ell+1}^{2\ell+2}R_{i}^{m}+R_{2\ell+2}^{m-1}\prod\limits_{i=1,i\neq 2\ell+2}^{2\ell+2}R_{i}^{m}+\sum_{j=1}^{2\ell}(R_{j}^{m-1}\prod\limits_{i=1,i\neq j}^{2\ell+2}R_{i}^{m})\right)\\ &=(\beta+1)^{\ell+1}\left(\sum_{j=1}^{2\ell+2}(R_{j}^{m-1}\prod\limits_{i=1,i\neq j}^{2\ell+2}R_{i}^{m})\right).\end{split}

Thus, we have proved the identity (13). Each term in the sum has a different multidegree and contains RmR^{m} at every position except one, where we have Rm1R^{m-1}. This implies A¯lC¯l\bar{A}_{l}\bar{C}_{l} is nonzero whenever β\beta is even, more precisely when m=2tm=2^{t} (see Remark˜3.2). Notice that, to get a lower bound for TCk(M¯α)\mathrm{TC}_{k}(\overline{\mathrm{M}}_{\alpha}) where k=2lk=2l, we consider the non trivial cohomology class A¯lC¯l\bar{A}_{l}\bar{C}_{l} of length l(2m1)+(l1)=km1l(2m-1)+(l-1)=km-1, and thus TCk(M¯α)km\mathrm{TC}_{k}(\overline{\mathrm{M}}_{\alpha})\geq km.

For kk odd, say k=2l+1k=2l+1, consider the following product A¯lC¯lR¯2+1m.\bar{A}_{l}\bar{C}_{l}\bar{R}_{2\ell+1}^{m}.

A¯lC¯lR¯2l+1m=(β+1)l(j=12l(Rjm1i=1,ij2lRim))R¯2l+1m.\displaystyle\begin{split}\bar{A}_{l}\bar{C}_{l}\bar{R}_{2l+1}^{m}=(\beta+1)^{l}\left(\sum_{j=1}^{2l}(R_{j}^{m-1}\prod\limits_{i=1,i\neq j}^{2l}R_{i}^{m})\right)\bar{R}_{2l+1}^{m}.\end{split}

Observe that the product will contain (β+1)l(j=12l(Rjm1i=1,ij2lRim))R2l+1m,(\beta+1)^{l}\left(\sum_{j=1}^{2l}(R_{j}^{m-1}\prod\limits_{i=1,i\neq j}^{2l}R_{i}^{m})\right){R}_{2l+1}^{m}, which is non-zero since its every term is of distinct multidegree and contains RmR^{m} at every position except one, where we have Rm1R^{m-1}. ∎

4. The case of monogenic code of size 3

In this section, we will derive sharp bounds on TCk(M¯α)\mathrm{TC}_{k}(\overline{\mathrm{M}}_{\alpha}) when the genetic code of α\alpha is {a,a+b,n}\langle\{a,a+b,n\}\rangle. To proceed, we first recall some relevant notions and results from the proof of [7, Theorem 2.4]. Building on these, we extend the result to the higher setting.

We now recall the following notations from [7].

  1. (1)

    Y1:=RrViY_{1}:=R^{r}V_{i} with iai\leq a,

  2. (2)

    Y2:=RrViY_{2}:=R^{r}V_{i} with a<ia+ba<i\leq a+b,

  3. (3)

    Y1,1:=RrViVjY_{1,1}:=R^{r}V_{i}V_{j} with i<jai<j\leq a,

  4. (4)

    Y1,2:=RrViVjY_{1,2}:=R^{r}V_{i}V_{j} with ia<ja+bi\leq a<j\leq a+b and

  5. (5)

    YSY_{S} refers to RrVSR^{r}V_{S} for an appropriate value of rr and S[a+b]S\subseteq[a+b] is a subgee.

Let ϕw:=ϕ(Yw)\phi_{w}:=\phi(Y_{w}) and ψw:=ψ(Yw)\psi_{w}:=\psi(Y_{w}) for all possible subscripts ww (see Section˜2). Then it was shown in [7, Proposition 2.5] that ϕ\phi satisfies the following identities:

ϕ1,1=ϕ1,2=1,ϕ2=a1,ϕ1=a+b,ϕ0=(a1)b+(a12).\phi_{1,1}=\phi_{1,2}=1,~\phi_{2}=a-1,~\phi_{1}=a+b,~\phi_{0}=(a-1)b+\binom{a-1}{2}.

We now briefly recall Davis’s strategy to construct the uniform homomorphism ψ\psi and describe two equations that come from the relation (6). Note that if {i,j}\{i,j\} is a subgee of the genetic code {a,a+b,n}\langle\{a,a+b,n\}\rangle, then either 1i<ja1\leq i<j\leq a or 1ia<ja+b1\leq i\leq a<j\leq a+b. We denote by R1,1\textbf{R}_{1,1}, a relation (6) corresponding to the subgees of the first kind and by R1,2\textbf{R}_{1,2} for the subgees of another kind. Then counting subgees (or the cohomology classes of type Y1Y_{1}, Y2Y_{2}, Y1,1Y_{1,1} and Y1,2Y_{1,2}) of an appropriate type we obtain ψ(R1,1)\psi(\textbf{R}_{1,1}) as

ψ0+(a2)ψ1+bψ2+(a22)ψ1,1+(a2)bψ1,2=0\psi_{0}+(a-2)\psi_{1}+b\psi_{2}+\binom{a-2}{2}\psi_{1,1}+(a-2)b\psi_{1,2}=0 (14)

and ψ(R1,2)\psi(\textbf{R}_{1,2}) as

ψ0+(a1)ψ1+(b1)ψ2+(a12)ψ1,1+(a1)(b1)ψ1,2=0.\psi_{0}+(a-1)\psi_{1}+(b-1)\psi_{2}+\binom{a-1}{2}\psi_{1,1}+(a-1)(b-1)\psi_{1,2}=0. (15)

In what follows, we will assume that aa is even. The other cases will follow similarly. Let V¯i=Vi1+1Vi\bar{V}_{i}=V_{i}\otimes 1+1\otimes V_{i} for i=1,a+bi=1,a+b. Then third equation can be obtained by applying ϕψ\phi\otimes\psi to the (m,m1)(m,m-1)-bidegree expansion of V¯12m12tV¯a+bR¯2t1\bar{V}_{1}^{2m-1-2^{t}}\bar{V}_{a+b}\bar{R}^{2^{t}-1} and equating it to 11, where ϕ\phi is the Poincaré duality isomorphism and 2t1<m2t2^{t-1}<m\leq 2^{t}. The (m,m1)(m,m-1)-bidegree expansion of this cohomology class is given by

Y2Y1+Y1,2Rm1+(1+δm,2t)RmY1,2+Y1Y2,Y_{2}\otimes Y_{1}+Y_{1,2}\otimes R^{m-1}+(1+\delta_{m,2^{t}})R^{m}\otimes Y_{1,2}+Y_{1}\otimes Y_{2},

where δ\delta denotes the Kronecker delta. Then ϕψ\phi\otimes\psi acting on the above expansion and equating it to 11 gives the third equation:

ψ0+ψ1+bψ2+εψ1,2=1,\psi_{0}+\psi_{1}+b\psi_{2}+\varepsilon\psi_{1,2}=1, (16)

where ε2\varepsilon\in\mathbb{Z}_{2} is an irrelevant quantity in solving the system of linear equations formed by (14), (15), and (16) with indeterminate ψ\psi. Davis has shown the existence of ψ\psi by proving that the system of linear equations described above has a solution. This shows the non-zeroness of the cohomology class V¯12m12tV¯a+bR¯2t1\bar{V}_{1}^{2m-1-2^{t}}\bar{V}_{a+b}\bar{R}^{2^{t}-1} and consequently gives us the desired lower bound on TC(M¯α)\mathrm{TC}(\overline{\mathrm{M}}_{\alpha}) when the genetic code of α\alpha is {a,a+b,n}\langle\{a,a+b,n\}\rangle with aa even.

We now characterize the values of aa and bb for which the RmR^{m} is non-zero, where mm is the dimension of M¯α\overline{\mathrm{M}}_{\alpha}.

Lemma 4.1.

Let {a,a+b,n}\langle\{a,a+b,n\}\rangle be the genetic code of α\alpha with n>a+b>a>0n>a+b>a>0 and n6n\geq 6. Then Rm0R^{m}\neq 0 if and only if either of the following cases holds:

  1. (1)

    a3(mod4)a\equiv 3~(mod~4),

  2. (2)

    a0(mod4)a\equiv 0~(mod~4) and bb even,

  3. (3)

    a2(mod4)a\equiv 2~(mod~4) and bb odd.

Proof.

Recall that ϕ(Rm)=ϕ0\phi(R^{m})=\phi_{0}, where ϕ\phi is the Poincaré duality isomorphism. Now

ϕ0=(a1)b+(a1)(a2)2.\phi_{0}=(a-1)b+\frac{(a-1)(a-2)}{2}.
  1. (1)

    Suppose a3(mod4)a\equiv 3~(mod~4). Then (a1)b(a-1)b is even and (a1)(a2)(a-1)(a-2) has only one 22 as a factor. Hence (a1)(a2)2\frac{(a-1)(a-2)}{2} is odd.

  2. (2)

    Suppose a0(mod4)a\equiv 0~(mod~4) and bb even. Then (a1)b(a-1)b is even and (a1)(a2)2\frac{(a-1)(a-2)}{2} is odd.

  3. (3)

    Since 44 divides (a2)(a-2), (a1)(a2)2\frac{(a-1)(a-2)}{2} is even. Also bb being odd makes (a1)b(a-1)b odd.

It is easy to check that for all other combinations of values of aa and bb, ϕ0=0\phi_{0}=0. ∎

As a consequence of Lemma˜4.1, we obtain the following sharp bounds.

Theorem 4.2.

Let m=2tm=2^{t} for some non-negative integer tt and {a,a+b,n}\langle\{a,a+b,n\}\rangle be the genetic code of α\alpha such that aa and bb satisfy either of the conditions of Lemma˜4.1. Then

kmTCk(M¯α)km+1.km\leq\mathrm{TC}_{k}(\overline{\mathrm{M}}_{\alpha})\leq km+1.
Proof.

Since RmR^{m} is non-zero by the conditions of Lemma˜4.1, the proof follows using similar techniques as used in proving Proposition˜3.4. ∎

Next we use a higher analogue of the choice of the cohomology classes in [7, Theorem 2.4] to prove the following result.

Theorem 4.3.

Let {a,a+b,n}\langle\{a,a+b,n\}\rangle be the genetic code of α\alpha. Then

kmk2+1TCk(M¯α)km+1.km-\bigg{\lfloor}\frac{k}{2}\bigg{\rfloor}+1\leq\mathrm{TC}_{k}(\overline{\mathrm{M}}_{\alpha})\leq km+1. (17)
Proof.

Suppose that 2t1<m2t2^{t-1}<m\leq 2^{t}. Before we proceed, we set some notations that will be used throughout the proof. Define xi:=pi(Va+b)x_{i}:=p_{i}^{*}(V_{a+b}), x¯i=xi+xi1\bar{x}_{i}=x_{i}+x_{i-1} and Ywi:=pi(Yw){}_{i}Y_{w}:=p^{*}_{i}(Y_{w}) for 1ik1\leq i\leq k. We proceed with the following cases.

Case-1: Suppose aa is even.

Assume k=2lk=2l. Consider the element

A¯l:=j=1lAj,\bar{A}_{l}:=\prod\limits_{j=1}^{l}A_{j},

where Aj=w¯2j2m12tx¯2jR¯2j2t1A_{j}=\bar{w}_{2j}^{2m-1-2^{t}}\bar{x}_{2j}\bar{R}_{2j}^{2^{t}-1}. We apply (ϕψ)l(\phi\otimes\psi)^{\otimes l} on the (m,m1,m,m1,,m,m1)(m,m-1,m,m-1,\dots,m,m-1)-multidegree term in the expansion of A¯l\bar{A}_{l}. Recall the notation (ϕψ)il(\phi\otimes\psi)^{\otimes l}_{i} from (7). Then the identity (ϕψ)il(Ai)=ψ0+ψ1+bψ2+εψ1,2(\phi\otimes\psi)^{\otimes l}_{i}(A_{i})=\psi_{0}+\psi_{1}+b\psi_{2}+\varepsilon\psi_{1,2} holds for all ii (see (16)). Thus, we obtain the following expression:

(ϕψ)l(A¯l)=j=1l(ϕψ)jl(Aj)=(ψ0+ψ1+bψ2+εψ1,2)l.(\phi\otimes\psi)^{\otimes l}(\bar{A}_{l})=\prod_{j=1}^{l}(\phi\otimes\psi)^{\otimes l}_{j}(A_{j})=(\psi_{0}+\psi_{1}+b\psi_{2}+\varepsilon\psi_{1,2})^{l}.

Since (16)\eqref{eq: phipsi acting on expansion} has a solution as shown by Davis, (ϕψ)l(A¯l)=1(\phi\otimes\psi)^{\otimes l}(\bar{A}_{l})=1 has also has a solution. Therefore, A¯l0\bar{A}_{l}\neq 0 for ll\in\mathbb{N}. This gives us the left inequality of (17).

We use similar idea to show that the product A¯lx¯2l+1R¯2l+1m1\bar{A}_{l}\bar{x}_{2l+1}\bar{R}_{2l+1}^{m-1} is nonzero when k=2l+1k=2l+1.

Observe that the type of the (m,m1,,m,m1,m)(m,m-1,\dots,m,m-1,m)-multidegree term of A¯lx¯2l+1R¯2l+1m1\bar{A}_{l}\bar{x}_{2l+1}\bar{R}_{2l+1}^{m-1} is A¯lY22l+1\bar{A}_{l}~{}_{2l+1}Y_{2}. Since ϕ2=a1=1\phi_{2}=a-1=1 and (ϕψ)l(A¯l)=1(\phi\otimes\psi)^{\otimes{l}}(\bar{A}_{l})=1, we get

((ϕψ)lϕ)(A¯lY22l+1)=(ϕψ)l(A¯l)ϕ(2l+1Y2)=1.((\phi\otimes\psi)^{\otimes{l}}\otimes\phi)(\bar{A}_{l}~{}_{2l+1}Y_{2})=(\phi\otimes\psi)^{\otimes{l}}(\bar{A}_{l})\phi(_{2l+1}Y_{2})=1.

This proves A¯lx¯2l+1R¯2l+1m10\bar{A}_{l}\bar{x}_{2l+1}\bar{R}_{2l+1}^{m-1}\neq 0 and thus we get the left inequality of (17).

Case-2: aa is odd and m2t1+1m\neq 2^{t-1}+1.

Then for k=2k=2, Davis proved that V¯1m1V¯a+b2R¯m20\bar{V}_{1}^{m-1}\bar{V}_{a+b}^{2}\bar{R}^{m-2}\neq 0 by showing the existence of uniform homomorphism ψ\psi, and then applying ϕψ\phi\otimes\psi on the expansion of V¯1m1V¯a+b2R¯m2\bar{V}_{1}^{m-1}\bar{V}_{a+b}^{2}\bar{R}^{m-2} in the bidegree (m,m1)(m,m-1). We generalize this idea for general k2k\geq 2.

Assume k=2lk=2l. Consider the element

A¯l:=j=1lAj,\bar{A}_{l}:=\prod\limits_{j=1}^{l}A_{j},

where Aj=w¯2jm1x¯2j2R¯2jm2A_{j}=\bar{w}_{2j}^{m-1}\bar{x}_{2j}^{2}\bar{R}_{2j}^{m-2}. We apply (ϕψ)l(\phi\otimes\psi)^{\otimes l} on the (m,m1,m,m1,,m,m1)(m,m-1,m,m-1,\dots,m,m-1)- multidegree term in the expansion of A¯l\bar{A}_{l}. Similar to Case-1, we obtain the following expression:

(ϕψ)l(A¯l)=j=1l(ϕψ)jl(Aj)=(ψ1+m(b+1)ψ2+εψ1,2)l.(\phi\otimes\psi)^{\otimes l}(\bar{A}_{l})=\prod_{j=1}^{l}(\phi\otimes\psi)^{\otimes l}_{j}(A_{j})=(\psi_{1}+m(b+1)\psi_{2}+\varepsilon^{\prime}\psi_{1,2})^{l}.

In [7, Theorem 2.4], Davis has shown that ψ1+m(b+1)ψ2+εψ1,2=1\psi_{1}+m(b+1)\psi_{2}+\varepsilon^{\prime}\psi_{1,2}=1 has solution. Here, the value of ε\varepsilon^{\prime} is again irrelevant. Consequently, (ϕψ)l(A¯l)=1(\phi\otimes\psi)^{\otimes l}(\bar{A}_{l})=1 has a solution. Therefore, A¯l0\bar{A}_{l}\neq 0 for ll\in\mathbb{N}. This gives us the left inequality of (17).

We use a similar idea to show that the product A¯lw¯2l+1x¯2l+1R¯2l+1m2\bar{A}_{l}\bar{w}_{2l+1}\bar{x}_{2l+1}\bar{R}_{2l+1}^{m-2} is nonzero when k=2l+1k=2l+1. Observe that the type of the (m,m1,,m,m1,m)(m,m-1,\dots,m,m-1,m)-multidegree term of A¯lw¯2l+1x¯2l+1R¯2l+1m2\bar{A}_{l}\bar{w}_{2l+1}\bar{x}_{2l+1}\bar{R}_{2l+1}^{m-2} is A¯lY1,22l+1\bar{A}_{l}~{}_{2l+1}Y_{1,2}. Since ϕ1,2=1\phi_{1,2}=1 and (ϕψ)l(A¯l)=1(\phi\otimes\psi)^{\otimes{l}}(\bar{A}_{l})=1, we get

((ϕψ)lϕ)(A¯lY1,22l+1)=(ϕψ)l(A¯l)ϕ(2l+1Y1,2)=1.((\phi\otimes\psi)^{\otimes{l}}\otimes\phi)(\bar{A}_{l}~{}_{2l+1}Y_{1,2})=(\phi\otimes\psi)^{\otimes{l}}(\bar{A}_{l})\phi(_{2l+1}Y_{1,2})=1.

This proves A¯lw¯2l+1x¯2l+1R¯2l+1m20\bar{A}_{l}\bar{w}_{2l+1}\bar{x}_{2l+1}\bar{R}_{2l+1}^{m-2}\neq 0 and thus we get the left inequality of (17).

Case-3: aa is odd and m=2t1+1m=2^{t-1}+1.

Then for k=2k=2, Davis proved that V¯1mV¯a+bm10\bar{V}_{1}^{m}\bar{V}_{a+b}^{m-1}\neq 0 by constructing a uniform homomorphism ψ\psi and applying ϕψ\phi\otimes\psi to the expansion of V¯1mV¯a+bm1\bar{V}_{1}^{m}\bar{V}_{a+b}^{m-1} in the bidegree (m,m1)(m,m-1). We use this idea to prove our assertion for k2k\geq 2.

Assume k=2lk=2l. Consider the element

A¯l:=j=1lAj,\bar{A}_{l}:=\prod\limits_{j=1}^{l}A_{j},

where Aj=w¯2jmx¯2jm1A_{j}=\bar{w}_{2j}^{m}\bar{x}_{2j}^{m-1}. We apply (ϕψ)l(\phi\otimes\psi)^{\otimes l} on the (m,m1,m,m1,,m,m1)(m,m-1,m,m-1,\dots,m,m-1)- multidegree term in the expansion of A¯l\bar{A}_{l}. Similar to the previous cases, we obtain the following expression:

(ϕψ)l(A¯l)=j=1l(ϕψ)jl(Aj)=(ψ1+(b+1)ψ2)l.(\phi\otimes\psi)^{\otimes l}(\bar{A}_{l})=\prod_{j=1}^{l}(\phi\otimes\psi)^{\otimes l}_{j}(A_{j})=(\psi_{1}+(b+1)\psi_{2})^{l}.

In [7, Theorem 2.4], Davis has shown that ψ1+(b+1)ψ2=1\psi_{1}+(b+1)\psi_{2}=1 has solution. Consequently, (ϕψ)l(A¯l)=1(\phi\otimes\psi)^{\otimes l}(\bar{A}_{l})=1 has a solution. Therefore, A¯l0\bar{A}_{l}\neq 0 for ll\in\mathbb{N}. This gives us the left inequality of (17). The case when kk is odd is exactly the same as the odd case of Case 2. ∎

5. The genetic codes of type 2

In this section, we study the higher topological complexity of planar polygon spaces having genetic codes of Type 22. We refer the reader to Section˜2 for a brief description of these genetic codes.

It turns out that there are exactly 2727 Type 22 genetic codes (see [7, Table 2] for more details). To avoid repetition, we have opted not to include Table 2 from Davis’s paper. For these genetic codes, Davis has computed the zero-divisors-cup lengths. We now briefly describe Davis’s strategy.

Suppose m=n34m=n-3\geq 4 is the dimension of M¯α\overline{\mathrm{M}}_{\alpha}. The genetic codes are distributed into two cases:

Suppose m1m-1 is not a 22-power or if this is the last case of the [7, Table 2] which is the Type 22 genetic code {3,4,n},{2,5,n},{1,6,n}\langle\{3,4,n\},\{2,5,n\},\{1,6,n\}\rangle. Let V¯i=Vi1+1Vi\bar{V}_{i}=V_{i}\otimes 1+1\otimes V_{i} for i=1,2,3i=1,2,3. In this case, Davis proved that V¯1m1V¯22V¯3R¯m30\bar{V}_{1}^{m-1}\bar{V}_{2}^{2}\bar{V}_{3}\bar{R}^{m-3}\neq 0 by constructing a uniform homomorphism ψ\psi, and then applying ϕψ\phi\otimes\psi to the expansion of the cohomology class in bidegree (m,m1)(m,m-1), where ϕ\phi denotes the Poincaré duality isomorphism. The existence of ψ\psi was shown by solving a system of linear equations. We now explain his ideas briefly. The homomorphism ϕψ\phi\otimes\psi acting on the expansion of V¯1m1V¯22V¯3R¯m3\bar{V}_{1}^{m-1}\bar{V}_{2}^{2}\bar{V}_{3}\bar{R}^{m-3} in bidegree (m,m1)(m,m-1) gives us the following:

(ϕ2,3+ϕ1,2,3)ψ1+ϕ1,3(ψ2+ψ1,2)+(m1)ϕ1(ψ2,3+ψ1,2,3)\displaystyle(\phi_{2,3}+\phi_{1,2,3})\psi_{1}+\phi_{1,3}(\psi_{2}+\psi_{1,2})+(m-1)\,\phi_{1}(\psi_{2,3}+\psi_{1,2,3}) (18)
+{ϕ1ψ1,2,3if m is a 2-power,ϕ1,2,3ψ1+ϕ1,3ψ1,2if m1 is a 2-power,0otherwise.\displaystyle\quad+

Davis first computes the values of ϕS\phi_{S}, where ϕ\phi is a Poincaré duality isomorphism and SS is a subgee of a particular type. The key ingredient to achieve this in Davis’s method is to construct a matrix whose columns represent all subgees, including \emptyset, and rows represent all subgees except \emptyset. This matrix is actually a binary matrix in which the entry 11 appears in the places where the corresponding subgees for the row and column are disjoint. Each row corresponding to the subgee of this matrix, in fact, represents the relation 6. The reader is referred to [4, 5] for an illustration. After solving this system of linear equations using MAPLE, Davis obtained the values of ϕS\phi_{S} and observed that ϕ1,2,3=1\phi_{1,2,3}=1, while ϕ1,3=ϕ2,3=0\phi_{1,3}=\phi_{2,3}=0 in the first 2626 cases of the [7, Table 2].

A similar approach figures out the values of ψS\psi_{S}, except that the rows correspond only to subgees with more than one element, due to the |S|nd2|S|\geq n-d-2 constraint in (6). The uniform homomorphism ψ\psi in each case given in [7, Table 2] satisfies ψ1=1\psi_{1}=1 and ψ2,3=ψ1,2,3=0\psi_{2,3}=\psi_{1,2,3}=0. One can see that in this case (18) equates to 11 for the first 2626 genetic codes of Type 22 whose gees are described in [7, Table 2]. For the final case in [7, Table 2], again using MAPLE one can see that there is a uniform homomorphism ψ\psi for which the only nonzero values are ψ1\psi_{1} and ψ1,6\psi_{1,6}. Moreover, ϕi,j=1\phi_{i,j}=1 for all subgees {i,j}\{i,j\} and ϕ1,2,3=0\phi_{1,2,3}=0 as V1,2,3=0V_{1,2,3}=0. Since ϕ2,3ψ1=1\phi_{2,3}\psi_{1}=1, again (18) equates to 11. Thereby proving the nonzeroness of V¯1m1V¯22V¯3R¯m3\bar{V}_{1}^{m-1}\bar{V}_{2}^{2}\bar{V}_{3}\bar{R}^{m-3}.

If m1m-1 is a 22-power and it is not the last case of [7, Table 2] can be dealt similarly to the previous case. So we don’t repeat the explanation of his strategy.

Theorem 5.1.

Let α\alpha has the genetic code of type T2T_{2} in the Table˜1 and m4m\geq 4, then:

kmk2+1TCk(M¯α)km+1.km-\bigg{\lfloor}\frac{k}{2}\bigg{\rfloor}+1\leq\mathrm{TC}_{k}(\overline{\mathrm{M}}_{\alpha})\leq km+1. (19)
Proof.

Recall from Section˜2 that wj=pj(V1)w_{j}=p_{j}^{*}(V_{1}). Similarly, we define uj:=pj(V2)u_{j}:=p_{j}^{*}(V_{2}) and vj:=pj(V3)v_{j}:=p_{j}^{*}(V_{3}). Similar to Davis, we distribute the genetic codes into two cases.

Case 1: Suppose m1m-1 is not a 22-power or α\alpha has gees as in the final case in [7, Table 2].

Consider the element

A¯l:=j=1lAj,\bar{A}_{l}:=\prod\limits_{j=1}^{l}A_{j},

where Aj=w¯2jm1u¯2j2v¯2jR¯2jm3A_{j}=\bar{w}_{2j}^{m-1}\bar{u}_{2j}^{2}\bar{v}_{2j}\bar{R}_{2j}^{m-3}. We apply (ϕψ)l(\phi\otimes\psi)^{\otimes l} on the (m,m1,m,m1,,m,m1)(m,m-1,m,m-1,\dots,m,m-1)-multidegree term in the expansion of A¯l\bar{A}_{l}. When mm is a power of 22, we obtain the following:

(ϕψ)l(A¯l)=j=1l(ϕψ)jl(Aj)=((ϕ2,3+ϕ1,2,3)ψ1+ϕ1,3(ψ2+ψ1,2)+(m1)ϕ1(ψ2,3+ψ1,2,3)+ϕ1ψ1,2,3)l.\displaystyle\begin{split}&(\phi\otimes\psi)^{\otimes l}(\bar{A}_{l})=\prod_{j=1}^{l}(\phi\otimes\psi)^{\otimes l}_{j}(A_{j})\\ &=((\phi_{2,3}+\phi_{1,2,3})\psi_{1}+\phi_{1,3}(\psi_{2}+\psi_{1,2})+(m-1)\phi_{1}(\psi_{2,3}+\psi_{1,2,3})+\phi_{1}\psi_{1,2,3})^{l}.\end{split}

Then the above expression is equal to 11, since ϕ1,2,3=ψ1=1\phi_{1,2,3}=\psi_{1}=1 as explained at the beginning of this section.

Next, when m1m-1 is a power of 22 (and the gee is of the final type in [7, Table 2]), we obtain the following:

(ϕψ)l(A¯l)=j=1l(ϕψ)jl(Aj)=((ϕ2,3+ϕ1,2,3)ψ1+ϕ1,3(ψ2+ψ1,2)+(m1)ϕ1(ψ2,3+ψ1,2,3)+ϕ1,2,3ψ1+ϕ1,3ψ1,2)l.\displaystyle\begin{split}&(\phi\otimes\psi)^{\otimes l}(\bar{A}_{l})=\prod_{j=1}^{l}(\phi\otimes\psi)^{\otimes l}_{j}(A_{j})\\ &=((\phi_{2,3}+\phi_{1,2,3})\psi_{1}+\phi_{1,3}(\psi_{2}+\psi_{1,2})+(m-1)\phi_{1}(\psi_{2,3}+\psi_{1,2,3})+\phi_{1,2,3}\psi_{1}+\phi_{1,3}\psi_{1,2})^{l}.\end{split}

For the 2727-th gee in [7, Table 2], both the above two expressions (m1m-1 is a 22-power and not a 22-power) become 11, since there is homomorphism ψ\psi whose only non-zero values are ψ1\psi_{1} and ψ1,6\psi_{1,6}. Moreover, ϕi,j=1\phi_{i,j}=1 for all {i,j}\{i,j\} and ϕ1,2,3=0\phi_{1,2,3}=0. Therefore, only ϕ2,3ψ1\phi_{2,3}\psi_{1} survives.

Finally, in all other situations under this case, we get the following:

(ϕψ)l(A¯l)=j=1l(ϕψ)jl(Aj)=((ϕ2,3+ϕ1,2,3)ψ1+ϕ1,3(ψ2+ψ1,2)+(m1)ϕ1(ψ2,3+ψ1,2,3))l.\displaystyle\begin{split}&(\phi\otimes\psi)^{\otimes l}(\bar{A}_{l})=\prod_{j=1}^{l}(\phi\otimes\psi)^{\otimes l}_{j}(A_{j})\\ &=((\phi_{2,3}+\phi_{1,2,3})\psi_{1}+\phi_{1,3}(\psi_{2}+\psi_{1,2})+(m-1)\phi_{1}(\psi_{2,3}+\psi_{1,2,3}))^{l}.\end{split}

The above expression is equal to 11 for the first 2626 gees in [7, Table 2], since ϕ1,2,3=ψ1=1\phi_{1,2,3}=\psi_{1}=1. For the 2727-th gee, it is again 11, thanks to ϕ2,3ψ1\phi_{2,3}\psi_{1}.

Now we consider the cases when kk is odd, say k=2l+1k=2l+1. Recall the element A¯l\bar{A}_{l} that we used in this case of k=2lk=2l was A¯l:=j=1lAj,\bar{A}_{l}:=\prod\limits_{j=1}^{l}A_{j}, where

Aj=w¯2jm1u¯2j2v¯2jR¯2jm3.A_{j}=\bar{w}_{2j}^{m-1}\bar{u}_{2j}^{2}\bar{v}_{2j}\bar{R}_{2j}^{m-3}.

The nonzeroness of the product A¯lw¯2l+1u¯2l+1v¯2l+1R¯2l+1m3\bar{A}_{l}\bar{w}_{2l+1}\bar{u}_{2l+1}\bar{v}_{2l+1}\bar{R}_{2l+1}^{m-3} is given as follows. Consider the classes Ywi:=pi(Yw){}_{i}Y_{w}:=p^{*}_{i}(Y_{w}). Observe that the type of the (m,m1,,m,m1,m)(m,m-1,\dots,m,m-1,m)-multidegree term of A¯lw¯2l+1u¯2l+1v¯2l+1R¯2l+1m3\bar{A}_{l}\bar{w}_{2l+1}\bar{u}_{2l+1}\bar{v}_{2l+1}\bar{R}_{2l+1}^{m-3} is A¯lY1,2,32l+1\bar{A}_{l}~{}_{2l+1}Y_{1,2,3}. Since ϕ1,2,3=1\phi_{1,2,3}=1 in first 2626 cases of [7, Table 2] and (ϕψ)l(A¯l)=1(\phi\otimes\psi)^{\otimes{l}}(\bar{A}_{l})=1, we get

((ϕψ)lϕ)(A¯lY1,2,32l+1)=(ϕψ)l(A¯l)ϕ(2l+1Y1,2,3)=1.((\phi\otimes\psi)^{\otimes{l}}\otimes\phi)(\bar{A}_{l}~{}_{2l+1}Y_{1,2,3})=(\phi\otimes\psi)^{\otimes{l}}(\bar{A}_{l})\phi(_{2l+1}Y_{1,2,3})=1.

This proves A¯lw¯2l+1u¯2l+1v¯2l+1R¯2l+1m30\bar{A}_{l}\bar{w}_{2l+1}\bar{u}_{2l+1}\bar{v}_{2l+1}\bar{R}_{2l+1}^{m-3}\neq 0 and thus we get the left inequality of (19). For the 2727-th case we consider A¯lw¯2l+1u¯2l+12R¯2l+1m3\bar{A}_{l}\bar{w}_{2l+1}\bar{u}_{2l+1}^{2}\bar{R}_{2l+1}^{m-3}. Clearly, the type of the (m,m1,,m,m1,m)(m,m-1,\dots,m,m-1,m)-multidegree term of A¯lw¯2l+1u¯2l+12R¯2l+1m3\bar{A}_{l}\bar{w}_{2l+1}\bar{u}_{2l+1}^{2}\bar{R}_{2l+1}^{m-3} is A¯lY1,22l+1\bar{A}_{l}~{}_{2l+1}Y_{1,2}. Since ϕ1,2=1\phi_{1,2}=1, we obtain that

((ϕψ)lϕ)(A¯lY1,22l+1)=(ϕψ)l(A¯l)ϕ(2l+1Y1,2)=1.((\phi\otimes\psi)^{\otimes{l}}\otimes\phi)(\bar{A}_{l}~{}_{2l+1}Y_{1,2})=(\phi\otimes\psi)^{\otimes{l}}(\bar{A}_{l})\phi(_{2l+1}Y_{1,2})=1.

This gives us the left inequality of (19).

Case 2: Suppose m1m-1 is a 22-power and α\alpha has doesn’t have gees as in last case of [7, Table 2].

For k=2k=2, Davis proved that V¯1mV¯22V¯3R¯m40\bar{V}_{1}^{m}\bar{V}_{2}^{2}\bar{V}_{3}\bar{R}^{m-4}\neq 0 by constructing a uniform homomorphism ψ\psi, and then applying ϕψ\phi\otimes\psi to the expansion of the cohomology class in bidegree (m,m1)(m,m-1), where ϕ\phi denotes the Poincaré duality isomorphism. We generalize this idea for general k2k\geq 2.

First, assume that kk is even, say k=2lk=2l. Consider the element

A¯l:=j=1lAj,\bar{A}_{l}:=\prod\limits_{j=1}^{l}A_{j},

where Aj=w¯2jmu¯2j2v¯2jR¯2jm4A_{j}=\bar{w}_{2j}^{m}\bar{u}_{2j}^{2}\bar{v}_{2j}\bar{R}_{2j}^{m-4}. We apply (ϕψ)l(\phi\otimes\psi)^{\otimes l} on the (m,m1,m,m1,,m,m1)(m,m-1,m,m-1,\dots,m,m-1)-multidegree term in the expansion of A¯l\bar{A}_{l}. As a result, we obtain the following expression:

(ϕψ)l(A¯l)=j=1l(ϕψ)jl(Aj)=(ϕ1(ψ2,3+ψ1,2,3)+ϕ1,2,3ψ1+ϕ1,3ψ1,2)l.(\phi\otimes\psi)^{\otimes l}(\bar{A}_{l})=\prod_{j=1}^{l}(\phi\otimes\psi)^{\otimes l}_{j}(A_{j})=(\phi_{1}(\psi_{2,3}+\psi_{1,2,3})+\phi_{1,2,3}\psi_{1}+\phi_{1,3}\psi_{1,2})^{l}.

Note that, ψ1,2,3=ψ2,3=ϕ1,3=ϕ2,3=0\psi_{1,2,3}=\psi_{2,3}=\phi_{1,3}=\phi_{2,3}=0 and ϕ1,2,3=ψ1=1\phi_{1,2,3}=\psi_{1}=1 as explained in the beginning of this section. Hence, the above expression becomes 11, giving us the desired lower bound of (19).

Now we consider the cases when kk is odd, say k=2l+1k=2l+1. Recall the element A¯l\bar{A}_{l} that we used in case 11 of k=2lk=2l was A¯l:=j=1lAj,\bar{A}_{l}:=\prod\limits_{j=1}^{l}A_{j}, where

Aj=w¯2jmu¯2j2v¯2jR¯2jm4.A_{j}=\bar{w}_{2j}^{m}\bar{u}_{2j}^{2}\bar{v}_{2j}\bar{R}_{2j}^{m-4}.

We want to show that the product A¯lw¯2l+1u¯2l+1v¯2l+1R¯2l+1m3\bar{A}_{l}\bar{w}_{2l+1}\bar{u}_{2l+1}\bar{v}_{2l+1}\bar{R}_{2l+1}^{m-3} is nonzero. Consider the classes Ywi:=pi(Yw){}_{i}Y_{w}:=p^{*}_{i}(Y_{w}). Observe that the type of the (m,m1,,m,m1,m)(m,m-1,\dots,m,m-1,m)-multidegree term of A¯lw¯2l+1u¯2l+1v¯2l+1R¯2l+1m3\bar{A}_{l}\bar{w}_{2l+1}\bar{u}_{2l+1}\bar{v}_{2l+1}\bar{R}_{2l+1}^{m-3} is A¯lY1,2,32l+1\bar{A}_{l}~{}_{2l+1}Y_{1,2,3}. Since ϕ1,2,3=1\phi_{1,2,3}=1 and (ϕψ)l(A¯l)=1(\phi\otimes\psi)^{\otimes{l}}(\bar{A}_{l})=1, we get

((ϕψ)lϕ)(A¯lY1,2,32l+1)=(ϕψ)l(A¯l)ϕ(2l+1Y1,2,3)=1.((\phi\otimes\psi)^{\otimes{l}}\otimes\phi)(\bar{A}_{l}~{}_{2l+1}Y_{1,2,3})=(\phi\otimes\psi)^{\otimes{l}}(\bar{A}_{l})\phi(_{2l+1}Y_{1,2,3})=1.

This proves A¯lw¯2l+1u¯2l+1v¯2l+1R¯2l+1m30\bar{A}_{l}\bar{w}_{2l+1}\bar{u}_{2l+1}\bar{v}_{2l+1}\bar{R}_{2l+1}^{m-3}\neq 0 and we get the left inequality of (19).

6. Monogenic codes of size 4

In this section, we obtain bounds on the higher topological complexity of M¯α\overline{\mathrm{M}}_{\alpha}, where the genetic code α\alpha is {a,a+b,a+b+c,n}\langle\{a,a+b,a+b+c,n\}\rangle with n>a+b+c>a+b>a>0n>a+b+c>a+b>a>0.

Recall that, for a Poincaré duality isomorphism ϕ:Hm(M¯α;2)2\phi:H^{m}(\overline{\mathrm{M}}_{\alpha};\mathbb{Z}_{2})\to\mathbb{Z}_{2}, we have ϕ0=ϕ(Rm)\phi_{0}=\phi(R^{m}). In [7, Theorem 4.1], Davis has obtained the expression for ϕ0\phi_{0} as follows

ϕ0=(a2)(a+b+c1)+(a1)((b2)+(b1)(c1)).\phi_{0}=\binom{a}{2}(a+b+c-1)+(a-1)\left(\binom{b}{2}+(b-1)(c-1)\right).

The following result classifies values of a,b,ca,b,c in the genetic code {a,a+b,a+b+c,n}\langle\{a,a+b,a+b+c,n\}\rangle for which the values of ϕ0\phi_{0} are odd.

Lemma 6.1.

ϕ0\phi_{0} is odd if and only if any of the following conditions hold:

  1. (1)

    b+cb+c even, aa even, a+b0(mod4)a+b~\equiv 0~(\text{mod}~4);

  2. (2)

    b+cb+c even, aa even, c1(mod4)c~\equiv 1~(\text{mod}~4);

  3. (3)

    b+cb+c odd and a3(mod4)a~\equiv 3~(\text{mod}~4);

  4. (4)

    b+cb+c odd, b2,3(mod4)b~\equiv 2,3~(\text{mod}~4), and a0,2(mod4)a~\equiv 0,2~(\text{mod}~4);

  5. (5)

    a,b2(mod4)a,b~\equiv 2~(\text{mod}~4).

Proof.

(1 and 2) Under the assumption that both aa and b+cb+c are even, one can write ϕ0\phi_{0} as

(a2)+(b2)+bc+1+a(b+c).\binom{a}{2}+\binom{b}{2}+bc+1+a(b+c).

By [7, Lemma 4.2], the last expression is odd under the above-mentioned conditions.

(3) For a3(mod4)a~\equiv 3~(\text{mod}~4), (a1)((b2)+(b1)(c1))=0(a-1)\left(\binom{b}{2}+(b-1)(c-1)\right)=0. Moreover, we have (a2)=1\binom{a}{2}=1 and a+b+c1a+b+c-1 odd. Hence, the result follows.

(4) For b+cb+c odd, ϕ0\phi_{0} becomes

a(a2)+(a1)(b2).a\binom{a}{2}+(a-1)\binom{b}{2}.

When a0,2(mod4)a~\equiv 0,2~(\text{mod}~4), it becomes (b2)\binom{b}{2}, which is odd if and only if b2,3(mod4)b~\equiv 2,3~(\text{mod}~4).

(5) For a,b2(mod4)a,b~\equiv 2~(\text{mod}~4), both (a2)\binom{a}{2} and a1a-1 are odd. Again, (b2)\binom{b}{2} and b1b-1 both are odd. Consequently, there are two odd multiples of c1c-1 and the only surviving odd term is (a1)(b2).(a-1)\binom{b}{2}.

The previous lemma helps us to obtain the sharp bounds on TCk(M¯α)\mathrm{TC}_{k}(\overline{\mathrm{M}}_{\alpha}).

Theorem 6.2.

Let mm be a 22-power and {n,a+b+c,a+b,a}\langle\{n,a+b+c,a+b,a\}\rangle be the genetic code of α\alpha satisfying the conditions in Lemma˜6.1. Then

kmTCk(Mα)km+1.km\leq\mathrm{TC}_{k}(\mathrm{M}_{\alpha})\leq km+1.
Proof.

Note that if a,b,ca,b,c satisfies conditions in Lemma˜6.1, then Rm0R^{m}\neq 0. Then the proof of our assertion is similar to that of Proposition˜3.4. ∎

We denote the intervals [1,a][1,a], (a,a+b](a,a+b] and (a+b,a+b+c](a+b,a+b+c] by I1I_{1}, I2I_{2} and I3I_{3}, respectively and classify all possible subgees by tuples of sizes 0, 11, 22 and 33 (see the proof of [7, Theorem 4.1] for more details). The tuple (p)(p) of size one represents a subgee containing one element, and the location of this element is decided by the value of pp. The tuple (p,q)(p,q) of size two represents a subgee {i,j}\{i,j\} and the locations of ii and jj are decided by the values of pp and qq, respectively. For example, (1,2)(1,2) refers to a subgee of cardinality two, {i,j}\{i,j\} such that iI1i\in I_{1} and jI2j\in I_{2}. Similarly, the tuple (p,q,r)(p,q,r) represents a subgee {i,j,k}\{i,j,k\} such that the locations of i,j,ki,j,k are decided by the values of p,q,rp,q,r. For example, (1,2,2)(1,2,2) represents a subgee {i,j,k}\{i,j,k\} such that iI1i\in I_{1} and j,kI2j,k\in I_{2}. The following classification of types of elements that can be subgees of the genetic code {a,a+b,a+b+c,n}\{a,a+b,a+b+c,n\} has been given by Davis in the proof of [7, Theorem 4.1]:

𝒮={,(1),(2),(3),(1,1),(1,2),(1,3),(2,2),(2,3),(1,1,1),(1,1,2),(1,1,3),(1,2,2),(1,2,3)}.\mathcal{S}=\{\emptyset,(1),(2),(3),(1,1),(1,2),(1,3),(2,2),(2,3),(1,1,1),(1,1,2),(1,1,3),(1,2,2),(1,2,3)\}.

Next, we aim to obtain the bounds on TCk(M¯α)\mathrm{TC}_{k}(\overline{\mathrm{M}}_{\alpha}) when m>4m>4. For that purpose, we need subgees of size greater equal 22. Such sub-gees are described as follows:

𝒮={(1,1),(1,2),(1,3),(2,2),(2,3),(1,1,1),(1,1,2),(1,1,3),(1,2,2),(1,2,3)}.\mathcal{S^{\prime}}=\{(1,1),(1,2),(1,3),(2,2),(2,3),(1,1,1),(1,1,2),(1,1,3),(1,2,2),(1,2,3)\}.

Now for U𝒮U\in\mathcal{S}, let uiu_{i} be the number of ii’s in UU for 1i31\leq i\leq 3. For example, if U=(1,2,2)U=(1,2,2), then u1=1u_{1}=1 and u2=2u_{2}=2. For U𝒮U^{\prime}\in\mathcal{S^{\prime}}, applying ψ:Hm1(M¯α;2)2\psi:H^{m-1}(\overline{\mathrm{M}}_{\alpha};\mathbb{Z}_{2})\to\mathbb{Z}_{2} on the relation (3)(3) of Theorem˜2.4 we obtain the following expression

U𝒮(au1u1)(bu2u2)(cu3u3)ψU,\sum_{U\in\mathcal{S^{\prime}}}\binom{a-u_{1}^{\prime}}{u_{1}}\binom{b-u_{2}^{\prime}}{u_{2}}\binom{c-u_{3}^{\prime}}{u_{3}}\psi_{U}, (20)

where ψU=ψ(YU)\psi_{U}=\psi(Y_{U}).

We are now in a position to state our general result.

Theorem 6.3.

With the genetic code of α\alpha being as stated above, we have

kmk2+1TCk(M¯α)km+1km-\bigg{\lfloor}\frac{k}{2}\bigg{\rfloor}+1\leq\mathrm{TC}_{k}(\overline{\mathrm{M}}_{\alpha})\leq km+1 (21)

in the following situations:

  1. (1)

    m>4m>4, a,b1(mod4)a,b\equiv~1~(\text{mod}~4) and cc odd;

  2. (2)

    m>3m>3, a2(mod4)a\equiv~2~(\text{mod}~4), b4(mod4)b\equiv~4~(\text{mod}~4) and cc odd.

Proof.

In the 11-st situation, following the proof of [7, Proposition 4.3], we know that there exist a uniform homomorphism ψ:Hm1(M¯α;2)2\psi:H^{m-1}(\overline{\mathrm{M}}_{\alpha};\mathbb{Z}_{2})\to\mathbb{Z}_{2} which sends Yi,jY_{i,j} to 11 and other monomials to 0. Now to achieve our desired assertion, we will consider two cases depending on whether m2m-2 is not a 22-power.

Recall that we have wj=pj(V1)w_{j}=p_{j}^{*}(V_{1}) and xj=pj(Va+b)x_{j}=p_{j}^{*}(V_{a+b}). We now define yj:=pj(Va+b+c)y_{j}:=p_{j}^{*}(V_{a+b+c}). Case 1: Suppose m2m-2 is not a 22-power.
Consider the element

A¯l:=j=1lAj,\bar{A}_{l}:=\prod\limits_{j=1}^{l}A_{j},

where Aj=w¯2jm2x¯2j2y¯2j3R¯2jm4A_{j}=\bar{w}_{2j}^{m-2}\bar{x}_{2j}^{2}\bar{y}_{2j}^{3}\bar{R}_{2j}^{m-4}. We deal with the case when kk is even, say k=2lk=2l. We apply (ϕψ)l(\phi\otimes\psi)^{\otimes l} on the (m,m1,m,m1,,m,m1)(m,m-1,m,m-1,\dots,m,m-1)-multidegree term in the expansion of A¯l\bar{A}_{l}. Let us inspect that particular multidegree term of AjA_{j}, which can get mapped non-trivially.

i=1m3(m2i)(m4mi4)w2j1ix2j12y2j12R2j1mi4w2jm2iy2jR2ji+i=1m3(m2i)(m4mi3)w2j1ix2j12y2j1R2j1mi3w2jm2iy2j2R2ji1\displaystyle\begin{split}&\sum_{i=1}^{m-3}\binom{m-2}{i}\binom{m-4}{m-i-4}{w}_{2j-1}^{i}{x}_{2j-1}^{2}{y}_{2j-1}^{2}{R}_{2j-1}^{m-i-4}w_{2j}^{m-2-i}y_{2j}R_{2j}^{i}\\ &+\sum_{i=1}^{m-3}\binom{m-2}{i}\binom{m-4}{m-i-3}{w}_{2j-1}^{i}{x}_{2j-1}^{2}{y}_{2j-1}{R}_{2j-1}^{m-i-3}w_{2j}^{m-2-i}y_{2j}^{2}R_{2j}^{i-1}\end{split}

All terms of the above sum are of the type Y1,2,3Y1,3Y_{1,2,3}\otimes Y_{1,3}. In the first situation, it follows from [7, Theorem 4.1] that ϕ\phi sends each Yi,j,kY_{i,j,k} to 11 and all other monomials to 0. Now applying ϕψ\phi\otimes\psi to AjA_{j}, we obtain:

i=1m3(m2i)(m4mi4)ϕ(Y1,2,3)ψ(Y1,3)+i=1m3(m2i)(m4mi3)ϕ(Y1,2,3)ψ(Y1,3)=[i=1m3(m2i)(m4mi4)+i=1m3(m2i)(m4mi3)]ϕ1,2,3ψ1,3=(2m6m4)(m20)(m4m4)+(2m6m3)=(2m6m4)+(2m6m3)+1\displaystyle\begin{split}&\sum_{i=1}^{m-3}\binom{m-2}{i}\binom{m-4}{m-i-4}\phi(Y_{1,2,3})\psi(Y_{1,3})+\sum_{i=1}^{m-3}\binom{m-2}{i}\binom{m-4}{m-i-3}\phi(Y_{1,2,3})\psi(Y_{1,3})\\ &=\left[\sum_{i=1}^{m-3}\binom{m-2}{i}\binom{m-4}{m-i-4}+\sum_{i=1}^{m-3}\binom{m-2}{i}\binom{m-4}{m-i-3}\right]\phi_{1,2,3}\psi_{1,3}\\ &=\binom{2m-6}{m-4}-\binom{m-2}{0}\binom{m-4}{m-4}+\binom{2m-6}{m-3}=\binom{2m-6}{m-4}+\binom{2m-6}{m-3}+1\\ \end{split}

which is 1, by Lucas’s theorem. Therefore, we obtain:

(ϕψ)l(A¯l)=j=1l(ϕψ)j(Aj)=1.(\phi\otimes\psi)^{\otimes l}(\bar{A}_{l})=\prod_{j=1}^{l}(\phi\otimes\psi)_{j}(A_{j})=1.

Case 2: Suppose m2m-2 is a 22-power.
Consider the element

A¯l:=j=1lAj,\bar{A}_{l}:=\prod\limits_{j=1}^{l}A_{j},

where Aj=w¯2jm1x¯2j2y¯2j3R¯2jm5A_{j}=\bar{w}_{2j}^{m-1}\bar{x}_{2j}^{2}\bar{y}_{2j}^{3}\bar{R}_{2j}^{m-5}. Again, we start with kk even, say k=2lk=2l. We inspect the suitable multidegree term of AjA_{j}, which can get mapped non-trivially under ϕψ\phi\otimes\psi.

i=1m2(m1i)(m5mi4)w2j1ix2j12y2j12R2j1mi4w2jm1iy2jR2ji1+i=1m2(m1i)(m5mi3)w2j1ix2j12y2j1R2j1mi3w2jm1iy2j2R2ji2.\displaystyle\begin{split}&\sum_{i=1}^{m-2}\binom{m-1}{i}\binom{m-5}{m-i-4}{w}_{2j-1}^{i}{x}_{2j-1}^{2}{y}_{2j-1}^{2}{R}_{2j-1}^{m-i-4}w_{2j}^{m-1-i}y_{2j}R_{2j}^{i-1}\\ &+\sum_{i=1}^{m-2}\binom{m-1}{i}\binom{m-5}{m-i-3}{w}_{2j-1}^{i}{x}_{2j-1}^{2}{y}_{2j-1}{R}_{2j-1}^{m-i-3}w_{2j}^{m-1-i}y_{2j}^{2}R_{2j}^{i-2}.\end{split}

All the terms in the above sum are of the type Y1,2,3Y1,3Y_{1,2,3}\otimes Y_{1,3}. Hence, after applying ϕψ\phi\otimes\psi, it becomes, just similar to the last case:

i=1m2(m1i)(m5mi4)ϕ(Y1,2,3)ψ(Y1,3)+i=1m2(m1i)(m5mi3)ϕ(Y1,2,3)ψ(Y1,3)=[i=1m2(m1i)(m5mi4)+i=1m2(m1i)(m5mi3)]ϕ1,2,3ψ1,3=(2m6m4)+(2m6m3)=(2m6m4)\displaystyle\begin{split}&\sum_{i=1}^{m-2}\binom{m-1}{i}\binom{m-5}{m-i-4}\phi(Y_{1,2,3})\psi(Y_{1,3})+\sum_{i=1}^{m-2}\binom{m-1}{i}\binom{m-5}{m-i-3}\phi(Y_{1,2,3})\psi(Y_{1,3})\\ &=\left[\sum_{i=1}^{m-2}\binom{m-1}{i}\binom{m-5}{m-i-4}+\sum_{i=1}^{m-2}\binom{m-1}{i}\binom{m-5}{m-i-3}\right]\phi_{1,2,3}\psi_{1,3}\\ &=\binom{2m-6}{m-4}+\binom{2m-6}{m-3}=\binom{2m-6}{m-4}\\ \end{split}

which is 11, again by Lucas’s Theorem. Therefore, we obtain:

(ϕψ)l(A¯l)=j=1l(ϕψ)j(Aj)=1.(\phi\otimes\psi)^{\otimes l}(\bar{A}_{l})=\prod_{j=1}^{l}(\phi\otimes\psi)_{j}(A_{j})=1.

For both case-1 and case-2, when kk is odd, say k=2l+1k=2l+1, we consider the following element

A¯lw¯2l+1x¯2l+12y¯2l+1R¯2l+1m4.\bar{A}_{l}\bar{w}_{2l+1}\bar{x}_{2l+1}^{2}\bar{y}_{2l+1}\bar{R}_{2l+1}^{m-4}.

Then applying (ϕψ)lϕ(\phi\otimes\psi)^{\otimes l}\otimes\phi on an appropriate multidegree term of the above product, we get

(ϕψ)l(A¯l)ϕ(w2l+1x2+12y2l+1R2l+1m4)=ϕ1,2,3=1,(\phi\otimes\psi)^{\otimes l}(\bar{A}_{l})\phi(w_{2l+1}x_{2\ell+1}^{2}y_{2l+1}R_{2l+1}^{m-4})=\phi_{1,2,3}=1,

and thus completing the proof of the first situation.

Again in the second situation, following the proof of [7, Proposition 4.4] there exists a uniform homomorphism ψ:Hm1(M¯α;2)2\psi:H^{m-1}(\overline{\mathrm{M}}_{\alpha};\mathbb{Z}_{2})\rightarrow\mathbb{Z}_{2} which sends Yi,jY_{i,j} to 11 and other monomials to 0. Similarly, as in the first situation, we consider two cases depending on whether m2m-2 is a 22-power.

Case 1: Suppose m2m-2 is a 22-power.
Consider the element

A¯l:=j=1lAj,\bar{A}_{l}:=\prod\limits_{j=1}^{l}A_{j},

where Aj=w¯2j2x¯2j2y¯2jm1R¯2jm4A_{j}=\bar{w}_{2j}^{2}\bar{x}_{2j}^{2}\bar{y}_{2j}^{m-1}\bar{R}_{2j}^{m-4}. Start with kk even, say k=2lk=2l. We expand AjA_{j} in the suitable multidegree, which can get mapped non-trivially under (ϕψ)j(\phi\otimes\psi)_{j}.

Aj=i=1m2(m1i)(m4mi2)x2j12y2j1iR2j1mi2w2j2y2jm1iR2ji2.A_{j}=\sum_{i=1}^{m-2}\binom{m-1}{i}\binom{m-4}{m-i-2}{x}_{2j-1}^{2}{y}_{2j-1}^{i}{R}_{2j-1}^{m-i-2}w_{2j}^{2}y_{2j}^{m-1-i}R_{2j}^{i-2}.

Note that the above term is of the type Y2,3Y1,3Y_{2,3}\otimes Y_{1,3}. Thus applying ϕψ\phi\otimes\psi, we get:

i=1m2(m1i)(m4mi2)ϕ2,3ψ1,3=i=0m2(m1i)(m4mi2)=(2m5m2).\sum_{i=1}^{m-2}\binom{m-1}{i}\binom{m-4}{m-i-2}\phi_{2,3}\psi_{1,3}=\sum_{i=0}^{m-2}\binom{m-1}{i}\binom{m-4}{m-i-2}=\binom{2m-5}{m-2}.

Note that under the 22-nd situation we have ϕ2,3=1\phi_{2,3}=1 by [7, Theorem 4.1]. As m2m-2 is a 22-power, the binomial coefficient is of the form

(2t+2t1++2+12t).\binom{2^{t}+2^{t-1}+\cdots+2+1}{2^{t}}.

Hence, by Lucas’s theorem, it follows that (2m5m2)\binom{2m-5}{m-2} is odd. Therefore, we obtain:

(ϕψ)l(A¯l)=j=1l(ϕψ)j(Aj)=1.(\phi\otimes\psi)^{\otimes l}(\bar{A}_{l})=\prod_{j=1}^{l}(\phi\otimes\psi)_{j}(A_{j})=1.

Case 2: Suppose m2m-2 is not a 22-power.
Consider the element A¯l:=j=1lAj,\bar{A}_{l}:=\prod\limits_{j=1}^{l}A_{j}, where Aj=w¯2j2x¯2j2y¯2jm2R¯2jm3A_{j}=\bar{w}_{2j}^{2}\bar{x}_{2j}^{2}\bar{y}_{2j}^{m-2}\bar{R}_{2j}^{m-3}.

Start with kk even, say k=2lk=2l. We inspect the suitable multidegree term of AjA_{j}, which can get mapped non-trivially under ϕψ\phi\otimes\psi.

Aj=i=1m3(m2i)(m3mi2)x2j12y2j1iR2j1mi2w2j2y2jm2iR2ji1.A_{j}=\sum_{i=1}^{m-3}\binom{m-2}{i}\binom{m-3}{m-i-2}{x}_{2j-1}^{2}{y}_{2j-1}^{i}{R}_{2j-1}^{m-i-2}w_{2j}^{2}y_{2j}^{m-2-i}R_{2j}^{i-1}.

Applying ϕψ\phi\otimes\psi to the above Y2,3Y1,3Y_{2,3}\otimes Y_{1,3} type term we get:

i=1m3(m2i)(m3mi2)ϕ2,3ψ1,3=i=0m2(m2i)(m3mi2)(m2m2)(m30)=(2m5m2)+1.\displaystyle\begin{split}\sum_{i=1}^{m-3}\binom{m-2}{i}\binom{m-3}{m-i-2}\phi_{2,3}\psi_{1,3}&=\sum_{i=0}^{m-2}\binom{m-2}{i}\binom{m-3}{m-i-2}-\binom{m-2}{m-2}\binom{m-3}{0}\\ &=\binom{2m-5}{m-2}+1.\end{split}

Again, by Lucas’s Theorem, (2m5m2)\binom{2m-5}{m-2} is even. Therefore, we obtain:

(ϕψ)l(A¯l)=j=1l(ϕψ)j(Aj)=1.(\phi\otimes\psi)^{\otimes l}(\bar{A}_{l})=\prod_{j=1}^{l}(\phi\otimes\psi)_{j}(A_{j})=1.

Now for both case-1 and case-2 when kk is odd, say k=2l+1k=2l+1, we consider the following element :

A¯lx¯2l+12y¯2l+1R¯2l+1m3\bar{A}_{l}\bar{x}_{2l+1}^{2}\bar{y}_{2l+1}\bar{R}_{2l+1}^{m-3}

Then applying (ϕψ)lϕ(\phi\otimes\psi)^{\otimes l}\otimes\phi on an appropriate multidegree term of the above product, we get:

(ϕψ)l(A¯l)ϕ(x2l+12y2l+1R2l+1m3)=ϕ2,3=1.(\phi\otimes\psi)^{\otimes l}(\bar{A}_{l})\phi(x_{2l+1}^{2}y_{2l+1}R_{2l+1}^{m-3})=\phi_{2,3}=1.\qed

7. Genetic codes having two genes each of size 3 or genes of Type 1

In this section, we obtain sharp bounds on the TCk(M¯α)\mathrm{TC}_{k}(\overline{\mathrm{M}}_{\alpha}) when the genetic code of α\alpha is either having two genes each of size 33 or having genes of Type 11.

7.1. Two genes each of size 3

In this subsection we inspect the higher topological complexity of M¯α\overline{\mathrm{M}}_{\alpha} where the genetic code of α\alpha is {a+b,a+b+c,n},{a,a+b+c+d,n}\langle\{a+b,a+b+c,n\},\{a,a+b+c+d,n\}\rangle with a,b,c,d1a,b,c,d\geq 1.

We first show that the TCk(M¯α)\mathrm{TC}_{k}(\overline{\mathrm{M}}_{\alpha}) is either kmkm or km+1km+1 by classifying values of a,b,c,da,b,c,d of the genetic codes mentioned above for which ϕ0=1\phi_{0}=1. The expression for ϕ0\phi_{0} is given in [7, Proposition 5.1], which we describe now.

ϕ0=(a12)+(b2)+bc+(a+1)(b+c+d).\phi_{0}=\binom{a-1}{2}+\binom{b}{2}+bc+(a+1)(b+c+d).
Lemma 7.1.

We have ϕ0=1\phi_{0}=1 if and only if either of the following holds

  1. (1)

    a+b1(mod4)a+b\equiv~1~(\text{mod}~4) or a+b+2c2(mod4)a+b+2c\equiv~2~(\text{mod}~4), and (a+1)d0(mod4)(a+1)d\equiv~0~(\text{mod}~4)

  2. (2)

    a+b1(mod4)a+b\not\equiv~1~(\text{mod}~4) and a+b+2c2(mod4)a+b+2c\not\equiv~2~(\text{mod}~4), and (a+1)d0(mod4)(a+1)d\not\equiv~0~(\text{mod}~4)

Proof.

We split the expression of ϕ0\phi_{0} into two parts as follows:

[(a12)+(b2)+bc+(a1)(b+c)]+[(a+1)d].\left[\binom{a-1}{2}+\binom{b}{2}+bc+(a-1)(b+c)\right]+\left[(a+1)d\right].

To have ϕ0=1\phi_{0}=1, the two separated terms in the above expression must have different parity. By the if and only if condition given in [7, Lemma 4.2], the result follows. ∎

The proof of the following theorem is similar to that of Proposition˜3.4, and hence we omit it here.

Theorem 7.2.

Let mm be a 22-power and {a+b,a+b+c,n},{a,a+b+c+d,n}\langle\{a+b,a+b+c,n\},\{a,a+b+c+d,n\}\rangle and a,b,c,d1a,b,c,d\geq 1 be the genetic code of α\alpha. Then for the values of a,b,c,da,b,c,d given in Lemma˜7.1, we have

kmTCk(M¯α)km+1.km\leq\mathrm{TC}_{k}(\overline{\mathrm{M}}_{\alpha})\leq km+1.

We now obtain a weaker bound on TCk(M¯α)\mathrm{TC}_{k}(\overline{\mathrm{M}}_{\alpha}), generalizing Davis’s result from section 55 of [7].

Theorem 7.3.

Let α\alpha be the genetic code as described above and m=2t+mm=2^{t}+m^{\prime} with 2m2t+12\leq m^{\prime}\leq 2^{t}+1 for some positive integer tt. Then

kmk2+1TCk(M¯α)km+1.km-\bigg{\lfloor}\frac{k}{2}\bigg{\rfloor}+1\leq\mathrm{TC}_{k}(\overline{\mathrm{M}}_{\alpha})\leq km+1. (22)
Proof.

Using similar notations as in Section˜6, we consider the element

A¯l:=j=1lAj,\bar{A}_{l}:=\prod\limits_{j=1}^{l}A_{j},

where Aj=w¯2j2m3x¯2j2y¯2jR¯2j2t+11A_{j}=\bar{w}_{2j}^{2m^{\prime}-3}\bar{x}_{2j}^{2}\bar{y}_{2j}\bar{R}_{2j}^{2^{t+1}-1}. We first deal with the case when kk is even, say k=2lk=2l. Using [7, Proposition 4.5, Lemma 5.2] it follows that (ϕψ)j(Aj)=1(\phi\otimes\psi)_{j}(A_{j})=1. Consequently, A¯l\bar{A}_{l} is non-zero. For the case where k=2l+1k=2l+1, we will consider the element A¯lw¯2l+1x¯2l+1R¯2l+1m2\bar{A}_{l}\bar{w}_{2l+1}\bar{x}_{2l+1}\bar{R}_{2l+1}^{m-2}, which will be non-zero. Hence, the inequality (22) follows. ∎

7.2. Type 1

In this subsection we inspect the higher topological complexity of M¯α\overline{\mathrm{M}}_{\alpha} where the genetic code α\alpha is {1,1+b,1+b+c,n},{1,1+b+c+d,n}\langle\{1,1+b,1+b+c,n\},\{1,1+b+c+d,n\}\rangle with b,c,d1b,c,d\geq 1 .

Theorem 7.4.

Let α\alpha be the genetic code as described above. Then

kmk2+1TCk(M¯α)km+1km-\bigg{\lfloor}\frac{k}{2}\bigg{\rfloor}+1\leq\mathrm{TC}_{k}(\overline{\mathrm{M}}_{\alpha})\leq km+1 (23)

except in either of the following cases:

  1. (1)

    b1(mod4)b\equiv~1~(\text{mod}~4), cc odd, and dd even

  2. (2)

    m1m-1 is a 22-power, or

  3. (3)

    mm is a 22-power, b2(mod4)b\equiv~2~(\text{mod}~4), cc odd and dd even.

Proof.

We first define xj:=pj(V1+b)x_{j}:=p_{j}^{*}(V_{1+b}) yj:=pj(V1+b+c)y_{j}:=p_{j}^{*}(V_{1+b+c}) following the notations given in [7, Section 6]. Now consider the element A¯l:=j=1lAj,\bar{A}_{l}:=\prod\limits_{j=1}^{l}A_{j}, where Aj=w¯2jm1x¯2j2y¯2jR¯2jm3A_{j}=\bar{w}_{2j}^{m-1}\bar{x}_{2j}^{2}\bar{y}_{2j}\bar{R}_{2j}^{m-3}. We first deal with the case when kk is even, say k=2lk=2l. Using [7, Proposition 6.2] it follows that (ϕψ)j(Aj)=1(\phi\otimes\psi)_{j}(A_{j})=1. Consequently, A¯l\bar{A}_{l} is non-zero. For the case when k=2l+1k=2l+1, we will consider the element A¯lw¯2l+1x¯2l+1y¯2l+1R¯2l+1m3\bar{A}_{l}\bar{w}_{2l+1}\bar{x}_{2l+1}\bar{y}_{2l+1}\bar{R}_{2l+1}^{m-3}, which will be non-zero. Hence, we obtain the desired inequality (23). ∎

Acknowledgment. We thank the anonymous referee for the valuable suggestions which improved the paper in several aspects. The second author acknowledges the support of NBHM through grant 0204/10/(16)/2023/R&D-II/2789. The third author acknowledges the support of IISER Pune for the Institute Post-Doctoral fellowship IISER-P/Jng./20235445.

References

  • [1] Ibai Basabe, Jesús González, Yuli B. Rudyak, and Dai Tamaki. Higher topological complexity and its symmetrization. Algebr. Geom. Topol., 14(4):2103–2124, 2014.
  • [2] Octav Cornea, Gregory Lupton, John Oprea, and Daniel Tanré. Lusternik-Schnirelmann category, volume 103 of Mathematical Surveys and Monographs. American Mathematical Society, Providence, RI, 2003.
  • [3] Navnath Daundkar, Priyavrat Deshpande, Shuchita Goyal, and Anurag Singh. The Borsuk-Ulam theorem for planar polygon spaces. Topology Appl., 318:Paper No. 108204, 19, 2022.
  • [4] Donald M. Davis. www.lehigh.edu/ dmd1/27cases.pdf.
  • [5] Donald M. Davis. www.lehigh.edu/ dmd1/mapleresults.pdf.
  • [6] Donald M. Davis. On the zero-divisor-cup-length of spaces of oriented isometry classes of planar polygons. Topology Appl., 207:43–53, 2016.
  • [7] Donald M. Davis. Topological complexity of planar polygon spaces with small genetic code. Forum Math., 29(2):313–328, 2017.
  • [8] Donald M. Davis. Topological complexity of some planar polygon spaces. Bol. Soc. Mat. Mex. (3), 23(1):129–139, 2017.
  • [9] Donald M. Davis. On the cohomology classes of planar polygon spaces. In Topological complexity and related topics, volume 702 of Contemp. Math., pages 85–89. Amer. Math. Soc., [Providence], RI, [2018].
  • [10] M. Farber and D. Schütz. Homology of planar polygon spaces. Geom. Dedicata, 125:75–92, 2007.
  • [11] Michael Farber. Topological complexity of motion planning. Discrete Comput. Geom., 29(2):211–221, 2003.
  • [12] Michael Farber. Invitation to topological robotics. Zurich Lectures in Advanced Mathematics. European Mathematical Society (EMS), Zürich, 2008.
  • [13] J.-C. Hausmann and A. Knutson. The cohomology ring of polygon spaces. Ann. Inst. Fourier (Grenoble), 48(1):281–321, 1998.
  • [14] Jean-Claude Hausmann. Geometric descriptions of polygon and chain spaces. In Topology and robotics, volume 438 of Contemp. Math., pages 47–57. Amer. Math. Soc., Providence, RI, 2007.
  • [15] Lazar Lyusternik and Levi Šnirelmann. Méthodes topologiques dans les problèmes variationnels. Actualités scientifiques et industrielles, vol. 188, Exposés sur l’analyse mathématique et ses applications, vol.3, Hermann, Paris, 42, 1934.
  • [16] Yuli B. Rudyak. On higher analogs of topological complexity. Topology Appl., 157(5):916–920, 2010.