Higher topological complexity of planar polygon spaces having small genetic codes
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 , Davis showed that their topological complexity is either or . We extend these bounds to the setting of higher topological complexity. In particular, when is power of , we show that the -th higher topological complexity of these spaces is either or
Key words and phrases:
LS category, higher topological complexity, planar polygon spaces2020 Mathematics Subject Classification:
55M30, 58D29, 55R801. Introduction
A motion planning algorithm in a path-connected topological space is defined as a section of the free path space fibration where and denotes the free path space of , equipped with the compact open topology. To analyze the complexity of designing a motion planning algorithm for the configuration space of a mechanical system, Farber [11] introduced the notion of topological complexity. The topological complexity of a space , denoted by , is defined as the smallest positive integer for which can be covered by open sets , with each admitting a continuous local section of . The number represents the minimal number of continuous rules required to implement a motion planning algorithm in the space . Farber [11, Theorem 3] showed that is a numerical homotopy invariant of a space .
In [16], Rudyak introduced the higher analogue of topological complexity. For a path-connected space , consider the fibration defined by
(1) |
The higher topological complexity of , denoted by , is the smallest positive integer for which can be covered by open sets , such that each admits a continuous local section of . Note that when , the higher topological complexity coincides with . The above definition makes also sense for , but is always equal to (see [16]). Similar to , the invariant is associated with motion planning problems, where the input includes not only an initial and final point but also an additional 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 , denoted by , is the smallest positive integer such that can be covered by open subsets such that each inclusion is null-homotopic. The following inequalities were established in [1] shows how the LS category and higher topological complexity are related:
(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 be the diagonal map and denote the cup-length of , 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.
(3) |
We refer to the non-negative integer as the higher zero-divisors-cup-length of . On the other hand, if denotes the cup-length of a cohomology ring of . Then there is an inequality
(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:
(5) |
Moreover, an additional upper bound for 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 -gons with consecutive side lengths for some , 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 , denoted by , is the collection of all closed piecewise linear paths in the plane with side lengths , considered up to orientation-preserving isometries. Equivalently, we can describe as
where is the unit circle and the group of orientation-preserving isometries acts diagonally. The planar polygon space (associated with ) viewed up to isometries is defined as
where is the orthogonal group acting on , including all linear isometries.
The spaces and are also called moduli spaces of planar polygons. It is clear that is a double cover of . A length vector is called generic if . For such a length vector , the moduli spaces and are closed, smooth manifolds of dimension (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 are torsion-free. They also described the Betti numbers in terms of the combinatorial data associated with the length vector. The mod- cohomology ring of 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 is either or . The estimates for , 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 is determined by its genetic code . 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- cohomology ring of .
1.1. Structure of the article:
In Section˜2, we begin with the digression of genetic codes and the description of the mod- 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 when the genetic code of contains a gene of size . Our results Theorem˜3.1 and Theorem˜3.3 address the cases where has a monogenic code of size and where the genetic code includes a gene of size , respectively. In particular, in Theorem˜3.1 we identify cases for which is either or . In Proposition˜3.4, we explicitly consider the case of a genetic code having a gene of size and under some minor conditions we show that is either or .
Section˜4 consists of the cases of genetic code containing a gene of size . In Theorem˜4.2, we establish that is either or 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 . Specifically, our result Theorem˜5.1 extends [7, Theorem 3.1] to the sequential case.
In Section˜6, we obtain the bounds for where having monogenic codes of size . We begin this section by classifying (see Lemma˜6.1) such genetic codes for which the (see Theorem˜2.4 for the cohomology class ). This is crucial in identifying the genetic codes of for which we can have to be either or . 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 , with the former one being a strong bound and the latter one generalizing the result from section 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 by , for . There are two important combinatorial objects associated with the length vector .
Definition 2.1.
A subset is called -short if
and -long otherwise.
We may simply use "short" as a shorthand for -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 .
For a length vector , consider the collection of subsets of :
along with the partial order defined as follows:
Definition 2.2.
The genetic code of is the set of maximal elements of with respect to the partial order defined above.
If are the maximal elements of with respect to then the genetic code of is denoted by . Each ’s is called a gene and a subset is called gee for each (see [7] for more details).
Example 2.3.
Let (-tuple) be a length vector. Then the genetic code of is . Moreover, and .
Let be the genetic code of . Then it is easy to see that uniquely determines the collection . Consequently, it follows from [14, Lemma 1.2] that if the genetic codes of length vectors and are the same, then the corresponding planar polygon spaces are diffeomorphic.
We now recall the mod- cohomology algebra of 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 is generated by classes subject to the following relations:
-
(1)
, for .
-
(2)
, unless is a subgee.
-
(3)
For every subgee with ,
(6) where 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 of , the following set
forms a basis for , for .
We now set up some notations that will be used throughout the paper. Let be the projection onto the -th factor. There is an induced map on the cohomology ring
where is a field. Now considering and we set the following notions
-
(1)
-
(2)
-
(3)
-
(4)
Observe that and , where is the diagonal map.
Remark 2.6.
-
(1)
From Corollary 2.5, it follows that is linearly independent. Thus,
forms a linearly independent set in the cohomology algebra .
-
(2)
As part-(1) extends Corollary˜2.5, we repeatedly utilize similar canonical approaches to explore the relations among cohomology classes in 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 whose dimension is .
Proposition 2.7.
Let be a length vector. Then . Moreover, we have
Proof.
In most cases, to establish lower bounds for , Davis utilized two key group homomorphisms, and . Specifically, is the Poincaré duality isomorphism, and by choosing an uniform homomorphism such that , where a product of -many cohomology classes from the . Thus, one can conclude that .
For example, if the genetic code of a length vector is , then the uniform homomorphism in the sense of Davis is a homomorphism satisfying
-
•
if or if , and
-
•
if or if .
(Above, we have omitted writing powers of accompanying ’s as done by Davis in [7]).
Our main strategy is to determine lower bounds for as follows. We will construct classes when and when such that and . Here the map is given by the usual tensor product of maps as
defined by for , and for , . Also, we will use the following homomorphisms
(7) |
where the th and th entries are the maps and respectively, for .
In [7], Davis has explicitly shown that , for the genetic code having gene sizes from the table below. For genes of size up to , the following table depicts the number of possible genetic codes for the first few values of .
Gene Size | ||||
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 |
A genetic code is said to be of Type 1 if . For the case of 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 . Note that there are genetic codes of Type 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 cases. For example, is the genetic code of Type having genes of size , (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 when the genetic code of has a gene of size .
Since is a connected smooth manifold of dimension , we know that . In [7, Theorem 2.2], Davis has shown that the if the genetic code of is . We will now generalize this result in the higher setting.
Theorem 3.1.
Let be the genetic code of . Then
(8) |
Moreover, if for some non-negative integer , then
(9) |
Proof.
For , to obtain the lower bound on , Davis proved that by expanding it in bidegree as follows:
where , and using the linear independence of together with the fact that .
We now build on this idea to argue in the general case . We first prove the left inequality of (8) when . Consider the following product of cohomology classes:
We now expand in the multidegree , where and appear at -th and -th position, respectively. We then have
(10) |
By comparing the th position in (10) and applying Remark 2.6, the above product is non-zero. Now, has the following expression:
Observe that the above expansion contains the following -multidegree term
Every element in the above product is non-zero, which follows from the Remark˜2.6 and the fact . Moreover, this expression can not be annihilated by any other term, since no other term in the expansion of is devoid of in the odd positions.
Next, we obtain the desired result for . Consider the following product of cohomology classes:
Observe that the -multidegree term of the above product is , which is nonzero. This gives us the inequality
giving us the desired lower bound of (8).
Now, we give the sharp lower bound of , when for some non-negative integer . We will now prove the left inequality of (9) for the case .
We consider another product of cohomology classes:
for , and is the empty product. Thus for , we have .
Note that multiplying the class by the class brings into the -th position, keeping it non-zero. Our claim is
(11) |
We induct on . Suppose the statement is true for . Then is
In the penultimate step, we have used the induction hypothesis and the fact that is even. Therefore, whenever , it follows that , providing us :
We can impose conditions on to deduce when . For , Lucas’s theorem implies that the binomial coefficient is even. Thus, in this situation, we get
We now address the case when . Consider the element
Using (11) we have the expression:
The rightmost term expands as . We consider only the term from the expansion. Observe that and therefore , as the aforementioned term cannot be annihilated by any other term from the expansion of , due to positional differences. Hence,
Consequently, when for some . ∎
Remark 3.2.
For , by Lucas’s theorem mod when for some .
For genetic codes having a gene of size two of the length vector , Davis in [7, Theorem 2.3] has shown that . We will now prove a higher analogue of this result. In particular, we have:
Theorem 3.3.
Suppose the length vector has a genetic code with genes . Then
Proof.
First, we prove our assertion when . Note that, there exists a non-empty subset such that , as done in the proof of [7, Theorem 2.3], where is the Poincaré duality isomorphism. Define
Now using similar arguments as presented in the proof of [7, Theorem 2.3], we show that the product is nonzero. Recall from the proof of [7, Theorem 2.3] that there is a homomorphism which sends and to and all other monomials to zero.
In the following we apply on , using the homomorphisms (7).
This implies , thereby completing the proof of our assertion.
Now we consider the case . Consider the product:
Observe that the -multidegree term of the above product is . Applying on the previous product obtains:
This completes the proof. ∎
Next we compute sharp bounds on , when the genetic code of contains a gene . In particular, using the result [3, Example 4.15] for the genetic code we get the following.
Proposition 3.4.
Let be odd and be the genetic code of the length vector . If is a -power, then
Proof.
Using [3, Example 4.15] we get that is non-zero if is odd. Thus,
(12) |
is a non-zero cohomology class (using Remark 3.2), where the class .
Assume to be even, say . Consider the following product of cohomology classes:
-
•
, for (using (12)).
-
•
for , and is the empty product.
We inductively show that for each , the following identity holds.
(13) |
Consider the case for , the identity holds.
Assuming the required identity holds for some , we show that it also holds for the case of . Note that expands as
Thus, we have proved the identity (13). Each term in the sum has a different multidegree and contains at every position except one, where we have . This implies is nonzero whenever is even, more precisely when (see Remark˜3.2). Notice that, to get a lower bound for where , we consider the non trivial cohomology class of length , and thus .
For odd, say , consider the following product
Observe that the product will contain which is non-zero since its every term is of distinct multidegree and contains at every position except one, where we have . ∎
4. The case of monogenic code of size 3
In this section, we will derive sharp bounds on when the genetic code of is . 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)
with ,
-
(2)
with ,
-
(3)
with ,
-
(4)
with and
-
(5)
refers to for an appropriate value of and is a subgee.
Let and for all possible subscripts (see Section˜2). Then it was shown in [7, Proposition 2.5] that satisfies the following identities:
We now briefly recall Davis’s strategy to construct the uniform homomorphism and describe two equations that come from the relation (6). Note that if is a subgee of the genetic code , then either or . We denote by , a relation (6) corresponding to the subgees of the first kind and by for the subgees of another kind. Then counting subgees (or the cohomology classes of type , , and ) of an appropriate type we obtain as
(14) |
and as
(15) |
In what follows, we will assume that is even. The other cases will follow similarly. Let for . Then third equation can be obtained by applying to the -bidegree expansion of and equating it to , where is the Poincaré duality isomorphism and . The -bidegree expansion of this cohomology class is given by
where denotes the Kronecker delta. Then acting on the above expansion and equating it to gives the third equation:
(16) |
where is an irrelevant quantity in solving the system of linear equations formed by (14), (15), and (16) with indeterminate . Davis has shown the existence of by proving that the system of linear equations described above has a solution. This shows the non-zeroness of the cohomology class and consequently gives us the desired lower bound on when the genetic code of is with even.
We now characterize the values of and for which the is non-zero, where is the dimension of .
Lemma 4.1.
Let be the genetic code of with and . Then if and only if either of the following cases holds:
-
(1)
,
-
(2)
and even,
-
(3)
and odd.
Proof.
Recall that , where is the Poincaré duality isomorphism. Now
-
(1)
Suppose . Then is even and has only one as a factor. Hence is odd.
-
(2)
Suppose and even. Then is even and is odd.
-
(3)
Since divides , is even. Also being odd makes odd.
It is easy to check that for all other combinations of values of and , . ∎
As a consequence of Lemma˜4.1, we obtain the following sharp bounds.
Theorem 4.2.
Let for some non-negative integer and be the genetic code of such that and satisfy either of the conditions of Lemma˜4.1. Then
Proof.
Since 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 be the genetic code of . Then
(17) |
Proof.
Suppose that . Before we proceed, we set some notations that will be used throughout the proof. Define , and for . We proceed with the following cases.
Case-1: Suppose is even.
Assume . Consider the element
where . We apply on the -multidegree term in the expansion of . Recall the notation from (7). Then the identity holds for all (see (16)). Thus, we obtain the following expression:
Since has a solution as shown by Davis, has also has a solution. Therefore, for . This gives us the left inequality of (17).
We use similar idea to show that the product is nonzero when .
Observe that the type of the -multidegree term of is . Since and , we get
This proves and thus we get the left inequality of (17).
Case-2: is odd and .
Then for , Davis proved that by showing the existence of uniform homomorphism , and then applying on the expansion of in the bidegree . We generalize this idea for general .
Assume . Consider the element
where . We apply on the - multidegree term in the expansion of . Similar to Case-1, we obtain the following expression:
In [7, Theorem 2.4], Davis has shown that has solution. Here, the value of is again irrelevant. Consequently, has a solution. Therefore, for . This gives us the left inequality of (17).
We use a similar idea to show that the product is nonzero when . Observe that the type of the -multidegree term of is . Since and , we get
This proves and thus we get the left inequality of (17).
Case-3: is odd and .
Then for , Davis proved that by constructing a uniform homomorphism and applying to the expansion of in the bidegree . We use this idea to prove our assertion for .
Assume . Consider the element
where . We apply on the - multidegree term in the expansion of . Similar to the previous cases, we obtain the following expression:
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 . We refer the reader to Section˜2 for a brief description of these genetic codes.
It turns out that there are exactly Type 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 is the dimension of . The genetic codes are distributed into two cases:
Suppose is not a -power or if this is the last case of the [7, Table 2] which is the Type genetic code . Let for . In this case, Davis proved that by constructing a uniform homomorphism , and then applying to the expansion of the cohomology class in bidegree , where denotes the Poincaré duality isomorphism. The existence of was shown by solving a system of linear equations. We now explain his ideas briefly. The homomorphism acting on the expansion of in bidegree gives us the following:
(18) | ||||
Davis first computes the values of , where is a Poincaré duality isomorphism and 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 , and rows represent all subgees except . This matrix is actually a binary matrix in which the entry 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 and observed that , while in the first cases of the [7, Table 2].
A similar approach figures out the values of , except that the rows correspond only to subgees with more than one element, due to the constraint in (6). The uniform homomorphism in each case given in [7, Table 2] satisfies and . One can see that in this case (18) equates to for the first genetic codes of Type 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 for which the only nonzero values are and . Moreover, for all subgees and as . Since , again (18) equates to . Thereby proving the nonzeroness of .
If is a -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 has the genetic code of type in the Table˜1 and , then:
(19) |
Proof.
Recall from Section˜2 that . Similarly, we define and . Similar to Davis, we distribute the genetic codes into two cases.
Case 1: Suppose is not a -power or has gees as in the final case in [7, Table 2].
Consider the element
where . We apply on the -multidegree term in the expansion of . When is a power of , we obtain the following:
Then the above expression is equal to , since as explained at the beginning of this section.
Next, when is a power of (and the gee is of the final type in [7, Table 2]), we obtain the following:
For the -th gee in [7, Table 2], both the above two expressions ( is a -power and not a -power) become , since there is homomorphism whose only non-zero values are and . Moreover, for all and . Therefore, only survives.
Finally, in all other situations under this case, we get the following:
The above expression is equal to for the first gees in [7, Table 2], since . For the -th gee, it is again , thanks to .
Now we consider the cases when is odd, say . Recall the element that we used in this case of was where
The nonzeroness of the product is given as follows. Consider the classes . Observe that the type of the -multidegree term of is . Since in first cases of [7, Table 2] and , we get
This proves and thus we get the left inequality of (19). For the -th case we consider . Clearly, the type of the -multidegree term of is . Since , we obtain that
This gives us the left inequality of (19).
Case 2: Suppose is a -power and has doesn’t have gees as in last case of [7, Table 2].
For , Davis proved that by constructing a uniform homomorphism , and then applying to the expansion of the cohomology class in bidegree , where denotes the Poincaré duality isomorphism. We generalize this idea for general .
First, assume that is even, say . Consider the element
where . We apply on the -multidegree term in the expansion of . As a result, we obtain the following expression:
Note that, and as explained in the beginning of this section. Hence, the above expression becomes , giving us the desired lower bound of (19).
Now we consider the cases when is odd, say . Recall the element that we used in case of was where
We want to show that the product is nonzero. Consider the classes . Observe that the type of the -multidegree term of is . Since and , we get
This proves 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 , where the genetic code is with .
Recall that, for a Poincaré duality isomorphism , we have . In [7, Theorem 4.1], Davis has obtained the expression for as follows
The following result classifies values of in the genetic code for which the values of are odd.
Lemma 6.1.
is odd if and only if any of the following conditions hold:
-
(1)
even, even, ;
-
(2)
even, even, ;
-
(3)
odd and ;
-
(4)
odd, , and ;
-
(5)
.
Proof.
(1 and 2) Under the assumption that both and are even, one can write as
By [7, Lemma 4.2], the last expression is odd under the above-mentioned conditions.
(3) For , . Moreover, we have and odd. Hence, the result follows.
(4) For odd, becomes
When , it becomes , which is odd if and only if .
(5) For , both and are odd. Again, and both are odd. Consequently, there are two odd multiples of and the only surviving odd term is ∎
The previous lemma helps us to obtain the sharp bounds on .
Theorem 6.2.
Let be a -power and be the genetic code of satisfying the conditions in Lemma˜6.1. Then
Proof.
Note that if satisfies conditions in Lemma˜6.1, then . Then the proof of our assertion is similar to that of Proposition˜3.4. ∎
We denote the intervals , and by , and , respectively and classify all possible subgees by tuples of sizes , , and (see the proof of [7, Theorem 4.1] for more details). The tuple of size one represents a subgee containing one element, and the location of this element is decided by the value of . The tuple of size two represents a subgee and the locations of and are decided by the values of and , respectively. For example, refers to a subgee of cardinality two, such that and . Similarly, the tuple represents a subgee such that the locations of are decided by the values of . For example, represents a subgee such that and . The following classification of types of elements that can be subgees of the genetic code has been given by Davis in the proof of [7, Theorem 4.1]:
Next, we aim to obtain the bounds on when . For that purpose, we need subgees of size greater equal . Such sub-gees are described as follows:
Now for , let be the number of ’s in for . For example, if , then and . For , applying on the relation of Theorem˜2.4 we obtain the following expression
(20) |
where .
We are now in a position to state our general result.
Theorem 6.3.
With the genetic code of being as stated above, we have
(21) |
in the following situations:
-
(1)
, and odd;
-
(2)
, , and odd.
Proof.
In the -st situation, following the proof of [7, Proposition 4.3], we know that there exist a uniform homomorphism which sends to and other monomials to . Now to achieve our desired assertion, we will consider two cases depending on whether is not a -power.
Recall that we have and . We now define .
Case 1: Suppose is not a -power.
Consider the element
where . We deal with the case when is even, say . We apply on the -multidegree term in the expansion of . Let us inspect that particular multidegree term of , which can get mapped non-trivially.
All terms of the above sum are of the type . In the first situation, it follows from [7, Theorem 4.1] that sends each to and all other monomials to . Now applying to , we obtain:
which is 1, by Lucas’s theorem. Therefore, we obtain:
Case 2: Suppose is a -power.
Consider the element
where . Again, we start with even, say . We inspect the suitable multidegree term of , which can get mapped non-trivially under .
All the terms in the above sum are of the type . Hence, after applying , it becomes, just similar to the last case:
which is , again by Lucas’s Theorem. Therefore, we obtain:
For both case-1 and case-2, when is odd, say , we consider the following element
Then applying on an appropriate multidegree term of the above product, we get
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 which sends to and other monomials to . Similarly, as in the first situation, we consider two cases depending on whether is a -power.
Case 1: Suppose is a -power.
Consider the element
where . Start with even, say . We expand in the suitable multidegree, which can get mapped non-trivially under .
Note that the above term is of the type . Thus applying , we get:
Note that under the -nd situation we have by [7, Theorem 4.1]. As is a -power, the binomial coefficient is of the form
Hence, by Lucas’s theorem, it follows that is odd. Therefore, we obtain:
Case 2: Suppose is not a -power.
Consider the element
where .
Start with even, say . We inspect the suitable multidegree term of , which can get mapped non-trivially under .
Applying to the above type term we get:
Again, by Lucas’s Theorem, is even. Therefore, we obtain:
Now for both case-1 and case-2 when is odd, say , we consider the following element :
Then applying on an appropriate multidegree term of the above product, we get:
7. Genetic codes having two genes each of size 3 or genes of Type 1
In this section, we obtain sharp bounds on the when the genetic code of is either having two genes each of size or having genes of Type .
7.1. Two genes each of size 3
In this subsection we inspect the higher topological complexity of where the genetic code of is with .
We first show that the is either or by classifying values of of the genetic codes mentioned above for which . The expression for is given in [7, Proposition 5.1], which we describe now.
Lemma 7.1.
We have if and only if either of the following holds
-
(1)
or , and
-
(2)
and , and
Proof.
We split the expression of into two parts as follows:
To have , 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 be a -power and and be the genetic code of . Then for the values of given in Lemma˜7.1, we have
We now obtain a weaker bound on , generalizing Davis’s result from section of [7].
Theorem 7.3.
Let be the genetic code as described above and with for some positive integer . Then
(22) |
Proof.
Using similar notations as in Section˜6, we consider the element
where . We first deal with the case when is even, say . Using [7, Proposition 4.5, Lemma 5.2] it follows that . Consequently, is non-zero. For the case where , we will consider the element , which will be non-zero. Hence, the inequality (22) follows. ∎
7.2. Type 1
In this subsection we inspect the higher topological complexity of where the genetic code is with .
Theorem 7.4.
Let be the genetic code as described above. Then
(23) |
except in either of the following cases:
-
(1)
, odd, and even
-
(2)
is a -power, or
-
(3)
is a -power, , odd and even.
Proof.
We first define following the notations given in [7, Section 6]. Now consider the element where . We first deal with the case when is even, say . Using [7, Proposition 6.2] it follows that . Consequently, is non-zero. For the case when , we will consider the element , 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.