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

On Periodic Points in Covering Systems

Wang Yihan
[email protected]
Abstract

We study a system of intervals I1,,IkI_{1},\ldots,I_{k} on the real line and a continuous map ff with f(I1I2Ik)I1I2Ikf(I_{1}\cup I_{2}\cup\ldots\cup I_{k})\supseteq I_{1}\cup I_{2}\cup\ldots\cup I_{k}. It’s conjectured that there exists a periodic point of period k\leq k in I1IkI_{1}\cup\ldots\cup I_{k}. 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 ff: \mathbb{R}\to\mathbb{R}, and kk closed intervals I1,I2,,IkI_{1},I_{2},...,I_{k}. Throughout this paper, flf^{l} denotes the ll-th iterate of ff. And all the mappings are assumed to be continuous unless otherwise noted.

Definition 1.1.

We call (I1,I2,,Ik;f)(I_{1},I_{2},\ldots,I_{k};f) a coveringsystemcovering\ system if f(I1I2Ik)I1I2Ikf(I_{1}\cup I_{2}\cup\ldots\cup I_{k})\supseteq I_{1}\cup I_{2}\cup\ldots\cup I_{k}.

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 g(n)\leq g(n) in a covering system and established an upper bound for g(n)g(n) (g(n)2(n21)g(n1)n2)(g(n)\leq 2(n^{2}-1)g(n-1)^{n-2}). Moreover, they showed g(n)=ng(n)=n for n4n\leq 4 and conjectured in the remark that g(n)=ng(n)=n holds for all nn\in\mathbb{N}, which is proved in this paper as the main theorem.

Theorem 1.2.

For any covering system (I1,I2,,Ik;f)(I_{1},I_{2},\ldots,I_{k};f), there exists x0I1I2Ikx_{0}\in I_{1}\cup I_{2}\cup\ldots\cup I_{k} and lkl\leq k such that fl(x0)=x0f^{l}(x_{0})=x_{0}.

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 ff to get a cyclic permutation. And the proof of the theorem is completed by observing a property of cyclic elements of SnS_{n}.

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 k=1k=1).

Let f:If:I\to\mathbb{R} satisfy f(I)If(I)\supseteq I. Then there exists a fixed point of ff in II.

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 kk 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 ff be a cyclic permutation in the symmetric group SnS_{n}, i.e.i.e., ff can be written as (i1i2in)(i_{1}i_{2}\ldots i_{n}), meaning that f:i1i2ini1f:i_{1}\to i_{2}\to\ldots\to i_{n}\to i_{1}. And let Ai={i,i+1}A_{i}=\{i,i+1\} (i=1,2,,n1)(i=1,2,\ldots,n-1) be (n1)(n-1) particular sets. We define the characteristic number mim_{i} of each AiA_{i} as

mi=min{m|(convf)m(Ai)Ai}.m_{i}=\min\{m\arrowvert\ (convf)^{m}(A_{i})\supseteq A_{i}\}.

Here, convf(A)=convf(A)= convex hull of f(A)={minf(A),minf(A)+1,,maxf(A)}f(A)=\{\min f(A),\min f(A)+1,\ldots,\max f(A)\} for any finite set AA\subseteq\mathbb{N}.

And we define the characteristic sequence of ff to be

m1m2mn1,m_{1}^{\prime}\leq m_{2}^{\prime}\leq\ldots\leq m_{n-1}^{\prime},

where {mi}\{m_{i}^{\prime}\} is a rearrangement of {mi}\{m_{i}\}.

Example 1.5.

If f=(136245)S6f=(136245)\in S_{6}, then we can calculate m1=m3=m5=3m_{1}=m_{3}=m_{5}=3, m2=2m_{2}=2, m4=1m_{4}=1. The characteristic sequence is 123331\leq 2\leq 3\leq 3\leq 3.

Take m2m_{2} as an example. Since f(A2)=f({2,3})={4,6}f(A_{2})=f(\{2,3\})=\{4,6\}, convf(A2)=conv{4,6}={4,5,6}convf(A_{2})=conv\{4,6\}=\{4,5,6\}; and f({4,5,6})={1,2,5}f(\{4,5,6\})=\{1,2,5\}, conv{1,2,5}={1,2,3,4,5}conv\{1,2,5\}=\{1,2,3,4,5\}; therefore, (convf)2(A2)={1,2,3,4,5}A2(convf)^{2}(A_{2})=\{1,2,3,4,5\}\supseteq A_{2} and we have m2=2m_{2}=2.

Now we can state our main lemma.

Lemma 1.6.

For any cyclic permutation ff in the symmetric group SnS_{n}, the characteristic sequence {mi}\{m_{i}^{\prime}\} of ff satisfies miim_{i}^{\prime}\leq i,  i=1,2,,n1i=1,2,\ldots,n-1.

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 min1m_{i}^{\prime}\leq n-1, for i=1,2,,n1i=1,2,\ldots,n-1. 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 xk1<xk2<<xknx_{k_{1}}<x_{k_{2}}<\ldots<x_{k_{n}} is an orbit of periodic points of period nn under ff, then between any two adjacent points xkix_{k_{i}} and xki+1x_{k_{i+1}}, there exists a periodic point yiy_{i} of period min1m_{i}\leq n-1. Futhermore, these (n1)(n-1) numbers mim_{i} can be rearranged to have miim_{i}^{\prime}\leq i. In other words, all the periodic points of period i\leq i cannot lie only in (i1)(i-1) of the intervals [xkj,xkj+1][x_{k_{j}},x_{k_{j+1}}], 1jn11\leq j\leq n-1.

(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 SnS_{n} in combinatorics.

(4) It is necessary to require the permutation be cyclic. As a counterexmaple, consider f=(13)(2)S3f=(13)(2)\in S_{3}, then the characteristic numbers are m1=m2=2m_{1}=m_{2}=2.

We organize the paper as follows. In section 2, we introduce the discretization method and reduce theorem 1.2 in dynamics to a lemma in combinatorics. Then in section 3, we prove the equivalence of the lemma in section 2 and the one stated above. Finally in section 4, we prove lemma 1.6 .

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 ff 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 ff contains only part of another interval. What saves us here is that, we can consider perturbation of ff 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 (I1,I2,,Ik;f)(I_{1},I_{2},\ldots,I_{k};f) such that xI1Ik\forall x\in I_{1}\cup\ldots\cup I_{k}, and lk,fl(x)x,\forall\>l\leq k,f^{l}(x)\neq x, then δ>0\exists\>\delta>0 depending on ff, s.t.s.t., whenever gfC()<δ\lVert g-f\rVert_{C(\mathbb{R})}<\delta, it also holds that gl(x)xg^{l}(x)\neq x, for all lkl\leq k, xI1Ikx\in I_{1}\cup\ldots\cup I_{k}.

Proof.

Recall that ff has a periodic point of period k\leq k in I1IkI_{1}\cup\ldots\cup I_{k} is equivalent to that F(x)=(f(x)x)(fk(x)x)F(x)=(f(x)-x)\ldots(f^{k}(x)-x) has a zero in I1IkI_{1}\cup\ldots\cup I_{k}. Therefore, we only need to control |fl(x)gl(x)|\lvert f^{l}(x)-g^{l}(x)\rvert for 1lk1\leq l\leq k, xI1Ikx\in I_{1}\cup\ldots\cup I_{k}. And this will be done inductively. Indeed, we can write:

fl+1(x)gl+1(x)=f(fl(x))f(gl(x))+f(gl(x))g(gl(x))f^{l+1}(x)-g^{l+1}(x)=f(f^{l}(x))-f(g^{l}(x))+f(g^{l}(x))-g(g^{l}(x))

And if we assume we have proved |fl(x)gl(x)|\lvert f^{l}(x)-g^{l}(x)\rvert is sufficiently small (at least less than 1) for all xI1Ikx\in I_{1}\cup\ldots\cup I_{k}, then choose a closed interval [p,q][p,q] containing fl(I1Ik)f^{l}(I_{1}\cup\ldots\cup I_{k}). We will have gl(I1Ik)[p1,q+1]g^{l}(I_{1}\cup\ldots\cup I_{k})\subseteq[\,p-1,\,q+1\,]. Thus, we will obtain a control of |fl+1(x)gl+1(x)|\lvert f^{l+1}(x)-g^{l+1}(x)\rvert when xI1Ikx\in I_{1}\cup\ldots\cup I_{k}, since ff is uniformly continuous on [p1,q+1][\,p-1,\,q+1\,] and gfC()<δ\lVert g-f\rVert_{C(\mathbb{R})}<\delta. Finally, we repeat the procedure kk 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 Ij=[aj,bj]I_{j}=[a_{j},b_{j}] and f(Ij)=[cj,dj]f(I_{j})=[c_{j},d_{j}] in the given covering system (I1,I2,,Ik;f)(I_{1},I_{2},\ldots,I_{k};f), then we can assume f({aj,bj})={cj,dj}f(\{a_{j},b_{j}\})=\{c_{j},d_{j}\} without loss of generality, since we can replace IjI_{j} by a smaller subinterval if necessary.

Let a1<b1<a2<b2<<ak<bka_{1}<b_{1}<a_{2}<b_{2}<\ldots<a_{k}<b_{k} be all the end points of the intervals I1,,IkI_{1},\ldots,I_{k} and denote M0={a1,b1,,ak,bk}M_{0}=\{a_{1},b_{1},\ldots,a_{k},b_{k}\}. Consider their images f(M0)={f(a1),,f(bk)}f(M_{0})=\{f(a_{1}),\ldots,f(b_{k})\} and we eliminate the points which don’t belong to I1IkI_{1}\cup\ldots\cup I_{k} to obtain a set S1f(M0)S_{1}\subseteq f(M_{0}). Put M1=M0S1M_{1}=M_{0}\cup S_{1}. Inductively, in each step we consider the image of MiM_{i} under the mapping ff and we eliminate those points dropping out of I1IkI_{1}\cup...\cup I_{k} to get a set Si+1f(Mi)S_{i+1}\subseteq f(M_{i}). Then let Mi+1=MiSi+1M_{i+1}=M_{i}\cup S_{i+1}.

Finally we obtain a sequence of sets M0M1MiM_{0}\subseteq M_{1}\subseteq\ldots\subseteq M_{i}\subseteq\ldots, where each MiM_{i} contains the images of the initial end point set M0M_{0} under {id,f,f2,,fiid,f,f^{2},\ldots,f^{i}} except for those dropping out of I1IkI_{1}\cup\ldots\cup I_{k}.

Then we can find a sufficiently large integer NN such that xMNMN1\forall x\in M_{N}-M_{N-1}, dist(x,MN1)<δdist(x,M_{N-1})<\delta, where δ\delta comes from Lemma 2.1. Such NN exists because if not, there will be a sequence of disjoint points in [a1,bk][a_{1},b_{k}] with pairwise distance larger than δ\delta, which is absurd. Therefore, we can take small perturbation of ff to get a map f~\tilde{f} with f~f<δ\lVert\tilde{f}-f\rVert<\delta, and MN1~=MN~\widetilde{M_{N-1}}=\widetilde{M_{N}} for f~\tilde{f}. Indeed, we just move each point (x0,f(x0))MN1×MN(x_{0},f(x_{0}))\in M_{N-1}\times M_{N} on the graph of ff to the nearest point (x0,y0){x0}×MN1(x_{0},y_{0})\in\{x_{0}\}\times M_{N-1} preserving the continuity and the covering property of ff. For example, in a neighborhood of each x0MNx_{0}\in M_{N}, we can define f~(x)=f(x)±σϕ(xx0)\tilde{f}(x)=f(x)\pm\sigma\phi(x-x_{0}), where 0ϕ10\leq\phi\leq 1 is a cut-off function, ϕ(0)=1\phi(0)=1, σ<δ\sigma<\delta is chosen so that f~(x0)MN1\tilde{f}(x_{0})\in M_{N-1}. And for different x0x_{0} we choose different signs ±\pm so that (I1,I2,,Ik;f~)(I_{1},I_{2},\ldots,I_{k};\tilde{f}) is still a covering system. (See Figure 1 and 2.)

Thus, if we divide the intervals into small pieces using the points in MN~\widetilde{M_{N}}, then f~\tilde{f} will have the nice property that the image of each subinterval contains entirely another one or more subintervals. In other words, f~\tilde{f} can be regarded as a discrete mapping from {1,2,,n}\{1,2,\ldots,n\} to {subsets of {1,2,,n}\{1,2,\ldots,n\}}. Here each number jj represents a small piece of interval JjJ_{j}. And jf(j)j^{\prime}\in f(j) if f(Jj)Jjf(J_{j})\supseteq J_{j^{\prime}}. Note that the image of some interval may be empty, since we ignore the part outside I1IkI_{1}\cup\ldots\cup I_{k}. And we only care about the interval determined by {f(a),f(b)}\{f(a),f(b)\} if a,bMN~a,b\in\widetilde{M_{N}}, although the image f([a,b])f([a,b]) may be larger.

Example 2.2.

The two figures illustrate how we perturb ff locally to obtain f~\tilde{f} with f~\tilde{f} still continuous and inducing a covering system close to the initial one.

Refer to caption
Figure 1: the initial ff
Refer to caption
Figure 2: the construction of f~\tilde{f}

It should also be remarked here that the only useful information of ff on the so called “gaps” (see [1]) between two adjacent intervals IiI_{i} and Ii+1I_{i+1} is the behavior of the end points of each gap. In other words, for the image f(Jj)f(J_{j}) of each small subinterval JjJ_{j}, we only focus on the part f(Jj)(I1Ik)f(J_{j})\cap(I_{1}\cup\ldots\cup I_{k}). This is the reason why we eliminate the points outside I1IkI_{1}\cup\ldots\cup I_{k} in each step. We will see later that this information is enough for the proof.

To sum up, once we have a mapping ff satisfying the conditions in Lemma 2.1, we can find an f~\tilde{f} close enough to ff, which also satisfies the conditions but can be viewed as a discrete mapping. Such ff and f~\tilde{f} 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 ff: {1,2,,n}\{1,2,\ldots,n\} \to {subsets of {1,2,,n}}\{1,2,\ldots,n\}\} such that i=1nf(i)={1,2,,n}\bigcup_{i=1}^{n}f(i)=\{1,2,\ldots,n\}, and for any partition

I1={1,2,,i1},I2={i1+1,i1+2,,i2},,Ik={ik1+1,,n},I_{1}=\{1,2,\ldots,i_{1}\},I_{2}=\{i_{1}+1,i_{1}+2,\ldots,i_{2}\},\ldots,I_{k}=\{i_{k-1}+1,\ldots,n\},

there exists j{1,2,,k}j\in\{1,2,\ldots,k\} and r,sIjr,s\in I_{j} (r=sr=s is allowed) such that

(convf)l({r,s}){r,s},forsomelk.(convf)^{l}(\{r,s\})\supseteq\{r,s\},\quad for\ some\quad l\leq k.

Here, convf(A)=convf(A)= convex hull of f(A)={minf(A),minf(A)+1,,maxf(A)}f(A)=\{\min f(A),\min f(A)+1,\ldots,\max f(A)\} for any finite set AA\subseteq\mathbb{N}.

Remark 2.4.

We emphasize here again the difference between f([a,b])f([a,b]) and [f(a),f(b)][f(a),f(b)]. In general the former is larger. And one may worry that the discrete mapping ff induced from a covering system does not satisfy the corresponding covering property i=1nf(i)={1,2,,n}\bigcup_{i=1}^{n}f(i)=\{1,2,\ldots,n\}, since in the definition of ff, jf(j)j^{\prime}\in f(j) iff f(Jj)Jjf(J_{j})\supseteq J_{j^{\prime}}, where f(Jj)f(J_{j}) is understood in the sense of [f(a),f(b)][f(a),f(b)] rather than f([a,b])f([a,b]). 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 IjI_{j} are mapped to the end points of f(Ij)f(I_{j}). And it’s now easily checked that the perturbation and discretization preserve the covering property of ff, since the perturbation does not change the position of the end points of f(Ij)f(I_{j}) essentially.

Proof of Theorem 1.2 by lemma 2.3.

Suppose the covering system (I1,,Ik;f)(I_{1},\ldots,I_{k};f) is a counterexample to Theorem 1.2. Then as mentioned above, after perturbation, we can assume ff satisfies the conditions of lemma 2.3. Therefore the conclusion of lemma 2.3 gives us a subinterval of IjI_{j} represented by conv{r,s}conv\{r,s\} whose image under flf^{l} contains itself. Thus there must exist a periodic point of period lk\leq l\leq k in IjI_{j}, 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 ff be in lemma 2.3. We can assume further:

(1)ij\quad\forall i\neq j, f(i)f(j)=f(i)\cap f(j)=\emptyset.

(2)i\quad\forall i, f(i)f(i)\neq\emptyset.

(3)i\quad\forall i, f(i)f(i) contains exactly one element.

(4)f\quad f is a cyclic permutation in the symmetric group SnS_{n}, i.e.i.e., ff can be written as (i1i2in)(i_{1}i_{2}\ldots i_{n}).

Indeed, (1) can be realized since we can move out the common part of f(i)f(i) and f(j)f(j) from one of them without changing the property i=1nf(i)={1,2,,n}\bigcup_{i=1}^{n}f(i)=\{1,2,\ldots,n\}.

(2) is also satisfied because we can eliminate the number ii and restrict the mapping ff to {1,2,,n}{i}\{1,2,\ldots,n\}-\{i\} if necessary.

(3) is a consequence of (1), (2) and the condition i=1nf(i)={1,2,,n}\bigcup_{i=1}^{n}f(i)=\{1,2,\ldots,n\}. 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 fSnf\in S_{n} because of (3). Suppose ff is not cyclic. Then choose an orbit (i1im)(i_{1}\ldots i_{m}) of ff (m<nm<n) and restrict both ff and the partition I1,,IkI_{1},\ldots,I_{k} to {i1,,im}\{i_{1},\ldots,i_{m}\}.

In conclusion, we obtain:

Proposition 3.1.

It is sufficient to consider fSnf\in S_{n} and ff being cyclic in lemma 2.3.

Now we turn to the proof of the equivalence of lemma 2.3 and lemma 1.6. (See also Definition 1.4.)

Proposition 3.2.

Lemma 1.6 implies lemma 2.3.

Proof.

As mentioned above, we can assume fSnf\in S_{n} being cyclic, without loss of generality.

Suppose lemma 1.6 is true. Then for any partition

I1={1,2,,i1},I2={i1+1,,i2},,Ik={ik1+1,,n},I_{1}=\{1,2,\ldots,i_{1}\},\quad I_{2}=\{i_{1}+1,\ldots,i_{2}\},\quad\ldots,\quad I_{k}=\{i_{k-1}+1,...,n\},

the (k1)(k-1) characteristic numbers {mi1,mi2,,mik1}\{m_{i_{1}},m_{i_{2}},\ldots,m_{i_{k-1}}\} cannot cover {m1,,mk}\{m_{1}^{\prime},\ldots,m_{k}^{\prime}\}. Therefore, ti1,,ik1\exists\,t\neq i_{1},\ldots,i_{k-1}, and lkl\leq k such that mt=mlkm_{t}=m_{l}^{\prime}\leq k, since m1mkkm_{1}^{\prime}\leq\ldots\leq m_{k}^{\prime}\leq k in the characteristic sequence. In other words,

(convf)mt(At)At,whereAt={t,t+1}someIj.(convf)^{m_{t}}(A_{t})\supseteq A_{t},\quad where\quad A_{t}=\{t,t+1\}\subseteq some\ I_{j}.

Therefore, we can just choose r=t,s=t+1r=t,\ s=t+1 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 ff is a cyclic permutation in Definition 1.4, we can construct a directed graph Γf\Gamma_{f} with (n1)(n-1) vertices {A1,A2,,An1}\{A_{1},A_{2},\ldots,A_{n-1}\}, where Ai={i,i+1}A_{i}=\{i,i+1\} is defined in 1.4. And there is an edge from AiA_{i} to AjA_{j} if and only if convf(Ai)Ajconvf(A_{i})\supseteq A_{j}.

Example 3.3.

Let f=(136245)S6f=(136245)\in S_{6} as in example 1.5. Then the associated directed graph Γf\Gamma_{f} is drawn in figure 3.

Refer to caption
Figure 3: the directed graph Γf\Gamma_{f}

Thus, we see directly from the definition that the following proposition holds:

Proposition 3.4.

For a cyclic fSnf\in S_{n} and its associated directed graph Γf\Gamma_{f}, the characteristic number mim_{i} is equal to the minimal length of cycles starting from AiA_{i} in Γf\Gamma_{f}.

Using directed graphs, we can complete the proof of the equivalence of the two lemmas.

Proposition 3.5.

Lemma 1.6 is equivalent to lemma 2.3.

Proof.

Suppose lemma 2.3 is true. For a cyclic fSnf\in S_{n}, suppose on the contrary that there is a kn1k\leq n-1 such that mkk+1m_{k}^{\prime}\geq k+1, and kk is the minimal number with this property. Then there are only (k1)(k-1) characteristic numbers mi1,mi2,,mik1km_{i_{1}},m_{i_{2}},\ldots,m_{i_{k-1}}\leq k. Without loss of generality, assume i1<<ik1i_{1}<\ldots<i_{k-1}. And consider the partition:

I1={1,2,,i1},I2={i1+1,,i2},,Ik={ik1+1,,n}.I_{1}=\{1,2,\ldots,i_{1}\},\quad I_{2}=\{i_{1}+1,\ldots,i_{2}\},\quad\ldots,\quad I_{k}=\{i_{k-1}+1,\ldots,n\}.

By lemma 2.3, we can find j,lkj,l\leq k and r,sIjr,s\in I_{j} (r<sr<s) such that

(convf)l({r,s}){r,s}.(convf)^{l}(\{r,s\})\supseteq\{r,s\}.

Note that r=sr=s cannot happen in this case, since mr=msk+1m_{r}=m_{s}\geq k+1. Now if we extend ff to be a piecewise linear mapping from the interval [1,n][1,n] to [1,n][1,n], then we will obtain fl([r,s])[r,s]f^{l}([r,s])\supseteq[r,s]. Therefore, there exists x0[r,s]x_{0}\in[r,s] with fl(x0)=x0f^{l}(x_{0})=x_{0}. Note that since ff is cyclic as a permutation in SnS_{n} and lkn1l\leq k\leq n-1, x0x_{0} cannot belong to {1,2,,n}\{1,2,\ldots,n\}. Consequently, x0(t,t+1)x_{0}\in(t,t+1) for some integer t[r,s1]t\in[r,s-1] and it is similar for f(x0),f2(x0),,fl1(x0)f(x_{0}),f^{2}(x_{0}),\ldots,f^{l-1}(x_{0}). Since ff is piecewise linear, x0(t,t+1)x_{0}\in(t,t+1), f(x0)(t,t+1)f(x_{0})\in(t^{\prime},t^{\prime}+1) imply f([t,t+1])[t,t+1]f([t,t+1])\supseteq[t^{\prime},t^{\prime}+1], and it is similar for other pairs of points. So the orbit {x0,f(x0),,fl1(x0),fl(x0)}\{x_{0},f(x_{0}),\ldots,f^{l-1}(x_{0}),f^{l}(x_{0})\} gives a cycle of length lkl\leq k starting from the vertex AtA_{t} in the associated directed graph Γf\Gamma_{f}, which implies mtkm_{t}\leq k. But this tt together with i1,,ik1i_{1},\ldots,i_{k-1} gives kk characteristic numbers k\leq k, contradicting mkk+1m_{k}^{\prime}\geq k+1.

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.

Example 3.6.

(1) Let fSnf\in S_{n} be (12n1)(1\to 2\to\ldots\to n\to 1). And more generally, f(i)i+m(modn)f(i)\equiv i+m\pmod{n}, where (m,n)=1(m,n)=1. Then the characteristic sequence of ff is 12n11\leq 2\leq\ldots\leq n-1.

(2) Let fS2n+1f\in S_{2n+1} be of Stefan type (see [2] or [5] for more details), that is,

1(n+1)(n+2)n(n+3)(n1)2n2(2n+1)1.1\to(n+1)\to(n+2)\to n\to(n+3)\to(n-1)\to\ldots\to 2n\to 2\to(2n+1)\to 1.

Then the characteristic sequence of ff is

122442n22n22n.1\leq 2\leq 2\leq 4\leq 4\leq\ldots\leq 2n-2\leq 2n-2\leq 2n.

(3) In [6], the author gives all the directed graphs associated to cyclic permutations in S5S_{5}.

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 fSnf\in S_{n} be a cyclic permutation and Γf\Gamma_{f} be its associated directed graph. (See the discussion before propositon 3.4.) We can define TfT_{f} to be the (n1)×(n1)(n-1)\times(n-1) adjacency matrix of ff and Γf\Gamma_{f}. That is, Tij=1T_{ij}=1, if convf(Aj)Aiconvf(A_{j})\supseteq A_{i}, or equivalently, if there is an edge from vertex AjA_{j} to vertex AiA_{i} in Γf\Gamma_{f}. Otherwise, Tij=0T_{ij}=0.

Example 4.2.

Let f=(136245)S6f=(136245)\in S_{6} be the permutation in example 1.5 and example 3.3. Then the adjacency matrix TfT_{f} is

[0001100010100100101001100]\begin{bmatrix}&0&0&0&1&1&\\ &0&0&0&1&0&\\ &1&0&0&1&0&\\ &0&1&0&1&0&\\ &0&1&1&0&0&\end{bmatrix}

We denote the field {0,1}\{0,1\} with exactly two elements by 𝔽2\mathbb{F}_{2}. And we say AB(mod2)A\equiv B\pmod{2} for integer valued matrices AA and BB, when A=BA=B as matrices in M(𝔽2)M(\mathbb{F}_{2}).

Proposition 4.3.

For any cyclic permutation fSnf\in S_{n} and l+l\in\mathbb{N}_{+}, it holds that TflTfl(mod2)T_{f}^{l}\equiv T_{f^{l}}\pmod{2}.

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 f=(i1i2in)Snf=(i_{1}i_{2}\ldots i_{n})\in S_{n} cyclic, we first extend ff to be a continuous piecewise linear mapping from the closed interval [1,n][1,n] to itself. And recall we define Ai={i,i+1}A_{i}=\{i,i+1\}. Later on we will also regard AiA_{i} as the closed interval [i,i+1][i,i+1].

Now by the definition of TfT_{f} and the continuity of ff, (Tf)ij=1(T_{f})_{ij}=1 iff f(Aj)Aif(A_{j})\supseteq A_{i}. Consider the interval Ai=[i,i+1]A_{i}=[i,i+1] and its image fl(Ai)f^{l}(A_{i}). In general, we cannot expect fl(Ai)=[fl(i),fl(i+1)]f^{l}(A_{i})=[f^{l}(i),f^{l}(i+1)] to hold, where [a,b][a,b] denotes the closed interval whose end points are aa and bb. We can only say that fl(Ai)[fl(i),fl(i+1)]f^{l}(A_{i})\supseteq[f^{l}(i),f^{l}(i+1)] for continuous ff. However, if we count the multiplicity of each point in the image fl(Ai)f^{l}(A_{i}), we will find that fl(Ai)=[fl(i),fl(i+1)]f^{l}(A_{i})=[f^{l}(i),f^{l}(i+1)] actually holds in the sense of mod 2. Indeed, fl(Ai)f^{l}(A_{i}) can be viewed as a continuous curve in \mathbb{R} starting from the point fl(i)f^{l}(i) and ending at fl(i+1)f^{l}(i+1). Therefore, every point pfl(Ai)p\in f^{l}(A_{i}) is counted an even number of times in this curve, unless p[fl(i),fl(i+1)]p\in[f^{l}(i),f^{l}(i+1)].

Finally recall the definition of TfT_{f} and TflT_{f^{l}}, 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 SnS_{n}, then they may find that TfT_{f} defined above is just the matrix of the tautological representation of ff mod 2.

Corollary 4.5.

For cyclic fSnf\in S_{n}, we have TfnI(mod2)T_{f}^{n}\equiv I\pmod{2}, where I=In1I=I_{n-1} is the unit matrix.

Theorem 4.6.

For cyclic fSnf\in S_{n}, the characteristic polynomial of TfT_{f}, viewed as the determinant of a matrix in Mn1(𝔽2)M_{n-1}(\mathbb{F}_{2}), is exactly:

det(λITf)=1+λ+λ2++λn1(mod2).\det(\lambda I-T_{f})=1+\lambda+\lambda^{2}+\ldots+\lambda^{n-1}\pmod{2}.

We need the following lemma:

Lemma 4.7.

There exists a vector α𝔽2n1\alpha\in\mathbb{F}_{2}^{n-1} such that α,Tfα,,Tfn2α\alpha,T_{f}\alpha,\ldots,T_{f}^{n-2}\alpha are linearly independent over the field 𝔽2\mathbb{F}_{2}.

Proof.

We claim that if fSnf\in S_{n} is written as (i1i2in)(i_{1}i_{2}\ldots i_{n}), then the vector α=(0,,0,1,,1,0,,0)T=j=i1i21ej\alpha=(0,\ldots,0,1,\ldots,1,0,\ldots,0)^{T}=\sum_{j=i_{1}}^{i_{2}-1}e_{j} (or j=i2i11ej\sum_{j=i_{2}}^{i_{1}-1}e_{j}) satisfies the desired property, where ej𝔽2n1e_{j}\in\mathbb{F}_{2}^{n-1} is the unit vector whose jj-th component is 11.

We prove the claim by induction on nn. The case n=2n=2 is trivial. We assume the statement holds for all natural numbers <n<n, and we will prove it for nn.

Suppose

μ1α+μ2Tfα++μn1Tfn2α=0.\mu_{1}\alpha+\mu_{2}T_{f}\alpha+\ldots+\mu_{n-1}T_{f}^{n-2}\alpha=0.

By the definition of α\alpha, f=(i1i2in)Snf=(i_{1}i_{2}\ldots i_{n})\in S_{n}, as well as the argument in the proof of proposition 4.3, we can see that

Tfmα=j=im+1im+21ej,orj=im+2im+11ej,mn2.T_{f}^{m}\alpha=\sum_{j=i_{m+1}}^{{i_{m+2}-1}}e_{j},\quad or\sum_{j=i_{m+2}}^{{i_{m+1}-1}}e_{j},\quad\forall m\leq n-2.

In other words, if we view each vector eje_{j} as the closed interval [j,j+1][j,j+1] it represents, then we find that

α=[i1,i2],Tfα=[i2,i3],,Tfn2α=[in1,in].\alpha=[i_{1},i_{2}],\quad T_{f}\alpha=[i_{2},i_{3}],\quad\ldots\quad,T_{f}^{n-2}\alpha=[i_{n-1},i_{n}].

Now if i1=1i_{1}=1, take the inner product with e1e_{1}, and we obtain:

0=(μ1α+μ2Tfα++μn1Tfn2α,e1)=μ1,0=\left(\mu_{1}\alpha+\mu_{2}T_{f}\alpha+\ldots+\mu_{n-1}T_{f}^{n-2}\alpha,\>e_{1}\right)=\mu_{1},

because i2,,in2i_{2},\ldots,i_{n}\geq 2 and hence (Tfmα,e1)=0(T_{f}^{m}\alpha,e_{1})=0 for 1mn21\leq m\leq n-2. Similarly we can obtain μ1=0\mu_{1}=0 when i1=ni_{1}=n.

If 2i1n12\leq i_{1}\leq n-1, then since i1i_{1} does not appear as the end points of TfmαT_{f}^{m}\alpha, [i11,i1][i_{1}-1,i_{1}] and [i1,i1+1]orTfmα[i_{1},i_{1}+1]\subseteq or\not\subseteq T_{f}^{m}\alpha at the same time, where m1m\geq 1. Or equivalently,

(Tfmα,ei11)=(Tfmα,ei1),1mn2.(T_{f}^{m}\alpha,e_{i_{1}-1})=(T_{f}^{m}\alpha,e_{i_{1}}),\quad 1\leq m\leq n-2.

And then if we take the inner product with (ei11+ei1)(e_{i_{1}-1}+e_{i_{1}}), we obtain:

0=(μ1α+μ2Tfα++μn1Tfn2α,(ei11+ei1))=μ1(mod2),0=\left(\mu_{1}\alpha+\mu_{2}T_{f}\alpha+\ldots+\mu_{n-1}T_{f}^{n-2}\alpha,\>(e_{i_{1}-1}+e_{i_{1}})\right)=\mu_{1}\pmod{2},

Anyway, we find that μ1=0\mu_{1}=0 and consequently,

μ2Tfα++μn1Tfn2α=0.\mu_{2}T_{f}\alpha+\ldots+\mu_{n-1}T_{f}^{n-2}\alpha=0.

Now if we put g=(i2i3in)Sn1g=(i_{2}i_{3}\ldots i_{n})\in S_{n-1} and β=Tfα\beta=T_{f}\alpha with the i1i_{1}-th component dropped, we can see that Tfmα=Tgm1βT_{f}^{m}\alpha=T_{g}^{m-1}\beta, regardless of the i1i_{1}-th component of TfmαT_{f}^{m}\alpha. This is easily seen to be true since one can find Tfmα=Tgm1β=[im+1,im+2]T_{f}^{m}\alpha=T_{g}^{m-1}\beta=[i_{m+1},i_{m+2}] as intervals. And the following holds for β\beta and TgT_{g}:

μ2β+μ3Tgβ++μn1Tgn3β=0.\mu_{2}\beta+\mu_{3}T_{g}\beta+\ldots+\mu_{n-1}T_{g}^{n-3}\beta=0.

By induction hypothesis we conclude that μ2==μn1=0\mu_{2}=\ldots=\mu_{n-1}=0. Therefore, we have proved that the claim also holds for nn. This completes the proof of the lemma. ∎

Corollary 4.8.

The minimal polynomial g(λ)g(\lambda) of TfT_{f} has degree at least n1n-1, and therefore equals the characteristic polynomial.

Proof.

Suppose degg(λ)n2\deg g(\lambda)\leq n-2, then we have g(Tf)α=0g(T_{f})\alpha=0, which contradicts the lemma above. And the second statement follows by Cayley-Hamilton’s theorem. (Note that TfT_{f} is an (n1)×(n1)(n-1)\times(n-1) matrix.) ∎

With these results, we can compute the characteristic polynomial of TfT_{f}.

Proof of Theorem 4.6.

It follows from corollary 4.5 and 4.8 that

det(λITf)=g(λ)(λn1)=(1+λ)(1+λ+λ2++λn1)(mod2).\det(\lambda I-T_{f})=g(\lambda)\mid(\lambda^{n}-1)=(1+\lambda)(1+\lambda+\lambda^{2}+\ldots+\lambda^{n-1})\pmod{2}.

Moreover, since the only degree-one irreducible polynomials are λ\lambda and λ+1\lambda+1 in 𝔽2[λ]\mathbb{F}_{2}[\lambda], the right hand side can be further decomposed as

1+λ+λ2++λn1=(1+λ)mh(λ),1+\lambda+\lambda^{2}+\ldots+\lambda^{n-1}=(1+\lambda)^{m}h(\lambda),

where m0m\geq 0, and h(λ)h(\lambda) is a product of several irreducible polynomials of degree 2\geq 2. Hence, by comparing the degree we conclude that

det(λITf)=1+λ+λ2++λn1(mod2).\det(\lambda I-T_{f})=1+\lambda+\lambda^{2}+\ldots+\lambda^{n-1}\pmod{2}.

Proof of lemma 1.6.

We show that if the coefficient of λn1i\lambda^{n-1-i} in det(λITf)\det(\lambda I-T_{f}) is nonzero, then miim_{i}^{\prime}\leq i holds for the characteristic sequence in lemma 1.6.

Let us consider how the coefficients in det(λITf)\det(\lambda I-T_{f}) are computed. It is then easily seen that the coefficient of λn1i\lambda^{n-1-i} is nonzero implies the existence of an i×ii\times i principle matrix PiP_{i} of TfT_{f} whose determinant is nonzero. It then follows that there exists at least one diagonal of PiP_{i} whose elements are all equal to 1. Now it is easily checked that this diagonal give rise to several cycles of length i\leq i in the associated directed graph Γf\Gamma_{f}. And the vertices of these cycles correspond to the subscripts of the columns of PiP_{i}. Thus there exist at least ii vertices in Γf\Gamma_{f}, from whom the minimal length of cycles starting are all i\leq i. By proposition 3.4, this means that at least ii of the characteristic numbers i\leq i. So miim_{i}^{\prime}\leq i. ∎

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.