Sie sind auf Seite 1von 62

Combinatorics of Periodic Points of Interval Maps

Tyler London
Department of Mathematics
College of Liberal Arts and Sciences
Tufts University
Senior Honors Thesis
2009
Abstract
In 1964, Sharkovsky showed that there is a linear ordering
s
of the natural numbers
such that m
s
n if every continuous interval map having an n-periodic point has an
m-periodic point. This idea is generalized to get a partial ordering of cyclic permutations
called the forcing relation by taking into consideration the pattern of the periodic orbit.
We show that forcing can be reduced to the study of certain canonical piecewise linear
maps. Combinatorial algorithms of Baldwin and Coven et. al. that decide forcing are
presented, as well as a diagram of the forcing relation on over 190 cycles of length up to
8. The global structure of forcing is described by considering ips of cycles and unimodal
cycles. It is shown that the forcing relation restricted to unimodal cycles is a total ordering,
and this relation is diagrammed on 115 unimodal cycles. The local structure of forcing
is also explored by studying doublings and extensions. A new combinatorial proof of a
theorem describing the structure of doublings in the forcing order is given. Forcing-minimal
properties are examined by considering primary and simple cycles. Finally, using the
results proven about the forcing relation, we derive Sharkovskys theorem as a consequence
of the forcing theorem.
Key words and phrases. Interval maps, periodic orbit, Markov graph, Sharkovskys theo-
rem, cycle, forcing, doubling, block structure, extension, primary cycle, simple cycle.
Contents
Acknowledgments iii
0 Introduction 1
1 Background 3
1.1 Cyclic Permutations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.2 Continuous Maps of the Interval . . . . . . . . . . . . . . . . . . . . . . . . 5
1.3 The Theorem of Sharkovsky . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
2 The Forcing Relation on Cycles 11
2.1 The Combinatorial Data in a Periodic Orbit . . . . . . . . . . . . . . . . . . 11
2.2 The Forcing Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
3 A Computational Digression 20
3.1 Description of the Algorithm of Baldwin . . . . . . . . . . . . . . . . . . . . 20
3.2 Implementation of the Algorithm in GAP . . . . . . . . . . . . . . . . . . . 21
3.3 Eciency of the Algorithm . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
3.4 Properties of Forcing in the diagrammed data . . . . . . . . . . . . . . . . . 22
4 Global Structure of Forcing 25
4.1 The Flip Symmetry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
4.2 The Restriction of Forcing to Unimodal Permutations . . . . . . . . . . . . 26
5 Extensions of Cycles 32
5.1 Doublings . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
5.2 General Extensions of Cycles . . . . . . . . . . . . . . . . . . . . . . . . . . 39
6 Forcing-Minimal Cycles 44
6.1 Primary Cycles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
6.2 Simple Cycles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
7 A Forcing Proof of Sharkovskys Theorem 52
8 Conclusions 53
Bibliography 54
i
List of Figures
1.1 f

for = (1 5 2 6 3 4). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
1.2 f

for = (1 4 3 2 5). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.3 The Markov graph of = (1 4 3 2 5). . . . . . . . . . . . . . . . . . . . . . . 5
1.4 The Markov graph of q
4
(x) associated to the partition P. . . . . . . . . . . 8
1.5 The Markov graph associated to the partition J
1
, J
2
. . . . . . . . . . . . . . 9
2.1 Graphs of symmetric tent maps exhibiting dierent 5-cycles. . . . . . . . . . 12
2.2 Graphs of f

and f

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
2.3 f

for = (1 4 3 2 5). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
2.4 The signed Markov graph of = (1 4 3 2 5). . . . . . . . . . . . . . . . . . . 15
2.5 Markov graph of = (1 2 4 3) . . . . . . . . . . . . . . . . . . . . . . . . . . 16
3.1 Baldwins original diagram. . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
3.2 Computation of for n 8. . . . . . . . . . . . . . . . . . . . . . . . . . . 24
4.1 Illustration of the complexity of on some 4-cycles. . . . . . . . . . . . . . 25
4.2 The graph of f

for the unimodal cycle = (1 2 3 4 5). . . . . . . . . . . . . 27


5.1 The 6-cycle (1 5 3 2 6 4) as a doubling of (1 3 2). . . . . . . . . . . . . . . . . 33
5.2 The two possibilities for J
2k
J
2i
. . . . . . . . . . . . . . . . . . . . . . . . 35
5.3 The rst case: q
c
< q
i
< p
c
and p
c
a relative maximum. . . . . . . . . . . . 36
5.4 Tree of doublings starting at (1 2). . . . . . . . . . . . . . . . . . . . . . . . 38
5.5 A block structure over the 3-cycle (1 2 3). . . . . . . . . . . . . . . . . . . . 39
5.6 -extension of for = (1 3 2). . . . . . . . . . . . . . . . . . . . . . . . . . 41
6.1 The

Stefan 5-cycles and their characteristic spiraling out property. . . . . 46
6.2 The Markov Graph of a Stefan n-cycle. . . . . . . . . . . . . . . . . . . . . 47
6.3 The primary and simple 6-cycles. . . . . . . . . . . . . . . . . . . . . . . . . 48
6.4 The primary cycles of length 8. . . . . . . . . . . . . . . . . . . . . . . . . 51
ii
Acknowledgments
I want to express my gratitude to my advisor Zbigniew Nitecki for his mathematical
guidance over the past four years which has culminated in this thesis. His weekly meetings
and helpful editing were indispensable in the writing of this paper. I am grateful to Boris
Hasselblatt for his careful reading of this paper and many useful remarks and discussions.
I would also like to thank Kenneth Monks and Alexander Hulpke for teaching me to use
GAP, AJ Raczkowski for teaching me to count loops and Natasha Telfer, who helped me
prepare for the defense.
iii
Chapter 0
Introduction
This work was completed as a senior honors thesis for the 2008-2009 academic year at
Tufts University. It presents a relatively self-contained introductory survey of the com-
binatorial theory of periodic points of continuous maps of the interval and, in particular,
the forcing relation. There are 8 chapters. The rst chapter provides the necessary pre-
liminary denitions and results required for the rest of the paper. The concept of cyclic
permutations and their associated connect-the-dots maps and Markov graphs are covered
as well as several fundamental results from the theory of one-dimensional dynamics. This
chapter closes with a statement of Sharkovskys theorem on the coexistence of periodic
points of interval maps. The proof of this theorem is not given in Chapter 1, as one of
the goals of this paper is to give a proof of Sharkovskys theorem based on the theory
presented in this paper. This is done in Chapter 7.
Chapter 2 begins to analyze what happens when we consider not only the existence
of a periodic point, but how the points in a periodic orbit are permuted. In particular,
we see that to each periodic point we can associate a cyclic permutation. We dene the
forcing relation on cyclic permutations and show that it is a partial ordering. We show
that forcing is encoded in certain canonical maps and prove the Forcing Theorem, detailing
necessary and sucient conditions for one cycle to force another.
In Chapter 3, the computational aspects of forcing are explored. Chapter 2 provided
criteria for a cycle to force , but are these conditions computationally feasible? The
combinatorial algorithm of Baldwin that decides forcing is presented along with its im-
plementation in the computer algebra system GAP. The ineciency of this algorithm is
discussed, and we present other more ecient algorithms which decide forcing. Finally,
we present a signicantly improved diagram of the forcing relation on nearly 200 cycles of
length up to 8.
The next three chapters analyze various properties of forcing. Having identied forcing
as a partial order on the set of all cycles, what does its global structure look like? We
address this question in Chapter 4 by showing that there are two distinct global properties
of forcing: one is a sort of global symmetry between a cycle and its ip. The other global
property is that when we restrict the forcing order to unimodal cycles, it is a linear order.
The fth chapter takes a look at the local structure of forcing. Do there exist cycles
which are immediately above or immediately below a given cycle in the forcing order?
In Section 5.1, we show that immediately above each cycle lie all of its doubles. We
give a combinatorial proof of a theorem that characterizes three important properties
of doublings. Next, we analyze more general types of extensions of cycles. The Markov
graph of an extension has an easily described structure, and we prove several results about
1
extensions which give us even more insight into the local structure of forcing.
In Chapter 6, we study forcing-minimal properties. Given a xed n, what do the n-
cycles that force no other n-cycles look like? First, we discuss the notion of a primary
cycle. For any odd number q > 1, there are precisely two primary q-cycles, the

Stefan
cycles. However, for even numbers there can be very many primary cycles. Next, we
discuss simple cycles. The essential idea of a simple cycle is that all points left of some
middle division move to the right and all points right of the division move to the left.
We prove that every simple cycle is primary and sketch the proof that a primary cycle is
simple. To close the chapter, we give a diagram of all 33 primary cycles of length 8.
Finally, in Chapter 7, using the various results about the local, global and minimality
properties of forcing elucidated in earlier chapters, we give a proof of Sharkovskys theorem.
We conclude this paper with Chapter 8 where we discuss dierent generalizations of the
theory presented. Among the topics discussed are the forcing relation on generalized
patterns permutations, not just cycles and the generalization of the problem to two-
dimensional dynamical systems.
In many of the denitions and proofs it is very important to have an idea of how the
points in consideration are moving around. We have tried to provide numerous gures
throughout most of the paper to clear up any confusion.
2
Chapter 1
Background
The purpose of this chapter is to provide the reader with the necessary preliminary def-
initions and results required for the rest of the paper. In Section 1.1, we briey dene
cyclic permutations, establish our notation and give some of their properties of cycles.
Though most of the denitions presented in this section are not used in the remainder of
this chapter, they form an integral part of the remaining chapters. In Section 1.2, interval
maps and periodic points are dened. The main results of this section relate the existence
of periodic points to the existence of certain loops in Markov graphs of interval maps (Def-
inition 1.2.5). This result provides the foundation of our combinatorial study of periodic
points. Finally, Section 1.3 presents Sharkovskys beautiful theorem that describes the
coexistence of periodic points of interval maps. The proof of this theorem will be given
in Chapter 7 once the machinery of forcing has been suciently developed, but the main
idea of the standard proof is sketched by presenting a similar result of Li and Yorke.
For thorough expositions on interval maps and one-dimensional dynamics, the interested
reader is directed to [ALM93], [BS02, Chapter 7], [dMvS93], [BB04], [CE80], [BC92b]
and [Nit82].
1.1 Cyclic Permutations
Here we give a concise introduction to the concept of n-cycles and cyclic permutations
that will become the focus of this paper from Chapter 2 onward, where we consider sets
of points that are permuted cyclically under continuous maps.
Denition 1.1.1. A cyclic permutation of length n or simply an n-cycle is any bijection
: 1, 2, . . . , n 1, 2, . . . , n such that (1),
2
(1), . . . ,
n1
(1) are all distinct. We will
denote a given n-cycle using cycle notation by writing
= ((1)
2
(1) . . .
n1
(1)).
The length of a cycle will be denoted [[ and the set of all n-cycles will be denoted C
n
.
By a cycle, we mean any element of the set C =

nN
C
n
.
Example 1.1.2. Consider the 3-cycle dened by (1) = 2, (2) = 3, and (3) = 1.
Using cycle notation, we have = (1 2 3).
Given an n-cycle , one can naturally associate to two objects: a continuous map of
the interval [1, n] and a directed graph.
3
Denition 1.1.3. Let be a cyclic permutation of length n and for each i 1, . . . , n1,
let
i
= (i +1) (i). The connect-the-dots map or canonical -linear map f

: [1, n]
[1, n] is dened by
f

(i +x) = (i) +x
i

for 0 x 1.
Note that f

1, . . . , n
= . Simply put, f

is the piecewise linear map obtained by


connecting the points (i, (i)) and (i + 1, (i + 1)) with line segments for i = 1, . . . , n 1.
In addition, we say that k N, 1 < k < n is a maximum (resp. minimum) for the n-cycle
if (k) > max(k 1), (k + 1) (resp. (k) < min(k 1), (k + 1)); in either case,
k is a critical point or turning point for . This is equivalent to saying that f

is not
monotone on the set k 1, k, k + 1 or that k is a local extremum of f

.
Example 1.1.4. Consider the 6-cycle = (1 5 2 6 3 4). The graph of f

is shown in Figure
1.1, and k = 2 is a maximum of and k = 4 is a minimum of .
1 2 3 4 5 6
1
2
3
4
5
6
Figure 1.1: f

for = (1 5 2 6 3 4).
A directed graph is a pair G = (V, E) where V is a set whose elements are called vertices
and E is a set of ordered pairs of vertices called directed edges. If G is a directed graph,
then a path of length k in G is a k-tuple (v
1
, . . . , v
k
) such that each v
i
is a vertex and
(v
i
, v
i+1
) is a directed edge for 1 i k. We will primarily be concerned with paths in
directed graphs that begin and end at the same vertex. In particular, a path (v
1
, . . . , v
k
)
in G is called a closed path of length k or a loop if (v
k
, v
1
) is also a directed edge. In graph
theory, a closed path may also be referred to as a circuit.
Denition 1.1.5. Suppose is an n-cycle. Dene J
i
= [i, i + 1] for i = 1, . . . , n 1 and
let V = J
1
, . . . , J
n1
. The Markov graph associated to is the directed graph with n1
vertices V and an arrow (directed edge) drawn from J
k
to J
l
if and only if f

(J
k
) J
l
.
Example 1.1.6. Let = (1 4 3 2 5). By looking at the graph of f

(Figure 1.2), we see


that f

([1, 2]) [4, 5], f

([2, 3]) [2, 5], f

([3, 4]) [2, 3], and f

([4, 5]) [1, 3]. Therefore,


4
the Markov graph of has four vertices J
i
= [i, i + 1] for 1 i 4 with directed edges
from: J
1
to J
4
; J
2
to itself, J
3
and J
4
; J
3
to J
2
, and from J
4
to J
1
and J
2
. This graph is
illustrated in Figure 1.3.
1 2 3 4 5
1
2
3
4
5
Figure 1.2: f

for = (1 4 3 2 5).
J1 J2
J3 J4
Figure 1.3: The Markov graph of =
(1 4 3 2 5).
Lastly, we remark that if C
n
, then we will generally consider two dierent labellings
of the elements of P = p
1
, . . . , p
n
= 1, . . . , n. The rst is the spatial labeling where
p
i
< p
j
if and only if i < j. The second is the temporal labeling where p
i
=
i1
(p
1
)
for i = 2, . . . , n. Note that the above denition of the temporal labeling is equivalent to
saying p
i
= (p
i1
) for i = 2, . . . , n and p
1
= (p
n
).
1.2 Continuous Maps of the Interval
We aim to explore the combinatorial properties of periodic points of interval maps. By
an interval map, we mean a continuous function f : I I of an interval I R into itself.
Oftentimes, the family of all continuous self-maps of an interval I will be of interest, so
A(I) shall denote this set. Given f A(I), we shall examine the iterates f
n
of f, which
are dened inductively by f
0
= id and f
n
= f f
n1
for n N. For x I and f A(I),
the orbit of x under f is dened to be the set
O(x) = f
n
(x) : n = 0, 1, 2, . . ..
A point x is said to be periodic if for some k N, f
k
(x) = x. In this case, O(x) is nite
and the set x, f(x), . . . , f
k1
(x) is called a periodic orbit. The period
1
of a periodic point
x is the least positive integer n such that f
n
(x) = x; that is, f
n
(x) = x and f
k
(x) ,= x for
1 k n 1. We may also call a periodic point of period n an n-periodic point. Lastly,
if x is a 1-periodic point, f(x) = x, then we call x a xed point. Note that an n-periodic
point x is a xed point under the map f
n
, f
2n
, f
3n
, . . .; however, a xed point of the map
f
n
need not be a periodic point of period n.
1
In other literature, what we have dened to be the period of a point is sometimes called the least or
minimal period.
5
Example 1.2.1. Let be an n-cycle. Then it is clear from the denition of that
1, 2, . . . , n is a periodic orbit of period n for f

.
If P is a nonempty subset of an interval I, then P) will denote the closed convex hull
of P; that is, the smallest closed connected subset of I containing P. For example, in
R, 2, 3) = 2, 3) = [2, 3] and (1, 5)) = [1, 5]. Let f A(I) and P = p
1
< p
n

be a nite subset of I. We say that f is P-monotone if it is constant on each of the


two connected components of XP) and f is monotone on each interval [p
i
, p
i+1
] for
i 1, . . . , n 1. If I = [1, n] and P = 1, 2, . . . , n, then the simplest example of a
P-monotone map is the connect-the-dots map for some n-cycle .
Throughout this paper, a main strategy will be to associate certain directed graphs
to interval maps in a way so that the combinatorial properties of the graphs correspond
to certain properties of the periodic points of these maps. In particular, given f A(I),
we aim to show that every loop in the directed graph associated to f corresponds to a
periodic point of f.
Suppose I is an interval, f A(I) and J
1
and J
2
are closed subintervals of I. Then,
we say that J
1
f-covers J
2
and write J
1
J
2
if J
2
f(J
1
). Our rst goal is to show that
the existence of periodic points under a map f A(I) is related to how f acts on closed
subintervals of I.
Lemma 1.2.2. If f is a continuous map and I and J are closed intervals such that
f(I) J, then there exists a closed subinterval

I I such that f(

I) = J, f(int

I) = int J
and f(

I) = J.
Proof. Let J = [a, b]. By hypothesis, f(I) J, so there exist points a
0
, b
0
I such that
f(a
0
) = a and f(b
0
) = b. First, let us assume that a
0
< b
0
. Dene
a
1
= supx : a
0
x b
0
and f(x) = a.
Observe that by continuity, f(a
1
) = a and since f(b
0
) = b > a, a
1
is strictly less than b
0
.
Next, dene
b
1
= infx : a
1
x b
0
and f(x) = b,
again noting that f(b
1
) = b. Now, f(a
1
, b
1
) = a, b, and by denition of a
1
and b
1
, no
point of (a
1
, b
1
) maps to the boundary of J. Thus, f(int(a
1
, b
1
)) = intJ = (a, b) and so

I = [a
1
, b
1
] is the desired subinterval of I. The case when b
0
< a
0
is handled analogously,
interchanging the inmum and supremum.
We shall say that an interval J minimally f-covers an interval I if J satises the
conclusion of Lemma 1.2.2.
Lemma 1.2.3. If I
k
: k = 0, 1, . . . is a (nite of innite) family of nonempty closed
subintervals of I such that f(I
k
) I
k+1
(i.e. I
k
I
k+1
), then there exists a point x I
0
such that f
k
(x) I
k
for each k.
Proof. By Lemma 1.2.2, since f(I
0
) I
1
, there is a closed subinterval J
1
of I
0
that
minimally f-covers I
1
. Proceeding by induction, suppose that we have dened subintervals
J
1
J
2
J
n
in I
0
such that f
k
(J
k
) = I
k
and J
k
minimally f
k
-covers I
k
for every
1 k n. Then,
f
n+1
(J
n
) = f(f
n
(J
n
)) = f(I
n
) I
n+1
.
Appealing to Lemma 1.2.2, there exists a closed subinterval J
n+1
J
n
that minimally
f
n+1
-covers I
n+1
, thus f
n+1
(J
n+1
) = I
n+1
. We now have a nested sequence J
n

nN
of
6
non-empty closed intervals, so the intersection

n=1
J
n
is nonempty, and for any x in this
intersection, we have f
k
(x) I
k
.
Lemma 1.2.4. If J I is a closed subinterval such that f(J) J, then f has a xed
point in J.
Proof. Let a < b be the endpoints of J. Since f(J) J, there exist points , J such
that f() a and f() b. If we let g(x) = f(x) x, then
g() = f() f() a 0
and
g() = f() f() b 0.
As g is continuous, by the Intermediate Value Theorem, there exists a point x J such
that g(x) = 0; that is, x is a xed point of f.
Before we discuss the implications of the above lemmas, let us rst associate to each
interval map a directed graph which has the property that f-covering relations are encoded
as paths.
Given an interval I, a partition is a (nite or innite) collection of closed subintervals
J
k
of I, with pairwise disjoint interiors such that
k
J
k
= I. Given a partition J
k
of
I, an interval map f A(I) and a point x I, the itinerary of x of length n is the word
itin
n
(x) = (J
i1
, . . . , J
in
) in the letters J
1
, J
2
, . . ., such that f
k
(x) I
i
k
for 1 k n.
Itineraries give us information about where a point has been and where it might go,
without precisely telling us where the point is. Itineraries are very useful tools in symbolic
dynamics, and we will revisit their use in later chapters when unimodal permutations are
studied.
Denition 1.2.5. The Markov graph of f A(I) associated to the partition J
k
is
the directed graph with vertices J
k
, and a directed edge from J
l
to J
m
if and only if J
i
f-covers J
m
.
Example 1.2.6. The family of interval maps q
a
(x) = ax(1 x) : a R is referred
to as the quadratic or logistic family. Consider the partition P = J
1
= [0, 1/4], J
2
=
[1/4, 3/4], J
3
= [3/4, 1] of the interval [0, 1]. For a = 4, the quadratic map q
4
(x) =
4x(1x) maps [0, 1/4] to [0, 3/4], [1/4, 3/4] to [3/4, 1] and [3/4, 1] to [0, 3/4]. The Markov
graph of f associated to the partition P is shown in Figure 1.4.
Example 1.2.7. If we consider the partition of the interval [1, 5] given by P =
[1, 2], [2, 3], [3, 4], [4, 5], then for = (1 4 3 2 5), the Markov graph of f

associated to
the partition P is the same as in Example 1.1.6. See Figure 1.3. This illustrates that for
a connect-the-dots map, Denitions 1.1.5 and 1.2.5 coincide.
Given Denition 1.2.5, Lemma 1.2.3 reveals that every path in the Markov graph of
a partition corresponds to a point whose itinerary is exactly the sequence of vertices in
this path. Suppose, in particular, that the path is a closed path, i.e. a path of type
J
i0
J
i2
J
in1
J
i0
. Then J
i0
f
n
-covers itself, so Lemma 1.2.4 guarantees
the existence of a xed point x J
i0
of f
n
such that f
k
(x) J
i
k
and f
n
(x) = x. However,
if this loop is repetitive in the sense that it is formed by going around smaller loop several
times, then the period of the xed point may be strictly smaller than n. The above
discussion gives us the following results:
7
0
1
4
3
4
1
3
4
1
q4
J
3
J
1
J
2
Figure 1.4: The Markov graph of q
4
(x) associated to the partition P.
Corollary 1.2.8. If J
i0
J
i1
J
in1
J
n
= J
i0
is a loop in the Markov
graph of f of a partition, then there exists a xed point x J
i0
of f
n
with f
k
(x) J
i
k
for
0 k n 1.
Lemma 1.2.9. If J
i0
J
i1
J
in1
J
in
= J
i0
is a loop in the Markov graph
of f of a partition that is not formed by going p times around a shorter loop of length m
where mp = n, then the periodic point x of Corollary 1.2.8 has period n.
Any loop that cannot be formed by going multiple times around a shorter loop as in
the statement of Lemma 1.2.9 will be called a simple or nonrepetitive loop.
1.3 The Theorem of Sharkovsky
In the previous section, we showed that given an interval map and a partition of the
interval, by analyzing the Markov graph of a partition we can gain useful information
about the existence of periodic orbits. Suppose now we know that x
1
< < x
n
is a
periodic orbit of the interval map f A(I). Then, the set P = [x
1
, x
2
], , [x
n1
, x
n
]
is a partition of the interval [x
1
, x
n
]. By Lemma 1.2.3, closed paths in the Markov graph
of this partition correspond to xed points of some iterate of f in [x
1
, x
n
]. A very natural
question then arises: does the existence of a periodic orbit of period n guarantee the
existence of an orbit of some other period? This question was independently answered
by the Ukrainian mathematician Alexander Sharkovsky in two papers between 1964 and
1965 [Sha64, Sha65] and by Tien-Yien Li and James A. Yorke in 1975 [LY75].
To address this question, rst consider the following ordering
s
on the natural num-
bers dened by:
1
s
2
s
2
2

s
2
3

s

s
2
2
5
s
2
2
3
s

s
2
k
5
s
2
k
3
s

s
2 5
s
2 3
s

s
7
s
5
s
3.
This ordering is eponymously named the Sharkovsky ordering and encodes information
about which periodic orbits force other periodic orbits to exist. This is stated explicitly
in the following theorem:
8
Theorem 1.3.1 (Sharkovsky). Let m, n N. Then m
s
n if and only if every continuous
f : I I having a periodic point of period n has a periodic point of period m.
The proof of this theorem will be given at the end of the paper once the notions
of forcing have been built up suciently. Nevertheless, there are several nice proofs of
Theorem 1.3.1 that do not use forcing. The standard proof can be found in [BGMY80],
whereas slightly dierent and newer proofs can be found in [ALM93] and [BH]. For the
earlier versions, the reader is directed to [

Ste77] and [Sha64]. Moreover, both [Mis97],


[ALM93] and [BH] describe the rich history of Sharkovskys theorem and are interesting
reads.
Now, as was mentioned earlier, Li and Yorke came upon a version of Sharkovskys
result independently. In their paper, Period Three Implies Chaos [LY75], they proved
the following:
Theorem 1.3.2 (Period Three Implies Chaos). If an interval map has a periodic point
of period three, then it has a periodic point of period n for each n N.
Proof. Let f be an interval map and a, b, c be a period three orbit. Without loss of
generality, assume a < b < c. There are two possible cases to be considered: f(a) = b or
f(a) = c. First, let us suppose that f(a) = b. This implies that f(b) = c and f(c) = a.
Consider the partition J
1
= [a, b] and J
2
= [b, c] Since a < b and f(a) = b < c = f(b),
f(J
1
) J
2
and similarly f(J
2
) J
1
J
2
. The Markov graph of f associated to this
partition therefore has two vertices and directed edges from J
1
to J
2
and from J
2
to both
itself and J
1
, as in Figure 1.5. Since, f(J
2
) J
2
, there is a xed point of f by Lemma
1.2.4.
J
1
J
2
Figure 1.5: The Markov graph associated to the partition J
1
, J
2
.
Now, let us show that f has a point of period n for any n > 3. Consider the closed
path
J
2
J
2
J
2
J
2
J
1
J
2
with m > 3 edges. By Lemma 1.2.3, there exists x J
1
such that f
n
(x) = x and
f
k
(x) J
2
for 1 k n 1. If there were an m with 1 m < n such that f
m
(x) = x,
then we would have x = f
m
(x) J
2
. Thus, we would have x J
1
J
2
= b. We
need to show that x = b is impossible. If x = b, then f
2
(b) = f
2
(x) J
2
, contradicting
f
2
(b) = f
3
(a) = a. This contradiction shows that f
k
(x) ,= x for 1 k < n, and so x is a
periodic point of period n. Lastly, by considering the path J
1
J
2
J
1
, we also get a
periodic point of period two. Thus, f has periodic orbits for every period n N.
Observe that for the second case, f(a) = c, we need only relabel the graph, switching
J
1
and J
2
and the exact same argument works.
In the proof of Theorem 1.3.2, the way that the points of the period three orbit were
permuted under f was taken into consideration. In particular, for a < b < c, we had the
possibilities that a b c or a c b. In the rst case, we see that the periodic orbit
is permuted according to the 3-cycle (1 2 3). In the latter case, the orbit moves according
to (1 3 2). What additional information is gained by considering the cycle that an orbit
9
follows under an interval map? The aim of the next chapter is to make precise this relation
between periodic orbits and cycles.
10
Chapter 2
The Forcing Relation on Cycles
As was noted in Section 1.3, for interval maps, the existence of a point of period n implies
the existence of periodic points of dierent periods depending on where n is situated in
the Sharkovsky ordering. If we consider certain combinatorial data found in a periodic
orbit, what then, if anything, does this data force an interval map to exhibit? We begin
to address this question by rst precisely dening what this combinatorial data is and
then dening a forcing relation on based on this data.
2.1 The Combinatorial Data in a Periodic Orbit
Suppose f : I I is a continuous map of a closed interval I. The theorem of Sharkovsky
(Theorem 1.3.1) tells us that the existence of a point of period n under f implies the
existence of a point of period m for every m
s
n. The only information that this theorem
uses is the existence of a point of period n. However, we often know more. Suppose x is
a periodic point of period n. Label the points in the orbit of x from left to right, so that
O(x) = x
1
, . . . , x
n
with x
1
< < x
n
. Then, in addition to the fact that x is a point
of period n under f, we know how f permutes the set x
1
, . . . , x
n
; namely, there exists
an n-cycle such that f(x
i
) = x
(i)
for i = 1, . . . , n.
Denition 2.1.1. If is an n-cycle, f : I I a continuous map of a closed interval, and
X = x
1
, x
2
, . . . , x
n
I with x
1
< < x
n
, then we say X is an orbit of type (or X
exhibits ) if f(x
i
) = x
(i)
, 1 i n. In addition, f is said to exhibit the n-cycle if
there exist x
1
< < x
n
such that x
1
, . . . , x
n
is an orbit of type . We will denote the
set of all permutations exhibited by f by Perm(f).
Example 2.1.2. The symmetric family of tent maps T
a

a[0,2]
is given by T
a
(x) = ax for
x [0,
1
2
] and T
a
(x) = a(1x) for x [
1
2
, 1] [BB04]. For two values of a, a = a
1
1.72208
and a = a
2
1.51288, x
0
=
1
2
is a periodic point of period 5 for the tent maps T
a1
and T
a2
.
Denote the period 5 orbit for each map by x
0
, x
1
, x
2
, x
3
, x
4
with x
i
x
i+1
. Relabeling the
points in a linearly increasing manner, the permutations exhibited are obtained. As can
be seen in Figure 2.1 [BB04, Figure 3.3], T
a1
exhibits the permutation (1 2 4 3 5) whereas
T
a2
exhibits the permutation (1 3 4 2 5).
Denition 2.1.1 points towards a generalization of the Sharkovsky ordering, one that
considers orbit types. Example 2.1.2 shows that we may certainly have two distinct pe-
riodic orbits of period 5. In fact, a simple counting argument shows us that there are
11
x2 x3 x0 x4 x1
1
2
1 2 3 4 5
(1 2 4 3 5)
Ta
1
x2 x0 x3 x4 x1
1
2
1 2 3 4 5
(1 3 4 2 5)
Ta
2
Figure 2.1: Graphs of symmetric tent maps exhibiting dierent 5-cycles.
(n 1)! possible distinct orbit types for a periodic point of period n. In our more general
setting, what does the existence of a periodic orbit of type imply? This motivates the
following denition.
Denition 2.1.3. Let and be cycles. We say forces , denoted , if every
continuous map f : I I that has an orbit of type also has an orbit of type .
Equivalently, if f : Perm(f) f : Perm(f).
A priori, Denition 2.1.3 merely denes a relation, that is, a subset of C C. The
usefulness of this denition thus depends on how we can characterize this relation. First,
recall the denition of a partial order:
Denition 2.1.4. A partial order is a relation on a set X that is reexive, transitive and
antisymmetric. A set equipped with a partial order is called a partially ordered set.
As it turns out, forcing is a partial order on the set of cycles C but not a linear ordering.
Theorem 2.1.5 ([Bal87, Theorem 1.4]). Forcing is a partial order.
Proof. We need to show that is a reexive, transitive and antisymmetric relation. Let
, and be cycles. Reexivity is trivial: if a map exhibits , then it exhibits . Hence,
. For transitivity, suppose and . If a map exhibits , then it exhibits , and
since it exhibits , it exhibits . Thus, we have and so is a transitive relation.
Antisymmetry, on the other hand, requires a bit more work. Let n, m N and
C
n
, C
m
. Suppose that and . We aim to show that this implies = .
Following [Bal87] and [ALM93, p.53], let f be a polynomial of degree at least 2 which
passes through all of the points (i, (i)) for 1 i n, so that f has an orbit of type
(and therefore, by hypothesis, an orbit of type ). Now, as f is a polynomial, the equation
f
n
(x) = x has only nitely many solutions, so f can have only nitely many orbits of type
. Let P = x
1
, . . . , x
n
, x
1
< < x
n
, be one of these orbits having the property that
12
max P min P = x
n
x
1
is the least possible. Dene a continuous function g by
g(x) =

f(x
1
) : x x
1
f(x) : x [x
1
, x
n
]
f(x
n
) : x x
n
By construction, g has an orbit of type , namely P, and therefore, by hypothesis, has an
orbit of type , say Q = y
1
, . . . , y
m
, y
1
< < y
m
. Note that since g(P) = P and for
all x / [x
1
, x
n
], g(x) P, we have Q [x
1
, x
n
]. By the same construction as above, there
is a continuous map h given by
h(x) =

f(y
1
) : x y
1
f(x) : x [y
1
, y
m
]
f(y
m
) : x y
m
Since Q is an orbit of type for h, h must also have an orbit of type , say R = x

1
, . . . , x

n
.
But, R [y
1
, y
m
], so R is also an orbit of f of type . By the minimality of P, we have
R = P, so P = Q and thus = .
Theorem 2.1.6 ([Bal87, Theorem 1.5]). Forcing is not a linear ordering.
Proof. To show that forcing is not a linear ordering it suces to show that there exist
two cycles and such that , and ,. By denition of the forcing relation, it is
sucient to show that there exist two continuous maps, one exhibiting but not , and
the other exhibiting but not . To that end, let = (1 2 3) and = (1 3 2) and recall the
denition of the connect-the-dot maps f

and f

(Denition 1.1.3, see Figure 2.2).


First, suppose that f

exhibits . Consider the partition J


1
, J
2
of [1, 3] with J
i
=
[i, i + 1] and let P = x
1
< x
2
< x
3
be the periodic orbit of period three that exhibits
, i.e., x
1
x
3
x
2
x
1
. Clearly P ,= 1, 2, 3, otherwise would equal . Since
x
1
< x
2
and f

(x
1
) = x
3
> x
1
= f

(x
2
), we must have that x
1
J
2
. Similarly, as x
2
< x
3
and f

(x
2
) = x
1
< x
2
= f

(x
3
), x
2
must be in J
1
. But, as x
2
,= 2, this contradicts our
assumption that x
1
< x
2
. Thus, f

is a continuous map that exhibits but not , so


,. This amounts to the fact that there are no three points x
1
< x
2
< x
3
such that
f

(x
2
) < f

(x
3
) < f

(x
1
).
On the other hand, suppose that f

exhibits . The reasoning is almost identical


to the rst case. Using the same partition of [1, 3], this time let Q = y
1
< y
2
< y
3

denote the period three orbit that exhibits , so that y


1
y
2
y
3
y
1
. Again, note
that Q ,= 1, 2, 3 otherwise would be the same permutation as . Now, y
1
< y
2
and
f

(y
1
) = y
2
< y
3
= f

(y
2
) so y
1
J
2
. However, y
2
< y
3
and f(y
2
) = y
3
> y
1
= f(y
3
), so
y
2
J
1
. This again contradicts the assumption that y
1
< y
2
. Therefore, f

is a continuous
map that exhibits but not , so ,. This concludes the proof of showing that forcing
is not a linear ordering.
2.2 The Forcing Theorem
Thus far, we have shown that the set C of cycles is partially ordered by the forcing
relation , and moreover, this relation is not a linear ordering. Nevertheless, a criterion
for deciding when a cycle forces another cycle has yet to be given. The main goal
of this section, therefore, is to state and prove the Forcing Theorem (Theorem 2.2.1)
which gives a necessary and sucient condition for a cycle to force another. First proven
13
1 2 3
1
2
3
1 2 3
1
2
3
f

Figure 2.2: Graphs of f

and f

.
in [Bal87, Main Theorem 2.4], this theorem and its proof plant the seeds for Chapter 3
where an explicit algorithm is presented that decides the forcing relation.
Theorem 2.2.1 (Forcing Theorem). If C
n
and C
m
, then if and only if
either = or there is a loop of length m in the Markov graph of with orbit type .
To begin with, it is clear that given some cyclic permutation , the family of interval
maps that exhibits can be extremely large. An interesting question is then whether or
not there exists some canonical interval map that exhibits only the cyclic permutations
forced by . If this were to be the case, then we could reduce forcing to the study of
properties of these maps alone. In particular, our approach will be to show that such
canonical maps exist indeed, they are the connect-the-dots maps from Chapter 1 and
that by studying the combinatorial properties of the Markov graphs of these maps, the
desired criterion for forcing can be derived.
Let C
n
and consider the interval map f

. Observe that to every vertex I in


the Markov graph of (equivalently, the Markov graph of f

associated to the partition


[1, 2], . . . , [n 1, n], see Example 1.2.7), we can assign a sign that is either +1 or 1
depending on whether f

is increasing or decreasing, respectively, on I. The function that


assigns to a vertex its sign will be denoted by z. Therefore,
z(I) =

+1 if f

increasing on I
1 if f

decreasing on I
Example 2.2.2. The graph of f

and the signed Markov graph of are shown in Figures


2.3 and 2.4, respectively, for = (1 4 3 2 5).
Now, in Chapter 1 it was shown that a simple loop of length n in the Markov graph
of an interval map corresponds to an n-periodic point of the map. This periodic point
necessarily has an orbit type of for some n-cycle , so the question arises how do we
recover the orbit type from the simple loop? The key lies in studying the signs of the
vertices of the loop. We aim to introduce an ordering <

on loops of the same length in a


given Markov graph in such a way that the order of the real line is reected in the order
of the loops.
14
1 2 3 4 5
1
2
3
4
5
Figure 2.3: f

for = (1 4 3 2 5).
J
1
,+ J
2
,
J
3
,+ J
4
,
Figure 2.4: The signed Markov graph of =
(1 4 3 2 5).
Denition 2.2.3. For a loop (J
i1
, . . . , J
in
), we dene the shift operation by (J
i1
, . . . , J
in
) =
(J
i2
, J
i3
, . . . , J
in
, J
i1
). Given two loops = (J
i1
, . . . , J
in
) and = (J
k1
, . . . , J
kn
) of the
same length, we dene <

if and only if

M1

j=1
z(J
ij
)

i
M
<

M1

j=1
z(J
ij
)

k
M
,
where M n is the least value of j for which i
j
,= k
j
. This is a linear ordering.
When we are dealing with the Markov graphs of cycles (resp. connect-the-dots maps)
it will be notationally easier sometimes to label the vertices as integers. Thus, instead
of labeling the n 1 vertices of the Markov graph of an n-cycle by J
1
, J
2
, . . . , J
n1
, we
will simply label them 1, 2, . . . , n 1, where the integer i corresponds to the subinterval
[i, i +1]. Then, if = (a
1
, a
2
, . . . , a
n
) and = (b
1
, b
2
, . . . , b
n
) are loops of the same length,
we have <

if and only if

M1

j=1
z(a
j
)

a
M
<

M1

j=1
z(a
j
)

b
M
.
where M n is the least value of j for which a
j
,= b
j
.
The most important aspect of this ordering is the following: suppose f A(I) is an
interval map, J
1
, . . . , J
N
is a partition of I (labeled so that min J
i
< min J
i+1
) and we
have loops = (J
i1
, . . . , J
in
) and = (J
k1
, . . . , J
kn
) in the Markov graph of f associated
to this partition. These loops can be thought of as signed itineraries of periodic points
in I, say x corresponding to and y corresponding to . We will see in the proof of the
forcing theorem that the order of loops tells us the order of the points to which the loops
correspond:
Proposition 2.2.4. <

implies x < y.
15
Example 2.2.5. In the Markov graph of = (1 2 4 3) (Figure 2.5), one can nd the
nonrepetitive loops (1, 2, 2), (2, 2, 1) and (2, 1, 2). Using the order from Denition 2.2.3,
observe that (1, 2, 2) <

(2, 2, 1) and (1, 2, 2) <

(2, 1, 2) because +1 < +2 as integers.


Because the sign of the vertex 2 in the Markov graph of is negative, we have (2, 2, 1) <

(2, 1, 2), since 2 < 1 as integers. Thus, we have (1, 2, 2) <

(2, 2, 1) <

(2, 1, 2).
1, + 2,
3, +
Figure 2.5: Markov graph of = (1 2 4 3)
Denition 2.2.6. Let be a nonrepetitive loop of length mand let W = , (),
2
(), . . . ,
m1
().
Order W by <

and let ord : W 1, . . . , m be the unique one-to-one order preserving


function. We dene the orbit type or loop type of W to be the cycle that is dened by
(i) = ord ord
1
, 1 i m.
If is any loop, then we dene the orbit type of to be the orbit type of W where W
is the smallest such that W and W is closed under . With this denition, the orbit
type of is equal to the orbit type of ().
Example 2.2.7. Let W = (1, 2, 2) <

(2, 2, 1) <

(2, 1, 2) be the ordered collection


of loops of length 3 given in Example 2.2.5. We want to decide the orbit type of
this ordered list of shift equivalent loops. Denition 2.2.6 tells us that for i = 1, 2, 3,
(i) = (ord ord
1
)(i). Thus, for i = 1, we have
(1) = ord((ord
1
(1))) = ord(((1, 2, 2))) = ord((2, 2, 1)) = 2.
Similarly, for i = 2 and i = 3, we get (2) = 3 and (3) = 1. Thus, the orbit type of W is
(1 2 3).
Now we are ready to prove half of the Theorem 2.2.1.
Proposition 2.2.8 ([Bal87, Lemma 3.1]). If C
n
and f

has an orbit of type , then


either = or the there is a loop in the Markov graph of that has orbit type .
Proof. Following [Bal87, Lemma 3.1]: Let X = x
1
, x
2
, . . . , x
m
be an orbit of type with
x
1
< x
2
< < x
m
and note that by denition of f

we must have 1 x
1
and x
m
n.
Assume ,= , so that none of x
1
, . . . , x
m
are integers. Dene the path = (a
1
, . . . , a
m
)
in the Markov graph of by
a
j
= i f
j

(x
1
) [i, i + 1],
16
and observe that is actually a loop in the Markov graph of . We rst want to show
that is nonrepetitive, so suppose that is a loop in the Markov graph of of length
k such that k divides m and a
j1
= b
j2
if and only if j
1
j
2
(mod k). Dene intervals
I
j
, 0 j m by backwards induction from m. Let I
m
= [a
m
, a
m
+ 1], and if I
j
has
been dened for j 1, let I
j1
= f
1

(I
j
) [a
j1
, a
j1
+ 1] (where we let a
0
= a
m
for convenience). Now a
jk
= a
j
, so we can show by induction that I
jk
I
j
whenever
k j m and that f
k

is a one-to-one linear function from I


j
onto I
j+k
(1 j, j +k m).
Since I
0
I
k
, f
k

: I
0
I
k
has a xed point y, and clearly f
m
(y) = y also, since k divides
m. However, x
1
I
0
(since by denition f
j

(x
1
) [a
j
, a
j+1
] for all j) and f
m

(x
1
) = x
1
,
but x
1
,= y (since x
1
has least period m and y has least period k < m). Thus f
m

: I
0
I
m
is one-to-one, onto, linear, and has two xed points, so I
0
= I
m
and f
m

is the identity
function. Furthermore, since [f

(x)[ 1 whenever f

(x) is dened and 1 < x < n, we


have I
j
= [a
j
, a
j
+ 1] for all j and f
k

: I
j
I
j+k
is a linear map of slope either 1 or 1.
Thus if x
1
= a
m
+
1
2
, then f
k

(x) = a
m
+
1
2
contradicting x
1
has least period m > k, and
if x
1
is some element of [a
m
, a
m
+ 1] other than a
m
+
1
2
, then it is easy to see that X has
type , contradicting ,= . Thus must be nonrepetitive.
Next, we must show that the orbit type of equals . Let (k) =
k
() with (k) =
(a
1
(k), a
2
(k), . . . , a
m
(k)) and note that a
j
(k) = i if and only if f
j+k

(x
1
) [i, i + 1].
Suppose (k
1
) <

(k
2
). Let j be least such that a
j
(k
1
) ,= a
j
(k
2
). Then if we look
at f
k1

(x
1
) and f
k2

(x
1
) and follow their progress as f

is repeatedly applied j times, we


see that they switch order if and only if z(a
j
(k
1
)) is 1 an odd number of times. But
since f
j

(f
k1

(x
1
)) [a
j
(k
1
), a
j
(k
1
) + 1] and f
j

(f
k2

(x
1
)) [a
j
(k
2
), a
j
(k
2
) + 1] we see that
f
k1

(x
1
) < f
k2

(x
1
).
Thus we have shown that
f
k1

(x
1
) < f
k2

(x
1
) (k
1
) <

(k
2
),
from which it follows that the orbit type of is .
Now we need to show the other direction: if there is a loop in the Markov graph of
of loop type , then any interval map with an orbit of type has an orbit of type . First,
we need to prove the following lemma:
Lemma 2.2.9 ([Bal87, Lemma 3.2]). Suppose that f : R R is continuous, a < b,
f(a) < f(b), and for some N N, (a
i
, b
i
) : 1 i N is a collection of intervals having
the properties that
(i) f(a) f(a
i
) < f(b
i
) f(b);
(ii) a a
i
< b
i
b;
(iii) (a
i
, b
i
) minimally f-covers itself;
(iv) If i < j, then either (a
j
, b
j
) (a
i
, b
i
) or (a
i
, b
i
) (a
j
, b
j
) = ;
(v) f is strictly increasing on the set a
i
: i < N b
i
: i < N.
If c < d, c, d [f(a), f(b)] are such that for all i either (c, d) (f(a
i
), f(b
i
)) or
(c, d) (f(a
i
), f(b
i
)) = , then there exist points a
N+1
and b
N+1
such that f(a
N+1
) =
c, f(b
N+1
) = d and (a
i
, b
i
) : 1 i N + 1 satises (i)-(v) with N replaced by N + 1.
17
Proof.
Case 1. (c, d) (f(a
i
), f(b
i
)) for some i.
Assume that i is the largest possible, so that (f(a
i
), f(b
i
)) (f(a
j
), f(b
j
)) or
(f(a
i
), f(b
i
)) (f(a
j
), f(b
j
)) = for all i ,= j (by condition (iv) and (v)). Let a
N+1
be the greatest element of [a
i
, b
i
] such that f(a
N+a
) = c and then let b
N+1
be the least
element of [a
N+1
, b
i
] such that f(b
N+1
) = d.
Case 2. (c, d) (f(a
i
), f(b
i
)) = for all i.
Now let a

be the greatest element of a b


i
: f(b
i
) c and let b

be the least
element of b a
i
: f(a
i
) d. Let a
N+1
be the greatest element of [a

, b

] such that
f(a
N+1
) = c and let b
N+1
be the least element of [a
N+1
, b

] such that f(b


N+1
) = d.
In either case the appropriate choice of a
N+1
and b
N+1
will work.
Proposition 2.2.10 ([Bal87, Theorem 3.3]). If f is an interval map having an orbit of
type C
n
and is a loop in the Markov graph of of loop type C
m
, then f has an
orbit of type .
Proof. Let P = p
1
< < p
n
be an orbit of type for f, and let = (a
1
, a
2
, . . . , a
m
)
be a nonrepetitive loop of length m in the Markov graph of . Let us dene integers
a
j
for j 0, m + 1, m + 2, . . . , 2m 1, 2m by a
0
= a
m
and a
j
= a
jm
; that is,
a
m+1
= a
1
, a
m+2
= a
2
, etc. For i 1, . . . , n 1, dene I
i
to be the interval [x
i
, x
i+1
].
We want to dene a collection of 2m + 1 intervals J
i
, 0 i 2m which satisfy the
hypotheses of Lemma 2.2.9. To do so, rst let J
2m
= I
am
. Suppose by induction that
J
j+1
, J
j+2
, . . . , J
2m
have been dened with J
i
I
ai
and that if j + 1 i < k 2m then
either J
i
J
k
or J
i
J
k
has at most one point. Assume also as part of the induction
hypothesis that J
i
minimally f-covers J
i+1
for j + 1 i 2m 1. If we consider the
intervals J
i
, j + 1 i 2m 1 such that J
i
I
ai
and J
i+1
I
aj+1
, then we see that
Lemma 2.2.9 applies (using either f or f, depending on whether f(x
aj
) < f(x
aj+1
) or
vice versa) in order to get J
i
I
aj
with f taking J
j
onto J
j+1
with the other claims of
the induction hypothesis holding. Therefore, we get that J
0
J
m
and f
m
takes J
0
onto
J
m
, so by Lemma 1.2.4, f
m
has a xed point y J
0
.
Case 1. For some 0 < i < j m, J
i
J
j
,=
Now, J
i
J
j
is impossible, since then we would have J
i+k
J
j+k
for all 0 k m
which would imply that a
i+k
= a
j+k
(since J
k
I
a
k
), contradicting that is nonrepeti-
tive. Thus J
i
and J
j
intersect at a single point, say J
i
J
j
= x, so by construction of
the J
k
s, J
i+k
J
j+k
is a single point for all 0 k 2mj, in particular J
i+2mj
J
2m
is
a single point, namely either a
a2m
or p
a2m+1
, so J
i+k
J
j+k
is a point of P for all such k.
Thus, since the induction used 2m steps and 2mj m, f
k
(p
1
) = x for some k m, so
there is a k such that J
i+k
J
j+k
= p
1
, which is impossible since J
i+k
J
j+k
[p
1
, p
n
].
Thus Case 1 leads to a contradiction.
Case 2. If 0 < i < j m, then J
i
J
j
= .
Then if J
i
J
j
I
k
for some xed k, then by the construction of J
i
, J
i+1
and J
j+1
stay in the same order as J i and J j if f(p
k
) < f(p
k+1
) and J
i+1
and J
j+1
switch
order compared to J
i
and J
j
if f(p
k
) > f(p
k+1
). The same argument used at the end of
Proposition 2.2.8 now gives that f
k
(y) : 1 k m is an orbit of f of type .
Thus, Propositions 2.2.8 and 2.2.10 prove both directions of the Forcing Theorem.
Before moving onto the next section, let us briey discuss the implications of Theorem
2.2.1. In Section 2.1, forcing was dened as a property of all interval maps exhibiting
a certain cycle: if and only if every continuous map that exhibits exhibits . In
18
this respect, Theorem 2.2.1 is remarkable: forcing can be reduced to simply studying
combinatorial properties of cycles and their Markov graphs. To decide whether or not a
given cycle forces another cycle , the only continuous function that we need to consider
is f

, and we need only consider this function so that we can construct the Markov graph
of . However, the function f

is merely a convenience, since we could easily construct


the Markov graph of directly from properties of . Therefore, in a sense, one could pose
forcing as an intrinsic order on the set of cyclic permutations, an order independent of any
relation to continuous maps and dynamical systems all together.
19
Chapter 3
A Computational Digression
Having dened the partial order on the set of all cyclic permutations and stated and
proven the forcing theorem (Theorem 2.2.1), we will now give an algorithm which describes
forcing. In [Bal87], Baldwin derives an explicit combinatorial algorithm which decides all
m-cycles that a given n-cycle forces. From this, one can clearly extend the algorithm
as follows: input an integer m and an n-cycle and return all k-cycles forced by for
1 k m. Additionally, one can make slight modications so that a user may input
cycles and and ask if . We give a thorough description of Baldwins algorithm
as well as our own implementation of the algorithm in the algebra system GAP and the
results obtained.
3.1 Description of the Algorithm of Baldwin
The proof of Theorem 2.2.1 essentially gives us the desired algorithm: a necessary and
sucient condition for to force is that there is a closed loop of length [[ in the Markov
graph of that has loop type . Therefore, we can describe the algorithm as follows:
1. Input a cyclic permutation C
n
and an integer m.
2. Construct the Markov graph of and nd all closed paths (loops) of length m in
this graph.
3. Partition the family of closed paths of length m into sets equivalent under the shift
operation.
4. Compute the orbit type of each shift equivalent set. This will give all C
m
forced
by .
Example 3.1.1. Let us nd all of the 3-cycles forced by = (1 2 4 3). To that end, we
need to rst construct the Markov graph of (Figure 2.5) and then nd all (non-trivial)
loops of length 3. Up to equivalence modulo the shift operation, there are four relevant
loops, namely L
1
= (1, 2, 2), L
2
= (3, 2, 2), L
3
= (1, 2, 3) and L
4
= (1, 3, 2). Let us analyze
these loops one at a time.
L
1
: From Example 2.2.5, the ordered list of loops that are shift equivalent to L
1
is
(1, 2, 2) <

(2, 2, 1) <

(2, 1, 2), and from Example 2.2.7, the orbit type of this
ordered list is (1 2 3). Therefore, we conclude that (1 2 3).
20
L
2
: Next, using the same strategy as above, we can order the list of loops that are shift
equivalent to L
2
by (2, 3, 2) <

(2, 2, 3) <

(3, 2, 2). Letting denote the orbit type


of this list, we have
(1) = ord((ord
1
(1))) = ord(((2, 3, 2))) = ord((3, 2, 2)) = 3,
and, by the same method, (2) = 1 and (3) = 2. Thus, = (1 3 2) and so (1 3 2).
Even though L
3
and L
4
are remaining, we note that since there are only two cyclic per-
mutations of length 3, and we have shown that (1 2 4 3) forces both of them, we are done.
Now, suppose that f A(I) has a periodic point of period 4. Sharkovskys theorem
tells us that the existence of this point only implies the existence of a point of period 2
and a xed point. An interesting consequence of Example 3.1.1, is that if this period 4
orbit exhibits the 4-cycle (1 2 4 3), then f has a periodic orbit of period 3 and so by Period
Three Implies Chaos, f also has a periodic orbit of period n for every n N. Thus, the
forcing relation reveals a wealth of information that Sharkovskys initially did not.
3.2 Implementation of the Algorithm in GAP
GAP is a system for computational discrete algebra that provides a programming language,
tools for computational group theory and large data libraries of algebraic objects, including
the classication of all small groups of order less than 2000 up to isomorphism class. With
regards to our analysis of the partial order on the set of of cycles, the most useful
aspect of GAP is its ease in using permutations. Here we give a brief description of our
implementation of Baldwins algorithm in GAP. The complete code for this can be found
at: http://tlondonforcing.webs.com/
Our implementation of the algorithm consists of four main functions which are called
from the main function forcedPermutations. To begin with, the user inputs a cycle
in cycle notation, its length n and the integer m for which one wants to nd all m-cycles
that forces. Next, the function MarkovGraph is called. This function takes two inputs:
and n and constructs the Markov graph of as a as follows: the graph is stored as
a list that has n 1 entries and each entry i of this list is itself a list containing all of
the vertices that i has a directed edge to. For example, the Markov graph of (1 4 3 2 5)
(Figure 1.3) would be stored as [[4], [2, 3, 4], [2], [1, 2]]. Now, with the Markov graph of
constructed, the function closedWalksOfGivenLength is called. This function, which
takes as arguments the Markov graph of and the integer m, nds all closed walks of
length m in the Markov graph. This operation is extremely time consuming as will be
described in Section 3.3. Given the list of all closed walks of length m, the function
shiftEquivWalks is called. This function partitions the set of all closed walks of length
m into lists of walks which are closed under the shift operation (Denition 2.2.3). Next,
the function baldwinOrderOnWalks orders each of these families according to the order
<

. Finally, for each of these ordered lists of walks closed under the shift operation, the
function baldwinForcedPermutation computes its loop type. The list of all loop types is
returned; that is, the list of all m-cycles forced by .
3.3 Eciency of the Algorithm
Theorem 2.2.1 states that a necessary and sucient condition for a cycle to force a
cycle is that there is a loop of length [[ in the Markov graph of that corresponds to
21
an orbit of type . Recall from Denition 2.1.3 that forcing is dened as a property of
all continuous maps exhibiting a certain cyclic permutation. In fact, there is an inherent
problem with using the connect-the-dots maps to determine forcing. The problem with
using f

as the model map is that, although the set of representatives of in f

is in
one-to-one correspondence with a subset of the closed paths of length [[ in the Markov
graph of , there is no ecient way known of determining whether has a representative
in f

. Baldwin solved this problem by ignoring it. In his algorithm, one examines all
closed paths of length [[ in the graph to see if any of them correspond to a representative
of . The number of such paths is exponential in [[ [BCMM92].
However, it is possible to speed up the algorithm considerably. In [BC92a], Bernhardt
and Coven present an eective polynomial-time algorithm for deciding forcing. This paper,
which can be seen as an extension of [BCMM92], considers horseshoe maps instead of
connect-the-dots maps.
Denition 3.3.1. Let be an n-cycle and let M be the number of nondegenerate intervals
I maximal with respect to f

is strictly monotone on I (such intervals are called laps).


The horseshoe map of is the map H

: [0, 1] [0, 1] dened as follows: H

(0) = 0
or 1 according to whether (1) < (2) or (1) > (2), and H maps each subinterval
[(i 1)/M, i/M] linearly onto [0, 1] for i = 1, . . . , M 1.
The essential idea to check whether forces is to rst nd a certain canonical repre-
sentative, say Q, of in H

. Then, forces if and only if H

has a representative P of
such that
min
qQ

q
i
M

min
pP

p
i
M

, i = 1, . . . , M 1.
This algorithm is similar to an earlier one given by Jungreis [Jun91] which used
itineraries of points.
3.4 Properties of Forcing in the diagrammed data
To begin with, Baldwins original diagram [Bal87, Figure 4], showing the forcing relation
on all 36 cycle of length up to 5, is shown in Figure 3.1. In Figure 3.2, the forcing relation
on almost 200 cycles of order up to 8 has been diagrammed. This diagram contains the 36
cycles from Baldwins original diagram of forcing on all cycles of length up to 5. We adopt
the convention that if and only if is above and there is a line segment between
and . There are several aspects of this diagram which warrant being pointed out, as
they will be described at length in later chapters. The rst clear property of Figure 3.2
is that it is symmetric about a vertical axis extending from (1 2). This symmetry could
simply be an artifact of the way in which the forcing relations were diagrammed; however,
we will show in Chapter 4.1 that this is not the case. In fact, in Proposition 4.1.4, we will
prove that for every cycle there exists a cycle

with the property that forces if and
only if

forces .
Though more subtle than the symmetry in Figure 3.2, one characteristic is that there
is a long vein of cycles on which the forcing relation is a linear ordering; that is, for any
two cycles in this vein, forces or forces . We will outline in Chapter 4.2 that the
forcing relation restricted to unimodal cycles is a linear ordering.
Lastly, of the cycles that have been diagrammed, if one looks at the lengths of the
cycles immediately above a given cycle (Denition 5.1.5), then the length turns out
to be 2[[. In Chapter 5.1, we will prove that immediately above any cycle lies all of
22
Figure 3.1: Baldwins original diagram.
its doubles. Consequently, we note that the forcing order on cycles is not locally nite:
between any two odd length cycles lie innitely many even length cycles.
23
1324
1426375
14253
16374285
1627435
1635274
1634275
17435286
17436285
153264
132
18527436
1742635
15324
1743526
18453627
163425
18452736
1423
16284537
15384627
142635
15384726
16273845
15273846
15483726
143625 152634
15482637 17362845
1437526
17284536
1527436
13254
15274386
16274385
1526374
1362745
15463827
15372846
15462837
12
123
14725638
1462537
13425
1453627
15463728
15472638
136245
14825637
13825647
1372546
1374625
1473625
14837256
13524
14836257
1452736
12534
13847256
1453726
1472536
indicates new data
indicates data from Baldwin's digram
indicates unimodalpermutation
143526
1243
135264 136425
134265 126435
1274536
1374526
1435276
1436275
13856247
13854627
12854637
14537286
14536287
14375286 13728546
14628537
14538627
1375246 1364275
13862547
13852746
13745286
13746285 14862537
14625387
14725386
14853726
14738526
14527386
14627385
12853746
13854726
15342 13245
14523 14352
1342
13452
15432 12345
14532
12543
12453 13542
12354
1432 1234
12435
15243
15423
164253
1752634
18635274
18627435
13647258
1362547
13725648
135246
163254
1734265
1735264
1726435
18435276
18364275
18436275
18362745
18274536
162534
14235
16357248
1536247
1546237
15647238
15637248
145236
1624537
17254638
152436
16372548
14325
153426
1543627
1635427
1643527
1726354
16453728
17436528
17364528
17453628
18274365
18273645
15234
18256347
1724536
18254637
162435
1734526
1725346
1625347
18263547
17256348
17263548
18354627
124536
12573648
1254637
12564738
1356247
13657248 18634275
1763425
18743526
165324
18742635
1753264
12467358
1235647
123546
12473658
1246357
13572468
12365748
18642753
1764253
18752634
165423
18763425
1765324
18753264
Figure 3.2: Computation of for n 8.
24
Chapter 4
Global Structure of Forcing
In application, one major dierence that the ordering has when compared to the
Sharkovsky ordering, is that forcing is only a partial ordering (Theorem 2.1.6). Whereas
any pair of elements in N are mutually comparable under the Sharkovsky ordering (that
is,
s
is a linear ordering), this need not hold for any pair of elements in C under . For
example, consider the 4-cycles (1 2 3 4) and (1 4 3 2). It can be shown (see Chapter 3) that
the only 4-cycle that (1 2 3 4) forces is (1 3 2 4), and the only 4-cycle that (1 4 3 2) forces
is (1 4 2 3). Thus, (1 2 3 4),(1 4 3 2) and (1 4 3 2),(1 2 3 4). However, the 4-cycle (1 2 4 3)
forces both (1 4 2 3) and (1 3 2 4), yet is incomparable with both (1 2 3 4) and (1 4 3 2)!
These relationships are illustrated in Figure 4.1.
(1 2 3 4)
(1 2 4 3)
(1 4 3 2)
(1 3 2 4) (1 4 2 3)
Figure 4.1: Illustration of the complexity of on some 4-cycles.
This one example begins to illustrate the complex global structure of (C, ). However,
there is some symmetry and nice structure to (C, ) when we either consider certain
pairs of cycles or restrict ourselves to certain types of cycles.
4.1 The Flip Symmetry
One distinctive characteristic of Figure 3.2 is that it is symmetric about a vertical axis
extending from (1 2). The aim of this section is to show that to every cycle we can asso-
25
ciate a cycle

with the property that if and only if

. Therefore, by diagramming
and

across from each other, we obtain a sort of global symmetry in (C, ).
Denition 4.1.1. Given C
n
, the ip of , denoted

, is dened to be the n-cycle such
that

(k) = n + 1 (n + 1 k).
Example 4.1.2. Consider the 4-cycle = (1 3 2 4). The ip of is dened by

(1) = 4 + 1 (4 + 1 1) = 4

(2) = 4 + 1 (4 + 1 2) = 3

(3) = 4 + 1 (4 + 1 3) = 1

(4) = 4 + 1 (4 + 1 4) = 2
Thus,

= (1 4 2 3).
Example 4.1.3. As Figure 3.2 would suggest, the ip of a cycle may equal . Indeed
for the cycles (1 2), (1 2 4 3) and (1 4 7 3 8 5 2 6) this is precisely the case.
We have chosen to call

the ip for the following reason: the graph of the connect-
the-dots map f

is a 180

rotation of the graph of the map f

. A consequence of this
observation is that the Markov graph of

is obtained from the Markov graph of by
relabeling the vertex i by n i.
Proposition 4.1.4 ([Bal87, Proposition 4.2], [ALM93, Proposition 2.4.6]). Let , C.
Then, if and only if

.
Proof. Let C
n
and suppose that for some cycle . Let h : [1, n] [1, n] be the
orientation-reversing homeomorphism h(x) = x +n + 1. Then h f

h
1
= f

. This
can be seen by the fact that for any k 1, . . . , n,
(h f

h
1
)(k) = h(f

(h
1
(k)))
= h(f

(n + 1 k)) def. of h
= h((n + 1 k)) def. of f

= h(n + 1

(k)) def. of

= (n + 1

(k)) +n + 1
= f

(k).
Since f

is linear on each interval of the form (k, k +1) for k 1, . . . , n1, we conclude
that h f

h
1
is also linear on each such interval and connects the dots (k,

(k)).
Thus, this map is indeed f

. Now, if X is the orbit of type for f

, then it follows from


the above construction that (h f

h
1
)(X) is an orbit of type for f

. Therefore,
by Theorem 2.2.1,

. By using the same h, but starting with f

, we get the other


direction. Thus, if and only if

.
Corollary 4.1.5. [ALM93, Corollary 2.5.7] A cycle either is equal to its ip or does not
force its ip.
4.2 The Restriction of Forcing to Unimodal Permuta-
tions
The aim of this section is to prove that forcing is a linear ordering when restricted to a
certain family of cyclic permutations called unimodal cycles. To fully develop the theory
26
necessary to show this would require a lot of new notation that is ultimately not used in
the remaining chapters, so this section is incomplete in that respect. Nevertheless, for any
result that is unproven, citations will be given where the interested reader may nd the
proofs.
By a unimodal map f, we mean an interval map that is continuous, has exactly one
critical point and is strictly increasing (resp. decreasing) left (resp. right) of the critical
point. Examples of unimodal maps include quadratic maps and tent maps of the interval.
The denition of a unimodal cycle follows naturally from the denition of a unimodal
map:
Denition 4.2.1. A unimodal cycle is a cycle such that f

is a unimodal map. Denote


the set of all unimodal n-cycles by U
n
.
Example 4.2.2. For any n N, the cycle (1 2 3 n 1 n) is unimodal. In particular,
the 5-cycle = (1 2 3 4 5) is a unimodal cycle and its connect-the-dots map f

is shown in
Figure 4.2.
1 2 3 4 5
1
2
3
4
5
f

Figure 4.2: The graph of f

for the unimodal cycle = (1 2 3 4 5).


One extremely powerful tool in the study of unimodal maps is itineraries. Just as we
considered signed itineraries in deciding forcing, we can consider itineraries of points to
classify orbits of unimodal maps. All of the results given below can be traced back to
Milnor and Thurstons paper [MT88] as well as [MSS73]. We will follow [CE80]. For now,
let us assume that we are considering a xed unimodal map f of the interval [1, 1] into
itself which has critical point at 0.
Denition 4.2.3 ([CE80, p.64]). Let f be a unimodal map as above. To each point
x [1, 1] we associate a nite or innite sequence of symbols L, C, R (for left, center and
right, respectively), called its itinerary I(x) as follows:
I1. I(x) is either an innite sequence of Ls and Rs, or a nite (or empty) sequence of Ls
and Rs, followed by C. The j
th
element of I(x) will be denoted I
j
(x), j = 0, 1, . . ..
27
I2. If f(x) ,= 0, (0 is the maximum of f), for all j 0, then I
j
(x) = L if f
j
(x) < 0 and
I
j
(x) = R if f
j
(x) > 0.
I3. If f
k
(x) = 0 for some k, then letting j denote the smallest such k, we set I
j
(x) = C
and I
l
(x) = L if 0 l < j and f
l
(x) < 0 and I
l
(x) = R if 0 l < j and f
l
(x) > 0.
Example 4.2.4. Consider the quadratic map f : [1, 1] [1, 1] given by f(x) =
2x
2
+ 1. Note that f(0) = 1 and f is strictly increasing (resp. decreasing) on [1, 0]
(resp. [0, 1]). Thus f is a unimodal map of [1, 1]. The itineraries of the xed points
x = 1 and x = 1/2 are
I(1) = LLL = L

and I(
1
2
) = RRR = R

.
For x .80901542, we have
I(x) = RLRLRL = (RL)

,
and lastly, for x .382683432, we have
I(x) = RRC.
We shall call a sequence I of symbols L, C, R admissible if either I is an innite sequence
of Ls and Rs or if I is a nite (or empty) string of Ls and Rs followed by C. We note
that (trivially) every itinerary is admissible. Just like in Chapter 2, we can use the shift
operation on itineraries of unimodal maps. Next, let us introduce an ordering <
u
on the
set of all admissible sequences.
Denition 4.2.5. We dene L < C < R. Suppose now that A ,= B are two admissible
sequences. Let i be the rst index for which A
i
,= B
i
. We say that A <
u
B if either
O1. There are an even number of Rs in A
0
A
1
A
i1
= B
0
B
1
B
i1
and A
i
< B
i
; or
O2. There are an odd number of Rs in A
0
A
1
A
i1
and A
i
> B
i
.
If none of these hold then we say B <
u
A. This ordering is called the unimodal order or
the parity-lexicographic order.
On the surface, this denition looks entirely new. However, suppose f is a unimodal
map and consider the partition J
1
= [1, 0], J
2
= [0, 1]. Note that the sign of J
1
is +1
and the sign of J
2
is 1. Rewrite A and B in terms of J
1
and J
2
; that is, replace each
L (resp. R) with a J
1
(resp. J
2
). Then, in Denition 4.2.5, condition O1. amounts to
saying that the product z(A
0
)z(A
1
) z(A
i1
) is positive and A
i
< B
i
, and condition
O2. amounts to saying that the product z(A
0
)z(A
1
) z(A
i1
) is negative and A
i
> B
i
.
Thus, looking back at Denition 2.2.3, it becomes clear that <
u
is equivalent to the order
<

for unimodal maps. The following lemma can easily be checked.


Lemma 4.2.6 ([CE80, Lemma II.1.1]). The relation <
u
is a linear ordering.
Like the order <

, we can use the unimodal order to show that the order of itineraries
of two points gives us the order of the two points.
Lemma 4.2.7 ([CE80, Lemma II.1.2]). If f is unimodal and I(x) <
u
I(y) then x < y.
28
Proof. We prove this by induction on the smallest index i for which I
i
(x) ,= I
i
(y). The
statement is true when i = 0. Suppose that i > 0 and that we have already shown the result
for i 1. When I
0
(x) = I
0
(y) = L, then if I(x) <
u
I(y), we nd that I(f(x)) <
u
I(f(y)).
But, I(f(x)) = I
1
(x)I
2
(x) , and hence I
i1
(f(x)) ,= I
i1
(f(y)). By the induction
hypothesis f(x) < f(y). Since I
0
(x) = I
0
(y) = L, we have x, y [1, 0) and since
f
[1, 0]
is strictly increasing, f(x) < f(y) implies x < y.
When I
0
(x) = I
0
(y) = R, then if I(x) < I(y) we nd I(f(x)) > I(f(y)) since I(f(x))
has one R less than I(x) before i. Hence f(x) > f(y). Since I
0
(x) = I
0
(y) = R, we have
x, y (0, 1] and since f
[0, 1]
is strictly decreasing, f(x) > f(y) implies x < y.
Lemma 4.2.8 ([CE80, Lemma II.1.3]). If f is unimodal and x < y then I(x)
u
I(y).
Proof. Since <
u
is a linear ordering, if x < y and I(x) > I(y), then Lemma 4.2.7 implies
a contradiction.
Denition 4.2.9 ([CE80, p.87]). Let f be a unimodal function and let A be an admissible
sequence. Suppose that for all k 0,

k
(A) < I(1) if I(1) is innite,

k
(A) < (DL)

if I(1) = DC and D is even,

k
(A) < (DR)

if I(1) = DC and D is odd.


Then we shall say that A is dominated by I(1) and we shall denote this by A I(1).
Theorem 4.2.10 ([CE80, Theorem II.3.8]). Let f be unimodal and assume A is an ad-
missible sequence satisfying
I(1) A I(1).
Then there is an x [1, 1] such that I(x) = A.
Theorem 4.2.10 allows us to nally show that when we restrict the forcing relation to
unimodal cycles, it is a linear ordering. We sketch the proof of this theorem as shown
in [Bal87].
Theorem 4.2.11 ([Bal87, Theorem 4.4]). Forcing is a linear ordering on the set of all
unimodal cycles.
Proof. Let , be unimodal cycles, say U
n
and U
m
. Then let I(n)C be the itinerary
of the point n in the function f

and dene I(m)C from in the same way, where I(n)


and I(m) are nite sequences of Ls and R

s. By symmetry assume I(m)C <


u
I(n)C,
and then one can show that either A = I(m)L or A = I(m)R satises the hypothesis of
Theorem 4.2.10 with f = f

or f = some conjugate to f

. The point x guaranteed by


this theorem is seen to generate an orbit of type . Thus f

has an orbit of type so by


Theorem 2.2.1, .
With the implementation of Baldwins algorithm in GAP, we were able to write a code
which diagrammed the forcing relation on all 115 unimodal cycles of length between 3 and
10. Table 4.1 should be read from top to bottom, left to right. Thus, the cycle at the
very top left, (1 2 3 4 5 6 7 8 9 10), forces every other cycle listed, and the cycle at the very
bottom right, (1 3 2 4), is forced by every cycle listed. There are some interesting empirical
observations to be noted. To begin with, there is one unimodal 3-cycle, (1 2 3), and there
are two unimodal 4-cycles, (1 2 3 4) and (1 3 2 4). When we diagram the 4-cycles with the
3-cycles, one of them forces (1 2 3) and the other is forced by (1 2 3). Next, there are three
29
unimodal 5-cycles, and when we diagram these with U
3
U
4
, we see that between every
two elements of U
3
U
4
lies a 5-cycle. This, in fact, appears to be the general trend:
suppose that we have diagrammed the forcing order on all unimodal cycles of length up
to n. When we adjoin to this list all unimodal cycles of length n + 1, between almost
every two cycles in U
n
lies a unimodal cycle of length n + 1. Essentially, the unimodal
n + 1-cycles ll in the gaps. For example, in the data below, note that almost every
other cycle is a 10-cycle. The only time this does not occur is when doublings appear.
The rst example of this is in the eighth and ninth rows of Table 4.1 where we have
(1 3 5 7 9 2 4 6 8 10)(1 2 3 4 5)(1 3 5 7 8 2 4 6 9)(1 2 4 7 9 3 5 8 6 10),
but note here that (1 3 5 7 9 2 4 6 8 10) is a doubling of (1 2 3 4 5).
(1 2 3 4 5 6 7 8 9 10) (1 2 3 4 5 6 7 8 9) (1 2 3 4 5 6 7 9 8 10) (1 2 3 4 5 6 7 8)
(1 2 3 4 5 6 8 9 7 10) (1 2 3 4 5 6 8 7 9) (1 2 3 4 5 7 9 6 8 10) (1 2 3 4 5 6 7)
(1 2 3 4 6 8 9 5 7 10) (1 2 3 4 5 7 8 6 9) (1 2 3 4 5 8 7 9 6 10) (1 2 3 4 5 7 6 8)
(1 2 3 4 6 9 5 8 7 10) (1 2 3 4 6 8 5 7 9) (1 2 3 5 7 9 4 6 8 10) (1 2 3 4 5 6)
(1 2 4 6 8 9 3 5 7 10) (1 2 3 5 7 8 4 6 9) (1 2 3 4 7 9 5 8 6 10) (1 2 3 4 6 7 5 8)
(1 2 3 4 7 8 6 9 5 10) (1 2 3 4 7 6 8 5 9) (1 2 3 5 8 7 9 4 6 10) (1 2 3 4 6 5 7)
(1 2 3 5 8 6 9 4 7 10) (1 2 3 5 8 4 7 6 9) (1 2 3 5 9 4 7 8 6 10) (1 2 3 5 7 4 6 8)
(1 2 4 6 9 3 5 8 7 10) (1 2 4 6 8 3 5 7 9) (1 3 5 7 9 2 4 6 8 10) (1 2 3 4 5)
(1 3 5 7 8 2 4 6 9) (1 2 4 7 9 3 5 8 6 10) (1 2 4 6 7 3 5 8) (1 2 3 6 9 4 7 8 5 10)
(1 2 3 6 8 4 7 5 9) (1 2 4 7 8 5 9 3 6 10) (1 2 3 5 6 4 7) (1 2 4 7 8 6 9 3 5 10)
(1 2 3 6 7 5 8 4 9) (1 2 3 7 6 8 5 9 4 10) (1 2 3 6 5 7 4 8) (1 2 3 7 6 9 4 8 5 10)
(1 2 4 7 6 8 3 5 9) (1 3 5 8 7 9 2 4 6 10) (1 2 3 5 4 6) (1 3 5 8 6 9 2 4 7 10)
(1 2 4 7 5 8 3 6 9) (1 2 4 8 5 9 3 7 6 10) (1 2 4 7 3 6 5 8) (1 2 4 9 3 7 6 8 5 10)
(1 2 4 8 3 6 7 5 9) (1 2 5 8 4 7 9 3 6 10) (1 2 4 6 3 5 7) (1 3 6 9 2 5 8 4 7 10)
(1 3 5 8 2 4 7 6 9) (1 3 5 9 2 4 7 8 6 10) (1 3 5 7 2 4 6 8) (1 2 3 4)
(1 3 6 9 2 4 7 8 5 10) (1 3 6 8 2 4 7 5 9) (1 4 7 9 2 5 8 3 6 10) (1 3 5 6 2 4 7)
(1 4 7 8 3 6 9 2 5 10) (1 2 5 8 3 6 7 4 9) (1 2 5 9 3 7 6 8 4 10) (1 2 5 7 3 6 4 8)
(1 2 6 8 4 9 3 7 5 10) (1 3 6 7 4 8 2 5 9) (1 2 4 5 3 6) (1 3 6 7 5 8 2 4 9)
(1 2 6 7 5 9 3 8 4 10) (1 2 5 6 4 7 3 8) (1 2 6 7 5 8 4 9 3 10) (1 2 6 5 7 4 8 3 9)
(1 3 7 6 8 5 9 2 4 10) (1 2 5 4 6 3 7) (1 3 7 6 8 4 9 2 5 10) (1 2 6 5 8 3 7 4 9)
(1 3 7 5 9 2 6 8 4 10) (1 3 6 5 7 2 4 8) (1 3 7 6 9 2 4 8 5 10) (1 2 4 3 5)
(1 3 6 4 7 2 5 8) (1 3 8 4 9 2 6 7 5 10) (1 3 7 4 8 2 6 5 9) (1 4 8 3 7 5 9 2 6 10)
(1 3 6 2 5 4 7) (1 4 8 3 7 6 9 2 5 10) (1 3 8 2 6 5 7 4 9) (1 3 9 2 6 7 5 8 4 10)
(1 3 7 2 5 6 4 8) (1 4 9 2 6 8 3 7 5 10) (1 4 7 3 6 8 2 5 9) (1 3 5 2 4 6)
(1 2 3) (1 5 9 2 6 8 3 7 4 10) (1 4 7 2 5 6 3 8) (1 4 9 2 6 7 5 8 3 10)
(1 4 8 2 6 5 7 3 9) (1 4 6 2 5 3 7) (1 5 7 3 8 2 6 4 9) (1 5 8 3 9 2 6 7 4 10)
(1 3 4 2 5) (1 5 6 4 8 2 7 3 9) (1 4 5 3 6 2 7) (1 5 6 4 7 3 8 2 9)
(1 6 5 7 4 8 3 9 2 10) (1 5 4 6 3 7 2 8) (1 6 5 7 4 9 2 8 3 10) (1 4 3 5 2 6)
(1 6 5 8 3 9 2 7 4 10) (1 5 4 7 2 6 3 8) (1 3 2 4)
Table 4.1: Forcing order on unimodal permuations of length 10.
Lastly, we remark that an interesting combinatorics question is to ask how many unimodal
cycles there are for a given n. If we denote this number by u(n), then the above tables
shows us that u(3) = 1, u(4) = 2, u(5) = 3, u(6) = 5, u(7) = 9, u(8) = 16, u(9) = 28 and
30
u(10) = 51. Weiss and Rogers [WR87] have shown that
u(n) =
1
n

d|n
d odd
(d)2
(n/d)1
,
where is the Mobius function.
Aside from the results presented in this chapter, there are very few other results avail-
able which convey a good conception of the global structure of forcing in the same way
that Sharkovskys theorem does for the relation
s
[Hal90]. However, in the next chapter
we will show that the local structure of forcing is completely describable.
31
Chapter 5
Extensions of Cycles
Whereas in the preceding chapter we analyzed the global features of (C, ), this chapter
takes a look at the local structure of forcing by considering certain cycles called extensions.
In Section 5.1, we will study doublings of cycles. It will be shown that doublings completely
describe the local structure of forcing: immediately above any n-cycle lies all of its 2
n1
doubles. Doublings are in fact only one particular type of an extension, and in Section 5.2,
we consider more general extensions of cycles such as block structures and -extensions.
These concepts will allow us to further characterize sucient conditions for one cycle to
force another and give us a better idea of the local structure of forcing.
5.1 Doublings
The basic idea of a doubling of a cycle C
n
is that we thicken each point 1, . . . , n
into a block containing two points, and then these n blocks are permuted according to .
Denition 5.1.1. If n is even, then a cyclic permutation of length n is a double (or
doubling)
1
if it cyclically permutes the n/2 sets 1, 2, . . . , n1, n. In this case, is said
to be a double of the permutation dened by (i) = j if (2i 1, 2i) = 2j 1, 2j.
Example 5.1.2. Let be the 6-cycle (1 5 3 2 6 4). Then cyclically permutes the pairs
1, 2, 3, 4, 5, 6 and is a double of the permutation (1 3 2). This is illustrated in Figure
5.1. Note that the boxes are permuted according to (1 3 2 ).
Let us rst show how one may go about constructing doublings of a given n-cycle
. First, write out the points 1, . . . , 2n on a horizontal line and draw a block around
each successive pair of points 2i 1, 2i. Label the n blocks by P
1
, P
2
, . . . , P
n
so that
P
i
= 2i 1, 2i. Starting in block P
1
with the point 1, draw an arrow to a point
in P
(1)
which we will call (1). From (1) draw an arrow to a point in P

2
(1)
and
call this point
2
(1). Continue this process n 2 times so that we have the n points

0
(1) = 1, (1),
2
(1), . . . ,
n1
(1) with
i
(1) P

i
(1)
. Since, P

n
(1)
= P
1
, we have no
choice but to draw an arrow from
n1
(1) to 2. Now we begin the process over again, only
this time we rst draw an arrow from 2 to the point in P

n+1
(1)
= P
(1)
which previously
did not have an arrow drawn to it and label this point
n+1
(1). After n 1 more steps,
1
We follow the language of [BCMM92, BC92a, MN91] by referring to 2-extensions as doubles or dou-
blings, as this seems to be the standard now. However, in [Ber87], Bernhardt refers to a doubling by
saying that a cycle is splittable, and in [Bal87], Baldwin says that a cycle is divisible by 2 if it is a double.
32
1 2 3 4 5 6
Figure 5.1: The 6-cycle (1 5 3 2 6 4) as a doubling of (1 3 2).
we are again back in the block P

n1
(1)
. To nish the construction, we draw an arrow
from the point
2n1
(1) back to 1. If we follow the arrow path that leaves the point
1, we observe that this path hits every number before hitting 1 again; consequently, the
permutation on the 2n points is in fact a cyclic permutation. Moreover, that is a
doubling of is clear from the above construction and Denition 5.1.1.
The above construction of a doubling leads one to the following question: how many
doublings are there of a given n-cycle? Observe that for the rst n 1 steps in the
construction of above, we had precisely two choices. However, at the n
th
step, an arrow
had to be drawn back to the point 2 and from then on, all of the choices were determined.
This argument proves the following result:
Proposition 5.1.3. For any n-cycle , there are 2
n1
doublings of .
One of the most important uses of doublings is in classifying immediate successors and
predecessors in (C, ). Intuitively, an immediate successor to a cycle is a cycle such
that and there is nothing else in between and . For example, if we consider the
integers partially ordered by the less-than-or-equal-to relation , then 4 is an immediate
successor to 3, since 3 4 and x Z : 3 x 4 = . The following denitions make
precise what we mean by immediate successors and predecessors.
Denition 5.1.4. If is a cycle, let For() denote the set of cycles forced by , i.e.
For() = C : .
Denition 5.1.5. Let , C. Then is said to be an immediate successor to (or
is an immediate predecessor of ) if ,= and For() = For() .
A priori, given an arbitrary partially ordered set, one need not have an immediate
predecessor for every element. For example, if we consider the reals instead of the integers
partially ordered by , then there is no immediate predecessor to 3. Nevertheless, with
regards to the forcing order, Bernhardt proved that the following are equivalent [Ber87,
Theorem 1.12]:
B1. has an immediate predecessor;
B2. is a doubling;
B3. there is only one nonrepetitive loop from J
1
to itself in the Markov graph of ;
B4. there does not exist a loop in the Markov graph of whose orbit type is .
33
We will prove a result similar to this, which classies the most important properties of
doublings.
Theorem 5.1.6 (Doubling Theorem). Let C
n
, P = p
1
< p
2
< < p
n
=
1, 2, . . . , n and consider the connect-the-dots map f

. The following are equivalent:


1. P is the only orbit of f

that exhibits .
2. There is a loop in the Markov graph of with loop type .
3. is not a doubling.
Lemma 5.1.7 ([Ber87, Lemma 2.4]). Suppose that C
2n
is a doubling of C
n
. There
is precisely one loop in the Markov graph of beginning (and ending) at J
1
. Moreover,
this loop has orbit type .
Proof. The Markov graph of has 2n 1 vertices J
1
, . . . , J
2n1
. Since is a doubling,
cyclically permutes the n pairs P
k
= 2k 1, 2k, k = 1, . . . , n. In particular, since is
a doubling of , we have (P
k
) = P
(k)
. This implies that for each odd index 2
k
(1) 1,
the vertex J
2
k
(1)1
has precisely one edge leaving it, and this edge goes to the vertex
J
2
k+1
(1)1
. Therefore, the only loop beginning at J
1
is the loop
J
1
J
2(1)1
J
2
n1
(1)1
J
1
.
This loop is of length n, has no arrows leaving the loop, and corresponds to an orbit of
type for the map f

.
As a result of Lemma 5.1.7, we have the following:
Corollary 5.1.8. If is a doubling of , then forces .
Lemma 5.1.9 ([Bal87, Lemma 4.7]). Suppose that C
2n
is a doubling of C
n
. The
subgraph of the Markov graph of consisting of all even-indexed vertices and all arrows
from a vertex of even index to another vertex of even index is graph isomorphic to the
Markov graph of .
Proof. Let k, i 1, . . . , n. By denition, a directed edge exist from J
2k
to J
2i
when
f

([2k, 2k + 1]) [2i, 2i + 1]. There are two possibilities for this to happen, either f

is
increasing on [2k, 2k +1] or f

is decreasing on [2k, 2k +1]. These cases are illustrated by


Figure 5.2.
Algebraically, this implies that we have J
2k
J
2i
if either
1. (2k) 2i + 1 and (2k + 1) 2i; or
2. (2k) 2i and (2k + 1) 2i + 1.
Now, if 2k P
k
, then (2k) is in (P
k
) = P
(k)
= 2(k) 1, 2(k). Similarly, if
2k + 1 P
k+1
, then (2k + 1) is in (P
k+1
) = P
(k+1)
= 2(k + 1) 1, 2(k + 1).
Therefore, there is an arrow from J
2k
to J
2i
if and only if one of the following holds:
1. (k) i and (k + 1) i + 1; or
2. (k + 1) i and (k) i + 1.
But, this is precisely the condition for there to be an arrow from the vertex of index j to
the vertex of index i in the Markov graph of . Thus the Markov graph of restricted to
the vertices with even subscripts is equivalent to the Markov graph of .
34
Case 1 Case 2
2i + 1
2i
2i + 1
2i
2k 2k + 1 2k 2k + 1
Figure 5.2: The two possibilities for J
2k
J
2i
.
For the next lemmas, assume that C
n
and P = p
1
< < p
n
= 1, . . . , n.
Recall that an orbit Q = q
1
< , c
n
exhibits if f

(q
i
) = q
(i)
. Trivially, P exhibits
f

. We will say that q


i
is next to p
i
for some i 1, . . . , n if Qq
i
, p
i
) = P q
i
, p
i
) = .
If a point p
c
is a critical point of f

(equivalently, c is a critical point of ), then we shall


call c a critical index. Note that a critical index may not be equal to be equal to 1 or n.
Lemma 5.1.10. For any i 1, . . . , n, there is no k dierent from i such that q
k
, p
k

q
i
, p
i
).
Proof. Suppose on the contrary that some q
k
q
i
, p
i
) for k ,= i. Then we either have
q
i
< q
k
< p
i
or p
i
< q
k
< q
i
. Since P and Q have the spatial labeling, in the rst case, we
must have p
i
< p
k
; hence, p
k
/ q
i
, p
i
). In the second case, we must have p
k
< p
i
and so
again p
k
/ q
i
, p
i
).
Lemma 5.1.11. No point of Q lies between q
c
and p
c
for any critical index c.
Proof, version 1, no cases. Suppose c is a critical index and that there exists some q
i
Q
that is between q
c
and p
c
. Note that both p
c
and q
c
are the same type of extrema either
both are relative maxima or both are relative minima. If this were not the case, then Q
and P would exhibit dierent cycles. Dene a function : [1, n] 1, 1, 0 by
(x, y) = sgn([x y][f

(x) f

(y)]) .
is a symmetric function, and locally, tells us whether f

is increasing or decreasing.
In particular, suppose that p is a critical point and x < p < y are two nearby points.
Then, if (p, x) = 1 and (p, y) = 1, by denition of , we see that p must be a relative
maximum (so long as x and y are close enough to p that there are no other extrema in
[x,y]). On the other hand, suppose that x
1
< x
2
< x
3
< < x
n
are points such that
(x
1
, x
2
) = (x
2
, x
3
) = (x
j
, x
j+1
),
but (x
j+1
, x
j+2
) is not equal to (x
1
, x
2
) for some 1 < j < n2. We see that essentially
detects the existence of critical points, in particular in the above example, there must be
a critical point in [x
j
, x
j+1
].
Now, since p
c
and q
c
are critical points of the same type and q
i
q
c
, p
c
), we must
have (q
i
, q
c
) = (q
i
, p
c
). But, only changes sign in the presence of a critical point,
35
and since all critical points of f

lie in P, there must exist some other critical point p


c
in
p
c
, q
c
). By Lemma 5.1.11, the corresponding point in the orbit of Q cannot also be in the
interval p
c
, q
c
); that is, q
c
/ q
c
, p
c
). We now have the situation with which we began: a
point of Q lies between q
c
and p
c
for a critical index c

. By the same reasoning as above,


there must exist another critical point p
c
in q
c
, p
c
) and consequently q
c
lies outside
of q
c
, p
c
). Note that if q
c
was left (resp. right) of q
c
, p
c
) then q
c
is also left (resp.
right) of q
c
, p
c
). Since f

has only nitely many critical points left (resp. right) of p


c
,
we eventually run out of critical points left (resp. right) of p
c
and reach a contradiction.
This shows that for any critical index i, no point of Q lies between q
i
and p
i
.
Proof, version 2, cases. The motivation for the rst proof came from the fact that even
though there are 4 separate cases to be handled, the argument used in each case is essen-
tially the same. The denition of seems to take care of the need for describing each of
these cases. The set up is the same as above: c is a critical index and we suppose there is
some q
i
q
c
, p
c
).
1. Suppose q
c
< q
i
< p
c
and p
c
(and thus q
c
) are relative maxima. Since p
c
is a
relative maximum of f

, in the presence of what we have right now, we should have


f

(q
c
) < f

(q
i
) < f

(p
c
). But, q
c
is a relative maximum of Q, so we should actually
have f

(q
c
) > f

(q
i
). Since all critical points of f

lie in P, there must be some other


critical point p
c
of f

that lies in [q
c
, p
c
]. From what we have, we cannot decide
whether q
c
< p
c
< q
i
or q
i
< p
c
< p
c
; however, this is not important. By Lemma
5.1.10, we cannot have q
c
[q
c
, p
c
]. But, since p
c
< p
c
, the spatial labeling on Q
requires that q
c
< q
c
(Figure 5.3). The argument begins again: q
c
, a point from Q,
q
c
q
i
p
c
q
c
q
i
p
c
q
c
p
c
p
c

Figure 5.3: The rst case: q


c
< q
i
< p
c
and p
c
a relative maximum.
separates critical points q
c
and p
c
. By the reasoning above, there must be a critical
point p
c
in [q
c
, p
c
]. This process will continue moving to the left, adding in critical
points until we have run out, which reaches a contradiction.
2. Next, if q
c
< q
i
< p
c
and p
c
(and thus q
c
) are relative minima, then note that the
argument is the same as above, except we should have f

(q
c
) > f

(q
i
) > f

(p
c
).
This does not aect the reasoning for the existence of a critical point to the left of
[q
c
, p
c
].
3. This leaves the case when p
c
< q
i
< q
c
, which is handled similarly as the above
cases.
36
Lemma 5.1.12. For each i, q
i
, p
i
) is mapped monotonically under f.
Proof. Suppose that for some i 1, . . . , n, q
i
, p
i
) is not mapped monotonically under
f

. Since all critical points of f

lie in P, there is some critical point p


c
q
i
, p
i
) for
c ,= i. By Lemma 5.1.11, there can be no point of Q between q
c
and p
c
, so we must have
q
c
q
i
, p
i
). But, now we have q
c
, p
c
q
i
, p
i
), contradicting Lemma 5.1.10. Therefore,
for each i, there is no critical point in q
i
, p
i
); hence q
i
, p
i
) is mapped monotonically under
f

for each i.
Now, we are ready to prove the Doubling Theorem:
Proof of Theorem 5.1.6.
2 3. Suppose that C
2n
is a doubling of C
n
and let = (J
i1
, . . . , J
i2n1
, J
i1
) be a
loop in the Markov graph of of length 2n. If J
1
is a vertex in , then by Lemma
5.1.7, modulo the shift operation, we have
= (J
1
, J
2(1)1
, , J
2
n1
(1)1
, J
1
, J
2(1)1
, , J
2
n1
(1)1
, J
1
)
and, moreover, does not correspond to an orbit of length 2n. Therefore, J
1
is not
a vertex in . Since no arrow in the Markov graph of goes from an odd-index
vertex to an even-index vertex, we must have consisting of only vertices with even
indices. By Lemma 5.1.9, the subgraph consisting of only even-index vertices is
graph isomorphic to the Markov graph of . Therefore, if were to correspond to
an orbit type , then there would be a loop in the Markov graph of of orbit type
, implying that . But, by Corollary 5.1.8, is a doubling of , so , and
therefore by the antisymmetry of forcing (Theorem 2.1.5), we must have = .
Contradiction. Thus, if is a doubling of , then there is no loop in the Markov
graph of having orbit type .
3 1. By contrapositive. Suppose that Q ,= P is another orbit of f

that exhibits . By
Lemma 5.1.12, f

is monotone on q
i
, p
i
) for each i, so, in fact, any iterate of f

is
monotone on q
i
, p
i
). Therefore, since P and Q are orbits of n-periodic points, f
n

takes 1 to 1, q
1
to q
1
and is monotone on [1, q
1
]. This shows that f
n

is the identity
on [1, q
1
], and so has derivative equal to 1 on the entire interval [1, q
1
]. We want
to show that, in fact, f
n

has derivative 1 on [1, 2]. If 2 [1, q


1
], then certainly
this holds. Otherwise, let x be largest such that f
n

has derivative 1 on [1, x] (and


therefore f
n

(x) = x). Then, f


k

(x) P for some iterate k strictly less than n since


the derivative of f
n

changes only at points that hit a critical point of f

(all of which
are in P) for such an iterate. Now, as f
k
(P) = P for all k 0, once an iterate of x
lands in P, it stays in P. Since k < n, this implies f
n

(x) = x P and as 1 < x 2,


it follows that x must equal 2. Thus, the derivative of f
n

is 1 on the interval [1, 2].


This implies that every iterate of f

has slope +1 or 1 on [1, 2]. Let f


k

be the
iterate that takes the point 2 to 1. The slope 1 condition tells us that f
k

(1)
must be equal to 2, and then we are done: since is a cycle, if the same iterate takes
1 to 2 and 2 to 1, then is a doubling.
1 2. Suppose that P is the only orbit that exhibits . Dene a loop in the Markov
graph of in the following way: choose small enough so that the length of the
interval f
k

([1, 1 + ]) is less than 1 for 0 k n. Then for each k there exists a


unique integer i(k) such that
f
k

([1, 1 +]) [i(k), i(k) + 1].


37
The loop is the path dened by J
i(k)
J
i(k+1)
for 0 k n + 1. Note that
since i(0) = 1 and i(n) = 1 this path is actually a loop [Ber87]. This loop is either
nonrepetitive or consists of repetitions of a shorter loop. Following [Ber87, Lemma
3.6], we will consider each case separately.
If consists of a nonrepetitive loop then the loop type of belongs to C
n
. Since (k)
is the unique integer in f
k
([1, 1 +]), it is clear that the loop type of is .
Suppose now that consists of repetitions of a shorter loop. Since f
k
([1, 1 + ]),
0 k n 1, contains all the integers from 1 to n, can repeat a vertex J
i
at
most twice. Thus the loop type of must be a cycle of length n/2. The loop starts
at J
1
, so this means that J
1
is repeated twice and so J
2
cannot be repeated, thus
J
2
is not in the path. Similarly, it can be shown that none of the even subscripted
vertices are in the loop and all of the odd indexed vertices are contained in the path.
This implies that is a doubling, and since we have shown 3 1, this contradicts
P being the only orbit exhibiting . Thus, if P is the only orbit exhibiting , then
there is a loop in the Markov graph of with loop type .
Before moving on to general extensions, observe Figure 5.4 which illustrates a sort of
tree of doublings. First, recall that there are 2
n1
doublings of any n-cycle (equivalently,
there are 2
n1
immediate successors to any n-cycle). Therefore, beginning with the unique
2-cycle (1 2), there are two doublings; namely, (1 3 2 4) and (1 4 2 3). Now, each of these
4-cycles has 2
41
= 8 doublings, all of which are diagrammed in Figure 5.4. The structure
of any of these 8-cycles is precisely what we would expect: any 8-cycle in Figure 5.4 has a
block structure over the 4-cycle that it doubles, and since this 4-cycles is a double of (1 2),
each 8-cycle also admits a division.
(1 2)
(1 3 2 4) (1 4 2 3)
(1 6 3 7 2 5 4 8) (1 8 3 6 2 7 4 5) (1 5 3 7 2 6 4 8) (1 7 4 6 2 8 3 5)
(1 5 4 7 2 6 3 8) (1 8 3 5 2 7 4 6) (1 6 3 8 2 5 4 7) (1 8 4 5 2 7 3 6)
(1 6 4 8 2 5 3 7) (1 7 3 6 2 8 4 5) (1 5 4 8 2 6 3 7) (1 8 4 6 2 7 3 5)
(1 5 3 8 2 6 4 7) (1 7 3 5 2 8 4 6) (1 6 4 7 2 5 3 8) (1 7 4 5 2 8 3 6)
Figure 5.4: Tree of doublings starting at (1 2).
38
5.2 General Extensions of Cycles
Having observed that much information about the local structure of forcing can be gained
by looking at doublings, we now consider more general extensions of cyclic permutations.
The rst type of extension that we will consider is that of a block structure. The
essential idea of a block structure is much like that of a doubling: we start with some
base cycle and then make each point of the cycle thick so that many points now
move according to . More formally, we have the following denition:
Denition 5.2.1. Let n N be such that n = km and pick C
n
and C
m
. We say
that has a block structure over if we can write P = 1, . . . , n as the disjoint union
of m blocks of size k, P = P
1
P
m
where P
i
= (i 1)k + 1, . . . , (i 1)k + k for
i = 1, . . . , m and (P
i
) = P
(i)
. Additionally, we shall call a reduction or the base cycle
of if has a block structure over .
One may also simply say that an n-cycle has a block structure if for each P
i
as above
there is a P
j
such that (P
i
) = P
j
. Thus, one need not start with a base cycle to dene
a block structure. Clearly, if a block structure is detected, then the base cycle is not
hard to nd: one need only dene P
(i)
= (P
i
) to retrieve .
Example 5.2.2. The 9-cycle = (1 9 4 3 7 6 2 8 5) has a block structure over the 3-cycle
= (1 3 2). Letting P
1
= 1, 2, 3, P
2
= 4, 5, 6 and P
3
= 7, 8, 9, we see that
(P
1
) = (1), (2), (3) = 9, 8, 7 = P
3
= P
(1)
(P
2
) = (4), (5), (6) = 3, 1, 2 = P
1
= P
(2)
(P
3
) = (7), (8), (9) = 6, 5, 4 = P
2
= P
(3)
This block structure is shown in Figure 5.5.
1 2 3 4 5 6 7 8 9
Figure 5.5: A block structure over the 3-cycle (1 2 3).
Our goal now is to see if we can obtain information about forcing by considering block
structures. In Section 5.1, we showed that the Markov graph of a doubling has an easily
identiable structure (Lemma 5.1.9). Let us say that a set of vertices in a Markov
graph is invariant if every edge emanating from a vertex in terminates in a vertex in .
Moreover, if we are dealing with a block structure and the points i, i + 1 are in distinct
blocks, then the interval [i, i +1] (and the vertex J
i
in the Markov graph) will be called a
P-gap.
Theorem 5.2.3 ([MN91, Theorem 4.1], [Bal87, Lemma 4.7]). Suppose that C
n
has a
block structure over C
m
.
1. The set
p
of vertices J
i
in the Markov graph of for which J
i
is not a P-gap is an
invariant set.
39
2. The set
g
of all P-gap vertices, together with all edges in the Markov graph of
which terminate (and hence originate) in a P-gap vertex, then
g
is isomorphic, as
a directed graph with an ordered set of vertices to the Markov graph of .
Proof. Suppose that n = k m so that has an m-block structure with k points in each
block. Observe that the the set of P-gap vertices is
g
= J
k
, J
2k
, . . . , J
(m1)k
and so

p
= J
i
: 1 i n 1, k i.
Proof of (1). The invariance of
p
follows directly from the denition of a block
structure since the block
P
i
= J
ki+1
, J
ki+2
, . . . , J
ki+k1
: 1 i m
moves to
P
(i)
= J
k(i)+1
, J
k(i)+2
, . . . , J
k(i)+k1
: 1 i m,
and as is an m-cycle, no arrow leaves
p
and goes to a vertex with an index divisible by
k.
Proof of (2). The map taking the index j of a vertex in
g
to j/k shows that
g
is
isomorphic to the Markov graph of .
As an immediate consequence of Theorem 5.2.3, we get the following result which will
be important in the forcing proof of Sharkovskys theorem:
Corollary 5.2.4. If C
n
has a block structure over C
m
, then .
In Theorem 2.1.5 we proved that forcing is a partial order, hence a transitive relation.
Therefore, as a result of Corollary 5.2.4, we also have
Corollary 5.2.5. Let , , C. If has a block structure over and , then .
The next type of extension that we will consider is a particular type of block structure.
Denition 5.2.6. Let be an n-cycle and be an m-cycle for some n, m N. A nm-cycle
is called an extension of by or a -extension of if
1. has a block structure over .
2. There is some block P
i
such that P
i
is an orbit of type under f
m

.
If in addition, satises
3. f

is monotone on each block except at most one,


then we will say that is a simple extension of by .
We will primarily be concerned with simple extensions. The essential idea of a simple
extension by is the following. Suppose that we have C
n
and C
m
. Take each
point i of 1, . . . , m and make it into a block P
i
of n points, so that we have m blocks
P
1
, . . . , P
m
each of size n. Pick some i 1, . . . , m and draw an (nm nm)-grid where
we will graph the connect-the-dots map of our extension. Connect the dots in the square
P
i
P
(i)
to resemble the graph of . On the remaining P
j
dierent from P
i
, connect the
dots of the square P
j
P
(j)
with either a line of slope +1 or -1. This is best understood
through an example.
40
Example 5.2.7. Let us describe a simple extensions of a 3-cycle by itself. Let = (1 2 3),
so that the simple -extension of will be a 3 3 = 9-cycle, which we will denote by
. Let us choose the block P
1
= 1, 2, 3 to satisfy condition 3 of Denition 5.2.6. This
means that we need to draw the graph of in the square P
1
P
(1)
= P
1
P
2
. This
amounts to connecting the points (1, 5) to (2, 6) and (2, 6) to (3, 4). Now, condition 2 in
Denition 5.2.6 tells us that f

must be monotone on the remaining blocks P


2
and P
3
.
Since (2) = 3, in the square P
2
P
3
, we connect the points (4, 9) to (5, 8) and (5, 8) to
(6, 7). Lastly, since (3) = 1, in the square P
3
P
1
, we connect the points (7, 3) to (8, 2)
and (8, 2) to (9, 1). The slope of the graph of f

in the blocks of P
2
and P
3
was chosen so
that satises Condition 1 of Denition 5.2.6, i.e., has a block structure over . The
graph of f

is shown in Figure 5.6. Observe that if we were to collapse each of the blocks
to a point, then this graph is the graph of f

. This is the general idea of an extension by


a cycle. By analyzing Figure 5.6, we see that = (1 5 8 2 6 7 3 4 9).
1 2 3 4 5 6 7 8 9
1
2
3
4
5
6
7
8
9
Figure 5.6: -extension of for = (1 3 2).
By denition, if is a simple extension of by , then there is at least one block that
is an orbit of type under f
n

. Can we deduce anything about the rst-return pattern of


the other blocks? Since we are dealing with simple extensions, the following proposition
tells us that we can.
Proposition 5.2.8 ([MN91, Remark 6.1], [ALM93, Proposition 2.10.5]). Suppose is a
simple extension of C
m
by C
n
. Then the orbit type of f
n

(P
i
) is either or

for
i = 1, 2, . . . , m.
Proof. Let j
0
denote the index of the block guaranteed by Denition 5.2.6 which is an
orbit of type under f
n

. For j ,= j
0
, the map
f
j
:= f
n


P
j
: P
j
P
(j)
41
is a homeomorphism. Therefore, the map
F = f

1
(j)
f

2
(j)
f
j0
f
(j)
f
j
: P
j
P
j
can be written in the form h
1
F
j0
h
2
for homeomorphisms h
1
and h
2
. Thus, since P
j0
is an orbit of type under f
n

, P
j
is an orbit of type or

depending on whether h
1
and
h
2
are orientation preserving or reversing.
Now, given , a simple -extension of some cycle , we want to know which cycles are
forced by .
Proposition 5.2.9 ([ALM93, Proposition 2.10.6]
2
). Let be a simple -extension of
and suppose that . Then either or is a

-extension of for some

forced by
or

. If ,= then in the last case both

,= and

,=

.
Proof. Assume that C
n
, C
m
and let P = 1, 2, . . . , n. Using the denition of a
-extension of , construct the m-blocks P
1
, . . . , P
m
. Since , there is an orbit R of f

which exhibits . Since f

is P-monotone, the set M =

m
i=1
P
i
) is f

-invariant. Hence
either R M = or R M.
Assume rst that R M = . We collapse each block P
i
) into a point and we
obtain from [1, n] an interval which can be stretched back to [1, n]. More precisely, there
exists a non-decreasing continuous map : [1, n] [1, n] such that is constant on
P
i
) and increasing on each component of [1, n]M. Then, the map

f dened by

f(x) =
f

(
1
(x)) is a continuous map of [1, n] and satises f

=

f . The set (M)
is an orbit of

f and by the denition of an extension, it exhibits the cycle . Moreover,

f
is (M)-monotone. Since is non-decreasing, (R) is an orbit of

f exhibiting the cycle
. Therefore, by the Theorem 2.2.1, .
Now, assume that R M. Denote R
i
= RP
i
). If f

(P
i
) = P
j
, then f

(P
i
)) = P
j
)
and hence R has a block structure over and f

is monotone on each block R


i
of R except
at most one. This shows that is a

-extension of for

equal to the rst-return pattern


of the block R
1
under f
m

. However, R
1
is an orbit of the P
1
-monotone map f
m


P
1
)
.
Therefore, by the forcing theorem, we conclude that the rst-return pattern of the block
P
1
under f
m

forces

. But by Proposition 5.2.8, the rst-return pattern of P


1
under f
m

is either or

, so we conclude that

or

.
Suppose that =

. By Corollary 4.1.5 and Theorem 2.2.1, we get that (P


1
, f
m

) and
(R
1
, f
m

) exhibit the same cycles. Now let P = p


1
, p
2
, . . . , p
n
and R = r
1
, r
2
, . . . , r
n

have the spatial labeling and let f

(p
j
) = p
1(j)
and f

(r
j
) = r
2(k)
for cyclic permutations

1
and
2
. If f

is monotone on P
i
), then
1
(j) =
2
(j) for j = (i 1)s + 1, (i 1)s +
2, . . . , (i 1)s + s where s = n/m. However, if f

is not monotone on a block P


i
, then
both
1
(j) and
2
(j) for j = (i 1)s + 1, (i 1)s + 2, . . . , (i 1)s +s are determined by

1
(k) and
2
(k) for k 1, 2, . . . , n(i 1)s + 1, (i 1)s + 2, . . . , (i 1)s + s (where
they are equal) and by the cycles (P
1
, f
m

) and (R
1
, f
m

) (which are also equal). Therefore,

1
=
2
, which implies = .
Lemma 5.2.10 ([ALM93, Lemma 2.11.4]). Let be a doubling of and suppose that
. Then either = or .
Proof. Assume that ,= . Since the unique 2-cycle (1 2) only forces itself and trivial cycle
(1), by Proposition 5.2.9, either or is a (1)-extension of . However, a (1)-extension
of is itself, so in every case .
2
Here dening and

to be equivalent would simplify the statement of this theorem (see Chapter 8).
42
Lemma 5.2.10 proves the earlier assertion that given a cycle , every doubling of is
an immediate successor to in the forcing order. The following is a sort of partial converse
to Proposition 5.2.9:
Proposition 5.2.11 ([ALM93, Lemma 2.10.8]). Let ,

, and be cycles and suppose


that is a simple -extension of and

or

. Then forces a simple

-extension
of .
Proof. Assume that C
n
and C
m
. Since is a simple -extension of , has a
block structure P
1
, . . . , P
m
over . Let us give these blocks the temporal labeling so that
f

(P
i
) = P
i+1
(where addition of indices is modulo m). Proposition 5.2.8 tells us that
the block P
1
has the rst-return pattern of either or

; that is, P
1
is an orbit of type
or

for the map f
m

. Now, by hypothesis forces

or

. Let us rst assume that


forces

. Then, there is an orbit Q


1
of f
m

that exhibits the cycle

and is such that


Q
1
) P
1
). For i = 2, . . . , m, dene Q
i
= f
i1

(Q
1
) and let Q =
m
i=1
Q
i
. Then we have
Q
i
) P
i
) for each i. Moreover the orbit Q exhibits a cycle that necessarily has a block
structure over . Since is a simple extension, the map f

is monotone on at east m1
intervals P
i
). Therefore, the cycle that Q exhibits is in fact a simple

-extension of .
By Theorem 2.2.1, forces this cycle. The above argument needs only slight changes if
we assume that

.
43
Chapter 6
Forcing-Minimal Cycles
This chapter aims to study minimality properties of forcing. The results presented here
provide the nal tool necessary to complete the forcing proof of Sharkovskys theorem
(Chapter 7). Looking ahead, to prove Sharkovskys theorem, we will begin with some
continuous map that has a periodic point of period n. This periodic point exhibits some
n-cycle, say , and it will be helpful to know if there exists some elementary n-cycle that
forces. If this were to be the case, then we could assume without loss of generality from
the beginning that the periodic point exhibits such an elementary cycle. We will show
that there are two notions of elementary cycles that of primary cycles (Denition 6.1.1)
and simple cycles (Denition 6.2.6) and will show that these notions actually coincide,
even though on the surface they are entirely dierent concepts. The general exposition of
this chapter very closely follows [ALM93, Chapter 2.9-2.11].
6.1 Primary Cycles
Looking at Figure 3.2, we see that among all 4-cycles, (1 3 2 4) and (1 4 2 3) are the only ones
that do not force any other 4-cycles. Considering 5-cycles now, we see that (1 3 2 4 5) and
(1 5 3 2 4) are minimal as well in this respect. This section aims to characterize properties
of such forcing-minimal cycles.
Denition 6.1.1. An n-cycle will be called primary
1
if it does not force any other n-cycle.
Equivale
Observe that for n = 1, the trivial cycle (1) is primary; for n = 2, (1 2) is the unique
primary cycle; and, for n = 3, (1 2 3) and (1 3 2) are the only 3-cycles and neither forces
the other, so they are primary. In the introduction, it was observed that there seem to be
two primary 4-cycles as well as two primary 5-cycles. In particular, note that (1 3 2 4) is
the ip of (1 4 2 3), and (1 3 2 4 5) is the ip of (1 5 3 2 4). The following proposition tells
us that this is always the case.
Proposition 6.1.2 ([ALM93, Proposition 2.9.1]). If is primary then

is primary.
Proof. Suppose that C
n
is primary. If

is not primary, then there exists some
C
n
, ,=

such that

. By Proposition 4.1.4, this implies that , and since
1
The notion of primary cycles, which we have adapted from [ALM93], is referred to as -minimal
in [Bal87].
44
is primary, = . But, since equal cycles have equal ips, we must have that

= .
Contradiction.
Thus, if we know that an n-cycle is primary, we know that there is at least one other
n-cycle that is primary. The next theorem will justify our interest in primary cycles by
showing that every n-cycle forces a primary cycle.
Proposition 6.1.3 ([ALM93, Theorem 2.9.4]). If C
n
, then there exists a primary
cycle C
n
such that .
Proof. Let C
n
. Theorem 2.1.5 tells us that is a partial order, and therefore,
restricted to the set of all n-cycles is also a partial order. This set is nite for any n N,
so there must exist some C
n
such that and forcing-minimal; that is, is
primary.
Before moving onto the concept of a simple cycle, we note that there is an interesting
corollary of Proposition 5.2.9 that relates the primarity of a cycle to extensions.
Corollary 6.1.4 ([ALM93, Corollary 2.9.5]). Let be a -extension of , say C
n
. If
is primary and does not force any n-cycles, then is primary.
6.2 Simple Cycles
Some of the earliest types of cycles on which the forcing relation was analyzed are the
so-called simple cycles. In [Ber84], Bernhardt described the structure of forcing on simple
permutations of a power two. If n is even, the essential idea of a simple cycle is that we have
a block structure consisting of two blocks P
1
= 1, . . . , n/2 and P
2
= n/2 + 1, . . . , n,
and P
1
is sent to P
2
and vice-versa. Thus, if we consider n = 2k points equally spaced on
a horizontal line, then one can draw a vertical line between the rst k points and the last
k points, and all points left (resp. right) of this line move to the right (resp. left) of the
line under an even simple cycle.
Simple cycles of an odd length are known as

Stefan cycles after the Czech mathemati-
cian Peter

Stefan who explicitly used them in his proof of Sharkovskys theorem [

Ste77].
Denition 6.2.1. Let q = 2k 1 for some k 2. A q-cycle is called a

Stefan cycle if
when expressed in cycle notation it is equal to either
(k (k 1) (k + 1) (k 2) (k + 2) (k j) (k +j) (2k 1))
or
(k (k + 1) (k 1) (k + 2) (k +j) (k j) (2k 1) 1) .
Equivalently, for q odd, q 3, we say that C
q
is a

Stefan cycle if there exists a
temporal labeling of P = p
1
, . . . , p
n
= 1, . . . , n such that either
p
n
< p
n2
< < p
5
< p
3
< p
1
< p
2
< p
4
< < p
n3
< p
n1
or
p
n1
< p
n3
< < p
4
< p
2
< p
1
< p
3
< p
5
< < p
n2
< p
n
The denition of a

Stefan cycle given above is far more complicated than the actual
concept of a

Stefan cycle. Essentially, for q = 2k 1, one starts at the point k, moves
left to the point k 1, then right to the point k + 1, then left to the point k 2, and so
45
on continuing in this spiraling out fashion until one reaches 2k 1, at which point the
cycle starts over at k. This construction can also be done by starting out moving from k
to k + 1, hence the two dierent forms of a

Stefan cycle given in Denition 6.2.1.
Example 6.2.2. Let us illustrate the characteristic spiraling out property of

Stefan
cycles described above. Let q = 5 = 2 3 1. According to Denition 6.2.1, there are two

Stefan cycles of length q, (3 2 4 1 5) = (1 5 3 2 4) and (3 4 2 5 1) = (1 3 4 2 5). These 5-cycles


are illustrated in Figure 6.1. Note that both cycles spiral out from 3.
1 2 3 4 5 1 2 3 4 5
Figure 6.1: The

Stefan 5-cycles and their characteristic spiraling out property.
The following results begin to show the usefulness of

Stefan cycles with regards to
proving Sharkovskys theorem.
Lemma 6.2.3 ([ALM93, Lemma 2.1.7]). The vertices of the Markov graph of a

Stefan
cycle of length n can be labeled so that
i. J
1
J
2
J
n1
J
1
ii. J
1
J
1
iii. J
n1
J
q
for q odd.
Proof. To show this we will use the second (equivalent) denition of a

Stefan cycle given
in Denition 6.2.1. Give P = 1, . . . , n = p
1
, . . . , p
n
the temporal labeling. Let J
1
=
[p
1
, p
2
] and J
i
= [p
i1
, p
i+1
] for i = 2, 3, . . . , n 1. Then this lemma follows from the
denitions of a

Stefan cycle and the Markov graph of a cycle. See Figure 6.2.
Proposition 6.2.4 ([ALM93, Lemma 2.1.8]). Let C
n
be a

Stefan cycle. Then for any
m
s
n there exists an m-periodic point of f

whose orbit is contained in (1, n).


Proof. Up to a relabeling of vertices, Lemma 6.2.3 completely describes the Markov graph
of .
Since is a

Stefan cycle, n is odd and n 3. Therefore, if m
s
n, then either m > n
or m 1, 2, 4, 6, 8, . . . , n1 (see the Sharkovsky ordering Section 1.3). Let us rst take
care of the case when m > n. Consider the nonrepetitive loop
J
1
J
2
J
n1
J
1
J
1
J
1
of length m in the Markov graph of . Since this loop is nonrepetitive, by Lemma 1.2.9,
it corresponds to an m-periodic point of f

.
Next, consider the case when m < n. We need to show that f

has an m-periodic point


for each m 1, 2, 4, 6, 8, . . . , n1. By Lemma 1.2.4, f

has a xed point. Now, for each


m ,= 1, take the nonrepetitive loop
J
n1
J
nm
J
nm+1
J
n2
J
n1
46
J1 J2
Jn1 J3
Jn2 J4

J5
Figure 6.2: The Markov Graph of a Stefan n-cycle.
of length m in the Markov graph of . By the same reasoning as above, since this loop is
nonrepetitive and of length m, it corresponds to an m-periodic point of f

.
Proposition 6.2.5 ([ALM93, Proposition 2.11.3]). Each Stefan cycle is primary.
Proof. Let C
n
be a

Stefan cycle. Lemma 6.2.3 describes the Markov graph of . In
particular, this lemma shows us that there are precisely two loops of length n in this graph:

1
= J
1
J
2
J
n1
J
1
J
1
and
2
= J
1
J
1
J
1
. Now,
2
is not
a simple loop so it does not correspond to an orbit of length n for f

. Therefore, the orbit


type of
2
is not an n-cycle. On the other hand,
1
is a simple loop and has orbit type .
Thus, by Theorem 2.2.1, the only n-cycle that forces is itself, so is primary.
Denition 6.2.6. A cycle of length q 2
k
for q odd and k 0 will be called simple
2
if it
can be obtained from a cycle of period 1 by making k 2-extensions and then, if q > 1, one

Stefan extension.
Observe that for q 3 and k = 0 (as above), there are precisely two simple cycles,
namely the

Stefan q-cycles. However, for k 1 there can be very many simple cycles.
We are using a notion of a simple cycle that is somewhat stronger than the notion found
in [Ber84] and [Bal87]. In these papers, a simple cycle need only have a 2-block structure
and no

Stefan mixing of the blocks is required. An interesting result of Bernhardt [Ber84,
Lemma 2.2] says that there 2
2
n
(n+1)
simple permutations of period 2
n
.
Example 6.2.7. Let us begin with some examples of simple 6-cycles. Since 6 = 3 2
1
,
Denition 6.2.6 tells us to extend the trivial cycle (1) by the two cycle (1 2), then extend
this by a

Stefan 3-cycle. Recall that all 3-cycles are

Stefan cycles, so let us extend (1 2)
by (1 3 2). This gives us either the 6-cycle = (1 6 3 4 2 5) or = (1 6 2 5 3 4). Now, by
analyzing Figure 3.2 in which all 6-cycles are diagrammed one sees that both and
are forcing-minimal, i.e., they forces no other 6-cycles. Thus, and are primary and so
2
The denition that Baldwin uses for simple [Bal87, Denition 4.8] refers only to a 2-extension. Thus,
for Baldwin, there exist simple cycles that are not primary.
47
by Proposition 6.1.2,

= (1 4 3 5 2 6) and = (1 5 2 4 3 6) should also be primary. Looking
again at Figure 3.2, we see that indeed

and are primary. Lastly, we note that had we
chosen to extend (1 2) by (1 2 3) instead of (1 3 2), we would have have produced

and .
Consequently, the above discussion gives all of the primary and simple 6-cycles. Figure
6.3 illustrates the fact that all of the simple 6-cycles have the characteristic two block
and spiraling

Stefan structure.
1 2 3 4 5 6 1 2 3 4 5 6
1 2 3 4 5 6 1 2 3 4 5 6
= (1 6 3 4 2 5) = (1 6 2 5 3 4)

= (1 4 3 5 2 6) = (1 5 2 4 3 6)
Figure 6.3: The primary and simple 6-cycles.
Example 6.2.7 points towards the most important theorem of this section, namely
Theorem 6.2.10, which says a cycle is primary if and only if it is simple. This will be
one of the keys to proving Sharkovskys theorem via forcing. We will rst work towards
showing that simple primary. To do so, wee need a couple of lemmas. The rst lemma,
Lemma 6.2.8, essentially tells us that the simple cycles of length 2
k
form a tree in the
forcing relation. The second lemma, Lemma 6.2.9, shows that any primary extension of a
simple cycle of length a power of 2 is itself primary.
Lemma 6.2.8 ([ALM93, Lemma 2.11.5]). Let be a simple cycle of length 2
k
for some
k 0. If forces , then is a 2
j
-cycle for some j k such that C
2
j .
Proof. This result follows almost immediately from the section on doublings (Section 5.1).
We prove this by induction on k. For k = 0, the result is trivial. Now, assume that the
result holds for simple cycles of length 2
k1
. If is a simple 2
k
-cycle, then is a doubling
of a simple cycle of length 2
k1
. Therefore, if forces , then by Lemma 5.2.10 and
the inductive hypothesis, we must have either is a 2
k
-cycle or is a 2
j
-cycle for some
j k 1.
Lemma 6.2.9 ([ALM93, Lemma 2.11.6]). Let be a simple 2
k
-cycle for some k 0. If
is a nontrivial extension of , then is primary.
Proof. The restriction that is not an extension guarantees that the length of is strictly
greater than the length of . Now, from the previous lemma, Lemma 6.2.8, it follows that
48
does not force any cycle of length larger than 2
k
. By Corollary 6.1.4, we conclude that
is indeed primary.
Theorem 6.2.10 ([ALM93, Theorem 2.11.7]). A cycle is primary if and only if it is
simple.
Proof that simple implies primary. Let be a simple cycle. First, if is of odd length,
say q 3, then is a

Stefan cycle. By Proposition 6.2.5, is primary. Next, suppose
that is of length 2
k
for k 0. As remarked after the denition of primarity, if k = 0
or k = 1 then must be primary. If, however, k > 1, then since (1 2) is primary, taking
the contrapositive of Lemma 6.2.9, we see that is primary. Lastly, suppose that is of
length q 2
k
with q 3 odd. Then, combining both Lemma 6.2.9 and Proposition 6.2.5,
we get that is primary.
To prove the reverse direction that a primary cycle is simple requires a considerable
amount of work that would amount to reproducing [BC86] or several sections of [ALM93].
Therefore, we will try to give a detailed sketch of how one shows that primary implies
simple.
Let us rst show how Alseda, LLibre and Misiurewicz proved the result in [ALM93].
First, they show that depending on the length of a cycle, a primary cycle is either a

Stefan
cycle or has a block structure over (1 2).
Proposition 6.2.11 ([ALM93, Proposition 2.11.8]). Let be a nontrivial primary cycle.
Then the following statements hold.
(a) If [[ is even, then has a block structure over (1 2).
(b) If [[ is odd, then is a

Stefan cycle.
The proof of this proposition requires the result that given any n-cycle , there exists
an interval map f such that f has an n-periodic point x, f is O(x)-monotone and O(x)
is the unique orbit of f that exhibits [ALM93, Proposition 2.5.4]. This can be proven
by using the pouring water construction [ALM93, p. 54]. Observe that this result is
almost guaranteed by the doubling theorem (Theorem 5.1.6) which says that so long as
is not a doubling, P = 1, 2, . . . , n is the unique invariant set of f

that exhibits .
Now, as an immediate consequence of Proposition 6.2.11 we have that for nontrivial
odd-lengthed cycles, primary implies simple. This is because given an odd-length primary
cycle , Proposition 6.2.11 implies is a

Stefan cycle, and by the denition of a simple cycle,
a

Stefan cycle is simple. What remains to be shown is that any even-lengthed primary
cycle (which by Proposition 6.2.11 has a block structure over (1 2)) is in fact a simple cycle.
To do so, requires very careful analysis of these block structures. Ultimately, one shows
that for a primary even-lengthed cycle , the rst-return patterns of the blocks are

Stefan
cycles of the same length and so is a

Stefan extension of some cycle [ALM93, Lemmas
2.11.9 - 2.11.16]. Moreover, they show that if is a primary n 2
k
-cycle, then has a block
structure over a simple 2
k
-cycle [ALM93, Proposition 2.11.17].
Block and Coppel [BC86] and Baldwin [Bal87, Theorem 4.9] (which heavily uses Block
and Coppels paper) proved the result that primary implies simple in a slightly dierent
way. Block and Coppel prove the following powerful theorem which essentially completes
the proof:
Theorem 6.2.12 ([BC86, Theorem 1]). If f has an orbit of period n, then it has a
(strongly) simple orbit of period n.
49
Now, suppose that is a primary cycle. By Theorem 6.2.12, f

exhibits a simple cycle,


say , of period n and so by the Forcing Theorem, . But, since is primary, i.e.
forcing-minimal, it forces no other distinct n-cycles and so we must have = ; that is
is simple. Thus, primary implies simple.
We end this chapter with a diagram of all primary cycles of length 8 (Figure 6.4).
Again, recall that forces in this diagram if and only if is above and there is a line
segment from to . For example, in Figure 6.4 we see that (1 2 3) forces (1 3 4 2 5), but
(1 2 3) does not force (1 7 4 3 5 2 6) . To nd all primary cycles of length less than or equal
to 8, we modied the algorithm from Chapter 3 as illustrated by the following pseudocode:
primaryCycles := [];
for k in {3,...,8} do
if k odd then
Append(primaryCycles,stefanCycles(k));
else
G:=cyclicPermutations(SymmetricGroup(k));
for g in G do
if g has a 2-block structure then
if the set of k-cycles forced by g is empty then
Append(primaryCycles,g);
return primaryCycles;
Earlier, it was stated that there are 2
2
n
(n+1)
cycles of length 2
n
that have a 2-block struc-
ture. Thus, this algorithm clearly becomes inecient for k even, k 10, but nevertheless
does the job for k 8.
50
(1 2)
(1 3 2 4) (1 4 2 3)
(1 4 2 5 3 6) (1 6 3 5 2 4)
(1 2 3) (1 3 2)
(1 3 4 2 5) (1 5 3 2 4)
(1 4 5 3 6 2 7) (1 7 4 3 5 2 6)
(1 4 3 5 2 6) (1 6 3 4 2 5)
(1 5 2 4 3 6) (1 6 2 5 3 4)
(1 4 3 6 2 5) (1 5 2 6 3 4)
(1 5 3 8 2 6 4 7) (1 7 3 5 2 8 4 6)
(1 6 3 8 2 5 4 7) (1 7 4 5 2 8 3 6)
(1 6 4 7 2 5 3 8) (1 7 4 6 2 8 3 5)
(1 5 4 7 2 6 3 8) (1 8 4 5 2 7 3 6)
(1 5 4 8 2 6 3 7) (1 7 3 6 2 8 4 5)
(1 6 3 7 2 5 4 8) (1 8 3 6 2 7 4 5)
(1 5 3 7 2 6 4 8) (1 8 3 5 2 7 4 6)
(1 6 4 8 2 5 3 7) (1 8 4 6 2 7 3 5)
Figure 6.4: The primary cycles of length 8.
51
Chapter 7
A Forcing Proof of
Sharkovskys Theorem
We will now prove Sharkovskys theorem (Theorem 1.3.1) by employing the various results
about the local, global and minimality properties of forcing elucidated in earlier chapters.
The forcing proof of Sharkovskys theorem that we will present rst appeared in [ALM89],
but the form which we give more closely follows [ALM93, Second proof of Theorem 2.1.5].
Proof of Sharkovskys theorem. We aim to show that if an interval map f has an n-periodic
point and m
s
n then f has periodic point of period m. If n = 1 then there is nothing to
prove. So, let us assume that that n 2. Let P denote the orbit of our n-periodic point,
and let be the n-cycle that P exhibits. The rst key point to note is that Proposition
6.1.3 allows us to assume that is primary because there exists a primary n-cycle that
forces. Moreover, in view of Theorem 6.2.10, we may also assume that is simple.
Now, suppose that n is odd. Since every odd-lengthed simple cycle is a

Stefan cycle,
is a

Stefan cycle. But, since is a

Stefan cycle, the result follows from Proposition 6.2.4
which shows that the Markov graph of has a nonrepetitive loop of length m for every
m
s
n.
Next, suppose that n is a power of 2. Then, the statement follows from the denition
of a simple cycle and the inductive use of Corollary 5.2.4. Therefore, we are left with
the case n = q 2
k
with k > 0 and q 3 odd. In this case, m
s
n implies that either
m = r 2
k
with r
s
q or m = 2
i
and i < k.
Since is simple, by Dention 6.2.6, it is a -extension of a simple cycle of period
2
k
where is one of the two

Stefan cycles of length q. (We will see that the choice of
does not matter). Let us rst assume that m = r 2
k
and r
s
q. By Lemma 6.2.4 and
Proposition 6.1.3, forces some cycle of period r. Then by Lemma 5.2.11, forces some
-extension of . This -extension of has period m and hence f has a cycle of period
m. Assume now that m = 2
i
with i < k. By Corollary 5.2.4, we get . Again by the
inductive use of Corollary 5.2.4 we get that forces some cycle of period m. Then
and f has a cycle of period m. This completes the proof.
52
Chapter 8
Conclusions
In this chapter, we reect on dierent ways in which the theory presented can be gener-
alized. To begin with, we have chosen to develop the theory of forcing from the point of
view of oriented cycles; that is we consider the cycles and

as distinct cycles (so long
as ,=

). However, as many of the proofs intimated, the theory of forcing can also be
developed by considering two n-cycles and equivalent if there exists a homeomorphism
h : [1, n] [1, n] such that f

h = h f

. This is an equivalence relation, call it and


the elements of C/ are called patterns. Note that

. In fact, many of the theorems
we have proven hold for patterns as well. For example, Proposition 4.1.4 is now trivial.
Some results like Proposition 5.2.8 are more easily stated in terms of patterns: if we extend
a cycle by , then the rst-return pattern of any block is the pattern of . In the proof of
Proposition 5.2.11, two dierent cases were required, one for when the rst-return pattern
was and one when the rst-return pattern was

. This would have been taken care of
in one case by considering only the pattern of . Essentially, considering only patterns
amounts to cutting Figure 3.2 in half and only considering one side.
In Chapter 2, the combinatorial data that we considered was the way in which the
points of a periodic orbit were permuted. This restricted our study of forcing to cyclic
permutations only. However, in [MN91], Misiurewicz and Nitecki consider general nite
invariant sets which allows the study of permutations as well as cycles. They prove a
combinatorial shadowing theorem, that describes the forcing of cycles which behave
as extensions of a given one for suciently long time [ALM93]. This allows much more
generality than the results that we have presented.
While this paper has addressed specically the case of a forcing relation for one-
dimensional maps, much more recent work has focused on forcing relations in higher
dimensional dynamical systems. Forcing relations provide information about when the
presence of certain dynamical features, such as the existence of periodic orbits of a particu-
lar type, imply the presence of other dynamical features [dCH89]. In the one-dimensional
case, we used the period and the induced permutation as specications of a periodic orbit:
what will take their place in two dimensions? A moments thought is enough to convince us
that the period alone is not a useful specication: given any set of positive integers which
includes 1, we can construct a dieomorphism f : D
2
D
2
having precisely this collection
of periods for its orbits [Hal90]. The two-dimensional analogue of the Sharkovsky order
on periods for maps of the interval restricts to a partial order on essential pseudo-Anosov
conjugacy classes in the mapping class group of the n-times punctured disk [Han97].
For more information on this theory, the reader is directed to [Hal91], [dCH89], [Hal90]
53
and [Han97].
54
Bibliography
[ALM89] Llus Alsed`a, Jaume Llibre, and Micha l Misiurewicz, Periodic orbits of maps
of Y , Trans. Amer. Math. Soc. 313 (1989), no. 2, 475538. MR MR958882
(90c:58145)
[ALM93] , Combinatorial dynamics and entropy in dimension one, Advanced
Series in Nonlinear Dynamics, vol. 5, World Scientic Publishing Co. Inc.,
River Edge, NJ, 1993. MR MR1255515 (95j:58042)
[Bal87] Stewart Baldwin, Generalizations of a theorem of Sarkovskii on orbits of con-
tinuous real-valued functions, Discrete Math. 67 (1987), no. 2, 111127. MR
MR913178 (89c:58057)
[BB04] Karen M. Brucks and Henk Bruin, Topics from one-dimensional dynamics,
London Mathematical Society Student Texts, vol. 62, Cambridge University
Press, Cambridge, 2004. MR MR2080037 (2005j:37048)
[BC86] L. S. Block and W. A. Coppel, Stratication of continuous maps of an in-
terval, Trans. Amer. Math. Soc. 297 (1986), no. 2, 587604. MR MR854086
(88a:58164)
[BC92a] Chris Bernhardt and Ethan M. Coven, A polynomial-time algorithm for de-
ciding the forcing relation on cyclic permutations, Symbolic dynamics and its
applications (New Haven, CT, 1991), Contemp. Math., vol. 135, Amer. Math.
Soc., Providence, RI, 1992, pp. 8593. MR MR1185081 (93j:58041)
[BC92b] L. S. Block and W. A. Coppel, Dynamics in one dimension, Lecture Notes
in Mathematics, vol. 1513, Springer-Verlag, Berlin, 1992. MR MR1176513
(93g:58091)
[BCMM92] C. Bernhardt, E. Coven, M. Misiurewicz, and I. Mulvey, Comparing periodic
orbits of maps of the interval, Trans. Amer. Math. Soc. 333 (1992), no. 2,
701707. MR MR1079051 (92m:58106)
[Ber84] Chris Bernhardt, Simple permutations with order a power of two, Ergodic
Theory Dynam. Systems 4 (1984), no. 2, 179186. MR MR766099 (86d:58092)
[Ber87] , The ordering on permutations induced by continuous maps of the
real line, Ergodic Theory Dynam. Systems 7 (1987), no. 2, 155160. MR
MR896787 (88h:58099)
[BGMY80] Louis Block, John Guckenheimer, Micha l Misiurewicz, and Lai Sang Young,
Periodic points and topological entropy of one-dimensional maps, Global
55
theory of dynamical systems (Proc. Internat. Conf., Northwestern Univ.,
Evanston, Ill., 1979), Lecture Notes in Math., vol. 819, Springer, Berlin, 1980,
pp. 1834. MR MR591173 (82j:58097)
[BH] Keith Burns and Boris Hasselblatt, The sharkovsky theorem: A natural direct
proof, Amer. Math. Monthly ?? (??), no. ?, ?
[BS02] Michael Brin and Garrett Stuck, Introduction to dynamical systems, Cam-
bridge University Press, Cambridge, 2002. MR MR1963683 (2003m:37001)
[CE80] Pierre Collet and Jean-Pierre Eckmann, Iterated maps on the interval as dy-
namical systems, Progress in Physics, vol. 1, Birkhauser Boston, Mass., 1980.
MR MR613981 (82j:58078)
[dCH89] Andre de Carvalho and Toby Hall, Decoration invariants for horseshoe braids,
arXiv:0808.3078v1 (1989), 142.
[dMvS93] Welington de Melo and Sebastian van Strien, One-dimensional dynamics,
Ergebnisse der Mathematik und ihrer Grenzgebiete (3) [Results in Mathe-
matics and Related Areas (3)], vol. 25, Springer-Verlag, Berlin, 1993. MR
MR1239171 (95a:58035)
[Hal90] Toby D. H. Hall, Coexistence theorems for periodic orbits of horseshoe type,
An essay submitted in competition for the Smith prise in mathematics (1990),
146.
[Hal91] Toby Hall, Unremovable periodic orbits of homeomorphisms, Math. Proc.
Cambridge Philos. Soc. 110 (1991), no. 3, 523531. MR MR1120486
(92i:58146)
[Han97] Michael Handel, The forcing partial order on the three times punctured disk,
Ergodic Theory Dynam. Systems 17 (1997), no. 3, 593610. MR MR1452182
(98i:57026)
[Jun91] Irwin Jungreis, Some results on the

Sarkovski partial ordering of permuta-
tions, Trans. Amer. Math. Soc. 325 (1991), no. 1, 319344. MR MR998354
(91h:58060)
[LY75] Tien Yien Li and James A. Yorke, Period three implies chaos, Amer. Math.
Monthly 82 (1975), no. 10, 985992. MR MR0385028 (52 #5898)
[Mis97] Micha l Misiurewicz, Remarks on Sharkovskys theorem, Amer. Math. Monthly
104 (1997), no. 9, 846847. MR MR1479989 (98f:58138)
[MN91] Michal Misiurewicz and Zbigniew Nitecki, Combinatorial patterns for maps
of the interval, Mem. Amer. Math. Soc. 94 (1991), no. 456, vi+112. MR
MR1086562 (92h:58105)
[MSS73] N. Metropolis, M. L. Stein, and P. R. Stein, On nite limit sets for trans-
formations on the unit interval, J. Combinatorial Theory Ser. A 15 (1973),
2544. MR MR0316636 (47 #5183)
[MT88] John Milnor and William Thurston, On iterated maps of the interval, Dynam-
ical systems (College Park, MD, 198687), Lecture Notes in Math., vol. 1342,
Springer, Berlin, 1988, pp. 465563. MR MR970571 (90a:58083)
56
[Nit82] Zbigniew Nitecki, Topological dynamics on the interval, Ergodic theory and
dynamical systems, II (College Park, Md., 1979/1980), Progr. Math., vol. 21,
Birkhauser Boston, Mass., 1982, pp. 173. MR MR670074 (84g:54051)
[Sha64] A. N. Sharkovsky, Co-existence of the cycles of a continuous mapping of the
line into itself, Ukrain. Math. Zh. 16 (1964), no. 1, 6171.
[Sha65] , On cycles and structure of a continuous map, Ukrain. Math. Zh. 17
(1965), no. 3, 104111.
[

Ste77] P.

Stefan, A theorem of

Sarkovskii on the existence of periodic orbits of con-
tinuous endomorphisms of the real line, Comm. Math. Phys. 54 (1977), no. 3,
237248. MR MR0445556 (56 #3894)
[WR87] A. Weiss and T. D. Rogers, The number of orientation reversing cycles in
the quadratic map, Oscillations, bifurcation and chaos (Toronto, Ont., 1986),
CMS Conf. Proc., vol. 8, Amer. Math. Soc., Providence, RI, 1987, pp. 703
711. MR MR909946 (89b:58174)
57

Das könnte Ihnen auch gefallen