On Periodic Points in Covering Systems
Abstract
We study a system of intervals on the real line and a continuous map with . It’s conjectured that there exists a periodic point of period in . In this paper, we prove the conjecture by a discretization method and reduce the initial problem to an interesting combinatorial lemma concerning cyclic permutations. We also obtain a non-concentration property of periodic points of small periods in intervals.
1 Introduction
We consider a continuous mapping : , and closed intervals . Throughout this paper, denotes the -th iterate of . And all the mappings are assumed to be continuous unless otherwise noted.
Definition 1.1.
We call a if .
This concept was first introduced by S. A. Bogatyi and E. T. Shavgulidze in their paper [1], when they tried to obtain an analogue of Sharkovskii’s theorem for an arbitrary tree. They proved the existence of a periodic point of period in a covering system and established an upper bound for . Moreover, they showed for and conjectured in the remark that holds for all , which is proved in this paper as the main theorem.
Theorem 1.2.
For any covering system , there exists and such that .
We will prove this theorem by a discretization method, which reduces the initial problem to a lemma in combinatorics. The main idea is to cut the intervals into small subintervals and perturb the mapping to get a cyclic permutation. And the proof of the theorem is completed by observing a property of cyclic elements of .
The proof of the main theorem 1.2 will be divided into several steps and discussed in the following sections. We now state a well-known proposition which will be frequently used in the proof of theorem 1.2.
Proposition 1.3 (the case ).
Let satisfy . Then there exists a fixed point of in .
This fact is an easy consequence of the intermediate value theorem in calculus so we omit the proof. And we will see later that for larger, most cases can be reduced to this fundamental one.
We would like to end this introduction with the statement of the main lemma used in the proof of theorem 1.2, which studies the properties of cyclic permutations. Actually, this lemma is an extension of a lemma proved in Sharkovskii’s celebrated paper [2]. See the remark below.
Definition 1.4.
Let be a cyclic permutation in the symmetric group , , can be written as , meaning that . And let be particular sets. We define the characteristic number of each as
Here, convex hull of for any finite set .
And we define the characteristic sequence of to be
where is a rearrangement of .
Example 1.5.
If , then we can calculate , , . The characteristic sequence is .
Take as an example. Since , ; and , ; therefore, and we have .
Now we can state our main lemma.
Lemma 1.6.
For any cyclic permutation in the symmetric group , the characteristic sequence of satisfies , .
Remark 1.7.
(1) For another form of this lemma more closely related to the initial problem, see lemma 2.3 in the next section.
(2) In theorem 7 of Sharkovskii’s paper [2], it is actually proved that , for . Therefore, this lemma is a generalization of that result.
Meanwhile, this lemma leads in some sense to a non-concentration property of periodic points with lower periods. Namely, it follows that if is an orbit of periodic points of period under , then between any two adjacent points and , there exists a periodic point of period . Futhermore, these numbers can be rearranged to have . In other words, all the periodic points of period cannot lie only in of the intervals , .
(3) One can see that the statement of lemma 1.6 doesn’t involve any notations in analysis. And hence it is of independent interest in the study of the symmetric group in combinatorics.
(4) It is necessary to require the permutation be cyclic. As a counterexmaple, consider , then the characteristic numbers are .
2 Discretization
In this section we show that Theorem 1.2 can be proved by solving another discrete problem related to it. The main idea is to divide the initial intervals into smaller subintervals, such that the image of each subinterval under contains the whole of some other subintervals, not just part of them. However, it may happen that no matter how you divide the intervals, there will be an interval whose image under contains only part of another interval. What saves us here is that, we can consider perturbation of if the intervals are cut into enough small pieces. Indeed, we will use the following lemma:
Lemma 2.1.
If there is a covering system such that , and then depending on , , whenever , it also holds that , for all , .
Proof.
Recall that has a periodic point of period in is equivalent to that has a zero in . Therefore, we only need to control for , . And this will be done inductively. Indeed, we can write:
And if we assume we have proved is sufficiently small (at least less than 1) for all , then choose a closed interval containing . We will have . Thus, we will obtain a control of when , since is uniformly continuous on and . Finally, we repeat the procedure times and get the conclusion. ∎
Now we can introduce the discrete problem related to Theorem 1.2. To begin with, we make a technical assumption on the initial covering system in Lemma 2.1. (See remark 2.4 for the reason). Namely, if and in the given covering system , then we can assume without loss of generality, since we can replace by a smaller subinterval if necessary.
Let be all the end points of the intervals and denote . Consider their images and we eliminate the points which don’t belong to to obtain a set . Put . Inductively, in each step we consider the image of under the mapping and we eliminate those points dropping out of to get a set . Then let .
Finally we obtain a sequence of sets , where each contains the images of the initial end point set under {} except for those dropping out of .
Then we can find a sufficiently large integer such that , , where comes from Lemma 2.1. Such exists because if not, there will be a sequence of disjoint points in with pairwise distance larger than , which is absurd. Therefore, we can take small perturbation of to get a map with , and for . Indeed, we just move each point on the graph of to the nearest point preserving the continuity and the covering property of . For example, in a neighborhood of each , we can define , where is a cut-off function, , is chosen so that . And for different we choose different signs so that is still a covering system. (See Figure 1 and 2.)
Thus, if we divide the intervals into small pieces using the points in , then will have the nice property that the image of each subinterval contains entirely another one or more subintervals. In other words, can be regarded as a discrete mapping from to {subsets of }. Here each number represents a small piece of interval . And if . Note that the image of some interval may be empty, since we ignore the part outside . And we only care about the interval determined by if , although the image may be larger.
Example 2.2.
The two figures illustrate how we perturb locally to obtain with still continuous and inducing a covering system close to the initial one.


It should also be remarked here that the only useful information of on the so called “gaps” (see [1]) between two adjacent intervals and is the behavior of the end points of each gap. In other words, for the image of each small subinterval , we only focus on the part . This is the reason why we eliminate the points outside in each step. We will see later that this information is enough for the proof.
To sum up, once we have a mapping satisfying the conditions in Lemma 2.1, we can find an close enough to , which also satisfies the conditions but can be viewed as a discrete mapping. Such and are counterexamples to Theorem 1.2. Therefore, if we argue by contradiction, we can easily see that Theorem 1.2 is a corollary of the following lemma, whose proof will be discussed later.
Lemma 2.3.
For any mapping : {subsets of such that , and for any partition
there exists and ( is allowed) such that
Here, convex hull of for any finite set .
Remark 2.4.
We emphasize here again the difference between and . In general the former is larger. And one may worry that the discrete mapping induced from a covering system does not satisfy the corresponding covering property , since in the definition of , iff , where is understood in the sense of rather than . However, this problem can be easily handled by considering the initial covering system to be minimal. (See [1] for more details.) That is, in particular, the end points of each are mapped to the end points of . And it’s now easily checked that the perturbation and discretization preserve the covering property of , since the perturbation does not change the position of the end points of essentially.
Proof of Theorem 1.2 by lemma 2.3.
Suppose the covering system is a counterexample to Theorem 1.2. Then as mentioned above, after perturbation, we can assume satisfies the conditions of lemma 2.3. Therefore the conclusion of lemma 2.3 gives us a subinterval of represented by whose image under contains itself. Thus there must exist a periodic point of period in , a contradiction. ∎
3 Equivalence of the Two Lemmas
Although lemma 2.3 reduces Theorem 1.2 to a discrete problem, it is not satisfactory since it involves something inconvenient to deal with, like convex hull and partition. In the following we will simplify the conditions in lemma 2.3 and find another description of it which only depends on the own property of a permutation. And the main goal of this section is to prove the equivalence of lemma 1.6 and lemma 2.3.
Let be in lemma 2.3. We can assume further:
(1), .
(2), .
(3), contains exactly one element.
(4) is a cyclic permutation in the symmetric group , , can be written as .
Indeed, (1) can be realized since we can move out the common part of and from one of them without changing the property .
(2) is also satisfied because we can eliminate the number and restrict the mapping to if necessary.
(3) is a consequence of (1), (2) and the condition . Indeed, one can apply the operations in (1) and (2) repeatedly until a minimal case. This couldn’t be an endless process because we are dealing with finite sets. And one can prove that the minimal set cannot be empty because the covering property is preserved.
For (4), note that because of (3). Suppose is not cyclic. Then choose an orbit of () and restrict both and the partition to .
In conclusion, we obtain:
Proposition 3.1.
It is sufficient to consider and being cyclic in lemma 2.3.
Proof.
As mentioned above, we can assume being cyclic, without loss of generality.
Suppose lemma 1.6 is true. Then for any partition
the characteristic numbers cannot cover . Therefore, , and such that , since in the characteristic sequence. In other words,
Therefore, we can just choose in lemma 2.3.
∎
Recall that in one-dimensional dynamics we often consider a directed graph associated to a periodic orbit (For more details, see [3] or chapter 1 of [4]). Namely, if is a cyclic permutation in Definition 1.4, we can construct a directed graph with vertices , where is defined in 1.4. And there is an edge from to if and only if .
Example 3.3.

Thus, we see directly from the definition that the following proposition holds:
Proposition 3.4.
For a cyclic and its associated directed graph , the characteristic number is equal to the minimal length of cycles starting from in .
Using directed graphs, we can complete the proof of the equivalence of the two lemmas.
Proof.
Suppose lemma 2.3 is true. For a cyclic , suppose on the contrary that there is a such that , and is the minimal number with this property. Then there are only characteristic numbers . Without loss of generality, assume . And consider the partition:
By lemma 2.3, we can find and () such that
Note that cannot happen in this case, since . Now if we extend to be a piecewise linear mapping from the interval to , then we will obtain . Therefore, there exists with . Note that since is cyclic as a permutation in and , cannot belong to . Consequently, for some integer and it is similar for . Since is piecewise linear, , imply , and it is similar for other pairs of points. So the orbit gives a cycle of length starting from the vertex in the associated directed graph , which implies . But this together with gives characteristic numbers , contradicting .
Combined with Proposition 3.2, the proof is completed. ∎
Therefore, we have reduced the initial problem on the existence of periodic points to a lemma in combinatorics on the properties of cyclic permutations. In the next section, we will prove the lemma and that will complete the proof of theorem 1.2.
We end this section with some examples of cyclic permutations whose characteristic sequences are known.
4 Proof of the Lemma in Combinatorics
We finally prove lemma 1.6 in this section. We first recall some definitions and notations from one-dimensional dynamics and linear algebra.
Definition 4.1.
Let be a cyclic permutation and be its associated directed graph. (See the discussion before propositon 3.4.) We can define to be the adjacency matrix of and . That is, , if , or equivalently, if there is an edge from vertex to vertex in . Otherwise, .
We denote the field with exactly two elements by . And we say for integer valued matrices and , when as matrices in .
Proposition 4.3.
For any cyclic permutation and , it holds that .
A proof of this proposition can be found in [4], Chapter 1, Propositon 20. And for readers’ convenience, we give a sketch of proof here.
Sketch of proof.
Given cyclic, we first extend to be a continuous piecewise linear mapping from the closed interval to itself. And recall we define . Later on we will also regard as the closed interval .
Now by the definition of and the continuity of , iff . Consider the interval and its image . In general, we cannot expect to hold, where denotes the closed interval whose end points are and . We can only say that for continuous . However, if we count the multiplicity of each point in the image , we will find that actually holds in the sense of mod 2. Indeed, can be viewed as a continuous curve in starting from the point and ending at . Therefore, every point is counted an even number of times in this curve, unless .
Finally recall the definition of and , and we get the conclusion of the proposition in the sense of mod 2. ∎
Remark 4.4.
If the readers are familiar with representation theory of the symmetric group , then they may find that defined above is just the matrix of the tautological representation of mod 2.
Corollary 4.5.
For cyclic , we have , where is the unit matrix.
Theorem 4.6.
For cyclic , the characteristic polynomial of , viewed as the determinant of a matrix in , is exactly:
We need the following lemma:
Lemma 4.7.
There exists a vector such that are linearly independent over the field .
Proof.
We claim that if is written as , then the vector (or ) satisfies the desired property, where is the unit vector whose -th component is .
We prove the claim by induction on . The case is trivial. We assume the statement holds for all natural numbers , and we will prove it for .
Suppose
By the definition of , , as well as the argument in the proof of proposition 4.3, we can see that
In other words, if we view each vector as the closed interval it represents, then we find that
Now if , take the inner product with , and we obtain:
because and hence for . Similarly we can obtain when .
If , then since does not appear as the end points of , and at the same time, where . Or equivalently,
And then if we take the inner product with , we obtain:
Anyway, we find that and consequently,
Now if we put and with the -th component dropped, we can see that , regardless of the -th component of . This is easily seen to be true since one can find as intervals. And the following holds for and :
By induction hypothesis we conclude that . Therefore, we have proved that the claim also holds for . This completes the proof of the lemma. ∎
Corollary 4.8.
The minimal polynomial of has degree at least , and therefore equals the characteristic polynomial.
Proof.
Suppose , then we have , which contradicts the lemma above. And the second statement follows by Cayley-Hamilton’s theorem. (Note that is an matrix.) ∎
With these results, we can compute the characteristic polynomial of .
Proof of Theorem 4.6.
Proof of lemma 1.6.
We show that if the coefficient of in is nonzero, then holds for the characteristic sequence in lemma 1.6.
Let us consider how the coefficients in are computed. It is then easily seen that the coefficient of is nonzero implies the existence of an principle matrix of whose determinant is nonzero. It then follows that there exists at least one diagonal of whose elements are all equal to 1. Now it is easily checked that this diagonal give rise to several cycles of length in the associated directed graph . And the vertices of these cycles correspond to the subscripts of the columns of . Thus there exist at least vertices in , from whom the minimal length of cycles starting are all . By proposition 3.4, this means that at least of the characteristic numbers . So . ∎
Acknowledgement
I would like to thank Prof. S. A. Bogatyi for introducing the problem to me. And I thank my supervisor, Prof. Tian Gang, for his encouragement as well as helpful suggestions. I’m also grateful to Dr. Ye Yanan and Dr. Zhao Yikai for writing computer programs to verify some claims.
References
- [1] S. A. Bogatyi and E. T. Shavgulidze. Periodic points of a map of a system of intervals. Mathematical Notes of the Academy of Sciences of the Ussr, 43(3):210–219, 1988.
- [2] A. N. Sharkovskii. Coexistence of cycles of a continuous map of the line into itself. International Journal of Bifurcation and Chaos, 5(5):1263–1273, 1995.
- [3] Uhland Burkart. Interval mapping graphs and periodic points of continuous function. Journal of Combinatorial Theory, Series B, 32(1):57–68, 1982.
- [4] L. S. Block and W. A. Coppel. Dynamics in One Dimension. Springer, 1992.
- [5] P. Štefan. A theorem of Šarkovskii on the existence of periodic orbits of continuous endomorphisms of the real line. Communications in Mathematical Physics, 54(3):237–248, 1977.
- [6] Philip D. Straffin. Periodic points of continuous functions. Mathematics Magazine, 51(2):99–105, 1978.