Elementary Fixed Point Theorems. (PDFDrive)
Elementary Fixed Point Theorems. (PDFDrive)
Elementary Fixed Point Theorems. (PDFDrive)
P. V. Subrahmanyam
Elementary
Fixed Point
Theorems
Forum for Interdisciplinary Mathematics
Editor-in-chief
P. V. Subrahmanyam, Indian Institute of Technology Madras, Chennai, India
Editorial Board
Yogendra Prasad Chaubey, Concordia University, Montreal (Québec), Canada
Jorge Cuellar, Principal Researcher, Siemens, Germany
Janusz Matkowski, University of Zielona Góra, Poland
Thiruvenkatachari Parthasarathy, Chennai Mathematical Institute, Kelambakkam, India
Mathieu Dutour Sikirić, Rudjer Boúsković Institute, Zagreb, Croatia
Bhu Dev Sharma, Jaypee Institute of Information Technology, Noida, India
The Forum for Interdisciplinary Mathematics series publishes high-quality
monographs and lecture notes in mathematics and interdisciplinary areas where
mathematics has a fundamental role, such as statistics, operations research,
computer science, financial mathematics, industrial mathematics, and
bio-mathematics. It reflects the increasing demand of researchers working at the
interface between mathematics and other scientific disciplines.
123
P. V. Subrahmanyam (emeritus)
Department of Mathematics
Indian Institute of Technology Madras
Chennai, Tamil Nadu, India
This Springer imprint is published by the registered company Springer Nature Singapore Pte Ltd.
The registered company address is: 152 Beach Road, #21-01/04 Gateway East, Singapore 189721,
Singapore
Preface
v
vi Preface
Dr. S. Alamelu Bai for proofreading the content and for the moral support. I am
thankful to Dr. N. Sivakumar (Texas A&M University) and Dr. Antony Vijesh
(IIT Indore) for getting several papers for reference. I am grateful to Profs. M. S.
Rangachari, G. Rangan, (Late) K. S. Padmanabhan and (Late) K. N. Venkataraman
of the Madras University, Chennai, for initiating me into different aspects of mathe-
matical analysis. Thanks are due to Mr. E. Boopal for typesetting the manuscript.
Suggestions for improvement and corrections of errors and misprints are earn-
estly solicited.
1 Prerequisites . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.1 Topological Spaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Metric Spaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.3 Normed Linear Spaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
1.4 Topological Vector Spaces . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
2 Fixed Points of Some Real and Complex Functions . . . . . . . ..... 23
2.1 Fixed Points of Continuous Maps on Compact
Intervals of R . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
2.2 Iterates of Real Functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
2.3 Periodic Points of Continuous Real Functions . . . . . . . . . . . . . 31
2.4 Common Fixed Points, Commutativity and Iterates . . . . . . . . . . 35
2.5 Common Fixed Points and Full Functions . . . . . . . . . . . . . . . . 41
2.6 Common Fixed Points of Commuting Analytic Functions . . . . . 44
2.7 Fixed Points of Meromorphic Functions . . . . . . . . . . . . . . . . . . 48
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
3 Fixed Points and Order . . . . . . . . . . . . . . . . .................. 51
3.1 Fixed Points in Linear Continua . . . . . .................. 51
3.2 Knaster–Tarski Principle . . . . . . . . . . . .................. 54
3.3 Tarski’s Lattice Theoretical Fixed Point Theorem
and Related Theorems . . . . . . . . . . . . . .................. 57
3.4 Some Applications . . . . . . . . . . . . . . . .................. 64
References . . . . . . . . . . . . . . . . . . . . . . . . . . .................. 70
4 Partially Ordered Topological Spaces and Fixed Points . . . . . . . . . 73
4.1 A Precis of Partially Ordered Topological Spaces . . . . . . . . . . . 73
4.2 Schweigert–Wallace Fixed Point Theorem . . . . . . . . . . . . . . . . 81
4.3 Set Theory, Fixed Point Theory and Order . . . . . . . . . . . . . . . . 86
4.4 Multifunctions and Dendroids . . . . . . . . . . . . . . . . . . . . . . . . . 90
ix
x Contents
xiii
Chapter 1
Prerequisites
This chapter is a precis of the basic definitions and theorems used in the sequel. It is
presumed that the reader is familiar with naive set theory (see Halmos [4]) and the
properties of real numbers and real functions (see Bartle [1]). Other mathematical
concepts and theorems relevant to specific sections of a chapter will be recalled
therein.
This section collects important concepts and results from topology. For proofs and
other details, Dugundji [3], Kelley [7], Munkres [9] and Simmons [13] may be
consulted.
Definition 1.1.1 Let X be a non-empty set. A collection of J of subsets of X is
called a topology on X , if
(i) φ, X ∈ T ,
G 1 ∩ G 2 ∈ T for G 1 , G 2 ∈ T and
(ii)
(iii) G∈F G ∈ T for any F ⊆ T .
Example 1.1.2 For a non-empty set X , the family {φ, X } is a topology on X called
the indiscrete topology on X , 2 X , the power set of X or the set of all subsets of X is
a topology on X called the discrete topology on X . The family of all subsets of X
© Springer Nature Singapore Pte Ltd. 2018 1
P. V. Subrahmanyam, Elementary Fixed Point Theorems,
Forum for Interdisciplinary Mathematics,
https://doi.org/10.1007/978-981-13-3158-9_1
2 1 Prerequisites
whose complements are finite sets together with the empty set is also a topology on
X called the co-finite topology on X .
Since the intersection of any collection of topologies on X is a topology on X ,
for any family F of subsets of X , there is the smallest topology on X containing F,
called the topology generated by F.
Example 1.1.3 A subset G of real numbers is called open if for each x ∈ G, an open
interval containing x lies in G. (Evidently the empty set is open.) This collection of
all open subsets of R, the real number system is a topology on R, called the usual
topology on R.
Definition 1.1.4 Let (X, T ) be a topological space. A neighbourhood of a point
x ∈ X is any subset of X containing an open subset G ∈ T , containing x. A neigh-
bourhood base or local base at x is a family Nx of neighbourhoods of x such that for
any neighbourhood N of x, there is a neighbourhood N x ∈ Nx such that x ∈ N x ⊆ N .
A topological space is called first countable if for each point there is a countable local
base. An interior point of A is a point a ∈ A such that A contains a neighbourhood
of a.
Definition 1.1.5 A subset F of a topological space (X, T ) is called a closed subset
of X if X − F is T -open. The closure of a subset A of X denoted by A is the
smallest closed set containing A. A subset S of X is said to be dense in X if S = X .
A topological space (X, T ) is called separable if it has a countable dense subset.
Remark 1.1.6 Let (X, T ) be a topological space and A, B ⊆ X . Then
(i) φ0 = φ, φ = φ, X 0 = X and X = X ;
(ii) A ⊇ A and A0 ⊆ A;
(iii) A ∪ B = A ∪ B, (A ∩ B)◦ = A◦ ∩ B ◦ ;
(iv) (A) = A and (A◦ )◦ = A◦ . Further A = {x ∈ X : every neighbourhood of x has
a non-void intersection with A}. A0 = {a ∈ A : a is an interior point of A}.
Definition 1.1.7 For a topological space (X, T ) B ⊆ T is called a base (or basis)
for T is for A1 , A2 ∈ B and x ∈ A1 ∩ A2 , there exists A3 ∈ B such that x ∈ A3 ⊆
A1 ∩ A2 . A subfamily S of T is called a subbase for T of B, if the family of
intersections of all finite subfamilies of S is a base for T . If the topology T has a
countable base, then the topological space is called second countable.
Remark 1.1.8 If S is a family of subsets of X with ∪{S : S ∈ S } = X , then S is
a subbase for a topology on X , for which B the family of subsets of X which are the
intersections of finite subfamilies of S is a base for this topology.
Remark 1.1.9 The family of all subintervals of the form [a, b), a < b, a, b ∈ R is
a base for a topology on R, called the lower limit topology on R. Similarly, the
family {(a, b] : a < b, a, b ∈ R} is a base for a topology on R called the upper limit
topology on R. The usual (standard) topology on R has the family of all open intervals
(a, b), a < b, a, b ∈ R as a base. R with the usual topology is separable and second
countable. However, R with the lower limit topology is separable and first countable
but is not second countable.
1.1 Topological Spaces 3
Definition 1.1.11 A partial order ≤ on a set X is called a linear order or total order
if for any pair of elements x, y ∈ X either x ≤ y or y ≤ x. A linearly ordered set is
also called a chain.
Definition 1.1.12 A partially ordered set (D, ≤) is called a directed set if for any
pair x, y ∈ X , there exists z ∈ D such that x ≤ z and y ≤ z.
In this section, basic concepts and theorems from the theory of metric spaces are
recalled. For details, in addition to the references cited in Sect. 1.1, Kaplansky [6]
may be consulted.
Definition 1.2.1 Let X be a non-void set. A map d : X × X → [0, ∞) (=R+ ) is
called a metric if
6 1 Prerequisites
Remark 1.2.11 In example 1.2.4, except the space described in (vi), the metric spaces
in examples (i)–(v) are complete.
Theorem 1.2.12 If (X, d) is a metric space, then d1 (x, y) = min{1, d(x, y)}, x, y ∈
X defines a metric on X and the topologies induced on X by these metrices are the
same.
Theorem 1.2.13 If (X n , dn ), x ∈ N is a sequence of metric spaces, then X = Xn
n∈N
is a metric space under the metric d defined by
∞
1 dn (xn , yn )
d(x, y) = ,
n=1
2 n 1 + dn (xn , yn )
Theorem 1.2.20 A regular T1 space is paracompact if and only if for each open
covering of X , there is a partition of unity subordinate to this covering. For every
compact T2 space, every open cover has a partition of unity subordinate to it.
Theorem 1.2.22 For a metric space (X, d), the following are equivalent:
(i) X is compact;
(ii) X is complete and totally bounded;
(iii) every sequence in X has a convergent subsequence;
(iv) X has the Bolzano–Weierstrass property, viz. for every infinite subset A of X
has a limit point x0 ∈ X , i.e. a point x0 such that every neighbourhood of x0
meets A.
Definition 1.2.24 For a non-void subset A of a metric space (X, d), the distance of
a point x from A is defined as D(x, A) = in f {d(x, a) : a ∈ A}.
Theorem 1.2.25 Let A be a non-void subset of a metric space (X, d). Then A =
{x ∈ X : D(x, A) = 0}. Further, |D(x, A) − D(y, A)| ≤ d(x, y) for any x, y ∈ X
and the map x → D(x, A) is a continuous map of X into R+ .
1.3 Normed Linear Spaces 9
Normed linear spaces, constituting the base of Functional Analysis are metric spaces
with a richer (algebraic) structure. They provide a natural setting for mathematical
modelling of many natural phenomena. Bollabos [2], Kantorovitch and Akhilov [5],
Lyusternik and Sobolev [8], Rudin [11, 12], Simmons [13] and Taylor [14] may
be consulted for a detailed exposition of the following concepts and theorems. It is
assumed that the reader is familiar with the concepts of groups, rings and fields.
Definition 1.3.1 A linear space or vector space over a field F is a triple (V, +, ·),
where + is a binary operation (called vector addition or simply addition) and · is a
mapping from F × V into V (called scalar multiplication) satisfying the following
conditions;
(i) (V, +) is a commutative group with θ (called zero vector) as its identity element;
(ii) for all λ ∈ F, x, y ∈ V λ.(x + y) = λx + λy;
(iii) for all λ, μ ∈ F and x ∈ V (λ + μ).x = λ.x + μ.x and λ · (μ · x) = (λ · μ) · x
(where λμ is the product of λ and μ under the multiplication in the field F);
(iv) 0 · x = θ, 1 · x = x for all x ∈ V , where 0 is the additive identity and 1 the
multiplicative identity of the field F.
Often 0 is also used to represent the zero vector and the context will clarify this
without much difficulty. If F is the field of real (complex) numbers then V is called
a real (complex) vector space. In what follows, we will be concerned only with real
or complex vector spaces. Also a linear subspace V1 of V is a subset of V which is
a linear space over F with vector addition and scalar multiplication of V restricted
to V1 .
Definition 1.3.2 A subset S of a linear space V over a field F is said to be linearly
n
independent if for every finite subset {s1 , . . . , sn } of S, αi si = θ implies αi = 0
i=1
n
for i = 1, 2, . . . , n where αi ∈ F. An element of the form αi si where αi ∈ F
i=1
and si ∈ S is called a finite linear combination of members of S.
Definition 1.3.3 A subset S of a linear space V over a field F is said to span V if
every element of V can be written as a finite linear combination of elements from S.
A maximal linearly independent subset of a linear space V over F is called a basis
for V .
Any two bases of a vector space have the same cardinality.
Definition 1.3.4 The cardinality of a basis of a linear space V is called the dimension
of the linear space. If a linear space has a finite dimension, then it is called a finite-
dimensional vector space. Otherwise the linear space is infinite-dimensional.
Definition 1.3.5 Let (X, +, ·) be a linear space over F = R or C. A map · :
X → R+ is called a norm if the following conditions are satisfied:
10 1 Prerequisites
Remark 1.3.8 The linear spaces in (ii)–(v) of Example 1.2.4 are Banach spaces
with the norms defined by x = d(x, θ) where d is the metric described in the
corresponding case, while (vi) of Example 1.2.4 is a normed linear space under the
b
norm a | f (t)|dt. However, this is not a Banach space.
Remark 1.3.11 A linear operator satisfying (iii) of Theorem 1.3.10 is called bounded.
In view of the theorem above, bound linear operators are precisely continuous linear
operators.
If T : X 1 → X 2 is a continuous linear operator where (X i , · i ), i = 1, 2 are
normed linear spaces, then
is finite and is called the norm of the linear operator and is denoted by T .
Theorem 1.3.29 Every incomplete inner product space can be isometrically embed-
ded as a dense subspace of a Hilbert space.
Theorem 1.3.30 A normed linear space (X, ) is an inner product space if and
only if the following parallelogram law is valid:
Example 1.3.38 (i) L 2 [0, 2π], the space of complex-valued Lebesgue measur-
able functions f on [0, 2π] which are square-integrable in the sense that
2π
| f (x)|2 d x < +∞ is a Hilbert space under the inner product < f, g >=
02π inx
0 f (x)g(x)d x. The set √e 2π : n = 0, ±1, ±2, . . . is an orthonormal basis.
14 1 Prerequisites
∞
(ii) L 2 (R), the space of all Lebegue-measurable functions for which −∞ | f 2 (x)|d x
∞
is finite, is also a Hilbert space under the inner product < f, g >= −∞ f (x)
x2
g(x)d x. {x n e− 2 , n = 0, 1, 2, . . . } gives rise to an orthonormal basis for L 2 (R)
via the Gram–Schmidt orthogonalization process (see Simmons [13]).
Among normed linear spaces, strictly convex spaces and the more specialized
uniformly convex spaces resemble the Euclidean spaces geometrically.
Remark 1.3.41 The above definition is equivalent to the requirement that for
xn , yn ≤ 1 and xn + yn → 2, xn − yn → 0 as n → ∞.
Clearly, every Hilbert space is uniformly convex. Also L p [0, 1] for p ≥ 2 is
uniformly convex. While every uniformly convex space is strictly convex, C[0, 1] is
not even strictly convex.
Theorem 1.3.45 If T ∈ B(H ), the space of all bounded linear operators mapping
H into itself, then T ∗ the adjoint of T is uniquely defined. Further,
(i) (T1 + T2 )∗ = T1∗ + T2∗ ,
(ii) (αT )∗ = αT ∗ ,
(iii) (T1 T2 )∗ = T2∗ T1∗ ,
(iv) (T ∗ )∗ = T ∗ ,
(v) T ∗ 2 = T 2 = T ∗ T .
1.3 Normed Linear Spaces 15
Definition 1.3.46 A linear operator T ∈ B(H ), the space of all bounded linear oper-
ators on a Hilbert space H is said to be self-adjoint if T = T ∗ .
Definition 1.3.47 For T ∈ B(H ), the spectrum of T is the set {λ ∈ C : T − λI is
not invertible}, I being the identity operator. An eigenvalue of T is a number λ ∈ C
such that there exists a nonzero vector x0 ∈ H with T x0 = λx0 and in this case x0 is
called an eigenvector (corresponding to the eigenvalue λ).
Theorem 1.3.48 For T ∈ B(H ), the space of all bounded linear operators on a
Hilbert space H is self-adjoint if and only if < T x, x > is real for all x ∈ H . So
the eigenvalues of a self-adjoint operator are real. Further σ(T ), the spectrum of
T lies in [m, M], where m = inf{< T x, x >: x ∈ H and x = 1} and M = sup{<
T x, x >: x ∈ H and x = 1}. Also, m, M ∈ σ(T ).
Definition 1.3.49 A linear operator P in B(H ) is called a projection if P is self-
adjoint and P 2 = P.
Remark 1.3.50 If P is a projection on a Hilbert space H , then P = M ⊕ M ⊥ where
M = {P x : x ∈ H }, the range of P and M ⊥ , the range of I –P. Further, every rep-
resentation of H as the orthogonal sum M + M ⊥ defines a unique projection of H
onto M.
Theorem 1.3.51 For any self-adjoint operators T in B(H ), there is a family {Pλ :
λ ∈ R} of projections on H satisfying the following conditions:
(i) if T C = C T for C ∈ B(H ), then Pλ C = C Pλ for all λ ∈ R;
(ii) Pλ Pμ = Pλ , if λ < μ;
(iii) Pλ−0 = lim Pμ = Pλ (i.e. Pλ is continuous from the left with respect to λ);
μ→λ−0
(iv) Pλ = 0 if λ ≤ m and Pλ = I for λ > M.
(Such a family of projections Pλ is called a resolution of identity generated by
T ).
Theorem 1.3.52 (Spectral theorem) For every self-adjoint operator T ∈ B(H ) and
any > 0,
M+
T = λd Pλ
m
where the Stieltjes integral is the limit of (appropriate) integral sums in the operator-
norm topology.
Definition 1.3.53 A linear operator T : N1 → N2 where N1 and N2 are normed
linear spaces is said to be a compact operator if T (U ) is compact in N2 for each
bounded subset U of N1 .
Theorem 1.3.54 Let T : B → B be a compact linear operator on a Banach space
B. σ(T ), the spectrum of T is finite or countably infinite and is contained in
[−T , T ]. Every nonzero number in σ(T ) is an eigenvalue of T . If σ(T ) is
countably infinite, then 0 is the only limit point of σ(T ).
16 1 Prerequisites
It is convenient to recall the definition of a topological group and list some of its
properties (see Kelley [7], Rudin [11] and Royden [10]).
Definition 1.4.1 Let (G, ·) be a group with the identity element e and for each
x ∈ G, x −1 denote the inverse of x (with respect to the binary operation ·). The triple
(G, ·, T ) is called a topological group where T is a topology on the group G with
the binary operation · such that the map (x, y) → x y −1 mapping G × G into G is
continuous. (Here G × G carries the product topology.)
If (G, ·) is a group and A, B ⊆ G, we write A · B = {a · b : a ∈ A, b ∈ B}.
Theorem 1.4.2 Let (G, ·, T ) be a topological group with the identity e. Then
(i) the map x → x −1 mapping G into G and the map (x, y) → x y mapping G × G
into G are continuous. Conversely if T1 is a topology on a group (G, ·) such
that x → x −1 and (x, y) → x y are continuous on G with the topology T1 , then
(G, ·, T1 ) is a topological group.
(ii) the inversion map i, defined by i(x) = x −1 is a homeomorphism of G onto
G; for each a ∈ G, L a (Ra ) called the left (right) translation by a, defined by
L a (x) = ax (Ra (x) = xa) are homeomorphisms;
(iii) a subset S of G is open if and only if for each x ∈ S, x −1 S (or equivalently
Sx −1 ) is a neighbourhood of e;
(iv) the family N of all neighbourhoods of e has the following properties:
(iv-a) for U, V ∈ N , U ∩ V ∈ N ;
(iv-b) for U ∈ N , V.V −1 ⊆ U for some V ∈ N ;
(iv-c) for U ∈ N and x ∈ G, x.U.x −1 ∈ N ;
(v) the closure of a (normal) subgroup of G is a (normal) subgroup of G;
(vi) every subgroup G 1 of G with an interior point is both open and closed and G 1
is closed or G 1 − G 1 is dense in G 1 ;
(vii) G is Hausdorff if it is a T0 space in the sense that for every pair of distinct
points, there is a point for which some neighbourhood does not contain the
other point.
A topological vector space can be defined in analogy with a topological group.
Definition 1.4.3 The quadruple (X, +, ·, T ) where (X, +, ·) is a vector space over
F = R or C and T is a topology on X is called a topological vector space (linear
topological space) if the following assumptions are satisfied:
(i) (X, T ) is a T1 -space;
(ii) the function (x, y) → x + y mapping X × X into X is continuous and
(iii) the function (α, x) → α.x mapping F × X into X is continuous.
Often, we simply say that X is a topological vector space (or t.v.s for short) when
the topology T on X and the vector space operations are clear from the context.
1.4 Topological Vector Spaces 17
Theorem 1.4.5 Let X be a t.v.s. For each a ∈ X and λ = 0 ∈ F define the transla-
tion operator Ta and the multiplication operator Mλ by the rules Ta (x) = x + a and
Mλ (x) = λ.x respectively for each x ∈ X . Then, Ta and Mλ are homeomorphism of
X onto X .
Further G ⊆ X is open if and only if Ta (G) is open for each a ∈ X . So the local
base at 0 completely determines the local base at any x ∈ X and hence the topology
on X .
Remark 1.4.6 Every normed linear space is a t.v.s.
Definition 1.4.7 A function p mapping a vector space X over F(=R or C) into F
is called a seminorm if
(i) p(x + y) ≤ p(x) + p(y) for all x, y ∈ X and
(ii) p(αx) = |α| p(x) for all x ∈ X and all α ∈ F.
A seminorm is a norm if p(x) = 0 for x = θ. A family P of seminorms is sepa-
rating if for each x = y, there is a seminorm p ∈ P with p(x − y) = 0.
Theorem 1.4.8 If P is a separating family of seminorms on a vector space V , then
V ( p, n) = {x ∈ X : p(x) < n1 }, p ∈ P is a local base of convex sets for a topology
T on X . Thus, (X, T ) is locally convex and each p is continuous. Also, E is bounded
if and only if p(E) is bounded for each p ∈ P.
Definition 1.4.9 For an absorbing subset A of a t.v.s. X , the map μ A : X → R
defined by μ A (x) = inf{t > 0 : t −1 x ∈ A} is called the Minkowski functional of A.
Listed below are some of the basic properties and features of a topological vector
space.
Theorem 1.4.10 Let X be a topological vector space
(i) if S ⊆ X , S = ∩{S + V : V is a neighbourhood of 0};
(ii) if S1 , S2 ⊆ X , S1 + S2 ⊆ S1 + S2 ;
(iii) if C ⊆ X is convex, so are C 0 and C;
(iv) if B ⊆ X is balanced, so is B and if in addition 0 ∈ B 0 , B 0 is balanced;
(v) the closure of a bounded set is also bounded;
(vi) every neighbourhood of 0 also contains a balanced neighbourhood of 0 and
so X has a balanced local base;
(vii) every convex neighbourhood of 0 contains a balanced convex neighbourhood
of 0;
∞
(viii) if V is a neighbourhood of 0 and rn ↑ +∞ where r1 > 0, X = rn V ;
n=1
18 1 Prerequisites
Theorem 1.4.12 If B is a local base for a t.v.s. (X, J ) comprising convex balanced
neighbourhood, then {μV : V ∈ B} is a family of continuous seminorms that are
separating (i.e. for x, y ∈ X , then there is a μV such that μV (x) = μV (y)). Further,
the topology having a local base generated by these seminorms of the form {x :
μV (x) < n1 }, V ∈ B, n ∈ N coincides with the topology on X .
Definition 1.4.13 A t.v.s is said to be locally convex if it has a local base of convex
sets. It is called an F-space if the topology is generated by complete translation-
invariant metric. A locally convex F-space is called a Frechet space.
is translation invariant.
Example 1.4.18 For 0 < p < 1, let L p [0, 1] be the linear space of all Lebegue-
1
measurable functions f on [0, 1] for which δ( f ) = 0 | f (a)| p d x < +∞. Then d,
defined by d( f, g) = δ( f − g) defines a translation-invariant metric on L p [0, 1] and
this metric is complete. Thus L p [0, 1] is an F-space. However, it is not locally
convex. Indeed L p [0, 1] is the only open convex set. So, 0 is the only continuous
linear functional on L p [0, 1] for 0 < p < 1 (See Rudin [12]).
(iii) μ(E) = sup{μ(K ) : K ⊆ E and K compact} is true for each open set E and
for any E ∈ S with μ(E) < +∞;
(iv) for E ∈ S with μ(E) = 0, A ∈ S for any A ⊆ E and μ(A) = 0.
When X is compact, μ can be chosen so that μ(X ) = 1, i.e. a Borel probability
measure.
Remark 1.4.24 In a Frechet space, for the convex hull H of a compact set, H is
compact and in a finite-dimensional space Rn , H itself is compact. Also if an element
x lies in the convex hull and a set E ⊆ Rn , then it lies in the convex hull of a subset
of E that contains at most n + 1 points.
Theorem 1.4.26 Let X be a t.v.s such that X ∗ separates points and μ be a Borel
probability measure on a compact Hausdorff space Q. If f : Q → X is continuous
and if the convex hull H of f (Q) has compact closure H in X , then the integral
y= f dμ
Q
Theorem 1.4.27 Let X be a t.v.s such that X ∗ separates points and Q, a compact
subset of X and H , the closed convex hull of Q be compact.
y ∈ H if and only if there is a regular Borel probability measure μ on Q such that
y= xdμ.
Q
References
1. Bartle, R.G.: The Elements of Real Analysis, 2nd edn. Wiley, New York (1976)
2. Bollobas, B.: Linear Analysis, An Introductory Course, 2nd edn. Cambridge University Press,
Cambridge (1999)
3. Dugundji, J.: Topology. Allyn and Bacon Inc., Boston (1966)
4. Halmos, P.R.: Naive Set Theory. D. Van Nostrand Co., Princeton (1960)
5. Kantorovich, L.V., Akilov, G.P.: Functional Analysis, Translated from the Russian by Howard
L. Silcock, 2nd edn. Pergamon Press, Oxford-Elmsford (1982)
6. Kaplansky, I.: Set Theory and Metric Spaces, 2nd edn. Chelsea Publishing Co., New York
(1977)
7. Kelley, J.L.: General Topology. D. Van Nostrand Company, Inc., Toronto (1955)
8. Lusternik, L.A., Sobolev, V.J.: Elements of Functional Analysis. Hindustan Publishing Corpo-
ration, Delhi (1974)
9. Munkres, J.R.: Topology: A First Course. Prentice-Hall, Inc., Englewood Cliffs (1975)
10. Royden, H.L.: Real Analysis, 3rd edn. Macmillan Publishing Company, New York (1988)
11. Rudin, W.: Functional Analysis. Tata McGraw-Hill Publishing Co. Ltd., New Delhi (1974)
12. Rudin, W.: Real and Complex Analysis, 3rd edn. McGraw-Hill Book Co., New York (1987)
13. Simmons, G.F.: Introduction to Topology and Modern Analysis. McGraw-Hill Book Co., Inc.,
New York (1963)
14. Taylor, A.E.: Introduction to Functional Analysis. Wiley, New York; Chapman Hall, Ltd.,
London (1958)
Chapter 2
Fixed Points of Some Real and Complex
Functions
This chapter highlights some fixed point theorems for certain real and complex
functions.
where ρ(x) = −1 for x < −1 and ρ(x) = 1 for x > 1 and ρ(x) = x for other real
numbers. Since g is continuous and ρ is continuous on R, clearly f is continuous and
t0 ∈ [−1, 1]. Further
maps [−1, 1] into itself. So by Theorem 2.1.5, f has a fixed point
(1−t0 )a (1+t )b
r −g + 20
t0 is neither −1 nor 1 and −1 < t0 < 1. So t0 = f (t0 ) = t0 −
2
g(b)−g(a)
.
Hence r = g (1−t20 )a + (1+t2 0 )b . In short, g has the intermediate value property.
Remark 2.1.8 Theorem 2.1.4 is not true if the interval is not compact. the map x →
x + 1 is continuous but has no fixed point in (−∞, ∞) or [0, ∞). The continuous
map x → 1+x 2
on [0, 1) has no fixed point in [0, 1). Theorem 2.1.4 fails even if f is
continuous everywhere on [a, b] except at a single point. For instance f : [0, 1] →
[0, 1] defined by
x
, x = 0
f (x) = 2
1, x = 0
Remark 2.1.9 F f , the set of fixed points of a continuous map on [a, b] is closed.
Indeed F f = {x ∈ [a, b] : f (x) = x} = g −1 (0) where g : [a, b] → R is defined by
g(x) = f (x) − x. Since {0} is a closed set and g is continuous g −1 {0} is a closed
subset. So F f is a closed subset of [a, b] ([a, b] being compact, F f is also compact).
Remark 2.1.10 Indeed we can prove that for each closed subset F of [0, 1] there
is a continuous map f : [0, 1] → [0, 1] for which F is the set of fixed points of
f . For proving this we can, without loss of generality, assume that 0, 1 ∈ F. So
[0, 1] − F = G is open and is a countable union of disjoint open intervals (ai , bi ),
i ∈ N. Now we consider the case when this collection is countably infinite, leaving
the case of finite collection as an exercise.
For n ∈ N define f n : [0, 1] → [0, 1] by
⎧ ∞
⎪
⎨x, x ∈ F ∪ i=n (a , b ),
ai +bi i i
f n (x) = ai , if x ∈ ai , 2 for i < n,
⎪
⎩ ai +bi
2x − bi , if x ∈ 2 , bi for i < n.
In this section, some theorems on the behaviour of iterates of real functions are
discussed. First, Krasnoselskii’s theorem on the convergence of special iterates of
non-expansive maps of [a, b], following Bailey’s [2] proof using elementary prop-
erties of subsequential limits is discussed in detail. Theorems 2.2.6–2.2.8 detail the
rates of convergence of iterates of special class of functions and are due to Thron
[30].
Theorem 2.2.1 (Krasnoselskii [20], Bailey [2]) Let f : I (= [a, b]) → I be a map
such that | f (x) − f (y)| ≤ |x − y| for all x, y ∈ I . For any x ∈ I , the sequence (xn )
defined recursively by xn+1 = 21 (xn + f (xn )), n = 1, 2, . . . , converges to some fixed
point of f .
26 2 Fixed Points of Some Real and Complex Functions
Proof Suppose that (xn ) does not converge to a fixed point. We show that this leads
to a contradiction. To this end, the proof is divided into several steps.
Step II. No subsequence of (xn ) converges to a fixed point of f . For, if (xni ) converges
to z and f (z) = z, then |z − xni +1 | ≤ |z − 21 (xni + f (xni )| ≤ 21 |z − xni | + 21 | f (z) −
f (xni )| (as z = 21 (z + f (z))) ≤ |z − xni | (since | f (x) − f (y)| ≤ |x − y|). This
shows that (xn ) itself converges to z, a fixed point of f , contradicting our assumption
that (xn ) does not converge to a fixed point of f .
Step III. Since (xn ) lies in the compact interval I = [a, b], it has a subsequential limit
p for which f ( p) > p. Otherwise for all subsequential limits p of (xn ), f ( p) ≤ p.
Let z be the infimum of all subsequential limits. Then z itself is a subsequential limit
of (xn ). So f (z) ≤ z. If f (z) < z, then f (z) < 21 ( f (z) + z) < z and 21 ( f (z) + z)
is a subsequential limit of (xn ) smaller than z, the smallest subsequential limit of
(xn ), we get a contradiction, unless f (z) = z. But by Step II above, f (z) cannot be
z. Thus, there is a subsequential limit p of (xn ) for which f ( p) > p.
Step IV. By Step II, there exists > 0 such that | f (x) − x| ≥ for all subsequential
limits x of (xn ). Otherwise, there is a sequence (wn ) of subsequential limits of (xn )
with |wn − f (wn )| < n1 for all n. This in turn implies that any subsequential limit
of (wn ), which is also a subsequential limit of (xn ) is a fixed point of f , contrary to
Step II.
Step V. Let w be the largest subsequential limit of (xn ) such that f (w) > w so
f (w) > Q = 21 ( f (w) + w) > w. Since Q is a subsequential limit exceeding w,
f (Q) < Q.
By Step IV, there is the least subsequential limit R of (xn ) such that f (R) < R
and w < R < f (w) (at least Q satisfies these conditions). Now f (R) < w.
Otherwise for A = 21 [R + f (R)], w < A < R. If f (R) ≥ w, then A = 21 (R +
f (R)) ≥ 21 (R + w) > 21 (w + w) = w and A = 21 (R + f (R)) < 21 (R + R) = R.
Since A is a subsequential limit greater than w, the largest subsequential limit less
than f (w), f (A) ≤ A. As A < R and R is the least subsequential limit with f (R) <
R, A ≤ f (A). Hence A = f (A) and this contradicts our assumption that no subse-
quential limit can be a fixed point of f . Hence f (R) < w. Consequently f (R) <
w < R < f (w) and |w − R| = R − w < | f (R) − f (w)| = f (w) − f (R). This is
a contradiction to the assumption on the map f that | f (x) − f (y)| ≤ |x − y| for all
x, y ∈ I . Hence (xn ) converges to a fixed point of f .
Remark 2.2.2 However, for any continuous map of I into itself, the sequence of iter-
ates defined in Theorem 2.2.1 may not converge. Let f : [0, 1] → [0, 1] be defined
by
2.2 Iterates of Real Functions 27
⎧
⎪ 3
⎨4 for 0 ≤ x ≤ 14
f (x) = 3 21 − x for 41 < x ≤ 21
⎪
⎩
0 for 21 ≤ x ≤ 1.
In this context, the following result due to Cohen and Hachigian [10] is pertinent.
Theorem 2.2.3 Let f : [−1, 1] → [−1, 1] be a continuous map such that f (−1) =
−1 and f (1) = 1. Then for each m = 0, 1, 2, . . . , f m+1 − I ≥ f m − I . Here I
denotes the identity map and g = sup{|g(x)| : x ∈ [−1, 1]) for any g ∈ C[−1, 1].
Proof If f ≡ I , the conclusion is obvious. So suppose that f = I . Let F = {x ∈
[−1, 1] : f (x) = x}. Since F is closed, the complement of F is open and so can be
written as a disjoint union of open subintervals Sα of [−1, 1]. For x ∈ Sα , f (x) <
x or f (x) > x. Clearly the conclusion is true for m = 0. Suppose the inequality
f k+1 − I ≥ f k − I is true for k = 1, 2, . . . , m. As [−1, 1] is compact and f m
is continuous, there exists p in [−1, 1] such that | f m ( p) − p| = f m − I .
Suppose without loss of generality f m ( p) > p. We claim that f ( p) > p. Clearly
f ( p) = p. If f ( p) < p, then for q = f ( p),
f m+1 − I ≥ f m − I , m = 0, 1, 2, . . . .
Cohen and Hachigian [10] have constructed an example of a continuous self-map
on the closed unit disc for which every point on the unit circle is a fixed point, with
the property that I − f > I − f k for some iterate f k of f .
For special real functions Thron [30] had obtained some interesting results on
the rates of convergence of iterates. Some of these are relevant to the solution of
Schroder’s functional equation. They provide useful estimates in approximating fixed
points by iterates.
28 2 Fixed Points of Some Real and Complex Functions
Definition 2.2.4 A map g : R → R is said to belong to the class H (a1 , k) if for some
x0 > 0, 0 < g(x) < x for x ∈ (0, x0 ] and g(x) = a1 x + x k+1 h(x) for x ∈ [0, x0 ]
where 0 ≤ a1 ≤ 1, k is a positive number and k is a continuous function on [0, x0 ]
with |h(x)| < M in [0, x0 ].
Remark 2.2.5 Clearly for g ∈ H (a, k), 0 is the unique fixed point of g and every
sequence (xn ) of g-iterates defined by xn+1 = g(xn ), n ∈ N and x1 ∈ (0, x0 ] con-
verges to 0.
Theorem 2.2.6 Let g ∈ H (a1 , k) where 0 < a1 < 1. Then for the sequence (xn ) of
g-iterates, there exists a constant K 1 (g, x) such that
xn
lim = K1
n→∞ a1n
1 − a1
0 < xnk M <
2
Sor xxn+1 < 1+a 2
1
< 1. Hence xn and xnk h(xn ) converge. So, the infinite product
∞
n
x k h(xn )
1+ n converges to a number L (say). Writing u n = axnn it follows that
a 1 1
n=1
u n+1 x k
h(x )
un
= ax1n+1
xn
= 1+ n a n .
n
xmk h(xm )
Since u n+1 = u 1 1+ , u n+1 converges to u 1 L. Hence u n = axnn
m=1
a 1 1
∞
If lim inf h(x) > 0, then (k + 1)−(m+1) log h(xm ) converges to a number
x→0
m=n 0
−n
K 3 (g, x1 ) − vn 0 , say. So (vn ) converges to log K 3 as n → ∞ or lim (xn )(k+1) =
n→∞
K3.
Suppose 0 < h(x) < M and that log h(xn ) could approach −∞ so that the series
(2.2.1) might not converge. Nevertheless, we have from (2.2.1)
n
vn+1 < (k + 1)−(m+1) log M + (k + 1)−n 0 log xn 0
m=n 0
−(n 0 +1) 1 − (k + 1)−n+n 0 +1
= log M(k + 1) + (k + 1)−(n 0 +1) log xnk+1
1 − (k + 1)−1 0
(2.2.2)
For large n 0 , the right-hand side of (2.2.3) or (2.2.4) as the case may be is negative
and is set as log K 2 (g, x1 ).
Now vn < log K 2 for n ≥ n 0 . So 0 < xn < K 2(k+1) .
n
Proof Since g(x) = x + x k+1 h(x), g(x) < x and |h(x)| < M, 0 ≤ −h(x) < M
for x ∈ [0, x0 ]. Hence B1 ≥ 0 and B2 ≤ M. Writing −h(xn ) = dn , xn+1 = xn +
xnk+1 h(xn ) becomes, for k = 1
xn+1 = xn (1 − xn dn )
and so
30 2 Fixed Points of Some Real and Complex Functions
1 1 1
= .
xn+1 xn (1 − xn dn )
∞
Choose n 1 (g, x1 , ) so that xn dn < 1, dnm xnm−1 < and B1 − < dn < B2 + .
m=2
3 3 3
For n ≥ n 1
∞
1 1
= + dn + dnm xnm−1 (by Binomial theorem)
xn+1 xn m=2
1 2
< + B2 + . (2.2.5)
xn 3
1
So xn 1 +m > .
2 1
m B2 + +
3 xn1
So for n ≥ n 1
1
xn >
n 1− n1
n
B2 + 2
3
+ 1
nxn 1
1
>
n B2 + 2
3
+ 1
nxn 1
Choose n 1 ≥ n 1 so that 1
nxn 1
< 3
for n ≥ n 1 . So we have for n ≥ n 1 ,
1
xn > .
n(B2 + )
1 1
> + B1 − .
xn+1 xn 3
1 1
> + (n − n 1 )(B1 − ) or
xn xn1
1
xn < . (2.2.6)
n 1 − n (B1 − ) + (nxn 1 )−1
n1
1
xn < .
n(B1 − )
For the case k = 1, define wn = xnk then xn+1 = g(xn ) = xn (1 + xnk h(xn )). So
1 k 1 k
wn+1 = g wnk = wn 1 + wn h wnk
= wn [1 + wn h 1 (wn )].
1
Since [g(wnk )]k is a function of wn , say g1 , it follows that g1 (w) ∈ h 1 (1, 1) for 0 ≤
w ≤ w0 = x0k . Also lim inf +
h 1 (w) = k B1 , lim sup −h 1 (w) = k B2 . The discussion
w→0 w→0+
now reduces the case k = 1 to the case k = 1 for g1 ∈ H (1, 1). It follows from the
previous discussion that for B1 > 0 and 0 < < B1 , there exists N ∈ N such that
for n > N
xn < [(B1 − )kn]− k
1
x
Remark 2.2.9 Since g(x) = sin( x2 ) ∈ H ( 21 , 2) in (0, 1), lim (2n sin n ( )) converges
n→∞ 2
for each x ∈ (0, 1) by Theorem 2.2.6.
Remark 2.2.10 Theorem 2.2.7 can be applied to g(x) = sin(x 1+ ) for any > 0 in
−n
(0, 1) to conclude that for any sequence (xn ) of iterates of sin(x 1+ ), lim (xn(1+) )
n→∞
converges.
This section treats Sharkovsky’s theorem on the existence of periodic points of con-
tinuous self-maps on a compact interval I ⊆ R. Sharkovsky published a fundamental
paper [27] on the existence of periodic points of continuous self-maps on compact
intervals in 1964, when he was about 27 years old. He introduced a new (total) order
on the set of natural numbers, often called Sharkovsky order. Interestingly, if a con-
tinuous map has a periodic point of period m, in the compact interval I (which it
maps into itself) it has periodic points of all periods ‘bigger than’ m (with respect to
32 2 Fixed Points of Some Real and Complex Functions
this order). The smallest natural number in this order is 3 and so it turns out that if
a continuous function mapping [a, b] into itself has periodic point of period 3, then
it has periodic points of all periods. Another implication of Sharkovsky’s theorem is
that if such a map has an odd periodic point then it has periodic points of all even
periods.
The more remarkable feature of Sharkovsky’s theorem is that its proof is essen-
tially based on the ingenious applications of the intermediate value theorem. The
paper by Li and Yorke [21] in 1975 proving a special case of Sharkovsky’s theorem
as well as May’s paper [22] highlighted the complicated behaviour of iterates of
simple functions and brought to limelight Sharkovsky’s work. The ‘simple proof’ of
Sharkovsky’s theorem presented below is due to Bau-Sen Du [14].
In the following, we assume that f : I → I is a continuous map, where I is a
compact interval in R. The following total ordering in N, the set of natural numbers
is called Sharkovsky’s ordering ≺. m ≺ n in the following ordering:
3 ≺ 5 ≺ 7 ≺ · · · ≺ 2.3 ≺ 2.5 · · ·
≺ 22 .3 ≺ 22 .5 ≺ 22 .7 ≺ · · · ≺ 23 .3 ≺ 23 .5 ≺ · · ·
≺ · · · ≺ 2n .3 ≺ 2n .5 ≺ · · ·
≺ · · · ≺ 23 ≺ 22 ≺ 2 ≺ 1
Lemma 2.3.1 Let a and b be points of I such that either f (b) < a < b ≤ f (a) or
f (b) ≤ a < b < f (a). Then there exists z, a fixed point of f < b, a 2-periodic point
y of f with y < z and a point v in (y, z) with f (v) = b and
Proof Whether f (b) < a < b ≤ f (a) or f (b) ≤ a < b < f (a), f (x) − x changes
sign in (a, b) and hence has a zero in (a, b). In other words, f has a fixed point z in
(a, b). As b ≤ f (a), a < z < b, and f (z) = z, there exists v ∈ [a, z) with f (v) = b.
If f (x) > z when min I ≤ x ≤ v, let u = min I ; otherwise let u = max{x : min I ≤
x ≤ v, f (x) = z}. Then f 2 (u) ≥ u and f (x) > z for u < x ≤ v. Since f 2 (v)(=
f (b)) ≤ a < v, f 2 has a fixed point in [u, v) or f has a 2 periodic point in [u, v).
If y is the largest 2-periodic point, then u ≤ y < v < z < f (y). Since f 2 (v) < v,
f 2 (x) < x for each x in (y, v].
Theorem 2.3.3 If f has a periodic point of least period m with m ≥ 3 and odd then
f has periodic points with least period n for each odd integer n ≥ m.
2.3 Periodic Points of Continuous Real Functions 33
Proof Let P be a periodic orbit of f with period m. By Lemma 2.3.1 and Remark
2.3.2. f has a fixed point z, a 2-periodic point y and a point v with y < v < z < f (y)
such that f (v) lies in P and f (x) > z and f 2 (x) < x when y < x ≤ v. Define
pm = v. As m is odd and y is a 2-periodic point of f , f m+2 (y) = f (y) > y and
because f 2 ( pm )(= f 2 (v)) is a period-m point of f , f m+2 ( pm ) = f 2 ( pm ) < pm .
So pm+2 = min{x : y ≤ x ≤ pm , f m+2 (x) = x} is well-defined and is an (m + 2)
periodic point of f . Since f m+4 (y) = f (y) > y and f m+4 ( pm+2 ) = f 2 ( pm+2 ) <
pm+2 (and it be noted that f 2 ( pm+2 ) cannot be pm+2 ). So pm+4 = min{x : y ≤ x ≤
pm+2 , f m+4 (x) = x} exists and is a periodic point of f with period (m + 4). Thus
proceeding, we obtain a decreasing sequence of points pm , pm+2 , . . ., pm+2k , . . . with
and f 2 (x) < x and f (x) > z for x in (y, v]. Write g = f 2 and let z 0 = min{t :
v ≤ t ≤ z, g(t) = t}. Then y and z 0 are fixed points of g such that y < v <
m+1
z 0 ≤ z < b = g 2 (v). Also g(x) < x and f (x) > z for y < x < z 0 . If g(x) <
z 0 for min I ≤ x ≤ z 0 , then g([min I, z 0 ]) ⊆ [min I, z 0 ] and this contradicts that
m+1
g 2 (v) = b > z 0 . Hence d = max{x : min I ≤ x ≤ y, g(x) = z 0 } is well defined
and f (x) > z > z 0 > g(x) for all x in (d, z 0 ). Define s = min{g(x) : d ≤ x ≤ z 0 }.
m+1
If s ≥ d, then g([d, z 0 ]) ⊆ [d, z 0 ]. But this contradicts that g 2 (v) = b > z 0 .
So s < d, [s, d] ∪ [d, z 0 ] are non-overlapping closed subintervals and f 2 [s, d] ∩
f 2 [d, z 0 ] ⊇ [s, d] ∪ [d, z 0 ]. Let
g : [d, z 0 ] → [d, z 0 ] be the map defined by g (x) =
max{g(x), d}. Clearly, g is continuous and onto and let t = min{x : d ≤ x ≤ z 0 ,
g(x) = d}. For each n ∈ N, define cn = min{x : d ≤ x ≤ t, g (x) = x}. It is not dif-
ficult to note that d < · · · < c4 < c3 < c2 < c1 ≤ y and that cn generates an n-period
orbit Q n ⊆ (d, z 0 ) of g . Clearly Q n is also an n-period orbit of g = f 2 . Since
x < z 0 ≤ z < f (x) for x in Q n , Q n ∪ f (Q n ) is 2n-period orbit of f . Thus f has
periods of all even orders.
Theorem 2.3.5 (Sharkovsky) Let f : I → I be a continuous map, where I is a
compact interval of real numbers. Then
(1) if f has a periodic point of period m and if m ≺ n (in the Sharkovsky order),
then f has also a periodic point of period n;
(2) for each positive integer n, there exists a continuous map g : I → I that has a
periodic point of period n but no point of period m ≺ n;
34 2 Fixed Points of Some Real and Complex Functions
Proof If f has j-periodic point with j ≥ 3 and odd, then by Theorem 2.3.3 f has
( j + 2) periodic point and by Theorem 2.3.4, f has a periodic point of period (2.3).
If f has (2. j) periodic point with j ≥ 3, and odd, f 2 has j-periodic point. So by
Theorem 2.3.3, f 2 has ( j + 2) periodic point and so f has either ( j + 2) periodic
point or period 2( j + 2) points. If f has ( j + 2) periodic point, then by Theorem
2.3.4, f has 2( j + 2) periodic point. In any case f has 2( j + 2) periodic point. If f 2
has j-periodic point, by Theorem 2.3.4, f 2 has 2.3 periodic point. So f has (22 .3)
k−1
periodic point. So if f has 2k . j periodic point, j ≥ 3 and odd and if k ≥ 2, then f 2
2k−1
has period 2. j points. So from what we have proved, we see that f has period
2( j + 2) points and period 22 .3 points. It follows that f has period (2k .( j + 2))
points and period (2k+1 .3) points, with j ≥ 3. If f has (2i . j) periodic points, j ≥ 3
i i −i
and odd and if i ≥ 0, then f 2 has j-periodic point. For ≥ i f 2 = ( f 2 )2 has
2
period j points. So by Lemma 2.3.1, f has period 2 points. So f has period 2+1
k−2
points for ≥ i. Finally when f has 2k -periodic points for some k ≥ 2, then f 2
2k−2
has 4 periodic point. Again by Lemma 2.3.1 f has 2 periodic points implying
that f has 2k−1 periodic points. Hence (1) is true.
For proving (2) and (3), without loss of generality, we can assume that I = [0, 1]
and T (x) = 1 − |2x − 1|, a map with a triangular graph having vertices at (0, 0),
( 21 , 1) and (1, 0). Then for each n ∈ N, T n (x) = x has exactly 2n distinct solutions
in I . So T has finitely many n-periodic orbits. Among these let Pn be an orbit
of the least diameter (= max Pn − min Pn ). Define Tn on I by Tn (x) = max Pn , if
T (x) ≥ max Pn , Tn (x) = min Pn , if T (x) ≤ min Pn and Tn (x) = T (x) for min Pn ≤
T (x) ≤ max Pn . Clearly Tx is continuous on I and Tx has exactly one-period n orbit,
i.e. Pn but has no m-periodic orbit for any m ≺ n.
Let Q 3 be any 3-periodic orbit of T of minimal diameter. Then [min Q 3 , max Q 3 ]
contains finitely many 6-periodic orbits of T . If Q 6 is one with smallest diameter,
then [min Q 6 , max Q 6 ] contains finitely many 12-periodic orbits of T . We choose
one, say Q 12 of minimal diameter and continue this process inductively. Define
q0 = sup{min⎧ Q 2i .3 : i ≥ 0} and q1 = inf{max Q 2i .3 : i ≥ 0}. Define T : I → I by
⎪
⎨q0 if T (x) ≤ q0
T (x) = q1 if T (x) ≥ q1 . Clearly T is continuous and has 2i -periodic
⎪
⎩
T (x) if q0 ≤ T (x) ≤ q1
point for i = 0, 1, 2, . . . but has no other periodic point. Thus (2) and (3) are
true.
If in addition O f (c) contains both a fixed point z and a point different from z, then
f has periodic points with all even periods. Arguments similar to those in Theorems
2.3.3 and 2.3.4 can be used.
Remark 2.3.7 Sharkovsky’s theorem cannot be generalized to continua (compact
2πi
connected subsets) of the plane. On the unit disc, the map z → ze 3 has 0 as the
only fixed point and all the other points are 3-periodic points. For each n ∈ N, the
2πi
map z → ze n has only one fixed point and the rest of the points are n-periodic
points. No point of fundamental period greater than n exists.
Sharkovsky’s result is definitely and unalterably one-dimensional (See
Ciesielski and Pogoda [8].) Nevertheless, there has been appropriate generaliza-
tion of Sharkovsky’s theorem to general topological spaces and more general maps
than continuous functions. See Schirmer [25].
It is natural to find out if two continuous real functions f, g : I (= [a, b]) → I have
a common fixed point. The maps x → x2 and x → 1 − x on [0, 1] have the only fixed
points 0 and 21 respectively. Since their compositions are 1−x2
and 1 − x2 , they do not
commute. If f, g : I → I have a common fixed point x0 , then x = f (x0 ) = g(x0 ) =
g f (x0 ) = f g(x0 ) and thus f and g commute at least on {x0 }. Ritt [24] showed that if
f and g are polynomials that commute, then they are within certain homeomorphisms
iterates of the same function, both power of x or both must be Chebyshev polynomials
and in both these cases, the commuting polynomials have a common fixed point. So
Dyer conjectured that if f, g : I (= [a, b]) → I are continuous real functions that
commute, then f and g have a common fixed point. However, Boyce [5] and Huneke
[17] had disproved the conjecture independently by constructing counter-examples
to point out that commuting continuous self-maps on a compact real interval may not
have a common fixed point. Isbell [18] first recorded this problem in a more general
form.
This section discusses some results that ensure the existence of common fixed
points of two commuting continuous functions f, g : I → I under suitable additional
assumptions. We recall the following definitions.
Definition 2.4.1 Let F be a family of maps from a topological space X into a
metric space (X, d). It is said to be equicontinuous at x0 ∈ X , if for each > 0,
there exists an open set O in X containing x0 such that for each x ∈ O and f ∈ F ,
d( f (x0 ), f (x)) < . F is said to be equicontinuous on X , if it is equicontinuous at
each x ∈ X .
Definition 2.4.2 If f : X → X is a map, a subset A ⊆ X is said to be f -invariant
or invariant (under f ) if f (A) ⊆ A.
An elementary proposition on invariant subsets of continuous maps on compact
intervals is given below.
36 2 Fixed Points of Some Real and Complex Functions
Proof Let C be a non-empty closed invariant subset of I and C be the family of all
closed invariant subsets of C. Clearly C ⊂ C . Let F be a chain of sets in C . Since F
is a subfamily of non-empty closed subsets of C which are indeed compact subsets of
I , F0 = ∩{F : F ∈ F } is non-empty and compact. Further f (F0 ) ⊆ f (F) ⊆ F for
all F ∈ F and hence f (F0 ) ⊆ ∩{F : F ∈ F } = F0 . Thus, F0 is an invariant closed
subset which is contained in each F ∈ F . Thus F0 is the least element of F in C.
So by Zorn’s Lemma, C has a minimal element C0 , which is a non-empty minimal
closed invariant subset of C.
Theorem 2.4.6 (Schwartz [26]) Every non-void closed invariant minimal subset of
the continuous function f : I → I is contained in the closure of P f , where P f =
{x ∈ I : f k (x) = x for some k ∈ N}, the set of periodic points of f .
Proof Let C1 ∪ {h} be any finite subset of F ∪ {h} of the form { f 1 , . . . , f n } ∪ {h} ∪
{g1 , . . . , gm } where f i , i = 1, 2, . . . , n ∈ F1 and {g1 , . . . , gm } ⊆ F2 . Since F fi is
n
a compact interval and f i ’s commute F fi is a non-empty compact interval, say
i=1
[c, d]. As h commutes with each f i ∈ C1 , h maps [c, d] into itself and so has a
fixed point z ∈ [c, d]. Now g1n (z) has a limit point z 1 in Pg1 by Theorem 2.4.6. As
Pg1 = Fg1 (by hypothesis (ii), and Fg1 is closed, Pg1 = Pg1 . Similarly g2n (z 1 ) has a limit
point z 2 in Pg2 = Fg2 = Pg2 and as Fg2 is closed z 2 ∈ Fg2 . Thus z 1 , z 2 ∈ [c, a]. Thus
proceeding, we see that {g nj (z j−1 )} has a limit point z j in Pg j for j = 2, . . . , m which
is fixed for f 1 , . . . , f n , h, g1 , . . . , gm . So ∩F f = φ for all f ∈ C1 ∪ {h}. It is also
easily seen that for any finite subset C2 of F1 , F f = φ as also F f = φ for
f ∈C2 f ∈C3
any finite subset C3 of F2 . Thus, the family of closed subsets {F f : f ∈ F ∪ {h}} of
[a, b] has finite intersection property and hence ∩{Fr : f ∈ F ∪ {h}} is non-empty,
in view of the compactness of [a, b].
Theorem 2.4.9 (Cano [6]) Let f : I (= [a, b]) → I be a continuous function such
that { f n : n ∈ N} is an equicontinuous family at each x ∈ I . Then
(1) F p , the fixed point set of f is a compact subinterval of I ;
(2) if F f is a non-degenerate interval, then F f = P f (P f being the set of periodic
points of f ).
Case (i) If f (x) < b) for all x ∈ (a0 , b0 ) then f n (x) ∈ (a0 , b0 ) for all n ∈ N and
f n (x) < f n+1 (x) < b0 and so it converges to a fixed point of f , which cannot
be in (a0 , b0 ) and hence has to be b0 . So given > 0, by the equicontinuity of
{ f n } at a0 , there exists δ > 0 such that |a0 − x0 | < δ such that for |a0 − x0 | < δ,
| f n (a0 ) − f n (x0 )| < . Since f n (a0 ) = a0 , for all n, this contradicts that f n (x0 )
converges to b0 .
Case (ii) Suppose for some x0 ∈ (a0 , b0 ), f (x0 ) ≥ b0 . Then there is a least num-
ber z in (a0 , b0 ) with f (z) ≥ b0 . In fact f (z) = b0 . Otherwise, there exists z < z
with f (z ) ≥ b0 by the continuity of f and this contradicts the definition of z. Thus
proceeding, we can find a non-increasing sequence (xn ) in (a0 , z] such that (xn ) con-
verges to a0 , x1 = z and f (xn ) = xn−1 , n = 2, 3, . . . . Since f n (xn ) = f n−1 (xn−1 ) =
· · · f (x1 ) = f (z) = b0 for all n, f n cannot be equicontinuous at a0 . (Note that as
(xn ) is non-increasing in (a0 , z) it converges to a number z ≥ a0 . z > a0 is a con-
tradiction as z = f (z ) and by assumption f has no fixed point in (a0 , b0 ).)
Suppose f (x) < x for all x ∈ (a0 , b0 ). We consider
Case (i) Suppose f (x) > a0 for all x ∈ (a0 , b0 ). Then for all x ∈ (a0 , b0 ), f n (x) >
f n+1 (x), n ∈ N and ( f n (x)) as in Case (i) converges to a0 . However the family of f
iterates cannot be equicontinuous at b0 .
Case (ii) If for some x ∈ (a0 , b0 ), f (x) ≤ a0 . Then there is a greatest element z
in (a0 , b0 ) with f (z ) ≤ a0 . In fact f (z ) = a0 . By this process, a non-decreasing
sequence (yn ) can be chosen in (z , b0 ] with y1 = z , f (yn ) = yn−1 , n = 2, 3, . . . .
So f n (yn ) = f (z ) = a0 . If (yn ) converges to w, then f (yn ) (= yn−1 ) converges
to f (w) and so w = f (w). As w ∈ / (a0 , b0 ), (yn ) converges to b0 . Since f n (yn ) =
f n−1
(yn−1 ) · · · = f (z ) = a0 . As yn converges to b0 , there is a contradiction to the
equicontinuity of f n at b0 .
Thus we have shown that F f is a non-void compact interval. If F f is non-
degenerate let F f = [a0 , b0 ] where a0 < b0 . Let f n (x) = x for some n and x ∈
[a, a0 ). (If x ∈ (b0 , b], then a similar argument can be provided). Since f n has
a fixed point and its iterates are equicontinuous at each point, f n (y) = y for
all y ∈ [x, a0 ] by what has been proved in (i) so far. Since f (y) > y for all
y ∈ [a, a0 ) and f (a0 ) = a0 , we can choose y from (x, a0 ) close to a0 , such that
a0 − k1 < y < f (y) · · · < f n−1 (y) < a and this implies f n (y) > y, a contradic-
tion. So a0 + k1 > f (y) > a0 > y > a0 − k1 . Then f (y) is a fixed point for f .
So f (y) = f 2 (y) and f n (y) = f n−2 ( f 2 (y)) = f n−1 (y). Thus proceeding, y =
f n (y) = f n−1 (b) · · · = f (y) contradicting f (y) > a > y. Thus if F f = [a0 , b0 ],
[a0 , a) has no periodic point. Similarly (b0 , b] has no periodic point.
Proof (i) =⇒ (ii). Suppose Fg is not a singleton and is [a1 , b1 ] where a1 <
b1 . Since for a1 < x < b1 , g n (x) = x for all n ∈ N, the continuity of g at x
implies that given > 0 with b − a > , there is a δ() > 0 such that (x − δ, x +
δ) ⊆ (a1 , b1 ) and |g(x) − g(x )| < for x ∈ (x − δ, x + δ). So |g n (x) − g n (x )| =
|g(x) − g(x )| < for x ∈ (x − δ, x + δ), proving the equicontinuity of {g n } on
(a1 , b1 ). We now show that {g n } is equicontinuous at a1 . Since g is continuous at
a1 , there exists δ() > 0 with > δ() for a given > 0 such that for a1 − δ < x <
a1 + δ, |g(x) − g(a1 )| = |g(x) − a1 | < . We now show by the principle of finite
induction that a1 − < g n (x) < a1 + for all x ∈ (a1 − δ, a1 + δ) for all n ∈ N.
Clearly, the inequality is true for n = 1. Suppose it is true for n = 1, 2, . . . , k.
Let x ∈ (a − δ, a). If a1 ≤ g k (x) < a1 + , then g k (x) ∈ Fg and so |g k+1 (x) −
a1 | = |g k+1 (x) − g k+1 (a1 )| = |g k (x) − g k (a1 )| = |g k (x) − a1 | < . If g k (x) < a1 ,
then g i (x) < a1 for i = 1, 2, . . . , k. Otherwise by induction hypothesis for some i,
1 ≤ i ≤ k and a1 ≤ g i (a) < a1 + or g i (x) ∈ Fg and so g k (x) ∈ Fg or g k (x) ≥ a1 ,
a contradiction. Since Fg = [a1 , b1 ], g(x) > x for x ∈ [a, a1 ). So g i (x) > g i−1 (x)
for i = 1, 2, . . . , k, implying that g k (x) > g k−1 (x) > · · · > x. As a1 − δ < x and
g k (x) < a1 , it follows that g k (x) ∈ (a1 − δ, a1 ). So |g(g k (x)) − g(a1 )| = |g k+1 (x) −
a1 | < . For x ∈ (a1 , a1 + δ) ⊆ [a1 , b1 ], |g n (x) − g n (a1 )| = |x − a1 | < . Thus g n
is equicontinuous at a1 . By a similar reasoning, (g n ) is equicontinuous at b1 .
(ii) =⇒ (i). This follows from the proof of Theorem 2.4.9 (i). In fact to prove
(i) of Theorem 2.4.9, it suffices to assume that { f n } is equicontinuous on F f .
δ) − g n ( 21 )| < for all n. Since g( 21 ) = 21 and choosing least n 0 such that 2n 0 δ > 14 ,
it follows that g n ( 21 + δ) = 0 for all n ≥ n 0 and |g n ( 21 + δ) − g n ( 21 )| = |0 − 21 | =
| 21 | < 41 , a contradiction. So (g n ) is not equicontinuous.
If f : [0, 1] → [0, 1] commutes with g at 21 , then f g( 21 ) = g( f ( 21 )) = f ( 21 ) (as
g( 2 ) = 21 ). Since f ( 21 ) is a fixed point of g and g has the unique fixed point 21 , f ( 21 ) =
1
1
2
. Thus, f and g have a common fixed point, even though {g n } is not equicontinuous.
This example points out that the hypothesis Fg is a singleton cannot be dropped
in Theorem 2.4.10.
The next theorem on the convergence of iterates, due to Coven and Hedlund [12],
was also obtained independently by Chu and Moyer [7].
Proof Let N be the least common period of the periodic points. Apply Theorem
2.4.12 to f N and that P f N = F f N . (It is to be observed that N must be a power of 2,
as can be seen from Sharkovsky’s theorem.)
2.4 Common Fixed Points, Commutativity and Iterates 41
Our next theorem characterizes functions f : I → I that are continuous and for
which P f = F f .
In this section, an existence theorem on the common fixed points for two commuting
continuous self-maps on a compact real interval, due to Cohen [9] is proved. This
supplements the theorems in Sect. 2.4. Without loss of generality we take I = [0, 1].
We need the following lemmata and definitions.
Proof Let a1 = max{inf f, inf g} and b1 = min{sup f, sup g}. Since f and g com-
I I I I
mute, f [0, 1] ∩ g[0, 1] = φ both f and g map [a1 , b1 ] into itself. Otherwise for
some x ∈ [a1 , b1 ], f (x) > b1 would imply that for some y ∈ [0, 1], g(y) = x and
g( f (y)) = f g(y) = f (x) > b1 . This implies that b1 < min{sup f, sup g}. Similarly
I I
f (x) < a1 for some x ∈ [a1 , b1 ] would imply that there exists y ∈ [0, 1] with g(y) =
x and g( f (y)) = f g(y) = f (x) < a1 . This means that a1 > max{inf I f, inf I g}, a
contradiction. Writing f 1 and g1 as the restrictions of f and g on J1 = [a1 , b1 ] respec-
tively, we can inductively define ai , bi and f i by ai = max{inf Ji−1 f, inf Ji−1 g} and
bi = min{sup Ji−1 f, sup Ji−1 g} where Ji−1 = [ai−1 , bi−1 ], i = 2, 3, . . . , and f i is the
restriction of f i−1 to Ji−1 . Since [ai , bi ], i = 1, 2, . . . , form a nested sequence of
compact subsets of [0, 1], they have a non-void intersection. If this intersection is
a singleton, then f and g have a common fixed point contrary to the assumption.
∞
Hence, [ai , bi ] is a non-degenerate compact interval [a, b] and the restriction f
i=1
and g of f and g respectively map [a, b] onto itself. If h is a homeomorphism of
[a, b] onto I = [0, 1]. Then, the continuous maps h f h −1 and hgh −1 map [0, 1] onto
itself but have no common fixed points by Lemma 2.5.1.
Similarly
44 2 Fixed Points of Some Real and Complex Functions
s−1 s−1 1
D( f 1 gs ) = , +
m m mn
s−1 1 s−1 2
D( f 2 gs ) = + , + ...,
m mn m mn
s−1 r −1 s−1 r
D( fr gs ) = + , +
m mn m mn
mn − 1 mn + 1
= ,
2mn 2mn
Thus D( fr gs ) = D(gs fr ). Since gs is continuous and onto [0, 1], its graph must
intersect the diagonal of I × I and gs has a fixed point z 1 . As D(gs ) ⊆ D( f 0 ),
z 1 ∈ D( fr ) and thus z 1 ∈ D( fr gs ) = D(gs fr ). So gs fr (z 1 ) = fr gs (z 1 ) = fr (z 1 ) and
z 2 = fr (z 1 ) is a fixed point of gs . Thus proceeding, we get a sequence z p of fixed
points of gs with z p+1 = fr (z p ). Since fr is monotone the sequence z p converges to
z 1 a fixed point of both f and g. The case when r is even and s is odd can be handled
similarly.
Remark 2.5.8 One can show that f is full if and only if f maps [0, 1] onto [0, 1]
and is an open map. For related work, Baxter and Joichi [3] may be referred.
We prove a theorem of Shields [28] on the common fixed points of analytic functions
in this section. We denote by G, a non-void bounded open connected set in the
complex plane. Let FG be the family of all analytic functions mapping G into itself.
Clearly FG is a semigroup under composition of mappings. We can consider H (G)
the linear space of all functions analytic on G and continuous on G, with the topology
of uniform convergence on compact subsets of G. This topology is a metric topology
and indeed it arises from a complete metric and so FG will inherit this metric topology.
The following lemma implies that FG is a topological semigroup (i.e. the composition
map is a continuous function from G × G into G).
Lemma 2.6.1 Let f n , gn ∈ FG and f n → f , gn → g in the topology of uniform con-
vergence on compact subsets of G. Then f n (gn ) → f (g) and so FG is a topological
semigroup.
Proof Let K be a compact subset of G and let U be an open set containing g(K )
with U compact and lying in G. Since gn → g uniformly on K , gn (K ) ⊂ U for all
n ≥ n 0 for some n 0 ∈ N. Now for all n
that f (gn (z)) → f g(z) and | f n (gn (z)) − f (gn (z))| ≤ sup | f n (w) − f (w)| → 0 as
w∈U
n → ∞. Hence ( f n gn ) converges uniformly on K to f g. Thus FG is a topological
semigroup.
A few facts from the theory of topological semigroups will be needed in the
sequel. For proofs and other details Numakura [23], Wallace [31] and Ellis [15] may
be consulted.
Definition 2.6.2 Let (S, ·) be a semigroup. An element e of S is called an idempotent
if e.e = e2 = e. An element 0 is termed zero if 0.x = 0 for all x ∈ S. 1 is called an
identity of S if 1.x = x = x.1 for all x ∈ S. In a semigroup S if ax = ay (xa = ya)
implies x = y for all a, x, y in S then S is called a semigroup satisfying the left
(right) cancellation law. If S satisfies both the left and right cancellation laws, it is
called a semigroup satisfying cancellation law.
The following is a basic result in the theory of topological semigroups and the
proof is essentially from Ellis [15].
Lemma 2.6.3 Let S be a compact Hausdorff topological semigroup. Then S has an
idempotent element.
Proof Let F be the family of all compact subsets K of S such that K 2 ⊆ K . F = φ,
as S ∈ F . F is partially ordered by set inclusion. As every chain in F has a lower
bound F has a minimal element A in F . If r ∈ A, then r A is a non-void compact
subset of S as r A is the image of the compact set A under the continuous map
x → r.x. So r A ∈ F and r A ⊆ A. Since A is minimal r A = A. So there exists
p ∈ A such that r p = r . Define L = {a ∈ A : ra = r }. Clearly p ∈ L and L is a
compact subset of A. Let 1 , 2 ∈ L. Then r 1 2 = r 2 = r and hence 1 ◦ 2 ∈ L.
So L 2 ⊆ L. Hence L ∈ F . As L ⊆ A and A is minimal L = A. Since r ∈ A = L,
r 2 = r from the definition of L. Thus S has an idempotent element.
We skip the proof of the following.
Lemma 2.6.4 Let S be a compact T2 topological semigroup which is commutative.
For x ∈ S and (x) = cl{x, x 2 , . . . , }, we have
(i) (x) contains exactly one idempotent;
(ii) if e is an identity for (x), then (x) is a group and x has an inverse in (x);
(iii) if e is a zero for (x), then xn → e.
The following lemma makes use of the basic properties of analytic functions.
Lemma 2.6.5 If the analytic function e ∈ FG is idempotent, then e(z) ≡ z on e(z)
is constant for all z ∈ G.
Proof If e(z) is constant for all z ∈ G, clearly it is an idempotent. Suppose e is a
non-constant analytic function on G, then f is an open mapping. So G 1 = e(G) is an
open set. Since e2 (z) = e(z), e(z) = z on G 1 . As G 1 is uncountable, and the analytic
functions, viz. identity function and e coincide on G 1 , e(z) must be z at each z
in G.
46 2 Fixed Points of Some Real and Complex Functions
We also recall some classical results from complex analysis (see Conway [11]
and Ahlfohrs [1].
Theorem 2.6.6 (Montel) Let H (G) be the linear space of analytic functions on the
open region G. A family F in H (G) is normal in the sense that every sequence in
F has a convergent subsequence if and only if F is locally bounded in H (G) (i.e.
for each compact subset K of G, there is a positive constant Mk with | f (z)| ≤ Mk
for all f ∈ F and z ∈ K ).
Theorem 2.6.7 (Hurwitz) Let A(G) be the linear space of all analytic functions
with the topology of uniform convergence on compact subsets of G. If ( f n ) converges
to f in H (G) and f n never vanishes on G for each n, then f ≡ 0 or f is non-zero
throughout G.
Lemma 2.6.8 Let D be the open unit disc in the complex plane C and f : D → D be
a bilinear (Mobius) transformation of D onto D. Then there arise three possibilities:
(i) f (z) = z on D;
(ii) f has exactly one fixed point in the closed unit disc;
(iii) f has two distinct fixed points in the unit circle and the iterates of f converge
to one of these fixed points.
(z−a)
Proof The general form of such a bilinear transformation is f (z) = α (1−a)z where
|α| = 1, |a| < 1.
If f is not the identity function the fixed points z = f (z) are given by
az 2 − (1 − α)z − αz = 0
As this equation is invariant under z → 1z , the fixed points of f (z) are inverses of
each other with respect to the unit circle. So there is a fixed point inside and another
outside the circle or there is a ‘double fixed point’ or two distinct fixed points on the
unit circle.
Lemma 2.6.9 Let f ∈ FG , be the subset of H (G) containing all analytic functions
mapping G into itself. Suppose f is not a homeomorphism of G onto itself. Then
there is a point z 0 in G and a subsequence { f ni } of f -iterates such that f ni (z) → z 0
uniformly on compact subsets of G.
Proof Write ( f ) = cl{ f n } in H (G). If ( f ) ⊆ FG , then ( f ) is a compact semi-
group under composition of functions and contains an idempotent element e(z) by
Lemma 2.6.3.
By Lemma 2.6.5 e(z) ≡ z for all z ∈ G or is a constant z 0 for all z ∈ G. If the
identity map belongs to ( f ), then by Lemma 2.6.4, ( f ) is a group and f ∈ ( f ) ⊆
FG would be invertible in F(G) contradicting that f is not a homeomorphism. Hence
e(z) ≡ z 0 , for all z ∈ G and is thus a zero for ( f ). So again by Lemma 2.6.4 f n (z)
converges to z 0 in the topology of FG .
Suppose g ∈ ( f ) does not belong to FG . Since f n (G) ⊆ G, g(G) ⊆ G. As
g∈/ FG , there is a point z ∈ G with g(z ) = z 0 ∈ / G. We claim that g(z) ≡ z 0 .
2.6 Common Fixed Points of Commuting Analytic Functions 47
Proof By Lemma 2.6.9, there exists z ∈ G with lim f ni (z) = z 0 in FG . For g ∈ C(G),
g(z 0 ) = g(lim f ni (z)) = lim f ni (g(z)) = z 0 .
Remark 2.6.11 If f is a bilinear map of the open unit disc D onto itself with two
distinct fixed points on the boundary, consider p a bilinear map, mapping D onto
the upper half-plane and taking these fixed points into 0 and ∞. For g = p f p −1 ,
0 and ∞ are fixed points of g and g maps the upper half-plane onto itself. Hence
g is a dilatation and is of the form g(z) = az, a > 0 and a = 1 as f (z) ≡ z. So
g n (z) = a n z tends to zero or to ∞. Thus the iterates of f converge to one of the fixed
points of f .
Remark 2.6.12 Wolff [32] and Denjoy [13] have shown independently in 1926 that
if f is analytic in D and f (D) ⊆ D, then either f is a bilinear map of D onto itself
with exactly one fixed point or f n converges to a constant C ∈ D.
We are now in a position to prove a theorem of Shields [28] on the fixed points
of commuting family of analytic functions on D.
Proof If F contains a constant function then that constant is the common fixed point.
Suppose it contains only non-constant continuous functions on D which are analytic
in D. So by the Maximum Modulus Theorem f (D) ⊆ D for each f ∈ F. Suppose
not all functions of F are bilinear maps of D onto D. So there exists f , different
from the identity map in F. Then Lemma 2.6.10 can be invoked to conclude that
there is a common fixed point for each f ∈ F. On the other hand if all the members
of F are bilinear, then if one of them has just one fixed point, then it is a common
fixed point for all. In case these have two fixed points then by Remark 2.6.11, the
iterates converge to one of the two fixed points and so invoking Lemma 2.6.10, we
conclude that for each f in F there is a common fixed point.
Remark 2.6.14 Theorem 2.6.13 due to Shields has been generalized to Hilbert spaces
by Suffridge [29].
48 2 Fixed Points of Some Real and Complex Functions
Theorem 2.7.1 (Picard (see Conway [11])) Suppose an analytic function f has an
essential singularity at a. Then in each neighbourhood of a, f assumes each complex
number, with one possible exception, infinitely many times.
Corollary 2.7.2 An entire function which is not a polynomial assumes every complex
number, with one exception infinitely many times.
In response to a question raised by Gross [16], Bergweiler [4] proved the following.
Theorem 2.7.3 (Bergweiler [4]) Let f be a meromorphic function that has at least
two different poles and let g be a transcendental entire function. Then the composite
function f ◦ g has infinitely many fixed points.
Lemma 2.7.5 Let f and g be meromorphic functions. Then f ◦ g has infinitely many
fixed points if and only if g ◦ f does.
Proof Let z 1 and z 2 be poles of f of order p1 and p2 . Using Lemma 2.7.4 choose the
functions h j for j ∈ {1, 2}. Let k1 (z) = h 1 (z p2 ) + z 1 and k2 (z) = h 2 (z p1 ) + z 2 . Now
f (k1 (z)) = f (k2 (z)) = z − p1 p2 for z = 0 in a neighbourhood of 0. Define u(z) =
2.7 Fixed Points of Meromorphic Functions 49
u(z) − k1 (z)
v(z) = .
k2 (z) − k1 (z)
0 is an essential singularity for u and v does not take the values 0, 1 and ∞ in a
punctured neighbourhood of 0. This contradicts Picard’s Theorem 2.7.1. Hence the
theorem.
Remark 2.7.6 It can be similarly shown that if f and g are transcendental meromor-
phic functions and if either f or g has at least three poles, then f ◦ g has infinitely
many fixed points.
References
1. Ahlfohrs, L.V.: Complex Analysis an Introduction to the Theory of Analytic Functions of One
Complex Variable, 3rd edn. McGraw-Hill Book Co., New York (1978)
2. Bailey, D.F.: Krasnoselski’s theorem on the real line. Am. Math. Mon. 81, 506–507 (1974)
3. Baxter, G., Joichi, J.T.: On functions that commute with full functions. Nieuw Arch. Wiskd.
3(XII), 12–18 (1964)
4. Bergweiler, W.: On the existence of fix points of composite meromorphic functions. Proc. Am.
Math. Soc. 114, 879–880 (1992)
5. Boyce, W.M.: Commuting functions with no common fixed point. Trans. Am. Math. Soc. 137,
77–92 (1969)
6. Cano, J.: Fixed points for a class of commuting mappings on an interval. Proc. Am. Math. Soc.
86, 336–338 (1982)
7. Chu, S.C., Moyer, R.D.: On continuous functions commuting functions and fixed points. Fun-
dam. Math. 59, 91–95 (1966)
8. Ciesielski, K., Pogoda, Z.: On ordering the natural numbers or the Sharkovski theorem. Am.
Math. Mon. 115, 159–165 (2008)
9. Cohen, H.: On fixed points of commuting function. Proc. Am. Math. Soc. 15, 293–296 (1964)
10. Cohen, H., Hachigian, J.: On iterates of of continuous functions on a unit ball. Proc. Am. Math.
Soc. 408–411 (1967)
11. Conway, J.B.: Functions of One Complex Variable, Springer International Student Edition.
Authorized reprint of the original edition published by Springer, New York, Narosa Publishing
House Reprint (2nd edn.) ninth reprint (1990)
12. Coven, E.M., Hedlund, G.A.: Continuous maps of the interval whose periodic points form a
closed set. Proc. Am. Math. Soc. 79, 127–133 (1980)
13. Denjoy, A.: Sur l’iteration des fonctions analytiques. C.R. Acad. Sci. Paris 182, 255–257 (1926)
14. Du, B.S.: A simple proof of Sharkovsky’s theorem revisited. Am. Math. Mon. 114, 152–155
(2007)
15. Ellis, R.: Distal transformation groups. Pac. J. Math. 8, 401–405 (1958)
16. Gross, F.: On factorization of meromorphic functions. Trans. Am. Math. Soc. 131, 215–222
(1968)
50 2 Fixed Points of Some Real and Complex Functions
17. Huneke, J.P.: On common fixed points of commuting continuous functions on an interval.
Trans. Am. Math. Soc. 139, 371–381 (1969)
18. Isbell, J.R.: Commuting mappings of trees, research problem # 7. Bull. Am. Math. Soc. 63,
419 (1957)
19. Jachymski, J.: Equivalent conditions involving common fixed points for maps on the unit
interval. Proc. Am. Math. Soc. 124, 3229–3233 (1996)
20. Krasnoselskii, M.A.: Two remarks on the method of sequential approximations. Usp. Mat.
Nauk 10, 123–127 (1955)
21. Li, T.Y., Yorke, J.A.: Period three implies chaos. Am. Math. Mon. 103, 985–992 (1975)
22. May, R.B.: Simple mathematical models with very complicated dynamics. Nature 261, 459–
467 (1976)
23. Numakura, K.: On bicompact semigroups. Math. J. Okayama Univ. 1, 99–108 (1952)
24. Ritt, J.F.: Permutable rational functions. Trans. Am. Math. Soc. 25, 399–448 (1923)
25. Schirmer, H.: A topologist’s view of Sharkovsky’s theorem. Houst. J. Math. 11, 385–394 (1985)
26. Schwartz, A.J.: Common periodic points of commuting functions. Mich. Math. J. 12, 353–355
(1965)
27. Sharkovsky, A.N.: Coexistence of cycles of a continuous mapping of the line into iteslf. Ukr.
Math. J. 16, 61–71 (1964)
28. Shields, A.L.: On fixed points of analytic functions. Proc. Am. Math. Soc. 15, 703–706 (1964)
29. Suffridge, T.J.: Common fixed points of commuting holomorphic maps of the hyperball. Mich.
Math. J. 21, 309–314 (1974)
30. Thron, W.J.: Sequences generated by iteration. Trans. Am. Math. Soc. 96, 38–53 (1960)
31. Wallace, A.D.: The structure of topological semigroups. Bull. Am. Math. Soc. 61, 95–112
(1955)
32. Wolff, J.: Sur l’iteration des functions holomorphe. C.R. Acad. Sci. Paris 182, 42–43 (1926).
200-201
Chapter 3
Fixed Points and Order
This chapter deals with fixed points of mappings on partially ordered sets (vide
Definition 1.1.10) under diverse hypotheses.
In this section, some elementary fixed point theorems on linear continua due to
Andres et al. [3] are discussed.
Definition 3.1.1 A partially ordered set (or a poset, for short) (X, ≤) is said to be
linearly ordered or totally ordered or a chain, if for any pair of elements x, y ∈ X ,
x ≤ y or y ≤ x.
Definition 3.1.3 A linearly ordered set (X, ≤) with more than one element is called
a linear continuum, if
(i) it is densely ordered or without gaps if for x, y ∈ X with x < y, there exists
z ∈ X such that x < z < y (a < b if a ≤ b and a = b) and
(ii) for each nonempty subset S of X bounded above there is a least upper bound
(l.u.b) in X (called l.u.b property).
© Springer Nature Singapore Pte Ltd. 2018 51
P. V. Subrahmanyam, Elementary Fixed Point Theorems,
Forum for Interdisciplinary Mathematics,
https://doi.org/10.1007/978-981-13-3158-9_3
52 3 Fixed Points and Order
Remark 3.1.4 If (X, ≤) is a linearly ordered set, then the sets of the form {x ∈ X :
x < a}, {x ∈ X : a < x}, a ∈ X is a subbase for a topology on X , called the order
topology on X .
X is connected and Hausdorff in the order topology if and only if X is a linear con-
tinuum in the sense of Definition 3.1.3. Moreover, the order topology is the smallest
topology on X under which < is continuous (see Kelley [10], pp. 57–58). Also in a
linearly ordered set with the order topology every closed and (order) bounded subset
of X is compact if and only if it has the l.u.b property (see Kelley [10], p. 162).
Definition 3.1.5 Let X be a linear continuum and 2 X denote the set of all subsets
(power set) of X . A map ϕ : X → 2 X − {φ} is called a multimap. An element x0 ∈ X
such that x0 ∈ ϕ(x0 ) is called a fixed point of ϕ.
Theorem 3.1.6 (Andres et al. [3]) Let (X, ≤) be a linear continuum and I =
[a, b] = {x ∈ X : a ≤ x ≤ b}, where a, b ∈ X . Let ϕ : I → X be a multimap on
I with a connected graph. If either I ⊆ ϕ(I ) or ϕ(I ) ⊆ I , then ϕ has a fixed point.
Theorem 3.1.6 has a simple corollary generalizing Theorems 2.15 and 2.1.7.
f has a connected graph, though f 2 (x) = { 61 , 13 } does not have a connected graph.
However f has a fixed point in view of Theorem 3.1.6 as f (I ) ⊆ I and so f 2 also
has a fixed point.
Proof As f n has a connected graph, by Theorem 3.1.6, f n has a fixed point. If this
is not a fixed point of f , then there is a nontrivial k ∈ N such that k factors n. So by
Proposition 3.1.9 f has a fixed point.
Now
⎧
⎪
⎨[0, 2 ], x = 0
1
f 2 (x) = x, x ∈ (0, 1) − { 21 }
⎪
⎩ 1
[ 2 , 1], x ∈ { 21 , 1}
and every point of [0, 1] is a fixed point of f 2 but f has no fixed point. Further the
graph of f is not connected. Hence in Corollary 3.1.11 the hypothesis that the graph
of f is connected cannot be dropped.
54 3 Fixed Points and Order
A set-theoretical fixed point theorem for maps in the power set of a set proved in
1927 by Knaster [11] and improved by Tarski germinated into the following theorem
referred as the Knaster–Tarski principle in the literature.
Theorem 3.2.2 Let (X, ≤) be a poset and f : X → X be an isotone map such that
(i) b ≤ f (b) for some b ∈ X ;
(ii) every chain in X 1 = {x ∈ X ; b ≤ x} has a supremum.
Then F f , the set of fixed points of f is nonempty and contains a maximal fixed point.
In this context the following theorem due to Bourbaki [5], also called Zermelo’s
theorem [19] is stated and the proof is left as an exercise.
Theorem 3.2.3 (Bourbaki–Zermelo) Let (X, ≤) be a poset in which every chain has
an upper bound. Let f : X → X be a map such that x ≤ f (x) for all x ∈ X . Then
f has a fixed point.
In fact Abian [1] and Moroianu [13] have pointed out that it suffices to assume
the completeness of each well-ordered subset of X instead of the completeness of a
chain in Bourbaki’s theorem. First, we recall
Remark 3.2.5 The set of natural numbers with the ‘natural order’ is well-ordered,
though the real number system is not well-ordered. While every nonempty subset of
[0, 1] has a least and a greatest element in [0, 1], these need not belong to the subset
concerned. So [0, 1] is not well-ordered. The Axiom of Choice is equivalent to the
hypothesis that every nonempty set can be well-ordered (see Kelley [10]).
We now prove the following version of Bourbaki’s theorem, due to Moroianu [13]
which is equivalent to the Axiom of Choice.
3.2 Knaster–Tarski Principle 55
Theorem 3.2.6 If every well-ordered subset A of a poset (X, ≤) has an upper bound
in X and f : X → X is a map such that x ≤ f (x) for all x ∈ X , then f has a fixed
point in X .
At this stage, it is convenient to introduce the following remark and a definition.
Remark 3.2.7 A map f : X → X , where (X, ≤) is a poset is called progressive or
expansive if x ≤ f (x) for all x ∈ X .
Definition 3.2.8 Let (X, ≤) be a poset and W be the set of well-ordered subsets of
X . If A ∈ W and x ∈ A the initial segment defined by x in A is the set A x = {y ∈
A : y < x}.
Definition 3.2.9 Let F : W → X be a map, where (X, ≤) is a poset and W , the set
of all well-ordered subsets of X . A well-ordered subset A of X is called an F-chain
if for each x ∈ A which is not the least element of A, f (A x ) = x.
We prove Theorem 3.2.6, making use of the following lemmata.
Lemma 3.2.10 For two distinct well-ordered subsets A and B of a poset (X, ≤),
the following statements are equivalent:
(i) one of A, B is an initial segment of the other;
(ii) if x ∈ A and y ∈ B are such that A x = B y then x = y.
Proof Let A be an initial segment of B. Let y1 = infimum B − A. Then A = B y1 . Let
x ∈ A and y ∈ B such that A x = {a ∈ A : a < x} = {b ∈ B : b < y} = B y . Since
A = B y1 , A x = {b ∈ B : b < x and b < y1 } = {b ∈ B : b < x ∗ } = Bx ∗ = B y . So
x ∗ = y. Hence x ∗ = min(x, y1 ) = y. Since each a ∈ A is less than y1 ∈ B − A and
x ∈ A, x ∗ = x. Thus x = y.
Suppose for two distinct well-ordered sets A and B there exist x in A and y
in B such that A x = B y . Clearly A and B have the same least element. Define
S = {x ∈ A ∩ B : A x = B y }. S is nonempty as it contains the least element of A and
B. Further, for any x ∈ S, S contains A x as also Bx . So S = A or A x for some x in
S. Since A and B are distinct, one can suppose that S = A. So S = A x for some x
in A. For similar reasons S = B or B y for some y in B. Thus S = A x = B y . Now
by hypothesis (ii) x = y and this means that A x = Bx and by the definition of S,
x ∈ S = A x implying that x < x. So A x = S = B. Thus (i) is true.
From the proof that (ii) implies (i) in Lemma 3.2.10 we have
Corollary 3.2.11 If A and B are two F-chains with the same least element then one
of them is an initial segment of the other.
Lemma 3.2.12 If τ (a) is the family of F-chains, having a ∈ X as the least element,
then C = ∪{A : A ∈ τ (a)} is an element of τ (a).
Proof By Corollary 3.2.11, C is a well-ordered subset of X with a as the least ele-
ment. If x ∈ C, then a ∈ A for some A ∈ C(a) and A x = C x . So f (C x ) = x from
the definition of A as an F chain. This implies that τ itself is an f chain.
56 3 Fixed Points and Order
Theorem 3.2.14 Let X be an ordered set such that every well-ordered subset A of X
has an upper bound in X . Let f : X → X be a (progressive) map such that x ≤ f (x)
for all x ∈ X . Then f has a fixed point.
Remark 3.2.15 Since each A ∈ W has an upper bound, U the set of upper bounds
of A is nonempty. For each A ∈ W , we can choose one from U A using the axiom
of choice. In case each A ∈ W has a supremum in X , then F(A) = f (sup A) is a
well-defined map of W into X (and there is no need to invoke the axiom of choice).
However, by invoking the Axiom of Choice to X it is clear that X has a maximal
element which will, of course, be a fixed point of f , the last element of C. Theo-
rem 3.2.14 is equivalent to the Axiom of Choice, as observed by Abian [2].
Theorem 3.2.16 (Abian [2]) The Axiom of Choice is equivalent to the statement
that every (progressive) map f : X → X , where (X, ≤) is a poset wherein every
well-ordered subset has an upper bound and x ≤ f (x) for all x ∈ X , has a fixed
point.
Proof The use of the Axiom of Choice in the proof of the fixed point theorem,
i.e. Theorem 3.2.14 was already noted. (Zorn’s Lemma, equivalent to the Axiom of
Choice can be invoked to obtain a maximal element z 0 and as z 0 ≤ f (z 0 ), z 0 is a
fixed point of f under the hypotheses of Theorem 3.2.14).
We now show that this fixed point theorem implies the Axiom of Choice. Let S
be any set and (W, ≤) be the set of all well-ordered subsets of S which are partially
ordered initial segment wise (i.e. W1 ≤ W2 if W1 is an initial segment of W2 , where
W1 , W2 ∈ W ). Let (N, ≤) be the set of all natural numbers with the usual order. Let
(W × N, ≤) be the cartesian product of W with N partially ordered lexicographically
(i.e. (W1 , n 1 ) ≤ (W2 , n 2 ) if W1 ≤ W2 and W1 = W2 and if W1 = W2 , n 1 ≤ n 2 . Define
f : W × N → W × N by f (w, n) = (w, n + 1). Clearly x ≤ f (x) for all x ∈ W ×
N and f has no fixed point. In view of Theorem 3.2.14 (W × N, ≤) has a well-ordered
subset without an upper bound (I).
However, every well-ordered subset of W has an upper bound. On the other hand,
not every (well-ordered) subset of N has an upper bound. (I) is impossible unless
(W, ≤) has a maximal element. This implies that S itself is a well-ordered set.
3.2 Knaster–Tarski Principle 57
Proof The set A = {x ∈ X : a ≤ x ≤ f (x) for all f ∈ F } is a poset with the partial
order inherited from (X, ≤) and every chain in A has a supremum in X . Since a ≤ x
for x ∈ A and f is isotone, f (a) ≤ f (x). As a ≤ f (a), we have a ≤ f (a) ≤ f (x)
for all x ∈ A and f ∈ F . For g ∈ F , a ≤ x ≤ f (x) implies a ≤ g(a) ≤ g(x) ≤
g( f (x)) = f g(x) for all f ∈ F and x ∈ A. So each g ∈ F maps A into itself.
Further every chain C in A has a supremum s in X and as a ≤ x ≤ s for x ∈ C,
a ≤ f (a) ≤ f (x) ≤ f (s) and x ≤ f (x) ≤ f (s) for each f ∈ F and x ∈ C. So
sup C = s ≤ f (s). Thus C has an upper bound in A. So there is a maximal element
x0 in A (by Zorn’s Lemma). Thus a ≤ x0 ≤ f (x0 ) for all f ∈ F . By the maximality
of x0 , it follows that x0 = f (x0 ) for all f ∈ F . Hence, there is a common fixed point
for all the functions in F .
The following well-known definitions figure in the original version of Tarski’s fixed
point theorem.
Definition 3.3.1 A poset (X, ≤) is called a lattice if it contains the minimum and
maximum of every pair of elements.
A lattice is said to be complete if every nonempty subset S of X has a supremum
and an infimum in X .
Example 3.3.2 (a) R, the set of all real numbers is a lattice in the usual order.
Though it is totally ordered, R is not a complete lattice. While N, the set of
natural numbers is well-ordered, R is not well-ordered.
58 3 Fixed Points and Order
(b) C[0, 1], the set of all continuous real-valued functions on [0, 1] with the partial
order defined by f ≺ g if and only if f (x) ≤ g(x) for all x ∈ [0, 1] is a lattice.
For f, g ∈ C[0, 1], min{ f, g} or max{ f, g} need not coincide with either f or
g. For instance for f (x) ≡ x, g(x) ≡ 21 for x ∈ [0, 1],
x, x ∈ [0, 21 ]
min{ f, g}(x) = 1
2
, otherwise
1
,x ∈ [0, 21 ]
max{ f, g}(x) = 2 .
x, otherwise
However, these are different from both f and g; That C[0, 1] is not a complete
lattice is left as an exercise.
(c) Let I be the closed unit interval in R and for n ∈ N, n > 1, let I n be the n-
fold cartesian product of I with itself. Define the partial order ≺ on I n be x =
(x1 , x2 , . . . , xn ) y = (y1 , y2 , . . . , yn ) for x, y ∈ I n if and only if xi ≤ yi for
all i = 1, 2, . . . , n. Though I n is not totally ordered it is a complete lattice for
n > 1.
(d) For a nonempty set X , 2 X , the power set of X with set-inclusion as a partial order
is a complete lattice.
Proof Let 0 = inf X and 1 = sup X and these exist as X is a complete lattice. So
v = sup{x ∈ X : f (x) ≥ x} is well-defined as f (0) ≥ 0. For any x ∈ X with x ≤
f (x), x ≤ u and by the isotonicity of f , f (x) ≤ f (u) and so x ≤ f (x) ≤ f (u).
As x ≤ u, for x ≤ f (x), u ≤ f (u). Again by the isotonicity of f , f (u) ≤ f ( f (u)).
So f (u) ∈ {x ∈ X : x ≤ f (x)}. Since u = sup{x ∈ X : x ≤ f (x)}, f (u) ≤ u. Thus
u = f (u) and u ∈ F f . Since for every fixed point x of f , f (x) ≥ x, u = sup F f .
The set X with the partial order , defined by x y if y ≤ x, called the dual-
order of ≤ is a complete lattice on which f is again an isotone map (in the dual
order). Further for any subset S of X , inf (≤) S = sup( ) S and sup(≤) S = inf ( ) S
where is the dual-order of the partial order ≤. So, by the first part of the theorem
proved so far, inf{x ∈ X : f (a) ≤ x} = sup{x ∈ X : x f (x)} is in F f . Clearly
inf F f = inf{x ∈ X : f (x) ≤ x}.
Let Y be any subset of F f . Then ([sup Y, 1], ≤) is a complete lattice, where
[a, b] = {x ∈ X : a ≤ x ≤ b}. If x ∈ Y , then x ≤ sup Y and so x = f (x) ≤
f (sup Y ) as f is isotone and Y ⊆ F f . So sup Y ≤ f (sup Y ). If sup Y ≤ z, then
sup Y ≤ f (sup Y ) ≤ f (z). So f , the restriction of f to [sup Y, 1] maps the com-
plete lattice [sup Y, 1] into itself and is isotone. So v = inf of all fixed points of f
(and indeed f ) in [sup Y, 1] is itself a fixed point of f by the preceding part of
the theorem. This is, indeed, the least fixed point of f , which is an upper bound of
3.3 Tarski’s Lattice Theoretical Fixed Point Theorem and Related Theorems 59
all elements of Y . Similarly, consider the complete lattice [0, inf Y ]. If x ∈ Y , then
inf Y ≤ x and f (inf Y ) ≤ f (x) ≤ x, as f is isotone and so f (inf Y ) ≤ inf Y . Thus
f ∗ , the restriction of f to [0, inf Y ] maps this complete lattice into itself. So u = sup
of all the fixed points of f ∗ in [0, inf Y ] is itself a fixed point and indeed the greatest
lower bound of all fixed points of f in Y . Thus, inf Y and sup Y are also fixed points
of f . Thus, (F f , ≤) is a complete lattice.
Remark 3.3.4 The existence of a fixed point in Tarski’s theorem can be proved along
the lines of Theorem 3.2.2 as well.
The next theorem is also due to Tarski [18] and its proof is left as an exercise.
Remark 3.3.6 A decreasing map on a complete lattice may not have a fixed point.
For example, the map f : [0, 1] → [0, 1] defined by
0, if x = 0
f (x) =
1, if x = 0
Remark 3.3.7 Theorem 3.3.5 is not true for a non-commutative family of isotone
maps. For instance, define f : [0, 1] → [0, 1] by f (x) = 1+x
2
and g : [0, 1] → [0, 1]
by g(x) = 2 for all x ∈ [0, 1]. While both f and g are isotone and non-commuting,
1
Remark 3.3.8 Any non-decreasing map of [0, 1] into itself (being isotone) will
always have a fixed point, even if it is not continuous. Contrast this with Theo-
rem 2.1.5 and Corollary 3.1.7.
Remark 3.3.9 The fixed point guaranteed by Tarski’s theorem is not unique. For the
map f : [0, 1] → [0, 1] defined by
x, x ∈ [0, 21 )
f (x) =
1, x ≥ 21 .
F f = [0, 21 ) ∪ {1} is the set of fixed points. Although 21 = sup[0, 21 ) in [0, 1], as a
subset Y = [0, 21 ) of the set of fixed points F f of f the supremum is 1!. F f is a
complete lattice.
60 3 Fixed Points and Order
Davis [7] had proved that the converse of Tarski’s theorem that a lattice in which
every isotone map has a fixed point is complete.
Based on Tarski’s observation that inf{x ∈ X : f (x) ≤ x} is a fixed point for f ,
an isotone self-map on a complete lattice X , Merrifield and Stein and Stein [12, 17]
proved some fixed point theorems supplementing/generalizing Tarski’s fixed point
theorem in complete lattices. A few of these results are highlighted in the sequel, as
these embellish the crux of the proof of Tarski’s theorem.
Remark 3.3.11 The transposition T (0) = 1 and T (1) = 0 on the lattice {0, 1} has
no fixed point. But T satisfies all the conditions of Theorem 3.3.10 for k = 2, n = 1
and hence T 2 has a fixed point.
If lim inf(xn ) = lim sup(xn ) = x, then (xn ) is said to converge to x (with respect
n→∞ n→∞
to the partial order ≤).
Theorem 3.3.14 Let (X, ≤) be a complete chain and T : X → X be a map such
that x ≤ y, x, y ∈ X implies that there exists a natural number N = N (x, y) such
that for all n ≥ N , T n (x) ≤ T n (y). Then T has a fixed point.
Proof Suppose T has no fixed point. Then the sets A = {x ∈ X : x ≥ T (x)} and
B = {x ∈ X : x ≤ T (x)} are disjoint and nonempty as 1 ∈ A and 0 ∈ B.
Let x ∈ A. If for some natural number p, T p (x) ≤ T p+1 (x), then by hypothe-
sis there exists N1 such that for n ≥ N1 , T p+n (x) ≤ T p+n+1 (x). As T x ≤ x, there
exists N2 such that for n ≥ N2 , T n+1 (x) ≤ T n (x). So for n ≥ max{N1 + p, N2 },
T n+1 (a) ≤ T n (x) ≤ T n+1 (x) or T n+1 (x) = T n (x). Thus T has a fixed point, viz.,
T n (a). Since we have assumed that T has no fixed point, for x ∈ A, T p (x) ≥ T p+1 (x)
for all p so that lim inf T n (x) ≤ x. By a similar reasoning lim inf T n (x) ≥ x for
n→∞ n→∞
x ∈ B. Let a = inf A. So for x ∈ A, a ≤ x. Hence by hypothesis, there exists N
such that for all n ≥ N , T n (a) ≤ T n (x). So lim inf T n (a) ≤ lim inf T n (x) ≤ x
n→∞ n→∞
for each x ∈ A. So lim inf T n (a) ≤ a = inf A.
If a ∈ B, then lim inf T n (a) = a implies a = T (a) = T 2 (a) . . . and T has a
n→∞
fixed point.
If T (a) < a, then T (a) ∈ A. Since a > T (a) ≥ T 2 (a) . . . , contradicting that a
is a lower bound for A. So T (a) = a. Thus T has a fixed point.
Remark 3.3.15 In a complete chain an operator T such that lim inf T n (x) ≤
n→∞
lim inf T n (y) whenever x ≤ y may not have a fixed point. The map T : [0, 1] →
n→∞
[0, 1] defined by
x
, for x = 0
T (x) = 2
1, for x = 0
Theorem 3.3.16 (Merrifield and Stein [12]) Let (X, ≤) be a complete lattice and
S, T : X → X be isotone maps. Suppose for each x ∈ X , there exist positive integers
p = p(x), q = q(x), i = i(x) and j = j (x) (depending on x) such that S p (x) ≤
T q (x) and T i (x) ≤ S j (x). Then S and T have a common fixed point.
The following is a result akin to the corollary above, also due to Merrifield and
Stein [12].
Theorem 3.3.18 Let S and T be isotone maps on the complete lattice (X, ≤) into
itself. If ST S n = T for some natural number n, then S and T have a common fixed
point.
While a decreasing or antitone self-map on a complete lattice may not have a fixed
point (vide Remark 3.3.6) Roth [14] noted that such a map has a fixed point under
additional assumptions. We conclude this section by providing two such results for
such maps, using the following.
Proof Since h is an isotone map on the complete lattice X , Fh the set of fixed points
of h in X is non-void and x0 = inf{x ∈ X : h(x) ≤ x} is in Fh , as this is the smallest
fixed point of h, in view of Tarski’s Theorem 3.3.3. Since x0 = inf Fh and g maps
Fh into itself h(x0 ) = x0 = inf Fh ≤ g(x0 ) as x0 ∈ Fh and x0 ≤ g(x) for all x ∈ Fh .
Thus, there is an element x0 with x0 = h(x0 ) and x0 ≤ g(x0 ).
Theorem 3.3.25 (Blair and Roth [4]) Let (X, ≤) be a complete lattice and u 1 , u 2 :
X → X be maps such that f = u 1 ◦ u 2 and g = u 2 ◦ u 1 are isotone. Then there exist
x, y ∈ X such that x = f (x) ≤ u 1 (y) and y = g(y) ≤ u 2 (x).
Proof Since f and g are isotone and X is a complete lattice F f and Fg the sets of fixed
points of the maps f and g, respectively, are nonempty and x = inf F f ∈ F f and
y = inf Fg ∈ Fg , by Tarski’s fixed point Theorem 3.3.3. So x = u 1 u 2 (x). So u 2 (x) =
u 2 u 1 (u 2 (x)) = g(u 2 (x)). So u 2 (x) ∈ Fg , as g = u 2 ◦ u 1 . By definition of y and g,
y = g(y) = u 2 u 1 (y) ≤ u 2 (x). Similarly, as y = u 2 u 1 (y), u 1 (y) = u 1 (u 2 u 1 (y)) =
(u 1 u 2 )(u 1 (y)) = f (u 1 (y)). Hence, u 1 (y) ∈ F f , f being u 1 ◦ u 2 . From the definition
of x, x = f (x) ≤ u 1 (y).
Remark 3.3.26 Theorem 3.3.25 and Corollary 3.3.24 have found applications in
Game Theory.
Tarski’s fixed point theorem central to lattice theoretical fixed point theorems is of
wide applicability. In this section, we use it to prove Schroder–Bernstein theorem a
basic result in set-theory, Cantor–Bendixon theorem, a representation theorem for
closed sets in general topology and existence theorems for a Cauchy problem for
a parabolic partial differential equation, a nonlinear complementarity problem and
certain formal languages.
Proof Let 2 A and 2 B be the power sets (sets of all subsets) of A and B, respectively.
2 A (and for that matter 2 B as well) is a complete lattice under set-inclusion as partial
order (see Example 3.3.2(d)). Define T : 2 A → 2 B by T (S) = A − g(B − f (S)) for
each subset S of A. For A1 , A2 ⊆ A and A1 ⊆ A2 , g(B − f (A1 )) ⊇ g(B − f (A2 ))
and so A − g(B − f (A1 )) ⊆ A − g(B − f (A2 )). In short T (A1 ) ⊆ T (A2 ) or T is
isotone on 2 A . So by Tarski’s fixed point Theorem 3.3.3, T has a fixed point S ∗ ⊆ A.
Thus S ∗ = T (S ∗ ) = A − g(B − f (S ∗ )).
Define h : A → B by
f (x), if x ∈ S ∗ ,
h(x) =
g (x), if x ∈ A − S ∗
−1
3.4 Some Applications 65
h(A) = h(S ∗ ∪ A − S ∗ )
= h(S ∗ ) ∪ h(A − S ∗ )
= f (S ∗ ) ∪ g −1 (g(B − f (S ∗ ))
= f (S ∗ ) ∪ B − f (S ∗ )
= B.
Remark 3.4.2 Schroder–Berstein theorem is quite useful in proving that two sets
have the same cardinality, for example (0, 1) ∪ (2, 3) and [−1, 0].
Example 3.4.5 For R with the usual topology, [0, 1] is a perfect set. So is the Cantor-
ternary subset of [0, 1]. N, the set of natural numbers is scattered. If X is a T1 -space,
then the derived set (of a set) is closed.
Theorem 3.4.7 (Schäfer [15]) Let E be a real Banach space and g : R × [0, T ] →
E be a bounded continuous function such that
for x, y ∈ R for each t ∈ [0, T ], an interval of real numbers, for some positive
constant L.
Then the function u defined by
t ∞
1 (x−ξ )2
u(x, t) = √ e− 4(t−τ ) g(ξ, t)dξ dt
0 −∞ 4π(t − τ )
Remark 3.4.8 Schäfer [15] observed that the above theorem can be proved along the
lines of proof of an existence theorem in Friedman [8].
We will use the above theorem to prove the following application of Tarski’s
theorem to the Cauchy problem for a parabolic equation.
Theorem 3.4.9 Let f : R × [0, T ] × ∞ (A) → ∞ (A) be a function with the fol-
lowing properties, A being a nonempty set, and ∞ (A) the Banach space of bounded
real functions on A.
(i) f is continuous;
(ii) for a constant L 1 > 0, for all (x, t, z), (y, t, z) ∈ R × [0, T ] × ∞ (A)
(iii) for some L 2 > 0, for all (x, t, z 1 ), (x, t, z 2 ) in R × [0, T ] × ∞ (A)
f (x, t, z 1 ) − f (x, t, z 2 ) ≤ L 2 z 1 − z 2 ;
(iv) for z 1 , z 2 ∈ ∞ (A) for (x, t) ∈ R × [0, T ] z 1 (a) ≤ z 2 (a) for all a ∈ A implies
f (x, t, z 1 ) ≤ f (x, t, z 2 );
(v) for a constant M > 0, for all (x, t, z) in R × [0, T ] × ∞ (A), f (x, t, z) ≤
M.
3.4 Some Applications 67
f (x, t, w(x, t)) − f (y, t, w(y, t)) ≤ f (x, t, w(x, t)) − f (y, t, w(x, t))
+ f (y, t, w(x, t)) − f (y, t, w(y, t))
≤ L 1 |x − y| + L 2 w(x, t) − w(y, t)
≤ (L 1 + L 2 L 3 )|x − y|. (vi)
are satisfied.
We now show that is a complete lattice and φ maps into itself and φ is isotone.
Since ∈ , is nonempty. Let S be a subset of which is non-void. For each s ∈ S,
−T M ≤ s(x, t)(a) ≤ T M for all a ∈ A and (x, t) ∈ R × [0, T ]. sup s(x, t)(a) for
a∈A
(x, t) ∈ R × [0, T ] lies between −T M and T M. Further, as s(x, t) − s(y, t ) ≤
L 3 |x − y| + L 4 t − t |, sup s(x, t) − sup s(y, t ) ≤ L 3 |x − y| + L 4 t − t |.
S S
Thus sup S ∈ . By a similar reasoning inf S ∈ . Hence is a complete lattice.
For w ∈ , since − ≤ w(x, t) ≤ it follows that − ≤ φ(w)(x, t) ≤ .
Further for s < t, s, t ∈ [0, T ]⎡φ(w(x, t)) − φ(w(x, s)) ≤⎤J1 + J2 , where J1 =
(x−ξ )2 (x−ξ )2
s ∞
1 e− 4(t−τ ) e− 4(s−τ ) ⎦
√ f (ξ, τ, w(ξ, τ ) ⎣ √ −√ dξ dτ and J2 =
0 −∞ π 4π(t − τ ) 4π(s − τ )
t ∞
1 (x−ξ )2
√ e− 4(t−τ ) f (ξ, τ, w(ξ, τ ))dξ dτ . Clearly J2 ≤ M|t − s| as
s −∞ 4π(t − τ )
68 3 Fixed Points and Order
So J1 + J2 ≤ {M + (L 1 + L 2 L 3 )2 T
π
}|t − s|. Thus φ(w)(r, t) − φ(w)(x, s) ≤
L 4 |t − s|. Also for each a ∈ A, φ(w(a))(x, s) − φ(w(a))(y, s) = ∂∂x φ(w)(a)(η, s)
(x − y) by the mean value √theorem, where η lies between x and y. So |φ(w)(a)(x, s)
− φ(w(a))(y, s)| ≤ 2M T |y − s| = L 3 |y − s|. Thus φ(w)(x, t) − φ(w)(y, s)
≤ L 3 |x − y| + L 4 |t − s|. Clearly − ≤ φ(w)(a) ≤ for w ∈ . So φ maps into
itself and is isotone. As is a complete lattice by Tarski’s fixed point Theorem 3.3.3
φ has a fixed point u. Clearly u satisfies (vii) and is thus a solution to the Cauchy
problem.
where b is a given element in L p [a, b] and T maps K (the cone in L p [a, b]) into
L p [a, b]. We have the following:
References
1. Abian, A.: Fixed point theorems of the mappings of partially ordered sets. Rend. Circ. Mat.
Palermo 20, 139–142 (1971)
2. Abian, A.: A fixed point theorem equivalent to the axiom of choice. Arch. Math. Log. 25,
173–174 (1985)
3. Andres, J., Pastor, K., Snyrychova, P.: Simple fixed point theorems on linear continua. Cubo
10, 27–43 (2008)
4. Blair, C., Roth, A.E.: An extension and simple proof of a constrained lattice fixed point theorem.
Algebra Universalis 9, 131–132 (1979)
5. Bourbaki, N.: Sur le theoreme de Zorn. Arch. Math. 2, 434–437 (1949–1950)
6. Chitra, A., Subrahmanyam, P.V.: Remarks on a nonlinear complementarity problem. J. Optim.
Theory Appl. 53, 297–302 (1987)
7. Davis, A.: A characterization of complete lattices. Pac. J. Math. 5, 311–319 (1955)
8. Friedman, A.: Partial Differential Equations of Parabolic Type. Prentice-Hall Inc., Englewood
Cliffs (1964)
9. Fujimoto, T.: Nonlinear complementarity problems in a function space. SIAM J. Control Optim.
18, 621–623 (1980)
10. Kelley, J.L.: General Topology. Springer, Berlin (1975)
11. Knaster, B.: Un theoreme sur les functions d’ensembles. Ann. Soc. Polon. Math. 5, 133–134
(1927–1928)
12. Merrifield, K., Stein, J.D.: Common fixed points of two isotone maps on a complete lattice.
Czechoslov. Math. J. 49(124), 849–866 (1999)
References 71
13. Moroianu, M.: On a theorem of Bourbaki. Rend. Circ. Math. Palermo 20, 139–142 (1971)
14. Roth, A.E.: A lattice fixed-point theorem with constraints. Bull. Am. Math. Soc. 81, 136–138
(1975)
15. Schäfer, U.: An existence theorem for a parabolic differential equation in ∞ (A) based on the
Tarski fixed point theorem. Demonstr. Math. 30, 461–464 (1997)
16. Stein, J.D.: Fixed points of inequality preserving maps in complete lattices. Czechoslov. Math.
J. 49(124), 35–43 (1999)
17. Stein, J.D.: A systematic generalization procedure for fixed point theorems. Rocky Mt. J. Math.
30, 735–754 (2000)
18. Tarski, A.: A lattice theoretical fix point theorem and its applications. Pac. J. Math. 5, 285–309
(1955)
19. Zermelo, E.: Neuer Beweis fur die Moglichkeit einer Wohlordnung. Math. Ann. 15, 107–128
(1908)
20. Zorn, M.: A remark on method in transfinite algebra. Bull. Am. Math. Soc. 41, 667–670 (1935)
Chapter 4
Partially Ordered Topological Spaces
and Fixed Points
A partial order on a set induces a natural topology on this set, and special properties of
the partial order influence this topology significantly. These aspects lead to new and
interesting fixed point theorems. The interconnections among partial order, topology
and fixed point property were systematically investigated by Wallace [11], Ward [12]
and Manka [6]. This chapter highlights these contributions to fixed point theory and
supplements the theorems detailed in the preceding chapter.
This section is a precis of the fundamental contributions of Ward Jr. [12], although
Wallace [11] had already pointed out the importance of partial order and the induced
topology for fixed point theorems. Using the definition of a quasi-ordered set (see
Definition 1.1.10), we define certain sets as in the following.
L(A) is called the set of predecessors of A and M(A), the set of successors of A.
We merely prove the last part of this lemma, leaving the other parts as exercises.
Proof Let (X, ≤) be a POTS with a continuous partial order. Let a and b be two
distinct points. If a b, by Lemma 4.1.8, there are disjoint neighbourhoods of a and
b. If a ≤ b then b a and again by Lemma 4.1.8, there are disjoint neighbourhoods
of b and a.
Theorem 4.1.11 (Wallace [11]) Every maximal chain in a QOTS is a closed set.
For the following basic theorem due to Wallace [11], a proof due to Ward is
sketched.
Proof Let L = {L(x) : x ∈ X }. L can be partially ordered with respect to set inclu-
sion. Clearly, if {L(xλ ) : xλ ∈ X, λ ∈ } is a chain in L, L(xλ ) is the intersection
λ∈
of
a family of closed sets with finiteintersection property in the compact space X . So
L(xλ ) is non-empty. Let x0 ∈ λ∈ L(xλ ). So L(x0 ) is a lower bound for each
λ∈
L(xλ ), λ ∈ . So by Zorn’s Lemma L has a maximal element L(x0 ). If x ≤ x0 and
x = x0 , then L(x ) ≥ L(x0 ) contradicting the maximality of L(x0 ). So x ∈ L(x0 )
or x0 ≤ x . Thus x0 is a minimal element. The proof for the upper semicontinuous
case is similar and left as an exercise, with the hint that one has to consider the closed
set M(x) instead.
The above theorem leads to an interesting proof of a theorem, due to Moore, on the
existence of non-cutpoints of a continuum. A few requisite definitions are recalled.
Remark 4.1.16 Generally, metrizable continua are studied in detail. Apart from those
continua defined in Definition 4.1.14, there are various classes of continua such as
tree-like continua, Peano continua and so on. Fixed point property for continuous
functions on such continua is a topic of active research.
For [a, b] in R, every interior point is a cutpoint while no point of the rectangle
[a, b] × [c, d] or a circle is a cut point.
The machinery of POTS developed so far can be used to prove the following
theorem due to Moore. The proof is due to Ward [12].
Theorem 4.1.17 (Moore, See Wilder [16]) A non-degenerate continuum has at least
two non-cutpoints.
Proof For the non-degenerate continuum X let N be the set of non-cut points. Sup-
pose N has at most one point. So there exists x0 ∈ X − N . As x0 is a cutpoint,
X − {x0 } = A ∪ B where A and B are non-empty separated sets. Without loss of
4.1 A Precis of Partially Ordered Topological Spaces 77
generality we can assume that N ⊆ B. So each point of A is a cut point. For each
x ∈ A, there is a decomposition X − {x} = A(x) ∪ B(x) where A(x) and B(x) are
non-void separated sets with x0 ∈ B(x) and so A(x) ⊆ A. We can define a partial
order on A by x ≤ y if and only if A(x) ⊆ A(y). Since A ∩ B(x) = φ, x ∈ A − A
cannot be a non-cutpoint different from x0 , possibly the only non-cutpoint lying in B.
So L(x) = A(x). Hence this partial order is lower semicontinuous. Since X is com-
pact, A is compact. So by Theorem 4.1.13, there is a minimal element p ∈ A. Since
p is minimal, A( p) is empty. Otherwise suppose q ∈ A( p). Then q ∈ A( p) = L( p).
So q ≤ p. Since p is a minimal element of A, q = p. So p ∈ A( p) and this contra-
dicts the fact that X − { p} = A( p) ∪ B( p). So A( p) is empty. This again contradicts
the construction that A(x) is non-empty for all x ∈ A. Hence our assumption, that
there is only one non-cutpoint, viz. x0 is wrong. So a non-degenerate continuum has
at least two non-cut points.
The concept of convexity given below is useful for the study of fixed point theo-
rems in QOTS.
Definition 4.1.18 A subset A of a QOTS is called convex if A = E(A). X is said to
be quasi-locally convex if for x ∈ X and E(x) ⊆ U , an open set, there is a convex
open set V such that E(x) ⊆ V ⊆ U . X is called locally convex, provided, for x ∈ X
and U an open set, there is a convex open set V with x ∈ V ⊆ U .
The next theorem due to Ward [12], an extension of a theorem of Nachbin on
compact POTS, is stated without proof.
Theorem 4.1.19 Let X be a compact QOTS with a continuous quasi-order and
b a, where a, b ∈ X . Then we can find a continuous order-preserving map f :
X → [0, 1] such that f (a) = 0 and f (b) = 1.
Using the above theorem, Ward proved
Theorem 4.1.20 A compact Hausdorff QOTS with a continuous quasi-order is
quasi-locally convex.
Proof Let X be a compact Hausdorff QOTS, x ∈ E(x) ⊆ U , where U is an
open subset of X . For t ∈ X − U , t x or x t. If t x, then by Theorem
4.1.19 above there is a continuous order-preserving map f t : X → [0, 1] with
f t (x) = 0 and f t (t) = 1. Let Ut = {y ∈ X : f t (y) < 21 }. Then Ut is a decreasing
open set containing x and t ∈ / U t . If x t, then Ut = {y ∈ X : 21 < gt (y)} is an
increasing open set containing x and t ∈ / U t , where gt : X → [0, 1] is a contin-
uous order-preserving map with gt (t) = 0 and gt (x) = 1, as guaranteed by The-
n
orem 4.1.19. Thus X − U ≤ U {X − U t : t ∈ X − U } with X − U ≤ X − U ti .
i=1
n
U= Uti is an open set containing x and disjoint from X − U . Since E(x) ⊆ Uti
i=1
as Uti is either increasing or decreasing, x ∈ E(x) ⊆ V ⊆ U . So X is quasi-locally
convex.
78 4 Partially Ordered Topological Spaces and Fixed Points
Corollary 4.1.21 A compact POTS with a continuous partial order is locally convex.
Definition 4.1.23 Let X be a topological space. A net (xλ , λ ∈ D), where (D, ≤1 )
is a directed set, is said to cluster at x0 ∈ X if for any open set U containing x0 and
λ ∈ D, there exists μ ∈ D with λ ⊆ μ such that xμ ∈ U .
Proof Without loss of generality, we assume that the net is monotone increasing.
Since X is compact, every net in X has a subset converging to some x0 ∈ X , and
hence clusters at x0 . This is also true of a monotone increasing net (xλ ). By Theorem
4.1.20 X is quasi-locally convex. So given an open set U containing E(x0 ), there is
an open convex set V such that E(x0 ) ⊆ V ⊆ U . Since (xλ ) clusters at x0 , xλ0 ∈ V
for some λ0 . For λ0 ≤1 λ, there exists λ with λ ≤1 λ such that xλ ∈ V (Here the net
is indexed over the directed set (D, ≤1 )). Since V is convex and xλ ∈ V , xλ ∈ V for
all λ ≥ λ . Suppose x ∈/ E(x0 ). Then there are disjoint open sets U1 and U2 such that
x ∈ U1 and x0 ∈ U2 , as X is Hausdorff. So U1 and V ∩ U2 are disjoint and hence (xλ )
cannot cluster at x as otherwise U1 and V ∩ U2 would intersect. Hence the cluster
points of (X λ ) are contained in E(x0 ).
Corollary 4.1.25 If X is a compact POTS with continuous order, then every mono-
tone net in X is convergent.
The following theorem gives a necessary and sufficient condition for a continuous
order-preserving map f on a Hausdorff QOTS so that there exists x in X that can
compare with f (x). This theorem is quite useful in obtaining fixed point theorems
in the setting of QOTS.
4.1 A Precis of Partially Ordered Topological Spaces 79
Theorem 4.1.28 (Ward [12]) Let X be a Hausdorff QOTS with compact maximal
chains and f : X → X , a continuous order-preserving map. A necessary and suf-
ficient condition that there exists a non-empty compact set K ⊆ E(x0 ) for some
x0 ∈ X such that f (K ) = K is that there exists x in X such that x and f (x) are
comparable.
Proof The necessity is obvious. Suppose x and f (x) are comparable. Then { f n (x) :
n ∈ N} is a chain, as f is order-preserving and is therefore contained in a maximal
chain, which by hypothesis is compact. So by Lemma 4.1.24 it clusters at some x0 and
all its cluster points are contained in E(x0 ). By Lemma 4.1.26 f (E(x0 )) ⊆ E(x0 ).
Define K = ∩{ f n (E(x0 )) : n ∈ N}. Then K is a non-empty compact subset of E(x0 )
and f (K ) = K .
Corollary 4.1.29 If X is a POTS with compact maximal chains and f : X → X is
a continuous order-preserving map, then a necessary and sufficient condition that f
has a fixed point is that there is an x ∈ X for which x and f (x) are comparable.
Proof If x = f (x), then x and f (x) are comparable. Conversely if x ≤ f (x) or
f (x) ≤ x, then { f n (x) : n ∈ N} is a chain which is contained in a maximal chain
which also clusters at some x0 ∈ X . Since f (E(x0 )) ⊆ E(x0 ) and E(x0 ) = {x0 }, X
being partially ordered x0 = f (x0 ).
It is possible to define the concept of boundedness in quasi-ordered spaces, as in
the following.
Definition 4.1.30 Let (X, ≤) be a quasi-ordered set with an element e ∈ X such
that e ≤ x for all x ∈ X . A subset A of X is said to be bounded away from e if there
is y ∈ X − E(e) with A ⊆ M(y).
Using this concept, some results on fixed points can be obtained.
Theorem 4.1.31 (Ward [12]) Let X be a Hausdorff QOTS with compact maximal
chains and suppose there exists e ∈ X such that e ≤ x for all x ∈ X . Let f : X → X
be a continuous order-preserving map satisfying
(i) for some x ∈ X − E(e), x and f (x) are comparable;
(ii) for x satisfying (i), either { f n (x) : n ∈ N} is bounded away from e or there
exists y ∈ X with x ∈ E( f (y)) and f (y) ≤ y.
Then there exists x0 ∈ X − E(e) and a non-void compact set K ⊆ E(x0 ) such that
f (K ) = K .
Proof Suppose (i) is satisfied for some x ∈ X − E(e). If { f n (x) : n ∈ N} is bounded
away from e, define K = ∩{ f n (E(x0 )) : n ∈ N} where { f n (x) : n ∈ N} clusters at
x0 and all its cluster points are in E(x0 ) (by Lemma 4.1.24) as in Theorem 4.1.28.
For this choice of K , f (K ) = K and K ⊆ E(x0 ). Clearly x0 ∈ / E(e).
If for no x satisfying (i), { f n (x) : n ∈ N} is bounded away from e, by (ii) there
is a y1 such that for x ∈ E( f (y1 )). f (y1 ) ≤ y1 . Inductively, (yn ) can be chosen for
n ≥ 2 by
80 4 Partially Ordered Topological Spaces and Fixed Points
x ≤ f (y1 ) ≤ y1 ≤ f (y2 ) ≤ y2 ≤ · · ·
Corollary 4.1.32 Let X be a POTS with compact maximal chains such that for some
e ∈ X , e ≤ x for all x ∈ X and f : X → X be a continuous order-preserving map
satisfying the following:
(i) for some x ∈ X − E(e), x and f (x) are comparable and
(ii) if x satisfies (i), then { f n (x) : n ∈ N } is bounded away from e or there exists
y ∈ X such that x ∈ E( f (y)) and f (y) ≤ y.
Then f has a fixed point different from e.
Corollary 4.1.33 Let (X, ≤) be a POTS which is Hausdorff and having compact
maximal chains and f : X → X an order-preserving continuous map. Suppose (i)
for some u ∈ X , L(u) = X and (ii) for x, y ∈ X , there exists z with x ≤ z, y ≤ z.
Then f has a fixed point.
Remark 4.1.36 Ward Jr. [12] has given an example to show that in a compact POTS
an order-preserving continuous surjection may not have a fixed point different from
e, without an additional hypothesis such as (ii) in Theorem 4.1.31.
4.2 Schweigert–Wallace Fixed Point Theorem 81
This section continues further the theory of partially ordered topological spaces
developed by Ward Jr. In particular, the concepts of end point and end element
and non-alternating maps and theorems relating to these are described. The section
culminates in the proof of Schweigert–Wallace fixed point theorem [9, 11] for home-
omorphisms on certain locally connected continua.
Remark 4.2.3 R, with the usual topology has no endpoint, while [a, b] of real num-
bers (a < b) has a and b as endpoints. On the other hand, the unit circle has no
endpoint.
We state below without proof, two results due to Wallace [10] for subsequent use.
Lemma 4.2.10 Let X be a locally connected continuum with an end element. Then
the relation ≤ defined on X by x ≤ y for x ∈ E, x = y or x separates E and y in X
is a semicontinuous quasi-order. If E is a singleton, then ≤ is a partial order.
4.2 Schweigert–Wallace Fixed Point Theorem 83
Proof The proof that ≤ is a quasi-order and that L(x) is closed for each x ∈ X can
be found in Whyburn [15].
If x ∈ E, M(x) = X and is closed. For x ∈ X − E, M(x) = {x} ∪ {y : x sep-
arates E and y in X }. Let C be the component of X − {x} containing E, then
M(x) = X − C. As X is locally connected and X − {x} is open, C is open. So
M(x) is closed.
f −m (x) ∩ U = φ = f −m (x) ∩ V.
4.2 Schweigert–Wallace Fixed Point Theorem 85
f (x, y, t) =
⎪
⎪ (0, 0, t), 1
≤ t ≤ 1,
⎩ 2
(0, 0, 2), 1≤t ≤2
Remark 4.2.18 Schweigert [9] originally proved the theorem under the assumption
that X was separable and semi-locally connected. Wallace [11] noted that these
hypotheses could be dropped in preference to local connectedness and also pointed
to the use of quasi-order in the proof. Wallace, in fact, proved a fixed point theorem
for homeomorphisms T such that T and T −1 preserve a transitive, reflexive binary
relation on the space X .
Remark 4.2.19 Wallace [11] has shown that if X has a cutpoint, then it has an end
element. He also noted that prime chains in Peano space are precisely cyclic elements,
considered by Whyburn [15]. Wallace [11] has further proved that if the end element
E of a continuum X has no cutpoint of X , then X − E is connected and if P is the
union of all the end elements of X not containing a cutpoint of X , then X − P is
connected and each component of P is an end element.
Remark 4.2.20 It was already stated that fixed points theorems generally prove that
for certain topological spaces every continuous self-map has a fixed point. Bing (see
[8]) has noted that for the proof of numerous fixed point theorems depend on the
‘dead-end method’ or the ‘dog-chase rabbit argument’. Roughly speaking for a given
mapping f , as x moves in X , f (x) moves ‘ahead’ of x ‘relative to some hidden order
structure’ till a special feature of the underlying space is exploited to locate a point
x below f (x) in the order and corner f (x) in a dead end. Ward Jr. [14] has captured
this ‘dead-end method’ in the following theorem, which is related to Theorem 4.1.31
and its corollaries.
In an insightful paper, Manka [6] described a connection between set theory and
fixed point theory via partial order. In this section, Manka’s approach is described,
based on the following definitions.
4.3 Set Theory, Fixed Point Theory and Order 87
Definition 4.3.1 A partial ordered set X is called inductive if for every totally
ordered subset of X there exists a least upper bound of this subset in X . Clearly
an inductive (partially ordered) set is non-empty. A partially ordered set X is called
acyclically ordered if for every p, r ∈ X with p ≤ r , the segment [ p, r ] = {x ∈ X :
p ≤ x ≤ r } is totally ordered.
Theorem 4.3.2 (Manka [6]) Let (X, ≤) be an inductively and acyclically ordered
poset and f : X → X be a map satisfying the following two conditions:
(i) p < f ( p) implies that for some q ∈ ( p, f ( p)] with q ≤ f (q);
(ii) q ≤ f (q) for all q ∈ Y ⊆ X implies that sup Y ≤ f (sup Y ).
Then f has a fixed point.
Manka used the above theorem to prove that certain compact connected topologi-
cal spaces have fixed point property for continuous functions. We need the following
concepts.
Definition 4.3.4 A continuum having exactly two points which do not disconnect it
is called an arc, including continua with only one point.
∪apτ = ab.
Remark 4.3.15 In Lemma 4.3.14, b = sup{ pt : t ∈ T } if and only if apt = ab.
t∈T
Clearly apt ⊆ ab for all t ∈ T so that apt ⊆ ab. But the closure of apt is a
t∈T t∈T
subarc ac of ab and apt ⊆ ac. So ac ⊆ ab. Hence c is an upper bound for { pt ∈ X :
t ∈ T }. Since b is the last upper bound, c = b.
In an arcwise connected continuum X , for every pair of arcs pq, pr with the same
initial point P, a binary relation ≺ can be defined by pq ≺ pr if pq ∩ pr = { p}.
In other words pq ≺ pr if pq ∩ pr is an arc non-degenerate to the point p. The
following lemma is left as an exercise.
Lemma 4.3.16 The binary relation ≺ is an equivalence relation in the family of all
non-degenerate arcs with the same initial point in X . If K is an arcwise connected
continuum of X and q, r ∈ K , then for p ∈/ K , pq ≺ pr as qr ⊆ K . Further ap ⊆
aq and pq ≺ pr imply ap ⊆ ar .
(For further details and related ideas Manka [5] may be consulted.)
Theorem 4.3.18 (Manka [6]) Every one arcwise connected nested continuum X with
the partial order ≤a is an inductively and acyclically ordered set. If f : X → X is
a map such that (a) f ( pq) is an arcwise connected continuum for each pq and (b)
if for each p = f ( p) there is { p} = pq ⊆ p f ( p) with pq ∩ f ( pq) = φ, satisfying
conditions (i) and (ii) of Theorem 4.3.2 in the order ≤a , then f has a fixed point.
In this section, we prove a fixed point theorem for dendroids following the techniques
developed by Manka, as described in Sect. 4.3. In fact, Manka’s fixed point theorem
subsumes that of Ward [13] and is proved in [7]. To this end, we need the following.
Definition 4.4.1 Let X be a topological space. A multifunction F : X → 2 X − {φ}
is called upper semicontinuous if F(x) is a closed set for each x ∈ X and F −1 (A) =
{x ∈ X : F(x) ∩ A = φ} is a closed subset of X for each closed subset A of X . F is
called lower semicontinuous, if F −1 (A) is open for each open subset A of X .
4.4 Multifunctions and Dendroids 91
Remark 4.4.2 For a compact metric space X , upper semicontinuity of F means that
limF(xn ) ⊆ F(lim xn ) for each convergent sequence (xn ) in X . Similarly, lower
semicontinuity of F : X → 2 X − {φ} means that F(lim xn ) ⊆ lim inf F(xn ) for
n n→∞
each sequence (xn ) converging in the compact metric space X . F is called continuous
if it is both lower and upper semicontinuous. Using the following version of Brouwer
reduction theorem, Manka [5] obtained an alternative proof of Ward’s fixed point
theorem for upper semicontinuous closed valued multifunctions on a dendroid.
Theorem 4.4.3 (Manka [5]) Every non-empty family P of closed subsets of a com-
pact metric space X which is closed with respect to the operation of closure of a
union of increasing sequences contains a maximal element.
We also recall the following results and their proofs are left as exercises.
Lemma 4.4.5 If a ∈
/ F(a), then for every d ∈ F(a), there exists ab ∈ Pa such that
ab ⊂ ad.
Proof Suppose d ∈ F(a) and that an arc ab ⊂ ad satisfies, by the upper semicon-
tinuity of F, ab ∩ F(ab) = φ. Now for each p ∈ ab − {b} and each q ∈ F( p), we
have p ∈ / F(ab) and d, q ∈ F(ab). Since F(ab) is a continuum (see Remark 4.4.4,
pq ≺ pd.
92 4 Partially Ordered Topological Spaces and Fixed Points
Lemma 4.4.6 If ab ∈ Pa , b ∈
/ F(b) and d ∈ F(b), then ab ⊂ ad.
Proof As b ∈/ F(b), it follows from the upper semicontinuity of F that for some p ∈
ab − {b}, p b ∩ F( p b) = φ. For each p ∈ p b − {b}, p ∈ ab − {b}. Hence ap ⊂
Proof For p ∈ ab − {b}, there exists n ∈ N such that p ∈ abn − {bn }. So pbn ≺ pb
since abn ⊂ ab and pq ≺ pbn for every q ∈ F( p), in view of abn ∈ Pa . From the
reflexivity of ≺, we have pq ≺ pb.
Theorem 4.4.9 (Ward Jr. [13]) If F : X → 2 X − {φ} is a continuum-valued upper
semicontinuous multifunction on a dendroid X , then F has a fixed point.
Proof (As in Manka [7]) If ab1 ⊂ ab2 ⊂ · · · is an increasing sequence of arcs in the
dendroid X , then abn is an arc ab for some b. Since abn is a continuum, for
n∈N n∈N
a proper subcontinuum this continuum containing a some b j will not be a member.
So abn is a continuum which is not the union of two proper subcontinua both
n∈N
containing a. So a is a point of irreducibility of abn . Thus abn = ab for some
n∈N n∈N
b. So by Lemmata 4.4.5, 4.4.8 and Theorem 4.4.3, for a ∈
/ F(a), there exists an arc ab
maximal in Pa . Now from Lemmata 4.4.5–4.4.7 it follows that b ∈ F(b). Otherwise,
if b ∈
/ F(b) then ab ⊂ ad for each d ∈ F(b) and by Lemma 4.4.5 there exists bc ⊂
bd such that bc ∈ Pb . So ab ∪ bc = ac ∈ Pa by Lemma 4.4.7 contradicting the
maximality of ab in Pa .
Corollary 4.4.10 (Borsuk [2]) Every dendroid has fixed point property for contin-
uous functions.
In this section, elementary methods of constructing spaces with fixed point property
are described. Relevant concepts are also presented.
4.5 Some Spaces with Fixed Point Property 93
Proposition 4.5.2 If X has fixed point property (for continuous functions), then any
retract of X also has fixed point property.
Remark 4.5.4 The unit circle S 1 in R2 with the usual topology is a connected, locally
connected compact metric space without fixed point property. For example, (x, y) →
(−x, −y) on S 1 has no fixed point.
Theorem 4.5.5 Let (X, d) be a compact metric space. Suppose for each > 0,
there is a continuous map f : X → X , where X is a subset of X with fixed point
property. If d( f (x), x) < for each x ∈ X , then X has the fixed point property.
Yet another useful idea is that of wedge of two spaces, defined below
Definition 4.5.8 Let X and Y be two disjoint spaces and let p ∈ X and q ∈ Y . The
wedge of X and Y at p and q, denoted by X ∨ p,q Y (or simply X ∨ Y ) is the quotient
space of X ∪ Y obtained by identifying p with q. (Clearly X ∨ Y has a natural copy
of X in X ∨ Y .)
Theorem 4.5.9 Let X and Y be T1 spaces with fixed point property. Then X ∨ Y
has the fixed point property.
p, p ∈ X
r ( p) =
w, p ∈ Y
X and Y are closed in X ∨ Y and r is continuous. Since X has the fixed point property,
r ◦ f has a fixed point, p say. Now r f ( p) = p. If w = f (w), then there is nothing to
prove. So let w = f (w). So p = w. So r f ( p) = w. So by definition of r , f ( p) ∈ X .
So r f ( p) = f ( p) = p. (Note the continuity of r is based on the T1 -hypothesis).
For these and similar results and examples, Nadler [8] may be consulted.
Connell [3] had given examples of noncompact plane sets U, V and W each having
fixed point property such that cl W , U 2 lack fixed point property and V is locally
contractible. Klee [4] had given an example of a space combining all these features.
For the sake of completeness, we give the following definitions.
Definition 4.6.1 Let X and Y be topological spaces and I , the closed unit interval
[0, 1]. A homotopy in a continuous map h : X × I → Y . We write h t to denote the
map from X into Y defined by h t (x) = h(x, t) for all x ∈ X for any fixed t ∈ I . A
continuous map f : X → Y is said to be homotopic to a continuous map g : X → Y ,
if there exists a homotopy h : X × I → Y such that h 0 = f and h 1 = g.
Example 4.6.4 Let Y be the set of real sequences (yn ) in the Hilbert space 2 such
that yi is non-zero for at most one i and 0 ≤ yi ≤ 1. If θ is the zero sequence in
2
and δn the sequence in 2 which is one in the nth place and zero elsewhere then
∞
Y = σn , where σn is the line segment joining θ to δn (and so σn = [θ, δn ]). Clearly,
n=1
Y is both contractible and locally contractible.
For each n, let rn be the retraction of Y onto σn which is identity in σn and maps
Y − σn onto θ . Let f : Y → Y be a continuous map. Suppose f (θ ) = θ . So for
some n, f (θ ) ∈ σn − {θ }. As rn f maps σn onto itself and rn f (θ ) = f (θ ) = θ and
σn being essentially a compact real interval has the fixed point property, rn f ( p) = p
for some p ∈ σn − {θ }. Since rn f ( p) = θ , f ( p) ∈ σn and so rn f ( p) = f ( p). So
f ( p) = p. Thus Y has the fixed point property.
In the space 2 × 2 , let P be the infinite polygon with vertices in the order
(θ, δ1 ), (δ1 , θ ), (θ, δ2 ), (δ2 , θ ), . . . , (θ, δn ), (δn , θ ), . . .. Clearly, P is closed in Y ×
Y and P is homeomorphic with [0, ∞). As Y × Y would admit a retraction onto P,
and P lacks the fixed point property, Y × Y cannot have the fixed point property, in
view of Proposition 4.5.2.
For each t ∈ [0, π ] and n ∈ N consider τn the arc consisting of all points
(xn (t), yn (t)) where xn (t) = (−1)n (1 + nt ) cos t and yn (t) = (1 + nt ) sin t. Each arc
τn has (1, 0) as an end point and X , the union of all the arcs τn is a homeomorphic
of y. But cl X contains the unit circle C and has a retraction onto C. But C does not
have the fixed point property. So cl X does not enjoy the fixed point property.
References
14. Ward Jr., L.E.: Monotone surjections having more than one fixed point. Rocky Mt. J. Math. 4,
95–106 (1974)
15. Whyburn, G.T.: Analytic Topology. American Mathematical Society, Providence (1942)
16. Wilder, R.L.: Topology of Manifolds. American Mathematical Society, Providence (1949)
Chapter 5
Contraction Principle
The contraction mapping principle proved independently by Banach [1] and Cac-
ciopoli [7] is a fundamental fixed point theorem, with an elementary proof. This
theorem has a wide spectrum of applications and is a natural choice in approximat-
ing solutions to nonlinear problems. According to Rall [18], the applications of the
contraction principle would fill volumes and Bollabos [4] calls it a doyen of fixed
point theorems. Charmed by both the simplicity and utility of this theorem, many
authors have generalized it in diverse directions. This chapter samples a few of these.
A simple proof of the contraction principle due to Palais [16] is given below. This is
preceded by a few definitions and remarks.
Definition 5.1.1 Let (X, d) be a metric space. A map T : X → X is said to be
a Lipschitz map with Lipschitz constant M, if for some M ∈ R+ , d(T x, T y) ≤
Md(x, y) for all x, y ∈ X . In this case M is called a Lipschitz constant for the map
T . If M < 1, then T is called a contraction (mapping) with contraction constant M.
If M = 1, T is called a non-expansive map. If d(T x, T y) = d(x, y) for all x, y ∈ X .
T is a distance-preserving map and is called an isometry.
Palais [16] proved the contraction principle, using the following contraction
inequality.
© Springer Nature Singapore Pte Ltd. 2018 97
P. V. Subrahmanyam, Elementary Fixed Point Theorems,
Forum for Interdisciplinary Mathematics,
https://doi.org/10.1007/978-981-13-3158-9_5
98 5 Contraction Principle
1
d(x, y) ≤ [d(x, T x) + d(y, T y)]
1−k
This corollary follows at once from Lemma 5.1.4 by choosing x and y as fixed
points of T .
Lemma 5.1.6 (Estimate for iterates) Let (X, d) be a metric space and T : X → X ,
a contraction mapping with contraction constant k. Then for any x ∈ X ,
kn + km
d(T n x, T m x) ≤ d(x, T x)
(1 − k)
For j ∈ N,
kn + km
d(T n x, T m x) ≤ d(x, T x)
(1 − k)
Theorem 5.1.7 (Contraction Principle) Let (X, d) be a complete metric space and
T : X → X , a contraction mapping with contraction constant k. Then, T has a
5.1 A Simple Proof of the Contraction Principle 99
kn
d(x ∗ , T n x) ≤ d(x, T x).
1−k
km + kn
d(T m x, T n x) ≤ d(x, T x).
(1 − k)
Proceeding to the limit in the above inequality as m tends to ∞, and noting that
{T m x} converges to x ∗ = T x ∗ , we get
kn
d(x ∗ , T n x) ≤ d(x, T x).
1−k
∗
Remark 5.1.8 In order that T x is at a distance less than (> 0) from x the fixed
n
kn
point of T , it suffices to choose n such that 1−k d(x, T x) < . In other words for
log +log(1−k)−log d(x,T x) ∗
N> log k
, d(x , T x) < . Thus in a specific situation, the fixed
N
Proof Since T n is a contraction on the complete metric space (X, d), it has a unique
fixed point x ∗ , say. Now T n+1 (x ∗ ) = T (T n x ∗ ) = T (x ∗ ) = T (T n x ∗ ) = T n (T x ∗ ).
Thus T x ∗ is a fixed point of T n . Since T n has the unique fixed point x ∗ , x ∗ = T x ∗ .
If y ∗ is another fixed point of T , then T n (y ∗ ) = y ∗ will be a fixed point of T n and
hence y ∗ = x ∗ . Thus T has a unique fixed point.
100 5 Contraction Principle
Remark 5.1.11 A contractive map T which is not a strict contraction may not have
a fixed point in a complete metric space. For example, the map x → x + x1 maps
[2, ∞) into itself has no fixed point in [2, ∞) which is complete with respect to the
usual metric. For 2 ≤ x < y, 0 < (y + 1y ) − (x + x1 ) = (y − x)(1 − x1y ) < y − x
and consequently this map is contractive.
5
for all x, y ∈ X , where ai ≥ 0 for i = 1, 2, . . . , 5 and ai < 1.
i=1
Then T has a unique fixed point and every sequence of T -iterates converges to
the unique fixed point.
So we get
a2 + a3
d(T x, T y) ≤ a1 d(x, y) + [d(x, T x) + d(y, T y)]
2
a4 + a5
+ [d(x, T y) + d(y, T x)] (5.2.1)
2
Letting y = T x in (5.2.1), we have
5.2 Metrical Generalizations of the Contraction Principle 101
a2 + a3
d(T x, T 2 x) ≤ a1 d(x, T x) + [d(x, T x) + d(T x, T 2 x)]
2
a4 + a5
+ [d(x, T 2 x)]
2
a2 + a3 + a4 + a5
≤ a1 d(x, T x) + d(x, T x)
2
a2 + a3 + a4 + a5
+ d(T x, T 2 x) (5.2.2)
2
5
Since 0 ≤ ai < 1, 0 ≤ k < 1. Further, d(T n x, T n+1 x) ≤ k n d(x, T x) for n ∈ N
i=1
and x ∈ X . For n, j ∈ N and x ∈ X
j
n
d(T x, T n+ j
x) ≤ d(T n+i−1 x, T n+i x)
i=1
j
≤ k n+i−1 d(x, T x)
i=1
n
k
≤ d(x, T x).
1−k
(a2 + a3 + a4 + a5 + 2a1 )
d(x ∗ , T x ∗ ) ≤ d(x ∗ , T x ∗ )
2
5
Since 0 ≤ ai < 1, it follows that x ∗ = T x ∗ .
i=1
If y ∗ is also a fixed point of T , then
5
Since 0 ≤ a1 + a4 + a5 ≤ ai < 1, x ∗ = y ∗ . Thus T has a unique fixed point and
i=1
every sequence of T -iterates converges to the unique fixed point.
Corollary 5.2.2 (Kannan [13]) If T : X → X is a map on a complete metric space
(X, d) such that
d(T x, T y) ≤ k1 d(x, T x) + k2 d(y, T y)
for all x, y ∈ x with k1 , k2 ≥ 0 and k1 + k2 < 1, then T has a unique fixed point and
every sequence of T -iterates converges to the unique fixed point.
Proof Set a1 = a4 = a5 = 0, a2 = k1 and a3 = k2 in Theorem 5.2.1.
Remark 5.2.3 A mapping satisfying the conditions of Corollary 5.2.2 (or Theo-
rem 5.2.1) need not be continuous, as seen from the following example.
x
, 0 ≤ x ≤ 21
The map T : [0, 1] → [0, 1] defined by T x = x4 1 is a discontinuous
5
, 2 <x ≤1
map with 0 as the unique fixed point. For x, y ∈ [0, 1], it can be shown that
3
|T x − T y| ≤ [|x − T x| + |y − T y|].
8
Corollary 5.2.4 Theorem 5.1.7 (Contraction Principle).
Proof Set a1 = k, a2 = a3 = a4 = a5 = 0 in Theorem 5.2.1.
In another direction, Boyd and Wong [6] generalized the contraction principle by
majorizing d(T x, T y) by ψ(d(x, y)) instead of kd(x, y), imposing suitable assump-
tions on the real-valued function ψ of the real variable. In this context we recall the
following.
Definition 5.2.5 Let ψ : [a, ∞) → R be a function, where a ∈ R. ψ is said to be
upper semicontinuous from the right at c ∈ [a, ∞) if lim+ sup ψ(t) ≤ ψ(c).
t→c
5.2 Metrical Generalizations of the Contraction Principle 103
Theorem 5.2.6 Let (X, d) be a complete metric space and T : X → X , a map such
that for all x, y ∈ X
d(T x, T y) ≤ ψ(d(x, y))
where ψ : P → [0, ∞) is upper semicontinuous from the right on P and ψ(t) < t
for all t ∈ P and t = 0, P being the range of d and P its closure in R+ . Then, T
has a unique fixed point and every sequence of T -iterates converges to this unique
fixed point.
Proof For x ∈ X , define cn = d(T n x, T n−1 x), n ∈ N with T 0 x = x. Clearly cn is
non-increasing and non-negative and hence converges to c ≥ 0. Since cn+1 ≤ ψ(cn )
for all n ∈ N, for c > 0,
and
d(T m(k)−1 x, T n(k) x) < 0 .
This can be done by choosing m(k) as the least natural number exceeding n(k) for
which dk ≥ 0 . Now
Since 0 > 0, this contradicts that 0 > ψ(0 ). Hence {T n x} is a Cauchy sequence
in X . As X is complete, it converges to an element x ∗ in X . Since for all x, y ∈ X ,
d(T x, T y) ≤ φ(d(x, y)) ≤ d(x, y), T is continuous. Since {T n+1 x} converges to
T x ∗ and is also a subsequence of {T n x}, it follows that x ∗ = T x ∗ . Since φ(t) < t
for all t > 0, it follows that the fixed point of T is unique.
Remark 5.2.7 Let X be (−∞, −1] ∪ [1, ∞) with the usual metric. Define T : X →
X by
− (x+1) , if x ≥ 1
T x = (1−x)2
2
, if x ≤ −1
Corollary 5.2.8 (Rakotch [17]) Let (X, d) be a complete metric space and T : X →
X an operator such that for all x, y ∈ X d(T x, T y) ≤ α(d(x, y))d(x, y), where
α : (0, ∞) → [0, 1) is a montonic decreasing function such that 0 ≤ α(t) < α(s)
for 0 < t < s. Then, T has a unique fixed point and every sequence of T -iterates
converges to the unique fixed point.
Proof Set ψ(t) = α(t)t for all t > 0 in Theorem 5.2.6. Since all the assumptions of
Theorem 5.2.6 are satisfied, the corollary follows.
Since 0 < α < 1, a = b. Thus f and g have a unique common fixed point.
Corollary 5.2.10 Let f and g be commuting mappings on a complete metric space
(X, d) into itself. Suppose f is continuous and g(X ) ⊆ f (X ). If for some α ∈ (0, 1)
and a positive integer k, d(g k x, g k y) ≤ αd(x, y) for all x, y ∈ X , then f and g have
a unique common fixed point.
Proof By Theorem 5.2.9, g k and f have a unique common fixed point a, say. Then
a = g k (a) = f a. So ga = g k ga = g f a = f ga, showing that ga is also a fixed point
for g k and f . By the uniqueness of the common fixed point for f and g k , it follows
that a = ga = f a.
Remark 5.2.11 In fact if f has a fixed point, then we can find a commuting map g with
a unique fixed point common with f , g(X ) = f (X ) and d(gx, g y) ≤ αd( f x, f y)
for all x, y ∈ X for some α ∈ (0, 1). This is readily seen by setting gx ≡ a, a fixed
point of f , and choosing any α ∈ (0, 1).
Example 5.2.12 f (x) = x 2 and g(x) = x 4 on [0, 21 ] satisfy all the assumptions of
Theorem 5.2.9 and 0 is the unique common fixed point.
Nadler [15] generalized the contraction principle for multivalued functions, involving
the Hausdorff metric. We need the following
Definition 5.3.1 Let (X, d) be a metric space and C B(X ) be the set of all non-empty
closed bounded subsets of X . For C ∈ C B(X ) define
Remark 5.3.2 H defines a metric on C B(X ), called the Hausdorff distance on the
space C B(X ). Further for x, y ∈ X , H ({x}, {y}) = d(x, y).
Nadler [15] proved a generalization of the contraction principle for mappings of
X into C B(X ), using the following definition and a lemma.
Definition 5.3.3 Let (X, d) be a metric space. A map F : X → C B(X ) is called
a multivalued contraction if there exists α ∈ (0, 1) such that for all x, y ∈ X ,
H (F x, F y) ≤ αd(x, y).
Lemma 5.3.4 Let (X, d) be a metric space and A, B ∈ C B(X ). Given > 0 and
a ∈ A, we can find b ∈ B such that d(a, b) ≤ H (A, B) + .
Proof Let r = H (A, B). For r = 0, the lemma is clear. For r > 0, by definition of
H (A, B), A ⊆ N (B, r + ). So for a ∈ A, there exists b ∈ B such that a ∈ B(b, r +
) or d(a, b) < r + = H (A, B) + .
Theorem 5.3.5 (Nadler) Let (X, d) be a complete metric space and F : X →
C B(X ) be a multivalued contraction with contraction constant α ∈ (0, 1). Then
F has a fixed point in X (i.e. an element x0 ∈ X with x0 ∈ F x0 ).
Proof For p0 ∈ X , since F( p0 ) ∈ C B(X ) for any p1 ∈ F( p0 ), for = α, it follows
from Lemma 5.3.4 above that there exists p2 ∈ F( p1 ) such that
d( p1 , p2 ) ≤ H (F( p0 ), F( p1 )) + α.
d( p2 , p3 ) ≤ H (F( p1 ), F( p2 )) + α2
So
j
d( pi , pi+ j ) ≤ d( pi+k−1 , pi+k )
k=1
j
≤ (αi+k−1 d( p0 , p1 ) + [i + k − 1]αi+k−1 ).
k=1
∞
∞
As αn and nαn converge, it follows that { pn } is a Cauchy sequence in the
1 1
complete metric space (X, d) converging to an element p ∗ in X . Writing d(a, B) =
inf{d(a, b) : b ∈ B}, it follows that
d( p ∗ , F p ∗ ) ≤ d( p ∗ , F pn ) + H (F p ∗ , F pn )
≤ d( p ∗ , pn+1 ) + αd( p ∗ , pn )
(as pn+1 ∈ F pn and F is a
multivalued contraction).
Proof The map x → {T x} maps the complete metric space (X, d) into C B(X )
and is a multivalued contraction, whenever T : X → X is a contraction, in view of
Remark 5.3.2. Hence by Nadler’s Theorem 5.3.5, this map has a fixed point, which
is clearly a fixed point T . The uniqueness can be proved independently.
Remark 5.3.7 A fixed point of the multivalued contraction, insured by Nadler’s the-
orem need not be unique. For example, for the map x → [0, 1] of R into C B(R),
every point of [0, 1] is a fixed point. This map, being a constant map, is clearly, a
contraction.
Remark 5.4.5 Clearly a gauge space is Hausdorff if and only if the gauge is sepa-
rating.
p
dλ (xn , xn+ p ) ≤ dλ (xn+k−1 , xn+k )
k=1
p
≤ dλ (T n+k−1 x0 , T n+k x0 )
k=1
p
≤ cλn+k−1 dλ (x0 , x1 )
k=1
j−1
(as dλ (x j−1 , x j ) ≤ cλ d(x0 , x1 ))
cλn
≤ dλ (x0 , x1 ).
1 − cλ
In this section, a converse to the contraction principle due to Bessaga [3] is proved
following a simplified approach due to Jachymski [11].
Theorem 5.5.1 (Bessaga) Let X be a non-empty set and T : X → X be a map and
k ∈ (0, 1). Then
(a) there exists a metric d on X such that d(T x, T y) ≤ kd(x, y) for all x, y ∈ X ,
whenever T n has at most one fixed point for each n ∈ N ;
110 5 Contraction Principle
(b) if in addition some T n has a fixed point, X has a complete metric d such that
d(T x, T y) ≤ kd(x, y) for all x, y ∈ X .
Jachymski’s proof [11] is based on the following
Lemma 5.5.2 Let T : X → X be a map and k ∈ (0, 1). The following statements
are equivalent:
(i) there exists a complete metric d on X such that d(T x, T y) ≤ kd(x, y) for all
x, y ∈ X ;
(ii) there exists a function ϕ : X → R+ such that ϕ(T x) ≤ kϕ(x) for all x ∈ X and
ϕ−1 {0} is a singleton.
Proof (i) ⇒ (ii). Since by the contraction principle T has a fixed point x ∗ , the map ϕ
defined by ϕ(x) = d(x, x ∗ ) is such that ϕ(T x) = d(T x, T x ∗ ) ≤ kd(x, x ∗ ) = kϕ(x)
for all x ∈ X and ϕ−1 {0} = {x ∗ }, a singleton. Thus (ii) is true.
(ii) ⇒ (i). Define d : X × X → R+ by
ϕ(x) + ϕ(y) for x = y
d(x, y) =
0, for x = y.
In particular
d(xm , x N (x) ) < λd(x, x ), ∈ N, m ≥ N (x).
By a similar reasoning for n ∈ N, there exists a least positive integer n = n (n) > n
such that
d(xm , xn ) < λd(xn , xn ), m ≥ n .
Define T : X → X by
x N (x) , if x ∈
/ A
T (x) =
xn , if x = xn ∈ A.
Clearly T has no fixed point, though it satisfies conditions (i) and (ii) above. Indeed
for T (x) = xn , T (y) = xm ,
λd(y, A − {y}), n ≥ m
d(xm , xn ) <
λd(x, A − {x}), n < m
and consequently (i) is true. This contradiction shows that (X, d) must be
complete.
Corollary 5.5.6 (Converse to Kannan’s fixed point Theorem (Corollary 5.2.2)) Let
(X, d) be a metric space. If for each λ ∈ (0, 21 ) every map T : X → X satisfying the
condition
d(T x, T y) ≤ λ[d(x, T x) + d(y, T y)], x, y ∈ X
Corollary 5.5.7 (Converse to Theorem 5.2.1) Let (X, d) be a metric space. If for
5
each {a1 , a2 , a3 , a4 , a5 } ⊆ [0, 1] with ai < 1, each map T : X → X satisfying
i=1
the inequality
(a) d(T x, T y) ≤ λ max{inf k∈N d(x, T k x), inf k∈N d(y, T k y)} x, y ∈ X
or
(b) d(T x, T y) ≤ λ max{inf k∈N d(x, T k y), inf k∈N d(y, T k x)} x, y ∈ X for fixed
λ > 0
and
(c) T (X ) is countable
has a periodic point.
Then (X, d) is complete.
Earlier Hu [10] proved the following converse of the contraction principle using
this argument.
Theorem 5.5.9 ([10]) Let (X, d) be a metric space. If for every closed non-empty
subset C of X , any contraction T : C → C has a fixed point, then (X, d) is complete.
In this section, a topological version of the contraction principle due to Kupka [14]
is presented. This fixed point theorem is proved for multifunctions which are feebly
topologically contractive, without involving any concept of completeness.
Definition 5.6.1 Let (X, T ) be a topological space. By a multifunction on X , we
mean a mapping of X into the set of all non-empty subsets of X . The graph of a
multifunction F on X is the set G ◦ F = {(x, y) : y ∈ F x, x ∈ X } and is denoted
by Gr F.
Proof Every contraction has a closed graph and is feebly topologically contractive
(as T has a unique fixed point x ∗ in a complete metric space X for a given pair
a, b ∈ X , G can be chosen as an open set containing x ∗ ).
In this section, Baranga’s proof [2] of the contraction principle using Kleene’s fixed
point theorem is presented. To this end, we need the following definitions and theo-
rems.
Definition 5.7.1 Let (P, ≤) be a partially ordered set. For an increasing sequence
(xn : n ∈ N), we denote the supremum of this sequence by ∨{xn : n ∈ N}. (P, ≤) is
said to be ω-complete if every increasing sequence (xn ) in P has a supremum in P.
Definition 5.7.2 Let (P, ≤) and (Q, ≤) be two partially ordered sets. A map f :
P → Q is said to be ω-continuous if for every increasing sequence (xn ) in P, such
that ∨{xn : n ∈ N} exists in P, also ∨{ f xn : n ∈ N} exists in Q and f (∨{xn : n ∈
N}) = ∨{ f (xn ) : n ∈ N}. (Clearly any ω-continuous function is increasing).
Proof Since (xn , kn ) ≤ (xn+1 , kn+1 ) for n ∈ N, d(xn , xn+1 ) ≤ kn − kn+1 for all n ∈
N. So (kn ) decreases in R+ and so converges to a non-negative number in R+ , say
5.7 Another Proof of the Contraction Principle 117
n
k. Clearly d(x j , x j+1 ) ≤ k1 − kn+1 . Allowing n to tend to +∞, it follows that
j=1
∞
∞
d(x j , x j+1 ) ≤ k1 − k. So d(xn , xn+1 ) converges. Clearly (xn ) is a Cauchy
j=1 1
sequence.
d(xn y , y) > 0 + kn y − k.
n−1
0 + kn y − k < d(xn y , y) ≤ d(xi , xi+1 ) + d(xn , y)
i=n y
≤ kn y − kn + d(xn , y)
≤ kn y − k + d(xn , y).
We are now in a position to provide Baranga’s proof [2] of the contraction principle
5.1.7 via Kleene’s fixed point Theorem 5.7.3.
Proof of Theorem 5.1.7 (Baranga [2]).
Let (X + , ≤) and f + : X + → X + be defined as in Proposition 5.7.7. For x0 ∈ X ,
we can find a > 0 such that (1 − c)a > d(x0 , f (x0 )) so that (x0 , a) ≤ f + (x0 , a).
By Kleene’s Theorem 5.7.3, (x̄, 0) = ∨{( f + )n (x0 , a) : n ∈ N} is a fixed point of f + .
So x̄ = lim f n (x0 ) is a fixed point of f . Clearly x̄ is independent of the choice of a.
Also f + (y, b) ≤ (y, b) if and only if b = 0 and f (y) = y. If y is a fixed point of f ,
one can choose a > 0 so that (x0 , a) ≤ (y, 0) and (x0 , a) ≤ f + (x0 , a). By the least
fixed point property it follows that (x̄, 0) ≤ (y, 0). This implies that d(x̄, y) = 0 or
f has a unique fixed point x̄.
References
1. Banach, S.: Sur les operations dans les ensembles abstraits et leur application aux equations
integrales. Fund. Math. 3, 133–181 (1922)
2. Baranga, A.: The contraction principle as a particular case of Kleene’s fixed point theorem.
Discret. Math. 98, 75–79 (1991)
3. Bessaga, C.: On the converse of Banach fixed point principle. Colloq. Math. 7, 41–43 (1959)
4. Bollabos, B.: Linear Analysis: An Introductory Course. Cambridge University, Cambridge
(1999)
5. Borwein, J.M.: Completeness and the contraction principle. Proc. Am. Math. Soc. 87, 246–250
(1983)
6. Boyd, D., Wong, J.S.W.: On nonlinear contractions. Proc. Am. Math. Soc. 20, 458–469 (1969)
7. Cacciopoli, R.: Un teorema generale sullesistenzadi elementi uniti in una transformazione
funzionale. Rend. Accad. Lincei 6(11), 749–799 (1930)
8. Dugundji, J.: Topology. Allyn and Bacon Inc., Boston (1966)
9. Hardy, G.E., Rogers, T.D.: A generalization of a fixed point theorem of Reich. Can. Math. Bull.
16, 201–206 (1973)
10. Hu, T.K.: On a fixed point theorem for metric spaces. Am. Math. Mon. 74, 436–437 (1967)
References 119
11. Jachymski, J.: A short proof of the converse to the contraction principle and some related
results. Topol. Methods Nonlinear Anal. 15, 179–186 (2000)
12. Jungck, G.: Commuting mappings and fixed points. Am. Math. Mon. 83, 261–263 (1976)
13. Kannan, R.: Some results on fixed points. Bull. Calcutta Math. Soc. 60, 71–76 (1968)
14. Kupka, I.: Topological conditions for the existence of fixed points. Math. Slovaca 48, 315–321
(1998)
15. Nadler, S.B.: Multi-valued contracting mappings. Pac. J. Math. 30, 475–488 (1969)
16. Palais, R.S.: A simple proof of the Banach contraction principle. J. Fixed Point Theory Appl.
2, 221–223 (2007)
17. Rakotch, E.: A note on contractive mappings. Proc. Am. Math. Soc. 13, 459–465 (1962)
18. Rall, L.B.: Computational Solution of Nonlinear Operator Equations. Wiley, New York (1969)
19. Subrahmanyam, P.V.: Fixed points and completeness. Monat. Math. 80, 325–330 (1975)
20. Tan, K.K.: Fixed point theorems for nonexpansive mappings. Pac. J. Math. 41, 829–842 (1972)
Chapter 6
Applications of the Contraction Principle
This short chapter offers a few samples of applications of the contraction principle.
It was already pointed out that the evergrowing list of applications of this fixed point
theorem would fill volumes.
Proof Define T1 (x) = T (x) + a for each x ∈ X . Clearly T1 maps X into itself and
T1 (x) − T1 (y) = T x − T y ≤ kx − y, where k = T < 1. Thus T1 is a con-
traction mapping X into itself. Since X is complete, T1 has a unique fixed point
x ∗ . Thus T x + a = x has the solution x = x ∗ . Again, by the contraction principle
{T1n (x0 )} converges to the unique solution of the equation x = T x + a.
n
n
n
has a unique solution (i) provided sup |ai j | < 1 or (ii) ai2j < 1.
i j=1 i=1 j=1
n
n
n
So A is a contraction on Rn , since ai2j < 1. Thus xi = ai j x j + bi ,
i=1 j=1 j=1
i = 1, 2, . . . , n has a unique solution, by Theorem 6.1.1.
Proof Let tX be the Banach space C[a, b] with the supremum norm. Then for each x ∈
C[a, b], a K (s, t)x(s)ds defines a continuous function on [a, b]. Further x(t) →
t
T (x(t)) = λ a K (s, t)x(s)ds is a linear operator on X .
For x1 , x2 ∈ X , λ ∈ R and t > 1
t
T x1 (t) − T x2 (t) = λ K (s, t)(x1 (s) − x2 (s))ds
a
So
t
|T x1 (t) − T x2 (t)| ≤ |λ| Mx1 − x2 ds
a
≤ |λ|Mx1 − x2 (t − a)
where M = Sup{|K (s, t)| : s, t ∈ [a, b]} and x1 − x2 = Sup{|x1 (t) − x2 (t)| : t ∈
[a, b]}}. Now for n ∈ N and t > a.
|λ|k M k
|T k x1 (t) − T k x2 (t)| ≤ (t − a)k x1 − x2
k!
124 6 Applications of the Contraction Principle
|λ|n+1 M n+1
T n+1 x1 − T n+1 x2 ≤ (b − a)n+1 x1 − x2
(n + 1)!
|λM(b − a)|n+1
= 0, for some n 0 ∈ N, |λM(b−a)|
n
Since lim < 1 for n ≥ n 0 . Thus
n→∞ (n + 1)! n!
n0
T is a contraction on X and for T1 = T + g, T1 is also a contraction on X . So T1
n0
has a unique fixed point in X = C[a, b] which is the solution of the given integral
equation. Thus the Volterra integral equation has a unique solution in C[a, b] for all
λ ∈ R.
Remark 6.1.5 We can supplement Corollary 6.1.3 on the eigen-value problem for
Fredholm integral equations. Suppose K : [a, b] × [a, b] → R and g : [a, b] → R
bb
are Lebesgue measurable functions such that g ∈ L 2 [a, b] and 0 < a a K 2 (s, t)
dsdt < +∞. Then we can show that the equation
b
x(t) = λ K (s, t)ds + g(t)
a
bb
has a unique solution in L 2 [a, b] for |λ| a a K 2 (s, t)dsdt < 1. For the proof we
use the Hilbert space L 2 [a, b] instead of C[a, b].
The next result insures that certain mappings on Banach spaces are surjections
and indeed homeomorphisms.
F T −1 x1 − F T −1 x2 = x1 − GT −1 x1 − x2 + G −1 x2
≥ x1 − x2 − GT −1 x1 − GT −1 x2
≥ (1 − αβ)x1 − x2
So F x1 − F x2 ≥ (1 − αβ)T x1 − T x2
or x1 − x2 ≥ (1 − αβ)T F −1 x1 − T F −1 x2
6.1 Linear Operator Equations 125
Example 6.1.8 For each (y1 , y2 ) ∈ R2 , we can find a unique (x1 , x2 ) ∈ R2 such that
x1 + x2 |x1 | + |x2 |
(2x1 − 3x2 , x1 − 2x2 ) − cos , = (y1 , y2 ).
50 50(1 + |x1 | + |x2 |
We prove below an existence theorem for the solution of an initial value problem for
a system of first-order ordinary differential equations.
Theorem 6.2.1 Let G be an open set in Rn+1 containing the point (t0 , x10 , x20 , . . . , xn0 )
and f i : G → R be continuous functions satisfying the Lipschitz condition
n
| f i (t, y1 , . . . , yn ) − f i (t, z 1 , . . . , z n )| ≤ M |y j − z j |
j=1
d xi
= f i (t, x1 (t), . . . , xn (t)) for i = 1, 2, . . . , n
dt
and xi (t0 ) = xi0 for i = 1, 2, . . . , n. Further, these are unique solutions and (t, x1 (t),
x2 (t), . . . , xn (t)) ∈ G for all t ∈ [t0 − h, t0 + h].
Proof Since G is open and (x10 , . . . , xn0 , t0 ) ∈ G, we can find a closed rectangle R =
[x10 − a, x10 + a] × . . . × [xn0 − a, xn0 + a] × [t0 − a, t0 + a], a > 0 in G. As each f i
is continuous on R and R is compact, we can find K > 0 such that | f i ( p)| ≤ K for
all i = 1, 2, . . . , n for all p ∈ R. In view of the continuity of each f i , i = 1, 2, . . . , n,
it is clear that xi (t) is a solution of
d xi
= f i (t, x1 (t), . . . , xn (t))
dt
xi (t0 ) = xi(0) , i = 1, 2, . . . , n
126 6 Applications of the Contraction Principle
if and only if t
xi (t) = xi (0) + f i (τ , x1 (τ ), . . . , xn (τ ))dτ .
t0
Choose h > 0 such that K h < a. Let (X, ρ) be the metric space of vector-valued
continuous real functions (x1 (t), . . . , xn (t)) defined on I = [t0 − h, t0 + h] with
n
the metric ρ(x, y) = sup |xi (t) − yi (t)|, where x = (x1 (t), . . . , xn (t)) and y =
t∈I i=1
(y1 (t), . . . , yn (t)).
Clearly (X, ρ) is complete. For x = (x1 , . . . , xn ) ∈ X , define T x = (y1 , . . . , yn )
where t
yi (t) = xi0 + f i (τ , x1 (τ ), . . . , xn (τ ))dτ .
t0
n
n
T (x) ∈ X whenever x ∈ X . Further, ρ(T x, T x ) = |yi − yi | ≤ Mρ(x, x )
i=1 i=1
h = n Mρ(x, x )h. If we further choose h such that n Mh < 1 then T is a contraction
on X and hence has a unique fixed point x = (x1 , . . . , xn ) which is the solution
of the initial value problem in [t0 − h, t0 + h] and |xi0 − xi (t)| ≤ K h ≤ a for all
t ∈ [t0 − h, t0 + h] and so lies in R ⊆ G. Thus for h < min{ Ka , n1M }, a unique local
solution exists in [t0 − h, t0 + h] for the initial value problem.
Corollary 6.2.2 Let G be an open set in R2 containing the point (t0 , x0 ) and f :
G → R be a continuous function satisfying the condition that | f (t, x1 ) − f (t, x2 )| ≤
M|x1 − x2 | for all (t, x1 ), (t, x2 ) ∈ G. Then there exists h > 0 such that the initial
value problem
dx
= f (t, x(t)), x(t0 ) = x0
dt
has a unique solution x(t) on [t0 − h, t0 + h] such that (t, x(t)) ∈ G for t ∈ [t0 −
h, t0 + h].
(This pertains to the case n = 1 in Theorem 6.2.1.
Remark 6.2.4 If the power series in m-real variables converges in S(a; ρ), then it
converges in S(a; ρ ) for 0 < ρ < ρ.
6.2 Differential Equations 127
du
= f (x, y), y(x0 ) = y0 ,
dx
If we choose 0 < < h such that < M1 , clearly d(T φ1 , T φ2 ) ≤ αd(φ1 , φ2 ), where
α = M < 1. Thus T is a contraction on the complete metric space (X, d) and hence
has a unique fixed point, which is the unique solution of the initial value problem in
[x0 − , x0 + ]. Further, this solution is a power series in x in this interval.
Remark 6.2.7 Clearly this theorem can be extended to a system of differential equa-
tions involving real-analytic functions of several real variables.
128 6 Applications of the Contraction Principle
Utz [12] raised the problem of determining conditions for the existence of a real
function y(x), not identically zero for which y (x) = ay(g(x))) where a is a given
constant and g(x) a given real function. Ryder [7] gave a solution to this problem
under suitable assumptions.
y (x) = Ay(g(x)), x ∈ Dg
y(0) = f 0
Proof Let S be the set of all functions f : Dg → Rn such that f (0) = f 0 and
f (x) − f 0 < L|x| for all x ∈ Dg for some L > 0. Define ρ : S × S → R+ by
If can be shown that ρ is a metric on S and in fact (S, ρ) is a complete metric space.
Define the operator T : S → S by
x
T ( f )(x) = f 0 + A f (g(s))ds.
0
So T ( f ) ∈ S for f ∈ S.
If f 1 , f 2 ∈ S, then f 1 (x) − f 2 (x) ≤ L|x| for x ∈ DG . Now for x ∈ Dg ,
x
T ( f 1 )(x) − T ( f 2 )(x) ≤ A L|g(s)|ds
0
≤ Ak L|x|
Consequently
ρ(T f 1 , T f 2 ) ≤ Akρ( f 1 , f 2 )
6.3 A Functional Differential Equation 129
d2x
= α f (t, x(t)), t ∈ (0, 1)
dt 2
x(0) = x(1) = 0.
α α
|T (x)(t)| ≤ f (s, x(s)) ≤ (K + Ma) < a
8 8
n
u t = A(t, z)u + B j (t, z)u z j + c(t, z) for (t, z) ∈ G (6.5.1)
j=1
C C
| f (z)| ≤ implies | f z j (z)| ≤ C p
d p (z) d p+1 (z)
6.5 An Elementary Proof of the Cauchy–Kowalevsky Theorem 131
1
|g (z)| ≤ max |g(z )|.
r |z−z |=r
1 C 1 C
| f z j (z)| ≤ max | f (z )| ≤ max p ≤
r |z−z |=r r d (z ) r (d − r ) p
Remark 6.5.2 The Cauchy problem (6.5.1) can be written as the equivalent integral
equation.
t
n
u(t, z) = g(t, z)+ [A(s, z)u(s, z) + B j (s, z)u z j (s, z)]ds (6.5.2)
0 j=1
t
where g(t, z) = φ(z) + c(s, z)ds (6.5.3)
0
α
|A(t, z)| ≤ , |B j (t, z)| ≤ β j ,
d(t, z)
η δ
|C(t, z)| ≤ p+1 , |φ(z)| ≤ in G;
d (t, z) d p (z)
α
(iii) p
+ (1 + 1p ) p+1 β j < η1 .
Then the Eq. (6.5.2) has a unique solution u in G satisfying |u(t, z)| ≤ C
d p (t,z)
for
(t, z) ∈ G.
Proof Let X be the Banach space of all functions u = (t, z) which are continuous
on G and taking values in Cm and analytic in z with the finite norm
132 6 Applications of the Contraction Principle
u = g + T (u)
Since
|t|
t
ds ds η
= p+1 < , g(t, z) ∈ X.
d p+1 (s, z) pd p (t, z)
0 0
d(s) − ηs
Applying Nagumo’s Lemma 6.5.1 to t with distance d(t, z) instead of and d(z)
we get,
u
|u z j (t, z)| ≤ C p p+1
d (t, z)
αu u
|Au| ≤ , |B j u z j | ≤ p+1 βjCp
d p+1 (t, z) d (t, z)
So for β = βj
t
ds
|T u(t, z)| ≤ u(α + βC p )
0 d (s, z)
p+1
1 ηu
≤ (α + βC p ) p .
p d (t, z)
(6.5.2) has a unique solution. Thus the Cauchy problem (6.5.1) with analytic data
has a unique solution.
Note 6.5.4 In Theorem 6.5.3, η > 0 can be chosen sufficiently small so that (iii) of
Theorem 6.5.3 holds.
xi
= f (ti , xi ), i = 0, 1, . . . , n, ux0 + vxn = w, u + v = 0 (6.6.1)
h
where 0 < h < Nn < N and ti = i h (i = 0, 1, . . . , n) are the grid points with xi =
xi+1 − xi , i = 0, . . . , n − 1 and u, v, w are constants. The problem is to find a vetor
x̃ = (x0 , x1 , . . . , xn ) ∈ Rn+1 satisfying (6.6.1).
Then the boundary value problem has a unique solution for 0 < h ≤ δ.
n 1
Proof Consider X = Rn+1 with the norm x = i=0 |x i |
p p
for p > 1. Clearly
X is a Banach space. The problem (6.6.1) is equivalent to the system of equations
n−1
w
xi = h G(i, j) f (t j , x j ) + , i = 0, . . . , n
j=0
u+v
where
u
, 0≤ j ≤i −1
G(i, j) = u+v
−u
u+v
, i ≤ j ≤n−1
Define T : X → X by
n−1
w
(T x̃)i = h G(i, j) f (t j , x j ) + , i = 0, . . . , n
j=0
u+v
134 6 Applications of the Contraction Principle
for x̃ = (x0 , . . . , xn ) ∈ X .
For x̃, ỹ ∈ X = Rn+1
n−1
|T (x̃)i − T ( ỹ)i | ≤ h |G(i, j)|| f (t j , x j ) − f (t j , y j )|
j=0
n−1
≤ Lh |G(i, j)||x j − y j |
j=0
⎛ ⎞ q1 ⎛ ⎞ 1p
n−1
n−1
≤ Lh ⎝ |G(i, j)|q ⎠ ⎝ |x j − y j | p ⎠
j=0 j=0
n−1
i|u|q + (n − i)|v|q
|G(i, j)|q = , i = 0, . . . , n.
j=0
|u + v|q
So
Lhx̃ − ỹ 1
|T (x̃)i − T ( ỹ)i | ≤ [i|u|q + (n − i)|v|q ] q , i = 0, . . . , n.
|u + v|
So
p
Lh p
|(T x̃)i − (T ỹ)i | ≤
p
x̃ − ỹ p [i|u|q + (n − i)|v|q ] q , i = 0, . . . , n.
|u + v|
So
n 1p
T x̃ − T ỹ = |(T x̃)i − (T ỹ)i | p
i=0
n 1p
Lh p
≤ x̃ − ỹ [i|u| + (n − i)|v| ]
q q q
|u + v| i=0
≤ αx̃ − ỹ
n 1p
Lδ p
where α = [i|u|q + (n − i)|v|q ] q < 1. So T is a contraction on X
|u + v| i=0
and so has a unique fixed point, which is a solution to the discrete boundary value
problem.
6.6 An Application to a Discrete Boundary Value Problem 135
The above inequality insures that the conditions of Theorem 6.6.1 are fulfilled.
Contraction principle has been a handy tool in the solution of a variety of functional
equations. As a matter of fact the implicit function theorem is a consequence of the
contraction principle.
1
(T y)(x) = y(x) − f (x, y(x)), x ∈ I.
D
Clearly T y is continuous for each y ∈ X . If y(x0 ) = y0 , then T y(x0 ) = y(x0 ) = y0
as f (x0 , y(x0 )) = f (x0 , y0 ) = 0. Now
1
T y0 − y0 = f (x, y0 )
D
1
< δ for |x − x0 | <
2
(by construction of I )
136 6 Applications of the Contraction Principle
T y − y0 ≤ T y − T y0 + T y0 − y0
1
≤ y − y0 + T y0 − y0
2
1 1
< δ+ =δ
2 2
Thus T maps X into itself and is a contraction (with constant 21 ). Since X is complete
T has a unique fixed point y0 (x) which is a solution of f (x, y0 (x)) = 0 for x ∈
[x0 − , x0 + ].
Remark 6.7.2 The above theorem can be extended to functions taking values in a
Banach space after suitable modifications.
Proof Without loss of generality we shall assume that 0 < |λ| < |μ| and |λ| < 1.
(In the second case F is replace by F −1 which exists in some neighbourhood of (0,
0)).
By the inverse function theorem there exists a continuously differentiable mapping
h defined in a neighbour V of(0, 0) satisfying h(x, g(x, y)) = y for (x, y) ∈ V . Also
h(x, y) = (0, μ1 ).(x, y) + o( x 2 + y 2 ), (x, y) → (0, 0).
Let L be any non-negative real number and c > 0 be such that D = {(x, y) ∈ R2 :
|x| ≤ c, |y| ≤ L|x|} ⊆ U ∩ V . Let X be the set of all functions ϕ : I = [−c, c] → R
such that ϕ(0) = 0 and |ϕ(s) − ϕ(t)| ≤ L|s − t| for all s, t ∈ I . It can be seen that
ρ : X × X → R+ defined by
ϕ(x) − ψ(x)
ρ(ϕ, ψ) = Sup
x∈I −{0} x
and
1
|h(x, y) − h(u, v)| ≤ |x − u| + + |y − v|
|μ|
Decreasing and hence c we can choose the Lipschitz constant in the above inequality
to be less than 1. So T is a contraction and hence has a unique fixed point. ϕ(x) ∈ X .
Thus
ϕ(x) = h(x, ϕ( f x, ϕ(x)) = T ϕ(x).
138 6 Applications of the Contraction Principle
Since h(x, g(x, φ(x)) = φ(x) and is inverse to g with respect to the second variable,
ϕ( f x, ϕ(x)) = g(x, ϕ(x)). Hence the existence of a Lipschitzian solution follows.
is a sequentially complete Hausdorff gauge space under the family F of all pseudo-
metrics Dx defined by
The proof of Theorem 6.7.4 is omitted. The next result is a fixed point theorem
obtained from Tan’s Theorem (see [9]) and Theorem 6.7.4.
Theorem 6.7.5 (see [8]) Let (Y, d) be a complete metric space and s : X → X a
function. Let P : S0 → S0 be an operator on a sequentially closed subset S0 of Y X
with the topology of pointwise convergence such that
(i) D X ( f 0 , P( f 0 )) < +∞ for some f 0 ∈ S0 and all x ∈ X where Dx is as defined
in Theorem 6.7.4;
(ii) for each x ∈ X there is an a(x) such that given f, g ∈ S0 there exists a non-
negative integer n = n(x) with d(P f (x), Pg(x)) ≤ a(x)d( f (s n (x)), g(s n (x))).
Further, for each x ∈ X ,
For the operator P mapping S0 into itself, (i) implies that P( f 0 ) ∈ S. Further, for
f ∈ S and x ∈ X ,
Dx ( f 0 , P( f )) ≤ Dx ( f 0 , P( f 0 )) + Dx (P( f 0 ), P( f ))
≤ Dx ( f 0 , P( f 0 )) + A(x)Dx ( f 0 , f ) (by (ii))
< +∞
where A(x) = sup{a(s k (x)) : k = 0, 1, 2, . . .}. So P maps S into itself. By (ii) again
for each x ∈ X and f, g ∈ S. As 0 < A(x) < 1, P is strictly contractive and it maps
the Hausdorff sequentially complete gauge space (S, F) into itself. Therefore P has
a unique fixed point g0 ∈ S, by Theorem 5.4.7. Also, as g0 ∈ S, for each x ∈ X
For this choice of the operator P and the space S, all the condition of Theorem 6.7.5
are satisfied. So the operator P has a unique fixed point in S. As any other solution
140 6 Applications of the Contraction Principle
of the functional equation (6.7.1) is in S in view of (i), (ii) and (iii), (6.7.1) has a
unique solution in B X .
has a unique solution in R R , the space of all real valued functions of a real variable,
where the function g is the one defined below:
⎧
⎪
⎨1/x, |x| > 1
g(x) = x, |x| ≤ 1, x irrational
⎪
⎩
x/2, otherwise.
In this section a short proof of the algebraic Weierstrass Preparation theorem due
to Gersten [2] is highlighted. To this end a few basic definitions are described. For
these Lang [5] and Zariski and Samuel [14] may be consulted.
Definition 6.8.1 A ring R is called a local ring if it is commutative and has a unique
maximal ideal.
Remark 6.8.2 If R is a local ring with the unique maximal ideal M, then x ∈ A − M
is a unit.
∞
Definition 6.8.3 For a ring R and an ideal I , suppose that I n = {0}. We can
n=1
define I n as a neighbourhood of 0 for each n. I n is the ideal generated by elements
of the form a1 a2 . . . an , ai ∈ I for i = 1, 2, . . . , n. In the same way we can say that a
sequence {xn } in R is Cauchy if given some power I k of I , there exists an integer M
such that for all m, n ≥ M, xm − xn ∈ I k . A sequence (xn ) in R is said to converge
to x in R if for each I k , there exists an integer M such that (xn ) ∈ x + I k for all
n ≥ M. R is said to be complete in the I -adic topology if every Cauchy sequence in
R converges.
Definition 6.8.4 Let R be a local ring and I = M the maximal ideal. R is called a
complete local ring if R is complete in the M-adic topology and we assume that the
M-adic topology is Hausdorff.
6.8 An Application to Commutative Algebra 141
Remark 6.8.5 Let k be a field and R = k[[X 1 , . . . , X n ]] the power series ring in n-
variables. Then R is a complete local ring. If M is the ideal generated by X 1 , . . . , X n
then R/M is isomorphic with k and so M is a maximal ideal. Any power series of
the form f (X ) = c0 − f 1 (X ) where c0 = 0 ∈ k and f 1 (X ) ∈ M is invertible as
f 1 (X ) ( f 1 (X ))2
(c0 − f 1 (X )) −1
= c0−1 1+ + + ··· .
c0 c02
So M is the unique maximal ideal and R is local. It can be readily verified that R is
complete.
For a ring A, the power series ring in n variables for n > 1 can be viewed
as the ring of power series in variable X n over the ring of power series in (n −
1) variables X 1 , . . . , X n−1 . Thus we have the identification A[[X 1 , . . . , X n ]] =
A[[X 1 , . . . , X n−1 ]][[X n ]]. When A is a field A[[X 1 , . . . , X n ]] is a complete local
ring. Moreover if R is a complete local ring, the power series ring R[[X ]] is a com-
plete local ring with the maximal ideal [M, X ], M being the maximal ideal of R. If
the power series an X n has unit constant a0 ∈ R − M, then the power series is a
unit in R[[X ]] as (a0 + h)−1 is a0−1 1 − ah0 + ah 2 − · · · .
2
Proposition 6.8.7 ([2]) Let A be a complete local ring with maximal ideal M which
is Hausdorff in the M-adic topology. Let B = A[[t]] be the power series ring in
∞
one variable t over A. Define d1 ( f, f ) = sup d(ak , ak ) for f = ak t k and f =
k∈Z + k=0
∞
ak t k where ak , ak ∈ A and d is as in Proposition 6.8.6. Then (B, d1 ) is a complete
k=0
metric space.
Definition 6.8.8 Let A be a complete local ring with maximal ideal M and B =
A[[t]]. A distinguished polynomial in B is of the form p0 + p1 t + · · · + pn−1 t n−1 +
t n where pi ∈ M.
The following is
∞
Theorem 6.8.9 (Algebraic Weierstrass Preparation Theorem) If f = ak t k ∈ B,
k=0
where ak ∈ A and if there exists n ∈ N such that ak ∈ M for k < n and an ∈ / M,
then f = up where u is a unit of B and p is a distinguished polynomial of degree n.
Also u and p are uniquely determined.
The proof of the algebraic preparation theorem is based on the following division
theorem for which Gersten [2] has given a proof using the contraction principle.
Theorem 6.8.10 ([2]) If f, b ∈ B and b ∈ M[t] and if n ∈ N, then f = q(t n +
b) + r where q, b ∈ B and r is a polynomial in t of degree < n. Further q and r are
uniquely determined.
Proof Define the operator E : B → B by x = p + E(x)t n , where p is a polynomial
in t of degr ee < n. Define T : B → B by T x = E( f − xb) for x ∈ B. So f −
xb = p + E( f − xb)t n . If x ∈ B, then f − x b = p + E( f − x b)t n . Noting that
T y = E( f − yb), and subtracting the latter equation from the former, we get −b(x −
x ) = p + (T x − T x )t n where p is a polynomial of degree < n. Note that the
coefficients of T x − T x involve only the coefficients of the left-hand side of degree
≥ n. So
1
d1 (T x, T x ) ≤ d1 (bx, bx ) ≤ d1 (x, x )
2
since b ∈ M[[t]]. Thus T is a contraction on B and hence has a unique fixed point
q ∈ B such that T q = q or f − qb = r + qt n where r is a polynomial of degree
< n. Thus f = q(t n + b) + r .
We can now deduce the algebraic Weierstrass preparation Theorem 6.8.9 from the
Division Theorem 6.8.10
∞
Proof Let f ∈ B with f = ak t k , where ai ∈ A and for some n ∈ N,ai ∈ M for
k=0
n−1
i < n with an ∈
/ M. Let b = ak t k . As ak ∈ M for k < n, b ∈ M[t]. Now by
k=0
Division Theorem 6.8.10, f = q(t n + b) + r where q, r ∈ B and r is a polynomial
in t of degree < n. Since b ∈ M[t], the distinguished polynomial p = t n + b is a
unit in B. So ( f − r ) p −1 = q or f p −1 = q + r p −1 = q ∈ b. So f = q p = pq
∞
with p being a distinguished polynomial. If q = bn t n , then a0 = b0 a0 so q is a
0
unit. The proof of uniqueness is left as an exercise.
6.9 A Proof of the Central Limit Theorem 143
Central limit theorem, a classical theorem of probability theory can be proved using
the contraction principle. Trotter [11] gave a proof based on Lindberg condition and
invoking the properties of certain linear non-expansive operators leading to a con-
traction. While Trotter’s proof completely avoids the use of characteristic functions,
Hamedani and Walter [3] gave a proof of the central limit theorem using the properties
of characteristic functions, for sub-independent identically distributed random vari-
ables, again using the contraction principle. Before describing the proof of Hamedani
and Walter [3], a few basic concepts of probability theory are stated below. See Kai
Lai Chung [1].
Definition 6.9.4 For r > 0 and a ∈ R given a random variable X with distribution
function F, the moment of X of order r about a is defined as
∞
E(X − a)r = (x − a)r μ(d x) = (x − a)r d F(x)
R −∞
μ being the probability distribution measure of X , provided the integral exists for
a = 0, r = 1 E(X ) is called the mean of X . The moments about the mean are called
central moments. The central moment of order 2 is called the variance of X and
is denoted by V ar (X ). The positive square root of V ar (X ) is called the standard
deviation of X , denoted by σ(X ).
Definition 6.9.7 For a random variable X with induced probability measure μ and
distribution function F, the characteristic function f (t) is defined as
f (t) = E(eit X ) = eit X (w) P(dw)
∞
= e μd(x) =
it x
eit x d F(x).
R −∞
Remark 6.9.8 For all t ∈ R, | f (t)| ≤ 1 = f (0) and f (−t) = f (t) (z being the com-
plex conjugate of z). f is uniformly continuous on R and for a, b ∈ R f a X +b (t) =
f X (at)eitb where f X denotes the characteristic function of X . For independent ran-
!
m !n
dom variables X i , i = 1, 2, . . . , n, E(e ) =
it Sn
E(e ) or f Sn =
it X i
f X i , Sn being
i=1 i=1
n
Xi .
i=1
and is written F = F1 ∗ F2 .
For these and the following theorem, Kai Lai Chung [1] may be consulted.
Theorem 6.9.11 If the distribution function F has a finite absolute moment of posi-
tive integral order k ≥ 1, then its characteristic function f has a bounded continuous
derivative of order k given by
∞
(k)
f (t) = (i x)k eit x d F(x)
−∞
Further
k
(i) j θk (k) k
f (t) = m ( j) t j + μ |t|
j=0
j! k!
6.9 A Proof of the Central Limit Theorem 145
where m ( j) is the moment of order j, μ(k) the absolute moment of order k and |θk | ≤ 1.
Definition 6.9.12 Two random variables X and Y on (, S , P) are said to be sub-
independent if the distribution of their sum is given by
where φ X,Y (t, s), φ X (t), φY (t) are characteristic functions corresponding to (X, Y ),
X and Y respectively. The random variables X 1 , X 2 , . . . , X n are said to be sub-
independent if for each subset {X i1 , . . . , X ik } of {X 1 , . . . , X n }
!
k
φ X i1 ,...X ik (t, t, . . . , t) = φ X i (t) for t ∈ R
i=1
Remark 6.9.13 The concept of sub-independence is more general than that of inde-
x 2
−t /2
e√
pendence. Further the random variable with distribution function F(x) = −∞ dt
2π
is called the standard normal distribution
For the subsequent discussion culminating in the proof of central limit theorem
we introduce the following class of random variables
In order to deduce the central limit theorem from the contraction principle we
need to set up appropriate complete metric spaces.
Proof Clearly the main issue in the proof is to show that dλ (F, G) is finite-valued
for F, G ∈ Mλ . If φ1 and ψ1 are the real parts of the characteristic functions of F
and G in Mλ , then for n = [λ].
146 6 Applications of the Contraction Principle
F being the distribution function if X . A similar estimate holds for the imaginary
parts of F and G. Thus dλ (F, G) < +∞.
m
(it) j
φ(t) − mj = t m h(t)
j=0
j!
Since Fk has the same jth moment for U = 0, 1, . . . , n, and φ is continuous at zero
Fk converges weakly to F in the sense that E(X k Y ) → E(X Y ) for each bounded
random variable Y , X k being the random variable with distribution function Fk and X
the random variable with distribution function X . Using Fatou’s lemma the existence
of jth moment of F follows and φ(t) is j times differentiable at 0. So its jth derivative
for j = 0, 1, 2, . . . n must be the same as that of φk . Then F ∈ Mn . That a Cauchy
sequence in Mλ for λ > n is a Cauchy sequence in Mn is left as an exercise.
6.9 A Proof of the Central Limit Theorem 147
The next proposition leads to the central limit theorem for s.i.i.d random variables.
λ
n+1
≤ (n + 1)− 2 d λ (X i , Yi )
i=1
d λ being the metric in Mλ . As {X n } and {Yn } are both sequences whose distribution
functions are bounded in Mλ , the right-hand side of the last inequality converges to
zero as n tends
n+1to infinity. n+1
Xi Yi
So d λ √ , √ → 0 as n → ∞. By Theorem 6.9.20
n + 1 i=1 n + 1
2m i=1
Zi 2m
Yi
dλ , → 0 as m → ∞ provided each Z i is a standard normal
i=1
2m/2 i=1 2m/2
variate. Choosing m to be the largest integer with 2m < n + 1, we get
λ
n+1
λ
(n + 1)− 2 d λ (Z i , Yi ) ≤ (λ + 1)− 2 (n − 2m )C
i=2m +1
≤ 2−m ( 2 ) (2m+1 − 2m )C
λ
(C, a constant)
n+1
Xi
So √ converges to Z in distribution.
i=1
n+1
References
7.1 Introduction
Siegel’s proof [11] is detailed below, based on a few definitions and lemmata,
φ, T and X being as in Theorem 7.2.1 above.
Definition 7.2.2 Let = { f : X → X with d(x, f x)) ≤ φ(x) − φ( f (x)) for x ∈
X }. Define T = { f ∈ : φ( f ) ≤ φ(T )}.
Thus f 1 ◦ f 2 ∈ T .
Let ( f n ) be a sequence in . For x ∈ X , define xn = f n ( f n−1 . . . ( f 1 x)) . . . ) =
f 1 ◦ f 2 · · · ◦ f n (x) for n ∈ N. Since d(xi , xi+1 ) ≤ φ(xi ) − φ(xi+1 ) for all i ∈ N and φ
is non-negative, {φ(xi )} is a non-increasing sequence of non-negative numbers which
is convergent. Consequently (xn ) is a Cauchy sequence in the complete metric space
(X, d) and so converges to some element x in X . Now d(xn , x) = lim d(xn , xk ) ≤
k→∞
φ(xn ) − lim inf φ(xk ) ≤ φ(xn ) − φ(x) (in view of the lower semicontinuity of φ).
k→∞
Thus for f n ∈ φ, n ∈ N, lim xn = lim f 1 ◦ . . . f n (x) = x exists for each x ∈ X .
n→∞ n→∞
n
Let x = f (x) = lim f i (x). Now
n→∞
i=1
Proof From the definition of r (Sx0 ), it follows that we can find f 1 ∈ such
that 0 ≤ φ( f 1 (x0 )) − r (Sx0 ) < 21 . Write x1 = f 1 (x0 ). As is closed under com-
positions, Sx1 ⊆ Sx0 and D(Sx1 ) ≤ 2(φ(x1 ) − r (Sx1 )) ≤ 2(φ( f 1 (x0 )) − r (Sx0 )) < 1.
Thus proceeding inductively we can obtain a sequence of functions f n such that
xn+1 = f n (xn ), Sxn+1 ⊆ Sxn and D(Sxn ) < n1 .
∞
k
If (a) is true, define f = f n and x = f (x0 ). Since x = lim f j (x j ),
k→∞
n=1 j=i+1
∞
x ∈ Sxi for each i. Since lim D(Sxn ) = 0, x ∈ Sxn . We claim that for g ∈ ,
n→∞
⎛ ⎞ n=1
∞
g(x) = x. Since g(x) = g ⎝ f i (xn )⎠, g(x) ∈ Sxi for each i ∈ N. g(x) = x as
j=i+1
lim D(Sxn ) = 0.
n→∞
Suppose (b) is true and x = lim f n f n−1 . . . f 1 (x0 ) = lim xn . Since xk ∈ Sn for
n→∞ n→∞
∞
k ≥ x, x ∈ S n for all n. As D(Sxn ) = D(S xn ), {x} = S xn . For g ∈ , g(xn ) ∈ Sxn
n=1
for each n and by the continuity of g, lim g(xn ) = g(x). So for any given > 0 we
n→∞
can find n 0 such that B(g(x), ) ∩ Sxn = φ for n > n 0 . So for n > n 0 , d(g(x), x) <
+ n1 . Thus d(g(x), x) ≤ . As > 0 is arbitrary g(x) = x.
satisfying the inequality of Caristi’s theorem has a fixed point, then the space is
complete.
Theorem 7.2.11 (Kirk [8]) Let (X, d) be a metric space and φ : X → R+ , any
lower semicontinuous map. If every map T : X → X satisfying d(x, T x) ≤ φ(x) −
φ(T (x)) has a fixed point, then (X, d) is complete.
Proof Let (xn ) be a non-convergent Cauchy sequence in (X, d). For each x ∈ X ,
define φ(x) = limn→∞ d(x, xn ). (Note that φ(x) is well-defined as d(x, xn ) is a
Cauchy sequence of real numbers and hence is convergent.) Given x ∈ X , define
the least natural number n(x) such that 0 < 21 d(x, xn ) < φ(x) − φ(xn(x) ). Define
T : X → X by T (x) = xn(x) and ψ : X → R+ by ψ(x) = 2φ(x), for x ∈ X . Since
|φ(x) − φ(y)| ≤ d(x, y), both φ and ψ are continuous. Clearly from the definition
of ψ, 0 < d(x, T x) ≤ ψ(x) − ψ(T x) for all x ∈ X . Clearly T has no fixed point.
Thus if X is not complete we can find a fixed point free map T : X → X satisfying
d(x, T x) ≤ ψ(x) − ψ(T (x)) for some lower semicontinuous map ψ : X → R+ .
Thus Caristi’s theorem is characteristic of completeness. The following simple
example shows that Caristi’s theorem applies even when T is not continuous.
Remark 7.2.13 It may be clarified that while proofs of Caristi’s theorem by Caristi
[4], Wong [16], Kirk [8] and Brondsted [3] invoke some form of the Axiom of
Choice such as Zorn’s Lemma, the constructive proofs such as those by Penot [10]
and Siegel [11] are indeed based on the axiom of choice for countable families. On
the other hand, Manka showed that Caristi’s theorem can be proved without choice
using Zermelo’s fixed point theorem for special posets. (See Kirk [9] and Jachymski
[7] for detailed comments and references.)
Jachymski [7] noted that Nadler’s fixed point Theorem 5.3.5 can be deduced from
Caristi’s Theorem 7.2.1.
Theorem 7.2.14 (Jachymski [7]) Let (X, d) be a complete metric space and T :
X → cl B X (the set of all bounded non-empty closed subsets of X ) such that for some
α ∈ (0, 1). H (T x, T y) ≤ αd(x, y), x, y ∈ X . Then T admits a selection g : X → X
which satisfies the conditions of Caristi’s Theorem 7.2.1 and hence has a fixed point.
So does T .
Proof Let β ∈ (0, 1) be such that α < β. For each x ∈ X , the set {y ∈ T x :
β(x, y) ≤ d(x, T x)} is non-empty. By the Axiom of Choice there exists a map
g : X → X such that gx ∈ T x and d(x, gx) ≤ d(x, T x). So d(gx, T x) ≤ H (T x,
T (g(x))) ≤ αd(x, g(x)). So
156 7 Caristi’s Fixed Point Theorem
1
d(x, g(x)) ≤ (βd(x, g(x)) − αd(x, g(x))
(β − a)
1
≤ [d(x, T x) − d(g(x), T g(x))]
(β − a)
≤ φ(x) − φ(g(x))
d(x, T x)
where φ(x) =
(β − a)
Thus g is a Caristi map, a selection from T and hence has a fixed point in X . So does
(T x,T y)
T . Further |φ(x) − φ(y)| ≤ d(x,y)+Hβ−α
≤ β−a
1+α
d(x, y).
Remark 7.2.15 Nadler noted that the multivalued map of his theorem may not admit
a selection which is a contraction. Let X be the unit circle in the complex plane. For
α α
each z = eiα , α ∈ [0, 2π], let T (z) = {ei 2 , ei(π+ 2 ) }, the set of square roots of z. For
z i = eiαi , i = 1, 2,
2 sin(α42 −α1 if α2 − α1 ≤ π
H (T z 1 , T z 2 ) =
2 cos(α42 −α1 if α2 − α1 > π
√ √
and H (T z 1 , T z 2 ) ≤ 22 d(z 1 , z 2 ) = 22 |2 sin(α42 −α1 |. Following Jachymski [7] if T has
a selection g which is a contraction and if g(1) = 1, then by the continuity of g at
α
1, for some α0 > 0, g(eiα ) = ei 2 for all α ∈ [0, α0 ). By the continuity of g at α∗ =
α α∗
sup{α0 ∈ (0, 2π) : g(eiα ) = ei 2 for all α ∈ [0, α0 )}. For α∗ < 2π, we get ei 2 =
α∗
ei( 2 +π) so that α∗ = 2π. While g(ei(2π− n ) ) → eiπ as n → ∞, by the continuity of g
1
α
at 1, ei(2π− n ) → −1 and g(ei 2 ) → 1. This is a contradiction. If g(1) = −1, a similar
1
Ekeland [6] obtained a theorem in the setting of metric spaces that finds wide use
in solving optimization problems and partial differential equations. His proof uses a
partial order employed earlier by Brondsted and Rockafeller and Bishop and Phelps.
Theorem 7.3.1 (Ekeland [6]) Let (X, d) be a complete metric space and F : X →
R ∪ {+∞}, a lower semicontinuous function ≡ +∞ bounded from below. For each
> 0 and u ∈ X satisfying
and every λ > 0 there exists v ∈ X such that F(v) ≤ F(u), d(u, v) ≤ λ and for
w = v, F(w) > F(v) − λ d(v, w).
7.3 Ekeland’s Variational Principle 157
Lemma 7.3.2 Let (X, d) be a complete metric space. Define the binary relation
on X × R by (x1 , a1 ) (x2 , a2 ) iff (a2 − a1 ) + αd(x1 , x2 ) ≤ 0 where α is any given
positive number. Then is a partial order on X × R such that {(x, a) : (x1 , a1 ) ≤
(x, a)} is closed in X × R for each (x1 , a1 ) in X × R. Further if S is a closed subset
of X × R such that for all (x, a) ∈ S, there exists m ∈ R such that a ≥ m. Then for
every (x1 , a1 ) ∈ S, there exists for the ordering an element (x, a) which is maximal
and greater than (x1 , a1 ).
1 1
|an+1 − m n+1 | ≤ |an − m n | ≤ n |a1 − m|.
2 2
So for (x, a) ∈ Sn+1 we get
1
|an+1 − a| ≤ |a1 − m|
2n
1 1
d(xn+1 , x) ≤ n |a1 − m|
2 α
Thus Diam(Sn ) → 0 as n → ∞.
∞
As X × R is complete, Sn = {(x, a)}. From the definition of (x, a), (xn , an )
n=1
(x, a) for each n and in particular for n = 1. If (x0 , a0 ) ∈ S is greater than (x, a),
∞
then by transitivity (xn , an ) (x0 , a0 ) for each n. So (x0 , a0 ) ∈ Sn and hence
n=1
(x0 , a0 ) = (x, a). Thus (x, a) is maximal.
F(u) ≤ inf{F(x) : x ∈ X } +
Then there exists v ∈ X such that (i) F(v) ≤ F(u); (ii) d(u, v) ≤ 1 and for all w = v,
F(w) + d(v, w) > F(v).
Corollary 7.3.4 Under the hypotheses of Corollary 7.3.3 given > 0, there exists
x ∈ X with f (x) < inf f + .
Proof Let T : X → X be an operator such that d(x, T x) ≤ φ(x) − φ(T (x)) for all
x ∈ X , where φ is a non-negative lower semicontinuous function.
Proof By Corollary 7.3.3 for = 1, there exists v ∈ X such that for all u = v
If v = T (v), then φ(T (v)) − φ(u) < d(u, T v) a contradiction. So v = T (v). Thus
T has a fixed point.
Proof As before set = 1 in Corollary 7.3.3. So there exists v ∈ X such that for all
y = v, φ(v) < φ(y) + d(v, y). We claim that v ∈ T (v). Otherwise for all y ∈ T (v),
d(v, y) > 0 and φ(v) − φ(y) < d(v, y). But by hypothesis d(v, y) ≤ φ(v) − φ(y).
Thus φ(v) − φ(y) < φ(v) − φ(y), a contradiction. So for some y ∈ T (v), d(v, y) =
0 or v ∈ T (v).
7.4 A Minimization Theorem 159
Takahashi [13] proved a minimization theorem which can be deduced from Ekeland’s
variational principle.
Theorem 7.4.1 (Takahashi [13]) Let (X, d) be a complete metric space and F :
X → R ∪ {+∞} a lower semicontinuous map bounded below. Suppose F satisfies
the (Takahashi) condition: there exists α0 > 0 such that for x ∈ X − Z , there exists
y = x with α0 d(y, x) ≤ F(x) − F(y) Z being the set {z ∈ X : F(z) = inf F} (the
X
set of possible minima of F}. Then there exists x0 ∈ X with F(x0 ) = inf F (or Z is
X
non-void).
Proof For 0 < α < α0 , by Ekeland’s principle (Corollary 7.3.4) there exists x ∈ X
such that F(x) < F(y) + αd(y, x), ∀ y = x. By Takahashi condition for z ∈ X −
Z , there exists y = z such that F(y) + αd(y, z) ≤ F(z). If x ∈/ Z , the last inequality
for the choice z = x contradicts its preceding inequality. So x ∈ Z is a minimizer of
F over X .
d(x0 , x) ≤ F(x0 ) − F(x) = F(x0 ) − inf f ≤
λ X
So d(xλ0 ,x) ≤ 1 or d(x0 , x) ≤ λ. We also have F(x0 ) ≤ F(x). We claim that for
x ∈ X 0 and y = x, F(y) > F(x) − λ d(x, y). Otherwise for some y ∈ X , y = x,
F(y) ≤ F(x) − λ d(x, y). Then
d(y, x0 ) ≤ d(x0 , x) + d(x, y)
λ λ λ
≤ F(x0 ) − F(x) + F(x) − F(y)
= F(x0 ) − F(y).
So y ∈ X 0 . So by Theorem 7.4.1 there exists x ∈ X such that F(x) = inf x∈X F(x).
This contradicts that F(y) < F(x0 ), a contradiction.
Theorem 7.4.3 Caristi’s Theorem 7.2.1 implies Ekeland’s Variational principle, viz.
Corollary 7.3.4.
160 7 Caristi’s Fixed Point Theorem
Proof Let the hypotheses of Corollary 7.3.4 hold. Clearly d1 (x, y) = d(x, y)
defines an equivalent metric on X . For x ∈ x, define T (x) = {y ∈ X : F(x) ≥
F(y) + d1 (x, y), y = x}. If the conclusion of Corollary 7.3.4 is false, then Clearly
T (x) = φ for all x ∈ X . T is a multivalued map of X into 2 X − {φ} satisfying
F(y) ≤ F(x) − d1 (x, y) for y ∈ T x. Since (X, d1 ) is complete, by Caristi’s theo-
rem, T has a fixed point x0 . However for x0 ∈ T x0 the definition of T (x) is violated.
This contradiction shows that Corollary 7.3.4 is true.
Remark 7.4.4 Thus both Caristi’s fixed point theorem and Takahashi’s minimization
theorem are equivalent to Ekeland’s variational principle. In other words Caristi’s
theorem, Ekeland’s principle and Takahashi’s theorem are equivalent.
Since Carisit’s theorem is characteristic of completeness, both Takahashi’s the-
orem and Ekeland’s principle are also characteristic of completeness. Sullivan [12]
proved that Ekeland’s principle implies the completeness of the metric space.
Theorem 7.4.5 (Sullivan [12]) Let (X, d) be a metric space and F : X → R ∪
{+∞}, F ≡ +∞ be any continuous map bounded below such that for each >
0, there exists x0 ∈ M satisfying F(x0 ) ≤ inf F + and for all x = x0 , F(x) >
F(x0 ) − d(x, x0 ).
Proof Suppose that (xn ) is a non-convergent Cauchy sequence in X . For each x ∈
F, the x → F(x) = lim d(x, xn ) is well-defined, continuous on X and bounded
n→∞
below by 0. Since (xn ) is Cauchy inf F = 0 and F ≡ +∞. Let 0 < < 1 then by
assumptions on F, we can find x0 ∈ X such that F(x0 ) ≤ and F(x) > F(x0 ) −
d(x, x0 ) for all x = x0 . As (xn ) is non-convergent F(x0 ) > 0.
Let η be any positive number less than (1−)F(x 4
0)
. Since (xn ) is Cauchy we can
find x N0 such that d(x N0 , xn ) ≤ η for all n ≥ N0 so that F(x N0 ) = lim (x N0 , xn ) ≤ η.
n→∞
Setting x = x N0 in the inequality
we get
or
(1 − )F(x0 )
0 < (1 − )F(x0 ) < 2η <
2
This contradiction shows that F(x) cannot be positive for all x ∈ X . In other words
(xn ) must converge in X or (X, d) is complete.
7.4 A Minimization Theorem 161
The next theorem shows that Takahashi’s minimization theorem also characterizes
completeness.
Theorem 7.4.6 A metric space (X, d) is complete if for every uniformly continuous
function f : X → R ∪ {+∞}, f ≡ +∞ and every x ∈ X with inf X f < f (x )
there exists z ∈ X with z = x and f (z) + d(x , z) ≤ f (x ), there exists x0 with
f (x0 ) = inf f .
X
Definition 7.5.1 For locally convex linear topological spaces X and Y and U ⊆ X
open, the map F : X → Y is said to be Gateaux differentiable at u ∈ U if for each
F(u + th) − F(u)
h ∈ X , lim exists. This limit, denoted by d F(u, h) is called the
t→0 t
Gateaux differential at u and is homogeneous.
Remark 7.5.2 If the Gateaux differential is linear and continuous, it is called Gateaux
derivative. Since the Gateaux differential of a discontinuous linear function is itself,
it follows that a Gateaux differential can be linear without being continuous.
Definition 7.5.3 A map F : X → Y where X and Y are normed linear spaces is said
to be Frechet differentiable at u if there exists a bounded linear function L such that
F(u + h) = F(u) + L(h) + o(h).
Proof From Ekeland’s Principle (Theorem 7.3.1), there exists x ∈ X such that
φ(x ) ≤ φ(x) + x − x for all x ∈ X . Let h ∈ X and t > 0. Setting x = x + th
in the above inequality we get
1
[φ(x ) − φ(x + th)] ≤ h
t
We can even ensure the existence of a critical point for φ which is a minimum
under an additional condition.
Proof Setting = 1
n
in Theorem 7.5.5 we get a sequence (xn ) ∈ X such that
1 1
φ(xn ) ≤ inf φ + and φ (xn ) ≤
X n n
References
1. Altman, M.: A generalization of the Brezis-Browder principle on ordered sets. Nonlinear Anal.
6, 157–165 (1982)
2. Brezis, H., Browder, F.E.: A general principle on ordered sets in nonlinear functional analysis.
Adv. Math. 21, 355–364 (1976)
3. Brondstedt, A.: On a lemma of Bishop and Phelps. Pac. J. Math. 55, 335–341 (1974)
4. Caristi, J.: Fixed point theorems for mappings satisfying inwardness conditions. Trans. Am.
Math. Soc. 125, 241–251 (1976)
5. De Figuiredo, D.G.: Lectures on the Ekeland Variational Principle with Applications and
Detours. T.I.F.R (1972)
6. Ekeland, I.: On the variational principle. J. Math. Anal. Appl. 47, 324–353 (1974)
References 163
7. Jachymski, J.R.: Caristi’s fixed point theorem and selections of set-valued contractions. J. Math.
Anal. Appl. 229, 55–67 (1998)
8. Kirk, W.A.: Caristi’s fixed point theorem and metric convexity. Colloq. Math. 36, 81–86 (1976)
9. Kirk, W.A.: Contraction mappings and extensions. In: Kirk, W.A., Sims, B. (eds.) Handbook
of Metric Fixed Point Theory. Springer, Berlin (2001)
10. Penot, J.P.: A short constructive proof of the Caristi fixed point theorem. Publ. Math. Univ.
Paris 10, 1–3 (1976)
11. Siegel, J.: A new proof of Caristi’s fixed point theorem. Proc. Am. Math. Soc. 66, 54–56 (1977)
12. Sullivan, F.: A characterization of complete metric spaces. Proc. Am. Math. Soc. 83, 345–346
(1981)
13. Takahashi, W.: Existence theorems generalizing fixed point theorems for multivalued mappings.
In: Baillon, J.B., Thera, M. (eds.) Fixed Point Theory and Applications, vol. 252, pp. 387–406.
Longman Scientific & Technical, Harlow (1991)
14. Turinici, M.: A generalization of Altman’s ordering principle. Proc. Am. Math. Soc. 90, 128–
132 (1984)
15. Turinici, M.: Remarks about a Brezis-Browder principle. Fixed Point Theory 4, 109–117 (2003)
16. Wong, C.S.: On a fixed point theorem of contractive type. Proc. Am. Math. Soc. 57, 282–284
(1976)
Chapter 8
Contractive and Non-expansive
Mappings
In this chapter, fixed points of contractive and non-expansive mappings are studied,
as also the convergence of their iterates.
A contractive mapping need not have a fixed point in a complete metric space as
seen from the following example.
Remark 8.1.2 The map x → e−x mapping R+ into itself has no fixed point. For 0 ≤
x < y, 0 < e−x − e−y = (= |e−x − e−y |) = e−ξ (y − x) < y − x = |x − y| by the
Mean-value Theorem.
In this connection the theorem below leads to the existence of a fixed point for
contractive mappings under a set of suitable conditions.
Proof While the existence follows from Theorem 8.1.3 the proof of the uniqueness is
left as an exercise. Since (xn k ) converges to u, given > 0, there exists N (), a posi-
tive integer such that d(u, xn k ) < . For all k ≥ N (). Now for n > n N () then for p =
n − n N () , d(u, xn ) = d(T p u, T p xn N ) ≤ d(T p−1 u, T p−1 xn N ) < d(u, xn N ) < . So
(xn ) converges to u.
Corollary 8.1.5 Let (X, d) be a compact metric space. Then every contractive self
map on X has a unique fixed point to which every sequence of iterates converges.
Remark 8.1.6 Reduction of the problem of finding the fixed point of the operator T
to that of finding the zero of the map x → d(x, T x) in the metric setting has been
suggested for instance in Dieudonne [3].
Even as Kannan’s Corollary 5.2.2 or the more general fixed point Theorem 5.2.1
apply to mappings which are not necessarily continuous, Theorem 8.1.3 has an
analogue for operators which need not be continuous.
a1 +a3 +a5
(ii) a2 + a3 < 1 and 1−a2 −a3
= 1.
Then T has a fixed point. If further a4
1−a3 −a5
< 1, then the fixed point is unique.
8.1 Contractive Mappings 167
So
(1 − a2 − a3 )d(y, T y) ≤ (a1 + a3 + a5 )r
1 +a3 +a5
As a1−a 2 −a3
= 1, it follows that d(y, T y) = r . If y = T y, then (ii) gives (1 − a2 −
a3 )d(T y, T y) < (a1 + a3 + a5 )d(y, T y). So d(T y, T 2 y) < d(y, T y) = r , contra-
2
1
d(T x, T y) < [d(x, T x) + d(y, T y)]
2
Then T has a unique fixed point.
Edelstein’s result (Corollary 8.1.4) may not be true for contractive maps on closed
bounded sets which are not compact, as shown by the following.
Example 8.1.9 (Ira Rosenholtz [22]) Let c0 be the Banach space of all null real
sequences with the supremum norm and S, the closed unit ball in c0 . Define T :
S → S by
168 8 Contractive and Non-expansive Mappings
T (x1 , x2 , . . . , xn , . . . ) = (1, a1 x1 , a2 x2 , a3 x3 , . . . , an xn , . . . )
where an is a sequence of positive real numbers such that each ak is less than 1 and
n
n
+1
Pn = a j is bounded away from zero. For example an = 22n +2 . Clearly T maps S
j=1
into the boundary of S as T x = 1 for x ∈ S. Also if x = (x1 , x2 , . . . , xn , . . . ) is a
fixed point of T , then T x = (y1 , . . . , yn , . . . ) where y1 = 1 and yn = an−1 xn−1 for
(1 + 1k )
n−1
1 + 2n−1
1
1
n > 1, so that yn = Pn−1 = 2
≥ ≥ . As (yn ) ∈
/ c0 , T cannot
k=1
(1 + 2k−1 ) 1 2 2
have a fixed point in S.
N
P
p
n m
B M C
Area of ABC = Area of ABM + Area of AMC. So
1 1 A 1 A
bc sin A = cm sin + bm sin
2 2 2 2 2
8.1 Contractive Mappings 169
or
2bc A
m= cos .
b+c 2
Since
a 2 = b2 + c2 − 2bc cos A
A
= b + c − 2bc 2 cos
2 2 2
−1
2
we get
⎫
m = (b+c)
1
bc[(b + c)2 − a 2 ⎪⎬
n = (c+a)
1
ca[(c + a)2 − b2 ] (8.1.1)
⎪
⎭
p = (a+b)
1
ab[(a + b)2 − c2 ]
Theorem 8.1.10 (Mironescu and Panaitopol [16]) The problem of finding the trian-
gle, given the lengths of internal angle bisectors is equivalent to a problem of finding
the fixed point of a suitable mapping.
Proof The set of Eq. (8.1.1) can be rewritten suitably. For instance the first equation
in (8.1.1) is
bc[(b + c)2 − a 2 ] − (b + c)2 m 2 = 0
It can be written as
(b + c)2 − (b − c)2
4m 2 = [(b + c)2 − a 2 ]
(b + c)2
or
(b − c)2 a 2
4m 2 = (b + c)2 + − [a 2 − (b − c)2 ]
(b + c)2
170 8 Contractive and Non-expansive Mappings
2
(b − c)a
= (b + c) ± − [a ± (b − c)]2
(b + c)
(b−c)a
Eliminating (b+c)
between those two equations one gets
2(b + c) = 4m 2 + (c + a − b)2 + 4m 2 + (a + b − c)2
Define x, y, z by
a = y + z, b = z + x, c = x + y
whence
b+c−a c+a−b a+b−c
x= , y= , z=
2 2 2
Using these in the above expression for m we get
1 2 1
x= [ m + y 2 − y] + [ m 2 + z 2 − z]
2 2
1 2 1
y = [ n + z 2 − z] + [ n 2 + x 2 − x]
2 2
1 2 1
z = [ p + x 2 − x] + [ p 2 + y 2 − y]
2 2
Since m, n, p > 0, x, y, z > 0. Further x < m, y < n and z < p. Thus the map
F : K → K where K = [0, m] × [0, n] × [0, p] defined√by F(x, y, z) = ( f m (y) +
f m (z), f n (z) + f n (z), f p (x) + f p (y)) where f α (t) = 21 [ α 2 + t 2 − t] on [0, α] for
α > 0 has a fixed point if and only if the system of Eq. (8.1.1) has a solution. a, b, c
in terms of m, n and p.
Remark 8.1.11 Mironescu and Panaitopol invoked Brouwer’s fixed point theorem
9.1 - to conclude that F has a fixed point which is the solution to the three internal
angle bisectors problem.
On the other hand Dinca and Mawhin [4] deduced it from the contraction principle
though it can also be obtained from Edelstein’s theorem (Corollary 8.1.4).
Theorem 8.1.12 (Dinca and Mawhin [4]) The map F : K → K defined by F(x, y,
z) = ( f m (y) + f m (z),√f n (z) + f n (x), f p (x) + f p (y)) where K = [0, m] × [0, n] ×
[0, p] and f α (t) = 21 [ α 2 + t 2 − t] is contractive and has a unique fixed point and
every sequence of F-iterates converges to the unique solution of the three internal
angle bisectors problem.
√
Proof f α (t) = 21 [ α 2 + t 2 − t] is continuous and f α (t) = 21 √α2t +t 2 − 1 is neg-
ative. Further, | f α (t)| < 21 . Now K is compact in R3 with the norm (x, y, z) =
max{|x|, |y|, |z|}. Further, | f α (t1 ) − f α (t2 )| < 21 |t1 − t2 | for t1 = t2 . So for (x, y, z)
= (x1 , y1 , z 1 ) F(x, y, z) − F(x1 , y1 , z 2 ) < max{|x − x1 |, |y − y1 |, |z − z 1 |}.
8.1 Contractive Mappings 171
Further as f α (t) < α2 and F maps K into itself. Thus F is a contractive self-map on
the compact space K . So by Edelstein’s theorem (Corollary 8.1.4) F has a unique
fixed point to which every sequence of F iterates converges. In other words there
is a unique triangle up to isometry having with prescribed lengths of internal angle
bisectors.
Remark 8.1.13 Dinca and Mawhin [4] proved Theorem 8.1.12 applying the contrac-
tion principle to a sequence of contractions converging pointwise to F.
In this section, some elementary results on the fixed points of non-expansive maps
are discussed.
Remark 8.2.2 A non-expansive mapping on a complete metric space may not have
a fixed point as is evident by considering the map x → x + a, a = 0 on R. Indeed
themap eit → ei(t+α) t ∈ [0, 2π ) where α ∈ (0, 2π ) in the unit circle in the complex
plane is an isometry without a fixed point on the compact connected locally connected
space S 1 .
For x = (xn ) in the closed unit sphere of c0 , the Banach space of all null sequences
x → T x = (1, x1 , x2 , . . . ), x = (xn ) is a fixed point free isometry.
Dotson Jr [5] proved a fixed point theorem for non-expansive maps in the setting
of star-shaped subsets of normed linear spaces.
Remark 8.2.4 While all convex sets are star-shaped about every one of its points a
star-shaped subset may not be convex as in evident by considering S = {0} × [0, 1] ∪
[0, 1] × {0} in R2 .
Proof Without loss of generality we can assume that S is star-shaped about 0 (i.e. 0 is
a star-centre of S). Then the map Tn defined by Tn (x) = (1 − n1 )T x (= n1 .0 + (1 −
1
n
)T x) maps S into itself for each n ∈ N. As Tn is a contraction on S and S being
compact is complete, Tn has a unique fixed point xn , say. So Tn (xn ) = (1 − n1 )T xn =
xn for each n ∈ N. As S is compact and xn ∈ S for each n ∈ N, there is a subsequence
172 8 Contractive and Non-expansive Mappings
While every element of X − S of a metric space X need not have a best approxima-
tion in S ⊆ X , the following proposition gives sufficient conditions for the existence
of a best approximation.
This corollary follows from Corollary 8.2.10 upon setting T1 : C[B] → C[B] by
T1 (g(x)) = A(g(T (x))).
Corollary 8.2.12 Let f be an even (odd) function in C[−1, 1] with the supremum
norm. If V is a finite dimensional subspace of C[−1, 1] such that whenever h(x) ∈ V ,
h(−x) also is in V , then f has an even (odd) function as best approximation in V .
As
1 u1 + u2 u1 + u2
u1 − u2 ≤ + u 2 − 1 u 1 + u 2 + F u 1 + u 2 ,
u 1 − 2 2
+F
2 2 2 2
the inequality
1
u i − 1 u 1 + u 2 + F u 1 + u 2 ≥
2 2 2 2 u1 − u2 (8.3.3)
is true for at least one of i = 1, 2. From Lemma 8.3.4 and the inequalities (8.3.1),
(8.3.2) and (8.3.3)
8.3 Browder–Gohde–Kirk Fixed Point Theorem 175
u1 + u2 u1 + u2
− F ≤ + 1 u1 − u2 η
2 2 2 + u 1 −u
2
2
≤ sup (ξ + )η
0<ξ < d(K ) +ξ
2
⎡ ⎤
⎣ ⎦
≤ max sup
√
( + ξ )η , sup ( + ξ )η
0<ξ ≤ − +ξ √
−<ξ ≤ d(K ) +ξ
2
√ d(K ) √
≤ max 2 , + η( ) = φ() (say).
2
So u 1 , u 2 ∈ C implies u 1 +u
2
2
∈ Cφ() . Clearly φ() → 0 as → 0.
For u 1 , u 2 ∈ D , vi ≤ a + for i = 1, 2 and as u 1 +u 2
2
∈ Cφ() , u 1 +u
2
2
≥
a(Cφ() )
Now by Lemma 8.3.4
a + − a(Cφ() )
d(D ) = sup u 1 − u 2 ≤ (a + )η
u 1 ,u 2 ∈D a+
∞
1 1
{t : 0 < g(t) < 2} = t : n ≤ g(t) < n−1
n=0
2 2
The centre c(A) for A is the set of all central points for A and r (A) is the radius of
A.
Proposition 8.4.2 ([18]) For a metric space (X, d) satisfying for each r > 0
and x, y ∈ X with x = y there exists δ > 0, z ∈ X such that B(x, r ) ∩ B(y, r ) ⊆
B(z, r − δ) given a bounded non-void subset A of X , c(A) contains at most one
point.
Proof If x, y ∈ c(A) and x = y, then for r = r (A), A ⊆ B(x, r ) ∩ B(y, r ) ⊆ B
(z, r − δ). This implies r (A) = r ≤ r − δ, a contradiction.
A centre of a set corresponds to Chebyshev centre of a set.
For developing fixed point theory Pasicki [18] introduced the following definition.
Definition 8.4.3 ([18]) Let Y be a non-void bounded subset of a metric space (X, d)
and F : Y → 2Y , a mapping with F(y) = φ for each y ∈ Y . x ∈ X is called a central
point for F if
The centre c(F) for F is the set of central points for F and r (F) is the radius of F.
8.4 A Generalization to Metric Spaces 177
Theorem 8.4.4 ([18]) Let Y be a non-void bounded subset of a metric space (X, d)
and f : X → X a map such that f (Y ) ⊆ Y and c( f |Y ) = {x}. If d( f (x), f (y)) ≤
d(x, y) for all x, y ∈ Y , then x is a fixed point for f .
Proof For f n−1 (Y ) ⊆ B(x, t), f n (Y ∩ B(x, t)) ⊆ B( f (x), t). As f is non-
expansive f (Y, B(x, t)) ⊆ B( f (x), t) and so f n (Y ) ⊆ B( f (x), t) implying that
f (x) ∈ c( f |Y ). As c( f |Y ) is a singleton, f (x) = x.
The above theorem leads to the formulation of a bead space and a discus space.
Definition 8.4.5 ([18]) A metric space is called a bead space if the following con-
dition holds:
for every r , β > 0, there exists a δ > 0 such that for each pair x, y ∈ X with
d(x, y) ≥ β, there exists a z ∈ X such that B(x, r + δ) ∩ B(y, r + δ) ⊆ B(z, r − δ).
Definition 8.4.6 ([17, 18]) A metric space is called a discus space if there exists a
map ρ : [0, ∞) × (0, ∞) → [0, ∞) such that
So η(β, [r0 , r0 + )) ⊆ (α, ∞). Finally one has η(β, (r0 − , r0 + )) ⊆ (α, ∞) and
η(β, ·) is lower semicontinuous. Also B(x, r ) ∩ B(y, r ) ⊆ B(x, r + δ) ∩ B(y, r +
δ) ⊂ B(z, r − δ) ⊆ B(z, r − δ + ) and so B(x, r ) ∩ B(y, r ) ⊆ B(z, ρ(d(x, y), r )
+ ). Thus (X, d) is a discus space.
Remark 8.4.9 If (X, d) is a metric space and r > 2 > 0, then B(x, r − 2) ⊆
B(x, r − ) ⊂ B(x, r )
So the definitions of bead and discus spaces can be formulated using closed balls.
We need the following lemma to decode the geometry of normed bead spaces.
Lemma 8.4.10 ([19]) In a normed linear space (X, ) the following conditions
are equivalent:
(i) for every r, β > 0, there exists δ > 0 such that for x, y ∈ X with x − y ≥ β,
there exists z ∈ X with B(x, r + δ) ∩ B(y, r + δ) ⊆ B(z, r − δ);
(ii) for every r, β > 0, there exists δ > 0 such that for x, y ∈ X with x − y ≥ β,
there exists z ∈ X such that B(x, r + δ) ∩ B(y, r + δ) ⊆ B(z, r );
(iii) for every r, β > 0, there exists δ > 0 such that for x ∈ X and 2 x ≥ β imply
B(−x, r ) ∩ B(x, r ) ⊆ B(0, r − δ).
Proof It is enough to prove the equivalence for y = −x. The set C = B(−x, r ) ∩
B(x, r ) is symmetric and therefore C ⊆ B(z, t) implies C ⊆ B(−z, t) ∩ B(z, t) ⊆
B(0, t). Thus we may let z = x+y 2
= 0 in (i), (ii) and (iii). Clearly (i) implies (ii)
and (iii). Choosing r = 1 in (iii) we get for β > 0 there exists δ > 0 such that for
u, x ∈ X with u + x , u − x < 1 and 2 x ≥ β > 0
8.4 A Generalization to Metric Spaces 179
(iv) u < 1 − δ
Suppose u + x , u − x < r + and 2 x ≥ (r + )β hold. Then u+x r +
,
u−x
r+
< 1 and 2 x
≥ β and in view of (iv) u
< 1 − δ or u < (r + )(1 − δ) <
r + r +
r 1 − 2δ for small values of . For y = −x, δ1 = min , r2δ , and 2 x ≥ β1 =
3rβ
2
> (r + )β, we get B(x, r + δ1 ) ∩ B(y, r + δ1 ) ⊆ B(0, r − δ1 ) which is (i).
Thus (i), (iii) and (iv) are equivalent. Now consider (ii). Then (ru+x , u−x < 1 + δ
−) (r −)
and (r2 −)
x
≥ β imply r −
u
< 1. (see (ii)). This relation can be written as u + x , u −
x < (r − )(1 + δ) = r + r δ − (1 + δ) and 2 x ≥ rβ > (r − )β implying
rδ
u < r − . Thus for < r δ − (1 + δ) or < 2+δ , u+x , u−x <r +
and 2 x ≥ rβ implying u < r − , verifying (i).
Theorem 8.4.11 ([19]) A normed linear space is uniformly convex if and only if
for each β > 0 there exists δ > 0 such that for u, x ∈ X , u + x , u − x < 1
and 2 x > β > 0 imply
(v) u < 1 − δ.
Proof From (iv) of Lemma 8.4.10 setting u = y+z 2
, x = y−z
2
one gets
for every β > 0, there exists δ > 0 such that y, z ∈ X , y , z < 1 and y −
z > β > 0 imply y+z 2
< 1 − δ. So X is uniformly convex.
If (v) is satisfied then writing y = u + x, z = u − x, then (iv) of Definition 8.3.1
is satisfied leading to uniformly convexity of X .
We now have
Theorem 8.4.12 ([19]) A normed linear (X, · ) is uniformly convex if and only
if any one of the conditions (i), (ii), (iii), (iv) of Lemma 8.4.10 and (v) of Theorem
8.4.11 are satisfied.
From Propositions 8.4.7, 8.4.8 and Theorem 8.4.11 we have
Theorem 8.4.13 ([19]) For any normed linear space the following conditions are
equivalent:
(i) X is a bead space;
(ii) X is a discus space;
(iii) X is a uniformly convex space.
Remark 8.4.14 ([19]) Each convex subset X of a uniformly convex normed linear
space satisfies the definition of a bead space with z = x+y
2
(see Definition 8.4.5).
We now proceed to provide Pasicki’s generalization of Browder-Gohde-Kirk The-
orem 8.3.3 based on the following lemmata.
Lemma 8.4.15 If (X, d) is a complete discus space, then (iv) of Definition 8.4.6 (of
Discus space) can be replaced by
for each x, y ∈ X and r > 0 there is a z ∈ X such that B(x, r ) ∩ B(y, r ) ⊆
B(z, ρ(d(x, y), r )).
180 8 Contractive and Non-expansive Mappings
Proof Let α = ρ(d(x, y), r ) and (αn ) ↓ α be such that there exist xn ∈ B(x, r ) ∩
B(y, r ) ⊆ B(xn , αn ). Suppose (xn ) is not Cauchy. Then there is β > 0 with d(xn , xk )
⊂ β for infinitely many k, n with k < n. Set 2η = α − ρ(β, α) = ρ(0, α) − ρ(β, α)
> 0 (by (i) of Definition 8.4.6). So B(x, r ) ∩ B(y, r ) ⊆ B(xn , αn ) ∩ B(xk , αk ) ⊂
B(xn , αk ) ∩ B(xk , αk ) ⊂ B(z n,k , ρ(d(xn , xk ), αk ) + η) for some z n,k ∈ X (by (iv)
of Definition 8.4.6). On the other hand, ρ(d(xn , xk ), αk ) ≤ ρ(β, αk ) (by ii) and
ρ(β, αk ) ≤ ρ(β, α) + η for sufficiently large k (by (iii) of Definition 8.4.6). Now one
gets B(x, r ) ∩ B(y, r ) ⊂ B(z n,k , ρ(β, α) + η) = B(z n,k , α − 2η + η) = B(z n,k ,
α − η) ⊂ B(z n,k , α). Thus the lemma is true. If (xn ) is a Cauchy sequence con-
vergent to z ∈ X , then B(xn , αn ) ⊂ B(z, α + β) for any β > 0 and all large n.
So B(x, r ) ∩ B(y, r ) ⊆ B(z, α + β) for all β > 0 and B(x, r ) ∩ B(y, r ) ⊆ B(z, α).
Since B(x, r ) ∩ B(y, r ) is open, the lemma follows.
Lemma 8.4.16 Let (X, d) be a complete discus space and let A ⊆ X be non-void
and bounded then c(A) is a singleton.
Proof Let (rn ) ↓ r = r (A) while A ⊆ B(xn , rn ). If (xn ) is not Cauchy, then d(xn , xk )
≥ β > 0 for infinitely many k, n with k < n. We have
So A ⊆ B(z n,k , r (A) − η (η being 21 (α − ρ(β, α))) as proceed in the course of the
preceding Lemma 8.4.15. This is a contradiction. Let (xn ) converge to x. Then for
any β > 0 B(xn , rn ) ⊆ B(x, r + β) for sufficiently large n implying A ⊆ B(x, r +
β) for all β > 0 and so x ∈ c(A). If x, y ∈ c(A) and d(x, y) ≥ β > 0, then by
Lemma 8.4.15 A ⊆ B(x, r ) ∩ B(y, r ) ⊆ B(z, ρ(β, r )) ⊆ B(z, r − η) for a η > 0,
a contradiction. So c(A) is a singleton.
Proof Let r = r (F) (refer Definition 8.4.3 for r (F) and c(F)). F n+1 (Y ) ⊆ F n (Y )
and so there exists rn ↓ r and a sequence (xn ) such that F n (y) ⊆ B(xn , rn ) for all n. If
(xn ) is not a Cauchy sequence, for infinitely many n, k with k < n d(xn , xk ) ≥ β > 0.
We have F n (Y ) ⊆ F n (Y ) ∩ F k (Y ) ⊂ B(xn , rn ) ∩ B(xk , rk ) ⊆ B(z n,k , ρ(β, rk )) and
so F n (Y ) ⊆ B(z n,k , r − η) for a η > 0 (as in the preceding proof), a contradiction.
Let (xn ) converge to x. We get F n (Y ) ⊆ B(x, r + β) for any β > 0 and sufficiently
large n. So x ∈ c(F). That c(F) is a singleton follows from Lemma 8.4.16.
Theorem 8.4.18 Let (X, d) be a complete bead (discus) metric space and f : X →
X a non-expansive map. If Y is a bounded non-empty subset of X with f (Y ) ⊆ Y ,
then c( f |Y ) is a singleton which is a fixed point of f .
8.5 An Application to a Functional Equation 181
In this section, Matkowski’s deduction [13] of a fixed point theorem from Pasicki’s
Theorem 8.4.4 is discussed as well as its application to a functional equation. It is in
the setting of a paranormed space.
Definition 8.5.1 Let X be a real linear space over R. A function p : X → R is called
a paranorm if the following conditions are satisfied:
(i) p(x) = 0 if and only if x = 0, p(x) = p(−x) for all x;
(ii) p(x + y) ≤ p(x) + p(y) for all x, y ∈ X ;
(iii) for tn , t ∈ R and xn , x ∈ X with tn → t and p(xn − x) → 0 as n → ∞,
p(tn xn − t x) converges to zero as n → ∞.
Remark 8.5.2 If (X, p) is a paranormed linear space, then d(x, y) = p(x − y)
defines a metric on X .
Definition 8.5.3 A paranormed space (X, p) is called uniformly convex if for
each r > 0 and ∈ (0, 2r ) there exists δ(r, ) ∈ (0, r ) such that for all x, y ∈ X
with p(x), p(y) ≤ r and p(x, y) ≥ , p( x+y2
) ≤ r − δ(r, ) (the function δ : →
(0, ∞) where = {(r, ) : r > 0, 0 < < 2r } is called the modulus of convexity
of (X, p).
The following theorem due to Matkowski [13] provides a class of examples of
paranormed spaces.
Theorem 8.5.4 Let ( , S , μ) be a measure space and S = S( , S , μ) the real
linear space of all μ-integrable simple functions x : → R. Let ϕ : [0, ∞) →
be an increasing bijection with ϕ(0) = 0. Then the functional pϕ (x) =
[0, ∞)
ϕ −1 ϕ(|x|)dμ is well-defined for each x ∈ S. If μ( ) = 1 and if there exists a set
A ∈ S with 0 < μ(A) < 1, then pϕ is a paranorm if and only if F : R+ × R+ → R+
defined by F(r, s) = ϕ(ϕ −1 (r ) + ϕ −1 (s)) r, z ∈ R+ is concave.
If μ( ) ≤ 1 and F is concave, then pϕ is a paranorm on S.
Remark 8.5.5 Let ϕ : R+ → R+ be twice differentiable with ϕ(0) = 0, ϕ (r ), ϕ
(r )
ϕ ϕ (r +s) ϕ (r )
> 0 for r > 0. If ϕ is super additive in (0, ∞) in the sense that ϕ (r +s) ≥ ϕ (r ) +
ϕ (s)
r, s > 0 then F : R+ × R+ → R+ defined by F(r, s) = ϕ(ϕ −1 (r ) + ϕ −1 (s))
ϕ (s)
,
is concave.
For the proofs of the following lemmata Matkowski [13] may be consulted.
Lemma 8.5.6 ([13]) Let ( , S , μ) be a measure space with μ( ) ≤ 1. Let ϕ :
R+ → R+ be an increasing bijection with ϕ(0) = 0 and F : R+ × R+ → R+ be
defined by F(r, s) = ϕ(ϕ −1 (r ) + ϕ −1 (s)), r, s ≥ 0. If F is concave, then pϕ defined
(in Theorem 8.5.4) is a paranorm on S = S( , S , μ). If further ϕ(r + s) + ϕ(|r −
s|) ≥ 2[ϕ(r ) + ϕ(s)] for all r, s ≥ 0 (i.e. ϕ is super quadratic), then for x, y ∈ S
d(u, x), d(u, y) < r + δ and d(x, y) > β. Then for s = u − x, t = u − y, p(s) <
r + δ, p(t) < r + δ and for s − t = y − x ∈ X − X ⊆ X , p(s − t) = p((u − x) −
(u − y)) = p(y − x) > β. So by assumption of Lemma 8.5.9, p(u − x+y 2
)=
u−x+u−y
p( 2
) = p( 2 ) < r − δ. Thus u ∈ B(z, r − δ) where z = 2 ∈ X . Thus
s+t x+y
Remark 8.5.10 The proof of the above lemma is due to Pasicki [20]. In the above
lemma X can be the whole space.
Remark 8.5.11 From Pasicki’s Theorem 8.4.4 it follows that a non-expansive map-
ping of a non-void bounded closed convex subset C of a complete uniformly convex
paranormed space X into itself has a fixed point, the modulus of convexity of X
being continuous.
Theorem 8.5.13 (Matkowski [13]) Let = [0, 1], S the σ -algebra of Lebesgue
measurable sets and μ the Lebesgue measure. Let ϕ : R+ → R+ be an increasing,
convex, superquadratic function such that F(r, s) = ϕ(ϕ −1 (r ) + ϕ −1 (s)) is concave,
r, s > 0. Let f : I → I be an increasing differentiable function, h : I × R → R
satisfy the Caratheodary conditions:
8.5 An Application to a Functional Equation 183
If fg ≤ 1, then the functional equation x(t) = h(t, x( f (t))) has a solution in
ϕ
S ( , S , μ), the completion of the paranormed space S = Sϕ ( , S , μ).
It may be mentioned that for the choice ϕ(t) = t p , p ≥ 1, the pϕ norms coincide
with L p norms.
While Banach’s contraction principle and its variants are concerned with the conver-
gence of the iterates of the operator under consideration it is also relevant to discuss
the summability of iterates. For instance, Krasnoselski’s theorem investigates the
behaviour of the sequence xn+1 = 21 (xn + T xn ) (vide section 2.2). Mann [12] gen-
eralized Krasnoselski’s theorem to more general sequences generated by regular
matrices. For discussing Mann’s theorem we need the following.
Theorem 8.6.2 (Silverman–Toeplitz, see Hardy [9]) An infinite matrix (cmn ) is reg-
ular if and only if the following are true:
(i) lim cmn = 0 for each n ∈ N ;
m→∞
∞
(ii) lim cmn = 1 and
m→∞
∞
n=1
(iii) sup |cmn | ≤ K < +∞ for some K > 0
m
n=1
Corollary 8.6.3 Let A be the infinite triangular matrix satisfying the following con-
ditions:
(a) ai j ≥ 0 for i, j ∈ N;
(b) ai j = 0 for all j > i
8.6 Convergence of Iterates in Normed Spaces 185
i
(c) ai j = 1 for all i ∈ N.
j=1
Then A is regular.
Proof Clearly (b) implies (i) and (c) and (a) imply both (ii) and (iii).
Theorem 8.6.6 (Mann [12]) Suppose neither {xn } nor {vn } (defined in Theorem
8.6.4) is convergent. Let X be the set of all limit points of {xn } and V the set of all limit
n
points of {vn }. If A satisfies additionally lim ann = 0 and lim |an+1k − ank | =
n→∞ n→∞
h=1
0, then X and V are closed connected sets.
Proof For a separation of V by two non-void closed sets A1 and A2 , we can find v1 ∈
A1 and v2 ∈ A2 such that d(v1 , v2 ) = inf{d(x, y) : x ∈ A and y ∈ B} = r > 0. For
186 8 Contractive and Non-expansive Mappings
ni
r
vi , we can find am i k vik ∈ B(vi , ) ∩ Ai , i = 1, 2, leading to a contradiction to the
k=1
3
choice of v1 and v2 with d(v1 , v2 ) = r , in view of the Silverman–Toeplitz conditions
n
and the additional hypotheses lim ann = 0 and lim |an+1k − ank | = 0. So V is
n→∞ n→∞
k=1
connected. Since T is continuous and X = T (V ), X is connected. V being compact,
so is X .
When T has a unique fixed point and E = [a, b] in R, the sequence (vn ) converges
to the unique fixed point.
Theorem 8.6.7 (Mann [12]) Let T : [a, b] → [a, b] be continuous with a unique
1
n
fixed point p. Define xn by xn+1 = T (vn ) where vn = xk where x1 ∈ [a, b].
n k=1
Then xn converges to p.
Proof Define A = (ank ) by
1
for k ≤ n
ank = n
0 for k > n
In this section, two basic results of Ishikawa [10] on the iterates of a non-expansive
mapping are detailed. These are in the setting of a Banach space. We make use of
the following definitions and lemmata.
Definition 8.7.1 Let D be a subset of a Banach space X , T : D → X a map and
x1 ∈ D. By M(x1 , tn , T ) we denote the sequence (xn ) defined by
xn+1 = (1 − tn )xn + tn T xn , where (tn ) is a real sequence.
x1 ∈ D and the real sequence (tn ) are said to satisfy condition (A)
∞
if 0 ≤ tn ≤ b < 1 for all n ∈ N , tn = +∞ and xn ∈ D for all n ∈ N .
n=1
8.7 Iterations of a Non-expansive Mapping 187
∞
Remark 8.7.2 If tn ∈ [a, b] for all n where 0 < a ≤ b < 1, then tn = +∞ and
n=1
0 ≤ tn ≤ b < 1.
Lemma 8.7.3 Let (sn ) be a real sequence and (u n ) a sequence in a Banach space
X . Then for any natural number M
M−1 M
si (1 − si )u i
i=1 i=1
⎧⎛ ⎞⎛ ⎞ ⎫
M ⎨
M−1
M−1
i ⎬
= 1− si u M − ⎝ s j ⎠ ⎝1 − s j ⎠ (u i+1 − si u i )
⎩ ⎭
i=1 i=1 j=i+1 j=1
(8.7.1)
When X is the real line and u i = 1 for all i, we have the special case
M−1 M
si (1 − si )
i=1 i=1
⎧⎛ ⎞⎛ ⎞ ⎫
M ⎨
M−1
M−1
i ⎬
=1− si − ⎝ s j ⎠ ⎝1 − s j ⎠ (1 − si )
⎩ ⎭
i=1 i=1 j=i+1 j=1
n
n
In this and what follows and are defined as 0 and 1 for n < m.
i=m i=m
Proof
∞
Since (xn ) is bounded and tn = +∞, we can find a natural number N such that
n=1
8.7 Iterations of a Non-expansive Mapping 189
N −1
N
r tm+i ≤ δ(M) + 1 ≤ r tm+i
i=1 i=1
and
N
xm+N +1 − xm+1 = {((1 − tm+i )xm+i + tm+i T xm+i ) − xm+i }
i=1
N
N
= tm+i (T xm+i − xm+i ) = (1 − si )u i
i=1 i=1
since si = 1 − tm+i ≥ 1 − b > 0, (8.7.1), the choice of m and Lemma 8.7.3 imply
190 8 Contractive and Non-expansive Mappings
N −1 −1
N
xm+N +1 − xm+1 ≥r (1 − si ) − r si
i=1 i=1
N −1 N
N
× 1− si − si (1 − si )
i=1 i=1 i=1
N −1 −1
N
≥r (1 − si ) − r si
i=1 i=1
N
N −1
=r tm+i − r (1 − tm+i )−1
i=1 i=1
N −1
≥ δ M + 1 − r (1 − tm+i )−1 (8.7.3)
i=1
N −1
N −1
(1 − tm+i )−1 = (1 + tm+i (1 − tm+i )−1 )
i=1 i=1
N −1
−1
= exp log(1 + tm+i (1 − tm+i ) )
i=1
N −1
−1
≤ exp tm+i (1 − tm+i )
i=1
N −1
−1
≤ exp (1 − b) tm+i
i=1
≤ exp{(1 − b−1 )(δ(M) + 1)r −1 } (8.7.4)
Proof Let D0 be the closure of the convex hull of T (D) ∪ {x1 }. By Mazur’s theorem
D0 is compact. Now M(x1 , tn , T ) lies in D0 . Condition (A) of Definition 8.7.1 along
with this implies that this sequence which indeed lies in D being in the compact set
D0 has a subsequence (xni ) converging to u ∈ D0 . Since (xni ) is in D and D is closed,
u ∈ D. From the boundedness of D0 and Lemma 8.7.4, lim xni − T xni = 0. As
i→∞
T is non-expansive
xn+1 − u = (1 − tn )xn + tn T xn − u
= (1 − tn )(xn − u) + tn (T xn − T u)
≤ xn − u for all n ∈ N.
Remark 8.7.9 Corollary 8.7.8 was proved for uniformly convex Banach spaces by
Krasnoselski.
If for some x1 ∈ D and (tn ) condition (A) is satisfied, then M(x1 , tn , T ) converges to
some fixed point of T in F.
xn+1 − u = (1 − tn )xn + tn T xn − u ≤ xn − u
xn − T xn ≥ f (d(xn , F)) ≥ f (r ).
Thus (u i ) is a Cauchy sequence in the closed set F of the Banach space X and hence
it converges to some u ∈ F. Given > 0 we can find i 0 > 0 such that 2−i0 < 2 and
u i0 − u < 2 . So for n > Ni0 ,
xn − u ≤ xn − u i0 + u i0 − u < + = .
2 2
Thus M(x1 , tn , T ) converges to a fixed point of T in D.
Condition (8.7.5) in the above theorem was originally considered by Senter and
Dotson.
8.8 A Generalization of the Contraction Principle Based on Combinatorics 193
Theorem 8.8.1 (Ramsey) Let S be an infinite set, n a natural number. Suppose that
every subset S of cardinality n is assigned one of a finite number of colours. Then
there exists an infinite subset T of S such that T is monochromatic (i.e. every subset
T of cardinality n has the same colour).
Lemma 8.8.2 Let m and n be natural numbers. Suppose G is a graph whose vertex
set is a disjoint union of countably many blocks, each of size m. Further suppose that
each edge has its two endpoints in distinct blocks and for any n blocks assume that
there is at least one edge having its endpoints in two of those blocks. Then there is
an infinite path in G, visiting no block more than once.
Proof Name the blocks as B0 , B1 , . . ., and number the vertices in each block Bi as
v(i, 1), v(i, 2), . . ., v(i, m). Colour the pairs of natural numbers with m 2 + 1 colours
by assigning as the colour of the pair (i, j) with i < j, either some pair ( p, q) so
that v(i, p) is adjacent to v( j, q) on G or if there is no edge between Bi and B j , the
special colour ‘none’ is assigned. By Ramsey’s Theorem 8.8.1 there is an infinite
set H of natural numbers, every pair of which has the same colour. By hypothesis,
this colour cannot be ‘none’. So let it be ( p, q). That is if i < j are both in H , then
v(i, p) is adjacent to v( j, q). Let h(i) be the ith element of H . The desired infinite
path is v(h(1), p), v(h(3), q), v(h(2), p), v(h(5), q), v(h(4), p), v(h(7), q), . . ..
Lemma 8.8.3 Let (X, d) be a metric space and 0 < α < 1, J a natural number
satisfying
min{d(T i x, T i y) : i = 1, 2, . . . , J } ≤ αd(x, y)
x) < 1−α
C
, we have for some k < J , d(T k x, T a(n)+k x) < 1−α C
. So by triangular
αC
inequality we have d(x, T a(n)+k
x) ≤ d(x, T x) + d(T x, T
k k a(n)+k
x) ≤ C + 1−α =
C
1−α
. Letting a(n + 1) = a(n) + k completes the proof.
194 8 Contractive and Non-expansive Mappings
Merrifield, Rothschild and Stein [15] proved the following generalization of the
contraction principle.
Proof Let x ∈ X and i, k = d(T i x, T k x), T i x = x and for a real number r , [r ]
denote the greatest integer less than or equal to r .
By Lemma 8.8.3, there is a sequence {n i : i = 1, 2, . . .} of natural numbers with
0, n i ≤ C and n i+1 − n i ≤ J . Applying the generalized contraction hypothesis to
each pair of points x and T ni x we get sequences {qi j : j = 1, 2, . . .} one sequence
) q *each i such that qi j , n i + qi j ≤ Cα and qi j+1 − qi ≤ J . If q = qi j , then j ≥
j
for
J
≥ q/J − 1, in which case q, n i + q ≤ Cα q/J −1 = C0 Q q where C0 = Cα and
1
Q = α J < 1. Note that if we have q, n i + q ≤ C0 Q q and q, n j + q ≤ C0 Q q for
two different integers i and j by the triangle inequality we have
n i + q, n j + q ≤ n i + q, q + q, n j + q
≤ 2C0 Q q
and N Q(Ak ) that can be expressed both in the form n i + q (an integer in N Q(A j ))
and n p + q (an integer N Q(Ak )).
Regard each integer as a vertex in a graph and partition the integers into a disjoint
union of blocks Bk = {(k − 1)J + 1, (k − 1)J + 2, . . . k J } for k = 1, 2, . . .. We say
that two vertices q and q in distinct blocks B j and Bk respectively are connected by
an edge if there are infinitely many integers that can be expressed in both the forms
n i + q and n p + q . From the above it is clear that for any collection of 2 J + 1
blocks, at least one edge has endpoints in two distinct blocks.
By Lemma 8.8.2 that there is an infinite path through the graph passing through
each block no more than once. Denote the vertices traversed in this path by
{r j : j = 1, 2, . . .}. Choose the sequences of integers {s j : j ∈ N}, and {t j : j ∈ N}
from {n j : j ∈ N} with the following three properties:
(i) if r j ∈ Bk , both r j + s j and r j + t j ∈ N Q(Bk );
(ii) r j + t j = r j+1 + s j+1 ;
(iii) r j + s j < r j+1 + s j+1
Consider the sequence of iterates with exponents r j + s j . Observe that
∞
∞
∞
r j + s j , r j+1 + s j+1 = r j + s j , r j + t j ≤ 2K 0 Q r j
j=1 j=1 j=1
Since the {r j : j ∈ N} are all distinct, the above series converges. So the sequence
of iterates is a Cauchy sequence. As X is complete, the resulting limit of this sequence
will be shown to be a fixed point of T . Let (T m i x) converge to z. We can even choose it
so that m i+1 > m i + J for all i. Since T is continuous, T m i +k → T k z for 1 ≤ k ≤ J .
Define L k = T k z for 0 ≤ k ≤ J . We claim that L j+1 = L j for some j < J . Since
T (L j ) = L j+1 it follows that L j is a fixed point of T .
By the generalized contraction hypothesis applied to x and y = T x, we con-
clude that for each i ∈ N, there exists an integer ji with 0 ≤ ji ≤ J − 1 and
d(T m j + ji x, T m i + ji +1 x) ≤ αri d(x, T x) where ri → ∞ (this can be seen by using
an argument in the second paragraph of the proof of this theorem). By the pigeon-
hole principle, there is an integer k with 0 ≤ k ≤ J − 1 and ji = k for infinitely
many i. For those i with ji = k for infinitely many i we have
As i → ∞, each of the three terms on the right-hand side of the above inequality
tends to zero. Thus L k = L k+1 or L k = T k z is a fixed point of T .
Remark 8.8.5 It is not known if the theorem is true even when T is not continuous
for all J .
196 8 Contractive and Non-expansive Mappings
References
1. Alspach, D.E.: A fixed point free nonexpansive map. Proc. Am. Math. Soc. 82, 423–424 (1981)
2. Browder, F.E.: Nonexpansive nonlinear operators in a Banach space. Proc. Natl. Acad. Sci.
U.S.A. 54, 1041–1044 (1965)
3. Dieudonne, J.: Foundations of Modern Analysis. Academic, New York (1962)
4. Dinca, G., Mawhin, J.: A constructive fixed point approach to the existence of a triangle with
prescribed angle bisector lengths. Bull. Belg. Math. Soc. Simon Stevin 17, 333–341 (2010)
5. Dotson Jr., W.J.: Fixed point theorems for nonexpansive mappings on star-shaped subsets of
Banach spaces. J. Lond. Math. Soc. 2(4), 408–410 (1972)
6. Edelstein, M.: On fixed and periodic points under contractive mappings. J. Lond. Math. Soc.
37, 74–79 (1962)
7. Goebel, K.: An elementary proof of the fixed point theorem of Browder and Kirk. Mich. Math.
J. 16, 381–383 (1969)
8. Gohde, D.: Uber Fixpunkte bei stetigen Selbst-abbildungen mit kompakten Iterierten. Math.
Nachr. 28, 45–55 (1964)
9. Hardy, G.H.: Divergent Series (Oxford, 1949)
10. Ishikawa, S.: Fixed points and iteration of a nonexpansive mapping in a Banach space. Proc.
Am. Math. Soc. 59, 65–71 (1976)
11. Kirk, W.A.: A fixed point theorem for mappings which do not increase distances. Am. Math.
Mon. 72, 1004–1006 (1965)
12. Mann, W.R.: Mean value methods in iteration. Proc. Am. Math. Soc. 4, 506–510 (1953)
13. Matkowski, J.: Fixed point theorem in a uniformly convex paranormed space and its application.
Topol. Appl. 160, 524–531 (2013)
14. Meinardus, G.: Approximation of Functions: Theory and Numerical Methods. Springer, Berlin
(1964)
15. Merrifield, J., Rothschild, B., Stein Jr., J.D.: An application of Ramsey’s theorem to the Banach
contraction principle. Proc. Am. Math. Soc. 130, 927–933 (2002)
16. Mironescu, P., Panaitopol, L.: The existence of a triangle with prescribed angle bisector lengths.
Am. Math. Mon. 101, 58–60 (1994)
17. Pasicki, L.: A basic fixed point theorem. Bull. Pol. Acad. Sci. Math. 54, 85–88 (2006)
18. Pasicki, L.: Bead spaces and fixed point theorems. Topol. Appl. 156, 1811–1816 (2009)
19. Pasicki, L.: Uniformly convex spaces, bead spaces and equivalence conditions. Czechoslov.
Math. J. 61(2), 383–388 (2011)
20. Pasicki, L.: A comment to Matkowski’s paper. Topol. Appl. 160, 951–952 (2013)
21. Reich, S.: Remarks on fixed points. Att. Acad. Naq. Lincei Rand. Sci. Fis. Mat. Matur. Serie
VIII, LII fasc 5, 689–697 (1972)
22. Rosenholtz, I.: On a fixed point problem of D.R. Smart. Proc. Am. Math. Soc. 55, 252 (1976)
23. Sims, B.: Examples of fixed point free mappings. In: Kirk, W.A., Sims, B. (eds.) Handbook of
Metric Fixed Point Theory. Kluwer Academic Publishers, Dordrecht (2001)
24. Van der Berg, Over dee bepaling van een drichoek, waarvan de deellijmen derdrie supplemen-
taire heoken gegevan zijn Niew Archiev voor Wiskunde, vol. 17, pp. 191–205 (1890)
Chapter 9
Geometric Aspects of Banach Spaces
and Non-expansive Mappings
9.1 Introduction
In this chapter, we outline the proof that a reflexive non-square Banach space has
fixed point property for non-expansive mappings on bounded closed convex sets. To
this end, some definitions are in order.
Definition 9.1.1 For a Banach space (X, · ), the closed unit ball and the unit
sphere in X are denoted by B X and S X , respectively. The Clarkson modulus of
convexity of X is a function δ X : [0, 2] → [0, 1] defined by
x + y
δ X () = inf 1 − : x, y ∈ S X , x − y ≥ .
2
ρ(t)
X is called uniformly smooth if ρ0 (X ) = lim+ = 0.
t→0 t
Remark 9.1.4 ρx is increasing, continuous and convex on R+ and ρ X (0) = 0
with ρ X (t) ≤ t. Also for a real Banach space√X with dim X ≥ 2 (or a complex
Banach space X with dim X ≥ 1), ρ X (t) ≥ 1 + t 2 − 1 = ρ2 (t). Lindenstrauss
[13] proved the important relations
t
ρ (t) = sup
X∗ − δ X (t) : ∈ [0, 2]
2
t
ρ X (t) = sup − δ X (t) : ∈ [0, 2] .
∗
2
So ρ X (t) = ρ X ∗∗ (t) and X is uniformly convex (uniformly smooth) if and only if its
dual X ∗ is uniformly smooth (uniformly convex). Also X is reflexive whenever it is
uniformly convex or uniformly smooth.
Proposition 9.1.6 (Kato [10]) Let X be a real Banach space with dim(X ) ≥ 2 (or
a complex Banach space wth dim(X ) ≥ 1). Then the following are equivalent.
(i) X is uniformly non-square;
(ii) δ X () > 0 for some ∈ (0, 2);
(iii) 0 (X ) < 2;
(iv) J (X ) < 2;
(v) ρ X (t0 ) < t0 for some t0 > 0;
(vi) ρ X (t) < t for all t > 0;
ρ X (t)
(vii) ρ X (0) = lim < 1;
∗
t→0 t
(viii) 0 (X ) < 2;
ρ X ∗ (t)
(ix) ρ X ∗ (0) = lim < 1.
t→0 t
9.1 Introduction 199
x − y x + y
> 1 − δ implies >1−δ
2 2
⇔ for some δ ∈ (0, 1), x, y ∈ S X and
x − y x + y
> 1 − δ implies 1 − ≥δ
2 2
⇔ for some δ ∈ (0, 1), δ X (2 − 2δ) ≥ δ
(i) ⇒ (ii) If X is uniformly non-square set = 2 − 2δ ∈ (0, 2) in the above equiva-
lences to get δ X () > 1 − 2 (= δ).
(ii) ⇒ (i) If for some 0 ∈ (0, 2), δ X (0 ) ≥ η0 > 0, where η0 ∈ (0, 1), then for 2 −
2δ = ∈ [0 , 2), δ ∈ 0, 1 − 20 and δ X (2 − 2δ) = δ X () = δ X (0 ) ≥ η0 > 0. This
means that for any x, y ∈ S X , with x − y ≥ 2 − 2δ necessarily 1 − x+y 2
≥ η0 .
If δ = min{δ, η0 ) then δ ∈ (0, 1). If x−y
2
≤ 1 − δ
, we are done. If x+y
2
> 1 − δ,
x+y x+y
then 1 − 2 ≥ η0 or 2 ≤ 1 − η0 ≤ 1 − δ . So X is uniformly non-square. (ii)
⇔ (iii) and (i) ⇔ (iv) follow from the definition. (v) ⇔ (vii) follows from the fact
that ρ Xt(t) is increasing. (vii) ⇔ (viii) and (ix) ⇔ (iii) follow from
ρx (t)
0 (X ∗ ) = 2ρ X (0) = 2 lim and
t→0 t
ρx ∗ (t)
0 (X ) = 2ρ X ∗ (0) = 2 lim .
t→0 t
We now show that (v) ⇔ (vi). If ρ X (t0 ) = t0 for some t0 > 0 then ρ X (t) = t for all
t > 0. Since ρ Xt(t) is increasing and ρ X (t) ≤ t, it follows that for t ≥ t0 , 1 = ρ Xt(t0 0 ) ≤
ρ X (t)
t
≤ 1 or ρ X (t) = t for t ≥ t0 . Let 0 < t < t0 and ρ X (t) < t. Since ρ X is convex
for t1 > t0
t0 − t t1 − t0
t0 = ρ X (t0 ) = ρ X t1 + t
t1 − t t1 − t
t0 − t t 1 − t0
≤ ρ X (t1 ) + ρ X (t)
t1 − t t1 − t
t0 − t t1 − t0
< t1 + t = t0 ,
t1 − t t1 − t
a contradiction. So ρ X (t) = t.
Remark 9.1.7 A uniformly non-square Banach space is super-reflexive (James [8])
if it has an equivalent norm ||| · · · ||| in which it is uniformly convex. The converse
is not true in the sense that δ(X,| |) () = 0 for 0 < < 2.
200 9 Geometric Aspects of Banach Spaces and Non-expansive Mappings
vex ( p , ||| · |||) and ( p , ||| · ||| ) are super-reflexive but not uniformly non- square
sine δ X () = 0 for all 0 < < 2. To see this take x = (1, 0, 0, . . .) and y =
(0, 1, 0, . . .).
From the above example, one can see that for any real Banach space X with
dim(X ) ≥ 2, there is an equivalent norm in which X is not uniformly non-square.
Garcia-Falset et al. [5] proved that in uniformly non- square Banach spaces, every
bounded closed convex subset has the fixed point property for non-expansive map-
pings. This is achieved by studying the properties of certain coefficients associated
with the geometry of Banach spaces.
Definition 9.2.2 (Dominguez Benavides [3]) For a Banach space X and a ≥ 0, the
parameter
Kirk [11] proved that every closed bounded subset having normal structure in
a reflexive Banach space has the fixed point property for non-expansive self-maps.
This can be deduced from the following lemma due independently to Goebel [6] and
Karlovitz [9], as shown in the proof of Theorem 9.2.5.
Lemma 9.2.3 (Goebel [6], Karlovitz [9]) Lex X be a Banach space, C0 , a weakly
compact convex subset of X and T : C0 → C0 a non-expansive map. Let C0 be a
minimal closed convex set that is invariant under T . (i.e. no proper closed convex
subset of C0 is invariant under T ). For each sequence of xn in C0 with lim ||xn −
n→∞
T xn || = 0 and for each x ∈ C0 lim ||xn − T xn || = diamC0 (such a sequence xn
n→∞
with lim ||xn − T xn || = 0 is called a sequence of approximate fixed points).
n→∞
Proof For y ∈ C0 , let s = lim sup ||y − xn ||. Let D = {x ∈ C0 : lim sup ||x − xn || ≤
n→∞ n→∞
s}. Clearly D is non-void closed and convex. D is invariant under T , since
m
of T (C0 ) = C0 . For u ∈ C0 , given > 0, we can find v = λk T xk with xk ∈ C0 ,
k=1
λk > 0 and λk = 1 with ||u − v|| ≤ . For w ∈ F, ||T w − u|| ≤ ||T w − v|| +
||u − v|| ≤ λk ||T w − T xk || + ||v − u|| ≤ λk ||w − xk || + ||v − u|| ≤ s + .
So T w ∈ F. Thus F is T -invariant. Hence the lemma.
At this stage we recall the definition of normal structure.
Definition 9.2.4 A convex subset K of of a Banach space X is said to have normal
structure if for each bounded convex subset K 1 of K containing more than one point,
there is a non-diametral point in the sense that for some x0 ∈ K 1 , sup{||x0 − k|| :
k ∈ K 1 } < diam K 1 .
With this we can prove Kirk’s theorem [11].
Theorem 9.2.5 Let K be a non-empty closed convex bounded subset of a reflexive
Banach space X such that K has normal structure. If T : K → K is non-expansive
then T has a fixed point in K .
Proof Since K is a bounded closed convex subset of the reflexive Banach space
X , K is weakly compact. By a standard application of Zorn’s lemma, K has a
minimal non-empty
closed convex subset C0 invariant under T . Let a ∈ C0 . Then
Tn (x) = an + 1 − n1 T xn , n ≥ 2 is a contraction mapping C0 into itself and hence
has a unique fixed point xn . Thus xn = an + 1 − n1 T xn for each n ≥ 2. Further
lim ||xn − T xn || → 0 as n → ∞. So by Goebel-Karlovitz Lemma 9.2.3. lim ||x −
n→∞
xn || = diam C0 for all x ∈ C0 . If C0 contains more than a singleton, it contains a
non-diametral point a such that sup ||a − x|| < diam C0 as C0 ⊆ K is a closed
convex subset of K and K has normal structure. This implies that lim ||a − xn || ≤
n→∞
sup ||a − xn || < diam C0 , contradicting Lemma 9.2.3. So C0 is a singleton which
n
necessarily is a fixed point of T as T C0 ⊆ C0 .
Remark 9.2.6 Lin [12] has generalized Goebel-Karlovitz lemma using certain non-
standard analytic considerations. Let X be a Banach space with norm || · ||, ∞ (X ),
the space of sequences (xn ) in X with norm sup{||xn || : n ∈ N} and C0 (X ) the
subspace of ∞ (X ) with null sequence (xn ) of X and [X ] the quotient space
∞ (X )/C0 (X ) with the norm ||[z n ]|| = lim sup ||z n || where [z n ] is the equivalent
class of {z n } ∈ ∞ . x ∈ X is identified with the class [x, x, . . .] and consequently X
can be considered a subset of [X ]. For a subset K of X the set [K ] = {[z n ] ∈ [X ] :
z n ∈ K for every n ∈ N}. If T : K → K is a map then [T ] : [K ] → [K ] is a map
defined in a natural way by [T ]([xn ]) = [T xn ].
Lemma 9.2.7 (Lin [12]) Let X be a Banach space and K a minimal weakly compact
convex subset of X , invariant under T . If [W ] is a non-empty closed convex subset
of [K ] which is invariant under [T ] then
for each x ∈ K .
We leave the details of the proof to the reader. The following theorem due to
Dominguez Benavides makes use of Lin’s lemma.
Theorem 9.2.8 (Dominguez Benavides [3]) If X is a Banach space for which
R(a, X ) < 1 + a for some a ≥ 0, then every non-empty bounded convex weakly
compact subset of X has fixed point property for non-expansive mappings.
Proof Suppose the theorem is false. Then, there is a non-empty weakly compact
convex subset K of X with diam(K ) = 1 and K is minimal invariant for a non-
expansive map T without a fixed point. Further there is a weakly null sequence of
approximate fixed points {xn } of T in X . Define
where t = 1+a 1
.
Clearly [W ] is a a closed convex [T ] invariant set. [W ] is non-empty as it contains
[t xn ]. So by Lin’s Lemma 9.2.7 it follows that
)
We can choose η > 0 such that η R(a, X ) < 1 − R(a,X
1+a
(i.e. R(a, X ) η + 1+a 1
<
n, we have ||yn − y|| ≤ t+ η. Also ||y|| ≤ lim||yn − xn || ≤ 1 −
1. For a large
yn yn −y y
t. So t+η = t+η + t+η ≤ R 1−tt
, X = R(a, X ). So lim||z n || = lim ||yn || ≤
R(a, X )(t + η) < 1, a contradiction to Lemma 9.2.7.
At this stage, we can introduce the concept of Banach–Mazur distance between
isomorphic Banach spaces, a useful concept in fixed point theory of non-expansive
maps in Banach spaces.
Definition 9.2.9 For isomorphic normed spaces X and Y , the Banach–Mazur dis-
tance between X and Y denoted by d(X, Y ) is defined as
Theorem 9.2.11 (Dominguez Benavides [3]) For isomorphic Banach spaces X and
Y
R(a, Y ) ≤ d(X, Y )R(a, X )
where a ≥ 0.
For a Banach space X , the coefficient M(X ) can be defined as in the following
Definition 9.2.12 If X is a Banach space, M(X ) is defined as sup R(a,X
1+a
)
:a≥0
Definition 9.3.1 Let X be a Banach space with a Schauder basis (en ) (that is each
∞
x ∈ X has a unique representation x = xn en , xn being scalars). X is called nearly
n=1
uniformly smooth (NUS) if for each > 0 there is δ > 0 such that if 0 < t < δ and
(xn ) is a basic sequence in B X there exists k > 1 so that ||x1 + t xk || < 1 + t.
X is called weakly near uniformly smooth (WNUS) if the above definition holds
for some > 0.
Garcia Falset [4] proved the following
9.3 Nearly Uniformly Smooth Spaces 205
Proof (b) ⇒ (c). Let (xn ) be a weakly null sequence in B X and x ∈ B X . Define (yn )
by y1 = x, yn+1 = xn for n ≥ 1. Then (yn ) is a weakly null sequence in B X . By (b),
there exists c ∈ (0, 1) and k1 > 1 with ||x + xk1 || ≤ 2 − c. Define another weakly
null sequence z n ∈ B defined by z 1 = x and z n = xk1 +n , n ∈ N. By (b) there exists
k2 > k1 such that ||x + xk2 || ≤ 2 − c. Thus, proceeding recursively, we can get a
subsequence xkn of weakly null sequence in B X such that ||x + xkn || ≤ 2 − c for all
kn and kn > 1 for all n. So lim||x + xn || ≤ 2 − c, for any weakly null sequence (xn )
in B X . So R(X ) < 2.
(c) ⇒ (b). Let (xn ) be a weakly null sequence in B X . As R(X ) < 2, R(X ) <
2 − c for some c ∈ (0, 1). So lim||x1 + xn || ≤ R(X ) < 2 − c. So for some k > 1,
||x1 + xk || ≤ 2 − c.
(a) ⇒ (b). By assumption, there exist , δ > 0 such that for all t ∈ (0, δ) and
any weakly null sequence (xn ) in B X there is k > 1 with ||x1 + xk || < 1 + t. Let
μ = min{1, δ}. Then for t < δ, ||x1 + xk || ≤ ||x1 + t xk || + (1 − t)||xk || ≤ 1 + t +
1 − t = 2 − t (1 − ) = 2 − c, where c = t (1 + ) ∈ (0, 1).
(b) ⇒ (a). By hypothesis, there exists c ∈ (0, 1) such that for each weakly null
sequence (xn ) in B X , there is k > 1 with ||x1 + xk || ≤ 2 − c. So for all t ∈ (0, 1),
Corollary 9.3.3 ([4]) For a Banach space X , the following are equivalent:
(a) X is WNUS;
(b) X is reflexive and R(X ) < 2.
Theorem 9.3.4 (Garcia Falset [4]) Let X be a Banach space such that R(X ) < 2.
Then, every non-empty weakly compact convex subset of X has fixed point property
for non-expansive mappings.
Proof We prove by the method of contradiction. Suppose the theorem is false. Then,
there is a weakly compact convex subset K of X with diam K = 1 which is minimal
for a non-expansive map T : K → K in the sense of Goebel-Karlovitz Lemma 9.2.3.
Let (xn ) be a weakly null sequence of approximate fixed points of T in K .
With the usual notation in Remark 9.2.6, define the subset [W ] of [X ] by [W ] :=
{[z n ] ∈ [K ] : ||[z n ] − [xn ]|| ≤ 21 , D([z n ]) ≤ 21 }. [W ] is seen to be a T -invariant
closed convex subset [X ]. By Lemma 9.2.3 x2n ∈ [W ]. So by Lin’s Lemma 9.2.7
206 9 Geometric Aspects of Banach Spaces and Non-expansive Mappings
z n k −y
Let ym = 1 − 1
m
. As (ym ) is a weakly null sequence in B X , it
max ||y||, lim ||z n m −y||
m→∞
follows from the definition of R(X ) that
y
R(X ) ≥ lim ym +
m→∞
max lim ||z n m − y||, ||y||
m→∞
So lim ||z n m || ≤ R(X ) max ||y||, lim ||z n m − y|| . Since (z n m − xn m ) converges
m→∞ m→∞
1
weakly to y, ||y|| ≤ lim ||z n m − xn m || ≤ lim ||z n − xn || = ||[z n ] − [xn ]|| ≤
. On
m→∞ m→∞ 2
the other hand, the weak limit of {z n m − y − (z n s − y)} = z n m − y as s → ∞.
So ||z n m − y|| ≤ lim ||z n m − y − (z n s − y)|| or lim ||z n m − y|| ≤ lim lim ||z n m
m→∞ m→∞ m→∞ s→∞
−y − (z n s − y)|| . As D(z n m − y) = D(z n m ) ≤ D(z n ) and D([z n ]) ≤ 21 , we have
1 R(X )
lim ||z n − y|| ≤ D([z n ]) ≤ . So ||[z n ] − 0|| = lim ||z n m || ≤ < 1, as
m→∞ m 2 m→∞ 2
1
max lim ||z n m − y||, ||y|| ≤ . This contradicts Lemma 9.2.7.
m→∞ 2
The following are proved using similar arguments.
Theorem 9.3.5 (Garcia Falset [4]) Let X and Y be isomorphic Banach spaces such
that d(X, Y )R(Y ) < 2. Then, every non-empty convex weakly compact subset of X
has the fixed point property for non-expansive maps, provided it satisfies weak Opial’s
condition. That is, lim inf ||xn || ≤ lim inf ||xn + x|| for each x ∈ X and each weakly
null sequence (xn ).
Proof Suppose false. Then, there is a convex weakly compact subset K of X
with diam(K ) = 1, which is minimal for a non-expansive map T : K → K by
Goebel-Karlovitz lemma. Thus, there is a weakly null sequence of almost fixed
points (xn ) for T in K . As in Remark 9.2.6 consider the subset [W ] of [X ] defined
by
1 1
[W ] = {[z n ] ∈ [K ] : ||[z n ] − [xn ]|| ≤ and for some x ∈ K , ||[z n ] − x|| ≤ }
2 2
9.3 Nearly Uniformly Smooth Spaces 207
−1
So for all k ≥ 1, 1 − k max lim||T (z n k ) − T (y)||, ||T y||
1
(T z n k − T y) and
k
−1
max lim||T (z n k ) − T (y)||, ||T y|| (T y) ∈ B y . So, from the definition of R(Y )
k
we get lim||T z n k || = R(Y ) max lim||T (z n k ) − T (y)||, ||T y|| .
k k
On the other hand since [z n ] ∈ [W ], there exists x0 ∈ K such that ||[z n ] − x0 || ≤ 21
and ||[z n ] − [xn ]|| ≤ 21 . So ||y|| ≤ lim||z n k − xn k || ≤ ||[z n ] − [xn ]|| ≤ 21 . As Y sat-
k
isfies weak Opial’s condition lim||T z n k − T y|| ≤ lim||T z n k − T x0 ||.
k k
So
lim||T z n k || ≤ R(Y )||T || max lim||z n k − x0 ||, ||y||
k k
||T ||
≤ R(Y ) .
2
The above conditions imply that
Definition 9.4.1 ([2]) The modulus of nearly uniform smoothness of a Banach space
X is the function X : [0, ∞) → R defined by
||x1 + t xn || + ||x1 − t xn ||
X (t) = sup inf − 1 : n > 1 : (xn )
2
is a basic sequence in B X . i.e., (xn ) is a Schauder basis for X.
Remark 9.4.2 ρ X (t) = sup ||x+t y||+||x−t
2
y||
: x, y ∈ B X is called the modulus of
uniform smoothness. Clearly ρ X (t) ≥ X (t) ≥ 0 for all t ≥ 0. When X is uniformly
ρ X (t) X (t)
smooth, ρ (0) = lim = 0 by definition. In this case lim = X (0) = 0.
t→0 t t→0 t
There is an equivalent characterization of near uniform smoothness for reflexive
Banach spaces, whose proof is available in [2].
As a result we have
Proposition 9.4.4 ([2]) A Banach space X is nearly uniformly smooth if and only
X (t)
if X is reflexive and lim = 0.
t→0 t
X (t)
Proof If lim = 0, then for > 0 there exists δ > 0 such that X (t) ≤ t for
t→0 t
t ∈ [0, δ). Since X is reflexive, we can find a basic sequence {xn } in B X as suggested
by Prus in §2 of [14] which is not norm convergent for which
9.4 Non-square Banach Spaces 209
1
||x1 + t xn k || ≤ {(1 + c)||x1 − 2t xn k || + ||x1 + 2t xn k ||}
2
(1+3t)
where c > 1, 1 + c < (1+2t) and (xn k ) is a subsequence of {xn }. Then for some k
||x1 + t xn k || ≤ (1 + )(1 + 2t) < 1 + 3t.
Conversely if X is nearly uniformly smooth then X is reflexive. Consider a weakly
null sequence (xn ). For each > 0, there is a δ > 0 with ||x1 + t z n || ≤ 1 + t for all
n > 1, where z n is a subsequence of (xn ) with z 1 = x1 . Since {x1 , −z 2 , −z 3 , . . . , }
is also weakly null ||x1 − t z n || ≤ 1 + t for some 0 < δ < δ for all t ∈ [0, δ). So
for 0 < t < δ
||x1 + t z n || + ||x1 − t z n ||
− 1 ≤ t
2
X (t)
So lim = 0.
t→0 t
For the proof that a non-square Banach space has the fixed point property for non-
expansive mappings on non-empty bounded closed convex subsets, Garcia-Falset et
al. [5] introduced the coefficient RW (X ) and M W (X ) relating them to the coeffi-
cients R(X ) and W (X ).
Definition 9.4.5 ([5]) Let X be a Banach space and a, a positive real number. Then
w
RW (a, X ) := sup min lim ||xn + x||, lim ||xn − x|| : xn ∈ B X , xn → 0, ||x|| ≤ a
n→∞ n→∞
1+a
M W (X ) := sup :a>0 .
RW (a, X )
Remark 9.4.6 For any Banach space X and a > 0, max{a, 1} ≤ RW (a, X ) ≤ 1 +
a. So 1 ≤ M W (X ) ≤ 2.
Lemma 9.4.7 ([5]) Let X be a Banach space. Given x ∈ X and a bounded sequence
(xn ) in X , there is a subsequence (xn k ) of (xn ) such that
and
lim ||(xn k − xn k+1 )|| ≤ D[(xn )]
k→∞
Proof Let
1
lim ||(xn 1 − xm ) + x|| > a −
m→∞ 2
1
lim ||(xn 1 − xm ) − x|| > b −
m→∞ 2
and
1
lim ||(xn 1 − xm )|| < D[(xn )] + .
m→∞ 2
Suppose n 1 < n 2 < · · · < n j have been defined such that for each k = 1, 2, . . . , j,
1
lim ||(xn k − xm ) + x|| > a −
m→∞ k+1
1
lim ||(xn k − xm ) − x|| > b −
m→∞ k+1
1
lim ||xn k − xm || < D[(xn )] +
m→∞ k+1
1
||(xn k − xn k+1 ) + x|| > a −
k+1
1
||(xn k − xn k+1 ) − x|| > b −
k+1
1
||xn k − xn k+1 || < D[(xn )] + .
k+1
From the above inequalities for k = j and the definition of a, b and D[(xn )], we can
find n j+1 > n j such that
1
||(xn j − xn j+1 ) + x|| > a −
j +1
1
||(xn j − xn j+1 ) − x|| > b −
j +1
1
||xn j − xn j+1 || < D[(xn )] +
j +1
9.4 Non-square Banach Spaces 211
1
lim ||(xn j+1 − xm ) + xm || > a −
m→∞ j +2
1
lim ||(xn j+1 − xm ) − xm || > b −
m→∞ j +2
1
lim ||xn j+1 − xm || < D[(xn )] + .
m→∞ j +2
Proof Let a, η > 0. From the definition of R(a, X ), we can find x ∈ X with ||x|| ≤ a
and a weakly null sequence (xn ) in B X with D[(xn )] ≤ 1 such that
w∗
Again for n ≥ 1, as f m → f , lim f m (xn ) = f (xn ), we have
m→∞
212 9 Geometric Aspects of Banach Spaces and Non-expansive Mappings
So
lim lim ||(xn − xm ) − x|| ≥ R(a, X ) − η,
n→∞ m→∞
W
as xn → 0.
Thus we have shown that
min{ lim lim ||(xn − xm ) + x||, lim lim ||(xn − xm ) − x|| ≥ R(a, X ) − η
n→∞ m→∞ n→∞ m→∞
min{ lim ||(xn k − xn k+1 ) + x||, lim ||(xn k − xn k+1 ) − x|| ≥ R(a, X ) − η
k→∞ k→∞
and lim ||xn k − xn k+1 || ≤ 1. So we can find k0 such that for all k ≥ k0 , ||xn k −
k→∞
xn k+1 || ≤ 1 + η. Define
xn k0 +k − xn k0 +k+1
yk := , k≥1
1+η
x
y :=
1+η
M(x) > 1 by Theorems 9.2.13, 9.4.8 implies that on non-void weak compact con-
vex subsets, non-expansive self-maps have fixed points.
We begin with
Proposition 9.5.1 ([5]) Let X be a Banach space and a > 0. Then
RW (a, X ) = sup{ inf (||ax1 + xn || ∧ ||ax1 − xn ||) : (xn ) a weakly null sequence in B X }
n>1
RW (a, X ) = sup{ inf (||ax1 + xn || ∧ ||ax1 − xn ||) : (xn ) a weakly null sequence in B X }.
n>1
Let η > 0, (xn ) a weakly null sequence in B X and x ∈ X with ||x|| ≤ a. Define (yn )
by
x/a, n = 1
yn =
xn , n ≥ 2
inf (||ay1 + yn || ∧ ||ay1 − yn ||) ≤ RW (a, X ).
n>1
||x + xn 1 || ∧ ||x − xn 1 || = ||ay1 + yn || ∧ ||ay1 − yn || < RW (a, X ) + η.
Suppose n 1 < n 2 . . . < n k have been found such that for j ∈ {1, 2, . . . , k}
||x + xn j || + ||x − xn j || < RW (a, X ) + η
inf (||az 1 + z n || ∧ ||az 1 − z n ||) ≤ RW (a, X )
n>1
||az 1 + z k || ∧ ||az 1 − z k || < RW (a, X ) + η.
||x + xn k+1 || ∧ ||x − xn k+1 || = ||az 1 + z k || ∧ ||az 1 − z k || < RW (a, X ) + η.
||x + xn k+1 || ∧ ||x − xn k+1 || < RW (a, X ) + η
We can get a subsequence (wm ) of (xn k ) for which lim ||x + wm ||, lim ||x − wm ||
m→∞ m→∞
exist.
Now
≤ RW (a, X ) + η
RW (a, X ) ≤ RW (a, X ).
So by definition of RW (a, X ),
Proof (i) Let t, η > 0 and (xn ) be a weakly null sequence in B X . By Proposition 9.5.1
1 1
inf (||x1 + t xn || ∧ ||x1 − t xn ||) = t inf {|| x1 + xn || ∧ || x1 − xn ||}
n>1 n>1 t t
1
≤ t RW ( , X ).
t
So for some k > 1
1
||x1 + t xk || ∧ ||x1 − t xk || < t RW ( , X ) + η
t
As ||x1 + t xk || ∨ ||x1 − t xk || ≤ 1 + t, we have
So
||x1 + t xn || + ||x1 − t xn || ||x1 + t xk || + ||x1 − t xk || − 1
inf −1 ≤
n>1 2 2
t (RW ( 1t , X ) + 1) − 1 + η
< .
2
t (RW ( 1 ,X )+1)−1+η t (RW ( 1 ,X )+1)−1
So X (t) < t
2
. As η → 0, X (t) ≤ t
2
.
(ii) Let a, η > 0 and (xn ) be a weakly null sequence in B X . So
||x1 + xn
|| + ||x1 − xn
|| 1
inf a a
− 1 ≤ X ( )
n>1 2 a
1
RW (a, X ) ≤ a(1 + X ( )).
a
Theorem 9.5.3 ([5]) If X is a reflexive Banach space for which X (0) < 1, then
M(X ) > 1. So every non-void bounded closed convex subset of X has fixed point
property for non-expansive maps.
Proof Clearly X (0) < 1 if and only if for some t > 0 sup{ Xs(s) : 0 < s ≤ t} <
1. So for some t > 0 X (t) < t. So by (ii) M W (X ) ≥ sup{ 1+ 1+t
X (t)
: t > 0}. Or
M W (X ) > 1. So by Corollary 9.4.9, and the reflexivity of X every non-void closed
bounded convex subset of X has fixed point property for non-expansive maps.
Corollary 9.5.4 ([5]) If X is a uniformly non-square Banach space, then M(X ) > 1
and so every non-void closed bounded convex subset of X has fixed point property
for non-expansive maps.
Proof Since X is uniformly non-square, so is X ∗ . By Proposition 9.1.6, 0 (X ∗ ) < 2.
So by Lindenstrauss formulae the modulus of smoothness ρ X satisfies ρ X (0) < 1
(again by Proposition 9.1.6). Since X (t) ≤ ρ X (t) for all t > 0, X (0) ≤ ρ X (0) < 1.
As X is reflexive (being non-square) by Theorem 9.5.3 every non-void bounded
closed convex subset of X has fixed point property for non-expansive mappings.
Remark 9.5.5 Garcia Falset et al. [5] have shown that in a reflexive Banach space
X , the following are equivalent:
(i) M W (X ) > 1;
(ii) for some a > 0, RW (a, X ) < 1 + a;
(iii) inf{ RW1+a
(a,X )
: a > 0} > 1;
(iv) X (0) < 1;
(v) for some t > 0, sup{ Xs(s) : 0 ≤ s ≤ t} < 1;
(vi) for some t > 0, X (t) < t.
Remark 9.5.6 If M W (X ) > 1, X may not be non-square although every non-void
closed bounded convex subset of X has fixed point property for non-expansive maps.
This is seen by the following example due to Garcia Falset et al. [5].
9.5 An Equivalent Definition of RW (a, X ) and Fixed … 217
Example 9.5.7 Let X = (R2 , || ||∞ ) and Y = (2 , || · ||2 ) and Z the product space
Z = X × Y with the norm ||(x; y)|| = max{||x||∞ , ||y||2 }, where x ∈ X = R2 and
y ∈ Y = 2 . For z 1 = ((1, 1); 0) and z 2 = (1, −1); 0), ||z 1 || = ||z 2 || = 1, ||z 1 +
z 2 || = ||z 1 − z 2 || = 2. Thus Z is not uniformly non-square. Let z n = (xn ; yn ) be a
W
weakly null sequence in Bz and z ∈ B Z , with z = (x; y). If z n → 0, then xn → 0 and
yn → 0 and lim ||xn + x||∞ = ||x||∞ ≤ ||z|| ≤ 1 and lim ||yn + y||22 =
n→∞
lim ||yn ||22
+ ||y|| ≤ lim ||z n || + ||z|| ≤ 2. Therefore, lim ||z n + z|| = lim
2 2 2
n→∞ n→∞ n→∞ n→∞
√ √
max{||xn + x||, ||yn + y||} ≤ 2. So RW (1, X ) ≤ 2 and M W (X ) > 1. So in X
every non-void closed convex bounded subset has fixed point property for non-
expansive maps.
References
10.1 Introduction
It is more than a century since Brouwer [4] proved a fixed- point theorem of great
consequence, in the setting of finite-dimensional Euclidean spaces. It was subse-
quently extended to normed linear spaces by Schauder [25], and later to locally
convex linear topological spaces by Tychonoff [31]. Brouwer’s theorem was gen-
eralized to multifunctions first by Kakutani [12], and later to locally convex linear
topological spaces by Glicksberg [8] and Ky Fan [6]. Brouwer’s theorem admits
of several proofs. Notable among them are those based on Sperner’s lemma [28]
or concepts of homotopy/homology from algebraic topology (see Dugundji [5] or
Munkres [17]) or concepts and results from Real analysis (see Milnor [16], Seki [26],
Rogers [23], Kannai [13], Traynor [30]). However, we provide here only the analytic
proof of Brouwer’s theorem and a proof based on Sperner’s lemma. Needless to state
that Brouwer’s theorem and its generalizations/variants find a wide range of appli-
cations in the solution of nonlinear equations, differential and integral equations,
mathematical biology and mathematical economics.
We collect in this section the basic theorems of analysis needed in the proof of
Brouwer’s theorem.
Theorem 10.2.1 (Weierstrass Approximation Theorem) If f is a continuous real-
valued function defined on a closed bounded subset S of Rn , then for any given
positive number , we can find a polynomial P of n variables x1 , . . . , xn such that
| f (x1 , . . . , xn ) − P (x1 , . . . , xn )| < for all (x1 , x2 , . . . , xn ) ∈ S.
Theorem 10.2.2 (Inverse Function Theorem) Let G be a non-empty open set in Rn
and f = ( f 1 , . . . , f n ) : G ⊆ Rn → Rn be a continuous mapping having continuous
© Springer Nature Singapore Pte Ltd. 2018 219
P. V. Subrahmanyam, Elementary Fixed Point Theorems,
Forum for Interdisciplinary Mathematics,
https://doi.org/10.1007/978-981-13-3158-9_10
220 10 Brouwer’s Fixed-Point Theorem
For these and other aspects of calculus in finite-dimensional spaces, Apostol [1]
may be consulted.
Theorem 10.3.1 (Brouwer) Every continuous function f mapping the closed unit
sphere Bn (of Rn ) into itself has a fixed point.
We deduce it from the no-retraction theorem via a lemma, following Rogers [23].
First, we recall the following.
∂gi
B n for all i, j = 1, 2, . . . , n, in view of the continuity of ∂x j
on B n . So by Theorem
10.2.4 for some k > 1,
C(Bn ) = C( f t (Bn ))
= |Det f t (x)|d x by Theorem 10.2.3
B
n
1
= Det ( f t (x))d x for t < ,
Bn k
which is a polynomial in t. However, left-hand side of the above equality has the
constant value C(B n ), and so this polynomial is constant for all t < k1 . But for all t ∈
[0, 1]. It = B n Det ( f t (x))d x is a polynomial in t, which is C(B n ) for 0 < t < k1 .
So It is constant for all t ∈ [0, 1]. Now, f 1 . f 1 = f 1 (x)2 = x2 = 1 for x ∈ B n ,
as f 1 = f is the retraction of B n onto S n−1 . So ∂∂xf1i . f 1 = 0 for 1 ≤ i ≤ n on B n . This
system of linear equations has non-trivial solutions on B n only when Det ∂∂xf11 = 0
where f 1 = ( f 11 , f 12 , . . . , f 1n ). So I (1) = 0 and this contradicts that C(B n ) > 0.
(1 − x2 )
w(x) = x − f (x)
(1− < x, f (x) >)
for x ∈ B n .
Clearly < x, f (x) > < 1 for x ∈ Bn . Otherwise < x, f (x) > ≥ 1 would imply
that 1 ≤ < x, f (x) > ≤ x f (x) ≤ 1 by Cauchy–Schwarz inequality leading
to 1 = < x, f (x) > = x f (x). This would mean that x = c f (x) for some
c = 0 and 1 = < x, f (x) > = c < f (x), f (x) > = c f x2 = 1c x2 with c > 0.
Therefore c = 1 with x = f (x), contradicting that f has no fixed point. Thus
< x, f (x) > < 1 for x ∈ B n .
1−x2 1−x2
Suppose w(x) = 0. Then x = 1−<x, f (x)>
f (x) = c f (x), where c = 1−<x, f (x)>
.
| f (x)
2 2
Since x = c f (x), c = 1−|c1−c f (x)2
, c − c 2 f (x)2 = 1 − |c |2 f (x)2 . So c =
1, contradicting that f has no fixed point. So w(x) = 0 for all x ∈ B n .
w(x)
Define g : B n → Rn by g(x) = w(x) . Clearly g(B n ) ⊆ S n−1 and for x ∈ S n−1 ,
w(x) = x and so g(x) = x. Thus g is a continuously differentiable retraction of B n
onto S n−1 , contradicting Lemma 10.3.3.
Hence every continuously differentiable map of B n into itself has a fixed
point.
Remark 10.3.6 From the no-retraction theorem we can deduce Brouwer’s theorem.
The next theorem points out that all compact convex subsets of Rn have the fixed-
point property for continuous functions and is a precursor to Schauder’s fixed-point
theorem.
10.3 Brouwer’s Fixed-Point Theorem 223
n+1
λi vi ∈ S and λi ≥ 0 . If ψ(y) = {i 0 , . . . , i k } then y is in the face [xi0 , . . . , xik ].
i=1
This face is called the carrier of y.
n+1
For the n-dimensional simplex S described above, if y ∈ S, then y = αi vi
i=0
and in this representation αi are uniquely defined. (α1 , . . . , αn+1 ) are called the
barycentric coordinates of y. Thus the carrier of y is well defined.
224 10 Brouwer’s Fixed-Point Theorem
Example 10.4.5
3
1 5 2
In the above 2-simplex [1, 2, 3], the subdivision {[1,3,5], [2,3,4], [2,4,5], [1,3],
[1,5], [2,3], [2,4], [2,5], [3,4], [4,5], 1, 2, 3, [4], [5]} is not simplicial for the inter-
section of the faces [1,3,5] with [2,3,4] is [3,4] and is not a common face. On the
other hand in the following figure:
3
1 6 2
for the simplex [1, 2, 3] {[1,3,4], [1,4,5], [1,5,6], [2,5,6], [2,4,5], [2,3,4], [1,4], {1},
[1,5], [4,5], [5,6], {5}, [2,4], [3,4], {2}, {4}} is a simplicial subdivision.
Definition 10.4.6 Let S = [x1 , . . . , xn+1 ] be an n-dimensional space which is sim-
plicially subdivided, with V being the set of all vertices of all sub-simplexes. A
10.4 A Proof of Brouwer’s Theorem from Sperner’s Lemma 225
n+1
n+1
where v = λi vi and f (v) = f i (v)vi .
i=1 i=1
This intersection is non-empty, for if f i (v) > λi for all i ∈ {i 1 , . . . , i }, then we
would have
n+1
n+1
1= f i (v) > vi = vi = 1
i=1 j=1 i=1
a contradiction, with the second inequality following from v ∈ [vi1 , . . . , vi ]. It may
be noted that we are using the representation of v, f (v) in barycentric coordi-
nates. Since λ is a labelling function satisfying Sperner’s lemma there exists a
completely labelled sub-simplex [ p1 , . . . , pm+1
] such that f i ( pi ) ≤ ( pi )i . As ↓ 0
there is a subsequence of simplexes such that pi → q as → 0 for each i =
1, 2, . . . , m + 1. Since f is continuous f i (q) ≤ qi , i = 1, . . . , m + 1. If f (q) = q,
f i (q) ≤ qi for all i and f k (q) < q for some k would contradict f k (q) = qk = 1.
So f (q) = q.
Using Sperner’s lemma, one can deduce a classical result due to Knaster Kura-
towski and Mazurkiewicz [14], called Knaster–Kuratowski–Mazurkiewicz Lemma
or simply KKM lemma.
Theorem 10.4.10 (KKM Lemma) Let be the simplex [e1 , . . . , em+1 ] in Rm+1
and F1 , . . . , Fm+1 be a family of non-empty closed subsets of such that for each
A ⊆ {1, 2, . . . , m + 1} the convex hull of {ei : i ∈ A} ⊆ Fi .
i∈A
m+1
Then Fi is non-empty and compact.
i=1
m+1
Proof Clearly Fi is closed and compact. For > 0 given subdivide into sub-
i=1
simplexes with mesh size ≤ . For a vertex v of the subdivision lying on the face
ei1 , . . . , eik+1 , by hypothesis there is an index i ∈ {i 1 , . . . , i k+1 } with v ∈ Fi . Labelling
all the vertices in this way. We observe that it satisfies all the conditions of Sperner’s
Lemma. So there is a completely labelled sub-simplex [ p1 , . . . , pm+1
] with pi ∈
Fi for each i. As ↓ 0, choosing a subsequence pi converging to q and noting
pi ∈ Fi for each i, it follows that z ∈ Fi for each i as Fi is closed for each i. Thus
m+1
z∈ Fi .
i=1
10.4 A Proof of Brouwer’s Theorem from Sperner’s Lemma 227
We can also prove that Brouwer’s theorem implies the KKM Lemma.
Aliter(Peleg [22]) Let F1 , . . . , Fm+1 satisfy the hypotheses of KKM Lemma. Define
gi (x) = d(x, Fi ) and f : → by f i (x) = 1+xi +g i (x)
m+1
g (x)
. Then f is continuous and
j=1 j
so by Brouwer’s theorem f has a fixed point x. So g j (x) = 0 for all j ∈ {1, . . . , m +
m+1
1}. So F j = φ.
j=1
For each y, F(y) = K − int U −1 (y) is a closed subset of K and hence is compact.
n n
For y ∈ co{x1 , . . . xn : xi ∈ K }, y ∈ F(xi ) otherwise y ∈
/ F(xi ) implies y ∈
/
i=1 i=1
n
n
(K − int U −1 (xi )) = K − int U −1 (xi ) or y ∈ int U −1 (xi ) ⊆ U −1 (xi ) for all
i=1 i=1
230 10 Brouwer’s Fixed-Point Theorem
V of Rn {x ∈ S : ψ(x) ⊆ V } is open in S.
Following Krasa and Yannelis [15], we can give an alternative proof of Theorem
10.6.3 using Brouwer’s theorem.
Aliter for Theorem 10.6.3 For each x ∈ K , define ϕ(x) = U −1 {x}. ϕ has open
lower sections. Suppose U −1 (x) = φ for all x. By assumption (ii) of Theorem
10.6.3 int U −1 {x} = φ. So {intϕ(x) : x ∈ K } is an open cover for K and as K
is compact it has a finite subcover {int U −1 (y1 ), int U −1 (y2 ), . . . , int U −1 (ym )}
say. Let gi (x) = dist (x, K − int ϕ(yi )} and αi (x) = mgi (x)
g j (x)
. Clearly each αi is
j=1
m
continuous, 0 ≤ αi ≤ 1, αi = 0 on K − int ϕ(yi ) and αi (x) = 1 for x ∈ K .
i=1
Thus {αi } is a partition of unity subordinate to the covering {int ϕ(yi )} for K .
m
The map f : K → K defined by f (x) = αi (x)yi is continuous on the com-
i=1
pact convex set K in Rn and so by Brouwer’s fixed-point theorem it has a fixed
m
point x0 = f (x0 ) = αi (x0 )yi in K . For all the αi (x0 ) = 0, x0 ∈ int ϕ(yi ) and
i=1
m
−1
x0 ∈ U (yi ) or yi ∈ U (x0 ). So x0 = αi (x0 )yi ∈ convex hull of U (x0 ) contra-
i=1
dicting the hypothesis of Theorem 10.6.3. Thus U has a maximal element.
Some consequences of Theorem 10.6.3 can now be deduced.
Theorem 10.6.6 Let U be a binary relation on K ⊆ Rm with values in Rm , K being
a non-empty compact convex subset. Suppose
1. x ∈
/ U (x) for all x ∈ K ;
2. U (x) is convex for all x ∈ K ;
3. {(x, y) : y ∈ U (x)} is open in K × K .
Then U has a maximal element and the set of all such maximal elements of U is
compact.
10.6 More on Brouwer’s Fixed-Point Theorem 231
p∗
λ( p∗ )
and β = λ( p ∗ )μ( p ∗ ) we get f i (π ∗ ) ≤ βπi∗ with equality unless πi∗ = 0. Since
S is the standard simplex and π ∗ , f (π ∗ ) ∈ S = {(π1 , . . . πn ) : i=1n
πi = 1, πi ≥ 0}
and so β = 1. So f i (π ∗ ) ≤ πi∗ with equality unless πi∗ = 0. This again implies that
f i (π ∗ ) = πi∗ , i = 1, 2, . . . , n. Thus π ∗ is a fixed point for f .
In other words, the equilibrium point is an n-tuple such that each player’s mixed
strategy maximizes his pay-off if the strategies of others are held fixed. So each
player’s strategy is optimal against those of the remaining players. A mixed strategy
si is said to use a pure strategy πiα if si = Ciβ πiβ and Ciα > 0. If s = (s1 , . . . , sn )
β
and si uses πiα , we say that s uses πiα . Since pi (s) is linear in si
Define piα (s) = pi (s; πiα ). Then a trivial necessary and sufficient condition
for s to be an equilibrium point is pi (s) = max piα (s). For s = (s1 , . . . , sn ), si =
α
ciα πiα , then pi (s) = ciα piα (s). So for the validity of pi (s) = max piα (s), we
α
α α
must have ciα = 0 whenever piα (s) < max piβ (s). This simply means that s does
β
not use πiα unless it is an optimal pure strategy. With these preliminaries Nash [19]
proved the following:
Theorem 10.6.11 (Nash [19]) A finite cooperative game has a symmetric equilib-
rium point.
Proof si = πiα / 1 has the property (si )φ = s j where j = i ψ so that the n-
α α
tuple s = (s1 , . . . , sn ) is fixed under any ψ. So any game has at least one symmetric
n-tuple. Clearly if s and t are symmetric, so is their convex combination. Thus the set
of symmetric n-tuples is convex. It is also closed. Consider the map T taking s to s
10.6 More on Brouwer’s Fixed-Point Theorem 235
defined in the proof of Theorem 10.6.10 and ψ is a symmetry on the game, then for
s = T s, (s )ψ = T (s ψ ). So T maps the closed convex subset of symmetric n-tuples
into itself and by the continuity of T and Brouwer’s theorem T has a fixed point in
the set of symmetric n-tuples. Thus there is a symmetric equilibrium point.
In an earlier paper, Nash [18] outlined the proof of the existence of the equilibrium
point using Kakutani’s fixed-point theorem. This seminal work culminated in the
award of a Nobel prize (jointly in economics) for Nash. For a perspective on the
impact of Nash equilibrium on social sciences Holt and Roth [10] may be referred.
Arnold [2] and later Niven [21] attempted a proof of the fundamental theorem of
algebra based on Brouwer’s fixed-point theorem. Subsequently, it was noted in [3]
that both the proofs contained errors. Later, Fort [7] salvaged it and we present
his proof first of Brouwer’s fixed-point theorem in the plane and then that of the
fundamental theorem of algebra.
By S we denote the set of all complex numbers z with |z| = 1. For z ∈ S, A(z) is
the set of all real numbers θ for which z = eiθ . Thus A(z) is the set of all arguments
of z. A continuous function defined on a subset X of the plane of complex numbers
and taking values in S is said to have a continuous logarithm on X if there exists a
real-valued continuous function φ of X such that f (z) = eiφ(z) for all z ∈ X . Two
basic properties of complex numbers, used in the sequel, are the following:
(a) for z 1 , z 2 ∈ S, |z 1 − z 2 | < 2 and θ1 ∈ A(z 1 ), then for a unique θ2 ∈ A(z 2 ) with
|θ1 − θ2 | < π.
(b) if θi ∈ A(wi ), i = 1, 2, and |θ1 − θ2 | < π, then |θ1 − θ2 | ≤ π|z 1 − z 2 |.
Theorem 10.7.1 If f : D → S is a continuous mapping, then f has a continuous
logarithm, D being a closed disc in the plane.
Proof Let D be the disc {z : |z − q| ≤ r }. Thus q is the centre of D and r its radius.
From the uniform continuity of f , it follows that there exists δ > 0 such that for
z 1 , z 2 ∈ D with |z 1 − z 2 | < δ, | f (z 1 ) − f (z 2 )| < 13 . Choose n ∈ N such that nr < δ.
Define Dk = {z : |z − q| ≤ rnk } for k ∈ N with 0 ≤ k ≤ n. Define φ on D by defining
it successively on D0 , . . . , Dn . φ(D0 ) is defined as φ(q) = θ ∈ A( f (q)) such that
0 ≤ θ < 2π. If φ is defined on Dk and z ∈ Dk+1 , let z be the nearest point of Dk to
z. Since |z − z | < δ, | f (z) − f (z )| < 13 . So by property (a) stated above, we may
define φ(z) to be the unique number A( f (z)) that differs from φ(z ) by less than π.
Let Sk be the statement: if z 1 , z 2 ∈ Dk and |z 1 − z 2 | < δ, then |φ(z 1 ) − φ(z 2 )| <
π. Clearly S0 is true as D0 contains only one point. Suppose Sk is true. For z 1 , z 2 ∈
Dk+1 with |z 1 − z 2 | < δ. Consider z 1 and z 2 the nearest points of z 1 and z 2 in Dk
respectively. From the definition of φ, it follows that |φ(z i ) − φ(z i )| < π, for i = 1, 2.
It is readily seen that |z 1 − z 2 | < δ and by inductive hypothesis |φ(z 1 ) − φ(z 2 )| < π.
So we have
236 10 Brouwer’s Fixed-Point Theorem
Proof Without loss of generality we may assume that p(z) = 0 for all z ∈ C and the
p(z) 1
coefficient of z n in p(z) is 21 . Now lim n = . So for some r > 0, p(z) and z n
z→∞ z 2
have arguments which
√ differ by less than π2 for |z| ≥ r and | p(z)| < |z|n . Now select
R > r such that | p(z)| + |z| < R for |z| ≤ r . Consider D the disc centred at 0 with
n
Hamilton [9] extended the Brouwer’s fixed-point theorem for peripherally continuous
maps, while Stallings [29] generalized Brouwer’s theorem for connectivity functions.
Whyburn [35] extended an intersection theorem due to Hurewicz and Wallman [11],
whence he deduced both the generalizations of Hamilton and Stallings. In this section,
Whyburn’s approach to these fixed-point theorems is described. See also [34].
Definition 10.8.1 A subset E of a topological space X is said to be quasi-closed or
of external dimension zero, if for each p ∈ X/E every neighbourhood of p contains
an open set having p, whose boundary does not intersect E. A subset G of X is
quasi-open if its complement is quasi-closed.
Hereinafter, we assume that the topological spaces are regular T1 spaces. The
following definitions are needed in the sequel.
Remark 10.8.5 A locally cohesive space is locally connected and has no local cut
point. If W is a canonical region in X about a ∈ X , any set K separating a and
Fr (W ) in W contains the boundary of a canonical region R lying in W . Thus in a
locally cohesive space X , given a closed subset E of X and an open set U containing
a ∈ X − E contains a canonical region R about a with R ⊂ U and E ∩ Fr (R) = φ.
Thus for G quasi-open, any open set U containing a ∈ G has a canonical region R
about a so that R ⊆ U and Fr (R) ⊆ G.
Definition 10.8.6 Two subsets A and B of a connected space X are weakly separated
in X by a set E provided no component of X − E meets both A and B.
The proof of this lemma is better understood on the basis of the following concepts
and propositions (see Whyburn [33]).
For the proofs of the following propositions Whyburn [33] may be consulted.
for each > 0, it contains at most a finite number of sets with diameter exceeding
) and G3 is a collection of singletons, then G is upper semicontinuous.
Proposition 10.8.16 Proposition 10.8.14 is true when X is locally compact and all
the sets in G are continua.
Definition 10.8.17 A collection G of subsets of a metric space X is called an upper
semicontinous decomposition of X if G = X , each set in G is compact and G is
G∈G
an upper semicontinuous collection.
Remark 10.8.18 If G is an upper semicontinuous decomposition of a metric space,
then a topology on G can be defined by declaring a neighbourhood of G ∈ G as a
subcollection U of G such that ∪{O : O ∈ U} is open in X and contains G. This
topology on G is called the hyperspace topology on G.
Lemma 10.8.19 ([35]) If X is a locally compact, locally cohesive metric space
and Y a regular T1 space, then any connectivity map f : X → Y is peripherally
continuous.
Proof For x ∈ X , let U and V be open sets containing x and f (x), respectively.
Without loss of generality let U be a canonical region with compact closure since X
is T1 , regular and locally compact. So the boundary B of U is connected and U is
unicoherent between x and B. Let U1 and V1 be open sets such that x ∈ U1 ⊆ U 1 ⊆ U
and f (x) ∈ V1 ⊆ V 1 ⊆ V . Let D = U 1 ∩ f −1 (V 1 ). If x is an interior point of A
comprising the component A0 of D containing x together with the union of all
components of U − A0 except the one containing B or x is separated in U from B
by a component H of D, we get an open set W ⊆ U with x ∈ W and Fr (W ) ⊆ A0
or Fr (W ) ⊆ H . In the former case, let W = int A and in the latter case, choose W
= component of U − H . In both the cases f (Fr (W )) ⊆ V .
Suppose x is neither in the interior of A nor is separated in U from B by any
single component D. So the decomposition of U into the sets A, B components of
D not contained in A and single points of U − A − D is upper semicontinuous.
If φ(U ) = M is the natural mapping of this decomposition, then φ is closed and
monotone (i.e. φ−1 (y) is a continuum for each y in the range of φ). So M is a
locally connected continuum. If a = φ(x) and b = φ(B) and N = C(a, b) the cyclic
element taken in M, then no point of φ(D) ∩ N is a cutpoint of N as no such point
can separate a and b in M or N . Since U is unicoherent between x and Fr (U ),
N is unicoherent. As φ(D) is totally disconnected R = N − φ(D) is connected. So
a ⊆ R. So φ−1 (R) = Q is connected as φ is monotone and closed. Also Q ⊇ {x},
as any region S in U1 must intersect Q. If S is not in A and a is not a cutpoint of
M. φ(S) ∩ N is non-degenerate and connected and so is not connected and is not
contained in φ(D). But then Q ∪ {x} is connected while (x, f (x)) is an isolated
point of the graph of f |Q ∪ x, because (q, f (q)) is not in U1 × V1 for q ∈ Q, since
f (q) ∈ Y − V1 for all q ∈ Q. This contradiction implies that x is either in the interior
of A or is separated in U from B by some single component of D. So f is peripherally
continuous.
242 10 Brouwer’s Fixed-Point Theorem
It may be added that Nash [20] raised the question of whether a connectivity map
on I n has a fixed point.
References
1. Apostol, T.A.: Mathematical Analysis, 2nd edn. Narosa Publishing House, New Delhi (1985)
2. Arnold, B.H.: A topological proof of the fundamental theorem of algebra. Am. Math. Mon.
56, 465–466 (1949)
3. Arnold, B.H., Nivan, I.: A correction. Am. Math. Montly 58(104) (1951)
4. Brouwer, L.E.J.: Uber abbildung von mannigfaltigkeiten. Math. Ann. 71, 97–115 (1912)
5. Dugundji, J.: Topology. Allyn and Bacon Inc., Boston (1966)
6. Fan, Ky.: Fixed point and minimax theorems in locally convex topological linear spaces. Proc.
N.A.S. 38, 121–126 (1952)
7. Fort, M.K.: Some properties of continuous functions. Am. Math. Mon. 59, 372–375 (1952)
8. Glicksberg, I.L.: A further generalization of the Kakutani fixed point theorem with application
to Nash equilibrium points. Proc. Am. Math. Soc. 3, 170–174 (1952)
9. Hamilton, O.H.: Fixed points for certain non continuous transformations. Proc. Am. Math.
Soc. 8, 750–756 (1957)
10. Holt, C.A., Roth, A.E.: The Nash equilibrium: a perspective. Proc. Natl. Acad. Sci. (USA) 101,
3999–4002 (2004)
11. Hurewicz, W., Wallman, H.: Dimension Theory. Princeton University Press, Princeton (1941)
12. Kakutani, S.: A generalization of Brouwer’s fixed point theorem. Duke Math. J. 8, 457–459
(1941)
13. Kannai, Y.: An elementary proof of the no-retraction theorem. Am. Math. Mon. 88, 264–268
(1981)
14. Knaster, B., Kuratowski, K., Mazurkiewicz, S.: Ein Beweis des Fixpunktsatzes fur n-
dimensionale simplexe. Fundam. Math. 14, 132–137 (1929)
15. Krasa, S., Yannelis, N.C.: An elementary proof of the Knaster-Kuratowski-Mazurkiewicz-
Shapely theorem. Econ. Theory 4, 467–471 (1994)
16. Milnor, J.: Analytic proofs of the “Hairy Ball thoerem” and the Brouwer fixed point theorem.
Am. Math. Mon. 85, 521–524 (1978)
17. Munkres, J.R.: Topology : A First Course. Prentice-Hall Inc., Englewood Cliffs (1975)
18. Nash, J.: Equilibrium points in N -person games. Ann. Math. 36, 48–49 (1950)
19. Nash, J.: Non-cooperative games. Ann. Math. 54, 286–295 (1951)
20. Nash, J.: Generalized Brouwer theorem. Bull. Am. Math. Soc. 62, 76 (1956)
21. Niven, I.: Extension of a topological proof of the fundamental theorem of algebra. Am. Math.
Mon. 57, 246–248 (1950)
22. Peleg, B.: Equilibrium points for open acyclic relations. Can. J. Math. 19, 366–369 (1967)
23. Rogers, C.A.: A less strange version of Milnor’s proof of Brouwer’s fixed point theorem. Am.
Math. Mon. 87, 525–527 (1980)
References 243
24. Scarf, H.: The approximation of fixed points of a continuous mapping. SIAM J. Appl. Math.
15, 1328–1343 (1967)
25. Schauder, J.: Der Fixpunktsatz in funktional raumen. Stud. Math. 2, 171–180 (1930)
26. Seki, T.: An elementary proof of Brouwer’s fixed point theorem. Tohoku Math. J. 9, 105–109
(1957)
27. Sonnenschein, H.F.: Demand theory without transitive preferences, with applications to the
theory of competitive equilibrium. In: Chapman, J.S., Hurwicz, L., Richter, M.K., Sonnen-
schein, H.F. (eds.) Preferences Utility and Demand: A Minnesota Symposium, pp. 215–233.
Harcourt, Brace, Jovanovich, New York (1971)
28. Sperner, E.: Neuer Beweis fur die invarianz der Dimensionszahl und des gebietes. Abh. Math.
Semin. Univ. Hambg. 6, 265–272 (1928)
29. Stallings, J.: Fixed point theorems for connectivity maps. Fundam. Math. 48, 249–263 (1959)
30. Traynor, T.: An easy analytic proof of Brouwer’s fixed point theorem. Atti Semin. Math. Fis.
Univ. Modena XLIV, 479–483 (1996)
31. Tychonoff, A.: Ein Fixpunktsatz. Math. Ann. 111, 767–776 (1935)
32. Uzawa, H.: Walras’ existence theorem and Brouwer’s fixed point theorem. Econ. Stud. Q. 3,
59–62 (1962)
33. Whyburn, G.T.: Analytic Topology. American Mathematical Society, vol. 28. Colloquium
Publications, AMS, Providence (1942)
34. Whyburn, G.T.: Connectivity of peripherally continuous functions. Proc. Natl. Acad. Sci.
(USA) 55, 1040–41 (1966)
35. Whyburn, G.T.: Quasi-closed sets and fixed points. Proc. Natl. Acad. Sci. (USA) 57, 201–205
(1967)
36. Whyburn, G.T.: Loosely closed sets and partially continuous functions. Mich. Math. J. 14,
193–205 (1967)
Chapter 11
Schauder’s Fixed Point Theorem
and Allied Theorems
11.1 Introduction
Example 11.1.1 B, be the closed unit ball in C0 the space of all null real sequences
x = (xn ) with the norm x = sup |xn | does not have the fixed point property.
n
For example, the map x → T x where T (x) = (1, x1 , x2 , . . . ), x being (x1 , x2 , . . . ,
xn , . . . ) maps B into itself. If T (x) = x, then x = (xn ) with xn ≡ 1 for all n contra-
dicting that xn is a null sequence.
Proof Given any > 0, by the compactness of K , we can find a finite number of
points x1 , . . . , x N in K such that each x ∈ K lies in an open ball centred at xi for
some i = 1, 2, . . . , N and of radius . Define g j : K → R+ by
− x − x j , if x − x j <
g j (x) =
0, if x − x j ≥ , j = 1, 2, . . . , N .
g j (x)
h j (x) = N .
i=1 gi (x)
N
Further, h j (x) = 1 ∀ x ∈ K and h j (x) = 0 if x − x j ≥ . The map x → V (x)
j=1
defined by
N
V (x) = h j (x)x j
j=1
N
x − V (x) = h j (x)(x − x j )
j=1
where the sum is only over those j for which x − x j < has positive contribution.
Thus for x ∈ K
x − V (x) ≤ h j (x)x − x j < .
V f (xn ) = xn .
We make use of the above criterion for compactness in the proof of Peano’s
Theorem.
11.1 Introduction 247
Sketch of Proof. This initial value problem is equivalent to solving the integral equa-
tion t
x(t) = x0 + f (s, x(s))ds.
t0
Even when a continuous function does not map a compact convex set into itself it
may have a fixed point under some additional assumptions. Rothe [16] obtained the
following fixed point theorem belonging to this category. For this purpose, we need
the following.
Theorem 11.1.8 (Rothe [16]) Let X be a normed linear space, B its closed unit
ball and S the unit sphere. Let T : B → X be a compact continuous map such that
T (∂ B) ⊆ B. Then T has a fixed point in B.
Proof Define r : X → B by
x, if x ∈ B
r (x) = x
x
, if x ∈
/ B.
11.2.1 Introduction
For I = [a, b], a < b, a, b ∈ R let C(I, R) be the Banach space of continuous real-
valued functions on I with supremum norm · . For M ≥ 0, we define
and for δ ≥ 0,
we have
Proof Let f ∈ Fδ (M). If M < 1, then f (x) − f (y) < x − y for x > y. Setting
x > y = a, it follows that f (x) < x as f (a) = a. This implies that f (b) = b < b, a
contradiction. So Fδ (M) = ∅. A similar argument shows that Fδ (M) = ∅ for δ > 1.
For M = 1 and f ∈ Fδ (M) and x > y, f (x) − f (y) ≤ x − y. For y = a we get
f (x) ≤ f (a) for all x. For x = b, y ≤ f (y) for all y, so that f (x) = x on I . A
similar argument for δ = 1 implies that Fδ (M) is a singleton containing only the
identity function.
Proof Clearly, Fδ (M) is a closed and convex subset of C[I, R]. Also | f (x)| ≤
max{|a|, |b|} for x ∈ I . As δ(x − y) ≤ f (x) − f (y) ≤ M(x − y) for x > y, x, y ∈
I , for f ∈ Fδ (M), Fδ (M) is uniformly bounded and equicontinuous. So by Arzela–
Ascoli Theorem Fδ (M) is compact.
For a proof, Zhang and Baker [24] may be consulted. The following proposition
implies that it suffices to solve the functional equation in [0, 1].
250 11 Schauder’s Fixed Point Theorem and Allied Theorems
where h(x) = a + x(b − a), μi (x) = λi (h(x)), Ri (x) = h −1 (Hi (h(x))), G(x) =
∞
h −1 (Fh(x)) and λi (x) ≥ 0 with λi (x) = 1.
i=1
∞
Noting that h and h −1 are affine continuous maps and μi (x) = 1 on [0, 1], for
i=1
x ∈ [0, 1]
∞
n
1
μi (x)Ri (g (x)) = lim
i
μi (x)Ri (g i (x))
n→∞
n
i=1
μi (x) i=1
i=1
1
n
= lim μi (x)h −1 (Hi hh −1 f i (h(x)))
n→∞
n
μi (x) i=1
i=1
n
μi (x)
= lim h −1 (Hi f i (h(x))
n→∞
n
i=1
μ j (x)
j=1
n
= h −1 lim μi (x)Hi f i (h(x))
n→∞
i=1
−1
=h F(h(x)) = G(x).
Lemma 11.2.6 If f, g ∈ Q(M), M > 1 and map I onto itself, then for i ∈ N
Mi − 1
f i − gi ≤ f − g.
M −1
∞
αi
K1 = + i Li M
i−1
i=1
δ
∞
L f (x) = λi ( f −1 (x))Hi ( f i−1 (x)), x ∈ I.
i=1
Lemma 11.2.8 Suppose that in addition to the hypotheses of the above theorem
f ∈ F(M). Then L f ∈ FK 0 (K 1 ) where K 0 and K 1 are as defined in that theorem.
Proof It is easy to see that L f (0) = 0 and L f (1) = 1. For x ≥ y, x, y ∈ I
∞
∞
L f (x) − L f (y) = λi ( f −1 (x))Hi ( f i−1 (x)) − λi ( f −1 (y))Hi ( f i−1 (y))
i=1 i=1
∞
= {[λi ( f −1 (x)) − λi ( f −1 (y))]Hi ( f i−1 (x))
i=1
252 11 Schauder’s Fixed Point Theorem and Allied Theorems
Similarly
∞
αi
L f (x) − L f (y) ≥ γi li δ i−1 − (x − y) = K 0 (x − y)
i=1
δ
∞
αi i L i (M
i−1
)
+ .
i=1
δ M −1
From Lemmata 11.2.3 and 11.2.6 and the definitions of λi (x) and Hi (x), one obtains
∞
|L f (x) − L g (x)| ≤ αi | f −1 (x) − g −1 (x)| + i Li f i−1 − g i−1
i=1
∞
αi Mi − 1
≤ + i Li f − g
i=1
δ M −1
1
T f (x) − T f (y) = L −1 −1
f (F(x)) − L f (F(y)) ≤ [F(x) − F(y)]
K0
≤ M(x − y).
Further,
1 1
T f (x) − T f (y) ≥ |F(x) − F(y)| ≥ K 1 δ(x − y) = δ(x − y).
K1 K1
Thus T f ∈ Fδ (M) and maps Fδ (M) into itself. For f, g ∈ Fδ (M) and x ∈ I
|T f (x) − T g(x)| = |L −1 −1
f (F(x)) − L g (F(x))|
≤ L −1 −1
f − L g (x)
Definition 11.3.1 Let (M, d) be a complete metric space and X a bounded subset
of M. Kuratowski measure of noncompactness of X denoted by α(X ) is defined by
Using the above definition, the following proposition can be proved easily.
Remark 11.3.6 When X is compact, the above result leads to Mazur’s theorem that
the closed convex hull of a compact set is compact. The calculation of the measure
of noncompactness is not always easy. In fact, the proof that α(B(0, 1)) = 2 for the
unit open ball B(0, 1) in a Banach space is not obvious.
Remark 11.3.7 α t X = t0 α(X ). This may be deduced from tX ⊆
0≤t≤t0 0≤t≤t0
conv[t0 X ∪ {0}].
Remark 11.3.8 One may alternatively cover a bounded subset X of a metric space
(X, d) by a finite number of open balls of radius smaller than > 0. This leads to the
Hausdorff (or ball) measure of noncompactness of X . This has been used by Gohberg,
Goebel, Nussbaum and others. This measure is closely related to the concept of
Hausdorff metric and is denoted by χ . It can be proved that ψ(X ) = inf{H (X, F) : F
a nonempty compact subset of M and H the Hausdorff distance on the space of all
closed nonempty bounded subsets of M}.
It is, therefore, possible to propose an abstract concept of a measure of noncom-
pactness based on an axiomatic approach.
Definition 11.3.9 A nonempty subfamily P of the family N of nonempty relatively
compact subsets of E is said to be the kernel (of a measure of noncompactness) if
the following conditions are satisfied:
(i) X ∈ P ⇒ X ∈ P;
(ii) X ∈ P, φ = Y ⊆ X ⇒ Y ∈ P;
(iii) X, Y ∈ P ⇒ λX + (1 − λ)Y ∈ P for λ ∈ [0, 1];
(iv) the family of compact sets in P is closed in the family of nonempty compact
sets with Hausdorff metric.
Definition 11.3.10 Let M be the family of non-void bounded sets of E. A function
μ : M → [0, ∞) is said to be a measure of noncompactness with the kernel P if it
satisfies the following:
(i) μ(X ) = 0 ⇔ X ∈ P;
(ii) μ(X ) = μ(X );
(iii) X ⊆ Y ⇒ μ(X ) ≤ μ(Y );
(iv) μ(conv X ) = μ(X );
(v) μ(λX + (1 − λ)Y ) ≤ λμ(X ) + (1 − λ)μ(Y ) for λ ∈ [0, 1];
(vi) for X n ∈ M and X n = X n and X n+1 ⊆ X n for all n ∈ N and lim μ(X n ) = 0
n→∞
∞
imply X n = φ.
n=1
Theorem 11.3.14 (Darbo [4]) Let C be a non-void closed convex set which is
bounded and μ be any measure of noncompactness with the kernel of the family
of bounded nonempty convex subsets of E. Let T : C → C be a μ-contraction in the
sense that μ(T (X )) ≤ kμ(X ) for any nonempty convex bounded subset X , where
0 ≤ k < 1. If T is continuous, then T has a fixed point and the set of fixed points of
T in C belongs to the kernel of μ.
respect to α has no fixed point. Thus Darbo’s theorem or Sadovskii’s theorem is not
true for 1-set contractions.
A fixed point theorem for mappings satisfying the so-called Leray-boundary con-
dition involving measures of noncompactness can also be proved.
Theorem 11.3.20 Let μ be a measure with kernel P having maximum property such
that {θ } ∈ P. Suppose C is an open and bounded neighbourhood of θ of a Banach
space E and T : C → E is a continuous k-set contraction with contractive constant
k (< 1) such that for any x ∈ ∂C, T x = λx for λ > 1. Then T has a fixed point in
C and the set of fixed points of T belongs to P.
Proof Let K = {x ∈ C : x = cT x for some c ∈ [0, 1]}. K is nonempty as θ ∈ K
and is obviously closed. As K ⊆ conv(T K ∪ {θ }) we have μ(K ) ≤ μ(T (K ) ∪
{0}) = μ(T (K )) ≤ kμ(K ) and K ∈ P. Clearly, K ∩ ∂C = φ. Since K is compact
and E − C is non-void closed set and disjoint from K , by Urysohn’s Lemma there
is a continuous function g : E → [0, 1] such that g(x) = 1 for x ∈ K and g(x) = 0
for x ∈/ C and 0 < g(x) < 1 for x ∈ C − K . Define the map F : E → E by
g(x)T (x), for x ∈ C
F(x) =
0, for x ∈
/C
Clearly, F maps each ball B(θ ; r ) (r > 0) containing C into itself. For any set X ,
F(X ) ⊆ conv(T (X ∩ C) ∪ {θ }) and so
theorem for a commuting family of affine continuous mappings, due to Markov [14]
and Kakutani [11].
Definition 11.4.5 A mapping f on a convex set C into a linear space is called affine
if f (αx + (1 − α)y) = α f (x) + (1 − α) f (y) for all x, y ∈ C and α ∈ [0, 1].
Proof By Tychonoff’s theorem (Corollary 11.4.4), F(T ), the set of fixed points, is
nonempty for each T ∈ F . As T is affine, F(T ) is closed and convex as well. If
S ∈ F , then S maps F(T ) into itself by the commutativity of S and T and again
by Tychonoff’s Theorem has a fixed point in F(T ). So F(T ) ∩ F(S) = φ. In fact,
the family {F(T ) : T ∈ F } is a family of nonempty compact convex subsets of K
with finite intersection property. Since K is compact, ∩{F(T ) : T ∈ F } = φ. Any
element in this intersection is a common fixed point for mappings in F .
Kakutani [11] has also given a more elementary proof of this theorem and it is
described below.
1
T(n) (x) = (x + t x + · · · + T n−1 (x))
n
for x ∈ M and n > 1. T k being the kth iterate of T is affine, maps M into
itself and commutes with each U ∈ A. Further, a fixed point of T is also a
fixed point of T(n) for all n > 1. So we may without loss of generality assume
that A is a semigroup containing such convex combinations of the iterates of T .
If S and T are in A then S(T (M)) ⊆ S(M) ⊆ M and S(T (M)) = T (S(M)) ⊆
T (M) ⊆ M. So ST (M) ⊆ S(M) ∩ T (M) ⊆ M. Since ST ∈ A ST (M) = φ. Thus
∩ T (M) = φ. If therefore follows that if A1 is a
S(M) finite subset of A, then
U (M) = φ. Since U (M) is compact for each U ∈ A, T (M) = φ. Let x ∗ ∈
U ∈A1 T ∈A
T (M). Then x ∗ = T(n) y for some y ∈ M. So T x ∗ − x ∗ = n1 (T n y − y) ∈ n1 M1
T ∈A
where M1 = {x − y : x, y ∈ M}. As M is compact, M1 is compact and bounded.
Since n ∈ N is arbitrary, T x ∗ = x ∗ . T being an arbitrary element of A, it follows
that x ∗ is a common fixed point for all T ∈ A.
Proof Suppose the lemma is false. Then the diagonal = {(x, x) : x ∈ K } and
the graph of T , viz = {(x, T x) : x ∈ K } are disjoint compact convex subsets of
E × E. So by the Hahn–Banach theorem there exist continuous linear functional 1
and 2 on E and real numbers α and β such that
Markov–Kakutani theorem also provides other applications, for example the exis-
tence of an invariant mean and the existence of Banach limits. These are described
below.
Definition 11.4.9 Let P(X ) be the set of all Borel probability measures on a compact
Hausdorff space X . Let μ ∈ P(X ). A μ-measurable map f : X → X is called mea-
sure preserving with respect to μ if μ(B) = μ( f −1 (B) for every Borel set B ⊆ X .
In this case μ is called an invariant measure for f .
Remark 11.4.10 For a compact Hausdorff space the dual of C(X ) can be identified
with M(X ), the space of complex regular Borel measures on X , (in view of the Riesz
Representation theorem) with μ = total variation of μ.
If f : X → X is continuous, then f˜μ defined by f˜μ (B) = μ( f −1 (B)) for each
Borel set B defines a probability measure on X .
We have
Hence we have
Theorem 11.4.12 For f ∈ C(X ), there exists μ ∈ P(X ) for which f is measure
preserving.
Proof The map f˜ : P(X ) → P(X ) defined by f˜μ (B) = μ( f −1 (B)) for any Borel
subset of X is continuous in the weak∗ topology by Lemma 11.4.11 and P(X ) is a
convex and closed subset of the closed unit sphere of M(X ). M(X ) is compact by
Alaoglu’s theorem. So by Tychonoff’s theorem f˜ has a fixed point μ∗ in P(X ) and
μ∗ (B) = f˜μ∗ (B) = μ∗ ( f −1 (B)) or f is measure preserving with respect to μ∗ .
11.4 Kakutani-Ky Fan–Glicksberg Fixed Point Theorem 263
(L t f )(s) = f (t ◦ s) for s ∈ S
≤ sup | f | = = 1
f ≤1
as L s f ≤ f . So TS (K ) ⊆ K . Further,
Ts Tt ( ) = Ts ( ◦ L t ) = ( ◦ L t ) ◦ L s
= ◦ L st = ◦ L ts
= Tt Ts ( ) ∀ ∈ B(S)∗ .
Thus {Ts } is a commuting family of linear operators mapping the compact convex
set into itself and so has a fixed point ∗ in K . Clearly, ∗ is left-invariant (and
right-invariant as S is commutative) and ∗ L s (1) = ∗ 1 = 1. We now show that
each element of K is positive. Suppose not. Then there exists f ∈ B(S), f ≥ 0 such
that for some ∈ K , f = β < 0. So for small > 0, we get
264 11 Schauder’s Fixed Point Theorem and Allied Theorems
So
1 < 1 − β ≤ |1 − β| = | (1 − f )|
≤ 1 − f ≤ 1,
The existence of a generalized limit or Banach limit for bounded sequences can
also be deduced from fixed point theorems.
We can also deduce the existence of a generalized or Banach limit of a bounded
sequence from Tychonoff’s theorem.
Remark 11.4.17 If a = (an ) ∈ m, then inf an ≤ L(a) ≤ sup an . Let m = inf an and
M = sup an . Clearly, m ≤ an ≤ M for n. So an − m ≥ 0. So L(an − m) ≥ 0, by
(iii). Since L(an − m) = L(an − m(1)) = L(an ) − m L(1, 1, . . . ) ≥ 0. So L(an ) ≥
m L(1) = m by (ii). Similarly L(an ) ≤ L(M) = M L(1, 1, . . . ) = M. Thus inf an ≤
L(a) ≤ sup an . This in turn implies |L(a)| ≤ sup |an | = a. Since L(1) = 1, L =
1. From (iv) and inf(an ) ≤ L(a) ≤ sup(an ) it follows that lim(an ) ≤ L(a) ≤ lim(an )
for a = (an ) ∈ ∞ (v). Also (i) and (v) imply (ii) and (iii).
Proof Define K = {L ∈ m ∗ : L satisfies (i), (ii), (iii) and (iv) of Definition 11.4.16}.
Clearly, K is non-void as L(a) = a1 for a = (an ) lies in K . Also K is con-
vex. Since K = ∩{L ∈ m ∗ : L(a) ≤ sup an } ∩ {L ∈ m ∗ : L(a) ≥ inf an } and {L :
L(a) ≤ ξ } and {L : L(a) ≥ η} are weak∗ closed in m ∗ , K is weak∗ closed. As
L = 1, K is a weak∗ closed subset of the unit ball in m ∗ ; K is weak∗ compact by
the Banach–Alaoglu theorem. K is thus compact and convex. Define the map T :
K → K by T (a) = T L(a) = L(a2 , a3 , . . . ) for a = (a1 , a2 , . . . ) T ∈ m ∗∗ . For a ∈
m and Q ∈ m ∗∗ , T −1 (N (Q; a)) = {L : T L ∈ N (Q, a)} = {L : |T L(a) − Q(a)| <
1} = {L : |L(a2 , a3 , . . . ) − Q(a1 , . . . , an , . . . )| < 1} = N (Q 1 ; (a2 , a3 , . . . )), Q 1
being in m ∗ for which Q 1 (a2 , a3 , . . . ) = Q(a1 , . . . ).
11.4 Kakutani-Ky Fan–Glicksberg Fixed Point Theorem 265
M f = { f λ : λ ∈ G}
is precompact in the uniform norm, f λ being the function f λ (x) = f (λx), λ, x ∈ G).
f is called left weakly almost periodic if the set M f defined above is precompact
in the weak topology in the linear space of bounded functions on G with uniform
norm.
Theorem 11.4.25 Let G be a group and f a bounded function and M f = { f λ : λ ∈
G, where f λ (x) = f (λx) for all x ∈ G}. Let K = coM f . If
(i) f is left almost periodic, then K contains constant functions;
(ii) if f is almost periodic and M f is precompact in the weak topology, then too K
contains constant functions.
Proof (i) K is a compact convex subset of B(G) the space of bounded functions on
G. The operators Tλ defined by Tλ g = gλ is a group of isometries mapping K into
itself. So by Kakutani’s Theorem 11.4.19. Tλ has a common fixed point f in M.
Thus f (x) = f (λx) ∀ λ ∈ G. Thus f (x) = f (e) for all x ∈ G or f is constant.
If M f is weakly compact, K is weakly compact and convex. Let Tλ be the group
of isometries mapping M into itself as before. Then by Ryll-Nardzewski Theorem
11.4.23. Tλ has a common fixed point which is a constant function again.
Next, following Rudin [17] we deduce the existence of left-invariant Haar measure
on a compact group. Let G be a compact topological group and C(G) the Banach
space of all continuous complex-valued functions with the supremum norm (Recall
that a topological group is a group with a Hausdorff topology such that (a, b) → ab−1
is continuous.
Lemma 11.4.26 Let G be a compact group, f ∈ C(G) and HL ( f ) the closed con-
vex hull of the left translates L s f of f . (Thus HL ( f ) = conv{L s f : L s f (x) =
f (sx), s ∈ G}. Then (a) f is uniformly continuous and (b) HL ( f ) is totally bounded
in C(G).
Proof f : G → C is called uniformly continuous if for each > 0 there is a neigh-
bourhood Ne of e the identity in G such that | f (x) − f (y)| < for x, y ∈ G with
x −1 y ∈ Ne .
As f is continuous on G, for each a ∈ G and > 0, there is a neighbourhood
Na of e such that | f (x) − f (a)| < 2 for all x ∈ a Na . From the definition of the
topology on G, we can find neighbourhoods Ua of e such that Ua Ua−1 ⊆ Na . As G is
k
compact and {aUa : a ∈ G} is an open cover for G, G ⊆ ai Uai for some k ∈ N.
i=1
k
Let U = Uai . Let x −1 y ∈ U . Choose ai , i = 1, 2, . . . , k such that y ∈ ai Uai .
i=1
Then | f (a) − f (ai )| < 2 . Now | f (ai ) − f (x)| < 2 as x ∈ yU −1 ⊆ ai Uai U −1 ⊆
a Na . So | f (x) − f (y)| ≤ | f (x) − f (ai )| + | f (ai ) − f (y)| < 2 + 2 = . Thus f
is uniformly continuous on G.
268 11 Schauder’s Fixed Point Theorem and Allied Theorems
For x −1 y ∈ U , | f (x) − f (y)| < . But for all s ∈ G, x −1 y = (sx)−1 (sy) and
|L s f (x) − L s f (y)| = | f (sx) − f (sy)| < as (sx)−1 (sy) = x −1 y ∈ U . Thus
HL ( f ) is an equicontinuous subfamily of C(G). Hence the lemma.
Theorem 11.4.27 Given a compact group G, there is a unique
regular Borel prob-
ability measure μ which is left-invariant. That is G f dμ = G (L s f )dμ for s ∈ G
and f ∈ C(G). This μ is also right-invariant and satisfies the relation
f (x)dμ = f (x −1 )dμ
G G
for f ∈ C(G).
Proof The operators L s on C(G) defined by L s f (x) = f (sx) for any given s ∈ G
and f ∈ C(G) for all x form a semigroup and indeed a group of isometries on
G. So it is equicontinuous. If f ∈ C(G), K f the closure of HL ( f ) is compact by
Lemma 11.4.26. Clearly, L s (K f ) = K f . So by Kakutani’s Theorem 11.4.19, there is
a common fixed point φ in K f for all L s . Thus L s φ = φ for all s ∈ G. Thus φ(x) =
φ(sx) for all s ∈ G so that φ(x) = φ(sx) for all s ∈ G so that φ(x) = φ(e) for all
x. Hence φ is constant. Since K f = cl HL ( f ). φ(e) can be uniformly approximated
by functions in HL ( f ). So for each f ∈ C(G), there exists a constant k which can
be uniformly approximated by convex combinations of left translates of f on G.
Similarly there is a constant k that can be uniformly approximated on G by convex
combinations of right translates of f . We will show that k = k. Let > 0 be any
prescribed
number. There exist finite sets A = {ai } and B = {bi } in G with αi , β j > 0
and αi = β j = 1 and
A B
(I) k − A αi f (ai x) < , x ∈ G and
(II) k − B β j f (b j x) < , x ∈ G
Setting x = b j in (I), multiplying (I) by β, and add over j to get
(III) k − i, j αi β j f (ai b j ) <
Similarly setting x = ai in (II), multiplying (II) by αi and add over i to get
(IV) k − i, j αi β j f (ai b j ) <
From (III) and (IV) it follows that k = k .
Thus for each f ∈ C(G) there is a unique number written as I f such that it can
be uniformly approximated by convex combinations of left translates of f (as well
as convex combinations of right translates of f ) with the following properties:
I f ≥ 0 for f ≥ 0,
I1 = 1
Iα f = α I f for any scalar α
I(L s f ) = I f = I(Rs f ) for each s ∈ G
11.4 Kakutani-Ky Fan–Glicksberg Fixed Point Theorem 269
We now show that I f +g = If + Ig . Given > 0 for a finite subset A = {ai } ⊆ G
and αi > 0 with αi = 1, I f − αi f (ai x) < for all x ∈ G. Define h(x) =
A
αi g(ai x), then h ∈ K g . So K h ⊆ K g . Since each of K f and K g contains unique
i
constant functions Ih and Ig , it follows that Ih = Ig . So there is a finite set B =
{b j } ⊆ G with β j > 0 and β j = 1 such that for all x ∈ G
Ig − β j h(b j x) < .
j
Kitamura and Kusano [12] proved the existence of oscillatory solutions for a first-
order nonlinear functional differential equation under suitable assumptions using
Tychonoff’s theorem. In what follows this existence result is described.
N
x (t) = qi (t) f i (x(gi (t))), t > a > 0 (11.1)
i=1
Suppose
(a) qi , gi : [a, ∞) → R are continuous functions, with qi (t) ≥ 0 and lim gi (t) =
t→∞
∞ for i = 1, 2, . . . , N ;
(b) f i : R → R is continuous, nondecreasing and t f i (t) > 0 for t = 0 for i =
1, 2, . . . , N ;
N ∞
(c) qi (t)dt < ∞;
i=1 a
N
t
x(t) = k + qi (s) f i (x(gi (s)))ds
i=1 T
N ∞
f i (2k) qi (s)ds < k
i=1 T
The next theorem follows similarly and its proof is left as an exercise.
Theorem 11.4.30 Under (a), (b) and (c) of Theorem 11.4.29, the equation
N
x (t) + qi (t) f i (x(gi (t))) = 0 (11.2)
i=1
Example 11.4.31 (Kitamura and Kusano [12]) For α > 0 and β ≥ 1, the equation
Remark 11.4.32 It has been observed in [12] that under the assumptions (a) and (b)
N ∞
of Theorem 11.4.29, along with qi (t)dt = ∞, for the equation
i=1 a
N
x (t) = qi (t) f i (x(gi (t))) (11.3)
i=1
every solution is oscillatory when gi (t) > t for i = 1, . . . , N ; for the equation
N
x (t) + qi (t) f i (x(gi (t))) = 0 (11.4)
i=1
(ii) g(x, ·) is concave for each x ∈ K (i.e. −g(x, ·) is a convex function for each
x ∈ K ).
Then there exists x0 ∈ K such that
Proof For each > 0, and a given x ∈ K we can find yx ∈ K and a neighbourhood
N x of x such that
Since K ⊆ N x and K is compact we can find a finite number of elements
x∈K
n
x1 , x2 , . . . , xn in K with K ⊆ N xi . Let {ϕ1 , . . . , ϕn } be a partition of unity for K
i=1
subordinate to the covering {N xi : i = 1, . . . , n}. Then the map f defined by
n
f (x) = ϕi (x)yxi
i=1
maps the closed convex hull of {yx1 , . . . , yxn } into itself and is continuous. Since X
is a linear topological space and the subspace topology on the closed convex hull of
{yx1 , . . . , yxn } is euclidean, f has a fixed point x ∗ by Brouwer’s fixed point theorem.
So
n
≥ ϕ(x ∗ ) sup g(xi , y) −
i=1 y∈K
for some x0 .
Consider a game involving n(≥ 2) players who pursue a strategy depending on the
strategies of other players. Let the strategy set of the ith player be denoted by K I and
K be the set K 1 × K 2 × · · · × K n . An element of K is called a strategy profile. For
n
each player let f i : K → R be the loss function of the ith player. If f k (x) = 0,
k=1
then this game is called zero-sum game.
(In other words this strategy profile minimizes the loss for each player.)
The existence of a Nash equilibrium for an n-person game can now be proved in
the setting of a locally convex topological vector space.
Proof Define g : K × K → R by
n
g(x, y) = f i (x) − f i (x1 , . . . , xi−1 , yi , xi+1 , . . . , xn )
i=1
for all xi ∈ K i .
274 11 Schauder’s Fixed Point Theorem and Allied Theorems
In the case of two person zero-sum game the hypotheses can be further weakened
and we get a theorem due to von Neumann.
Clearly, g(x1 , x2 ) = f 1 (x1 , x2 ) + f 2 (x1 , x2 ) = 0 or f 1 (x1 , x2 ) = − f 2 (x1 , x2 ).
References
1. Banas, J., Goebel, K.: Measures of Noncompactness in Banach Spaces. Marcel Dekker, Inc.,
New York (1980)
2. Browder, F.E.: The fixed point theory of multivalued mappings in topological vector spaces.
Math. Ann. 177, 283–301 (1968)
3. Cauty, R.: Solution du probleme de point fixe de Schauder. Fundam. Math. 170, 231–246
(2001)
4. Darbo, G.: Punti uniti in transformazioxi a condominio non compatio. Rend. Semin. Mat. Univ.
Padova 24, 84–92 (1955)
5. Day, M.M.: Amenable semigroups. Ill. J. Math. 1, 509–544 (1957)
6. Dugundji, J., Granas, A.: J. Math. Anal. Appl. 97, 301–305 (1983)
7. Fan, Ky.: Fixed point and minimax theorems in locally convex topological linear spaces. Proc.
N.A.S 38, 121–126 (1952)
8. Glicksberg, I.L.: A further generalization of the Kakutani fixed point theorem with application
to Nash equilibrium points. Proc. Am. Math. Soc. 3, 170–174 (1952)
9. Hahn, S.: A fixed point theorem. Math. Syst. Theory 1, 55–57 (1968)
10. Kakutani, S.: Two fixed point theorems concerning bicompact convex sets. Proc. Imp. Acad.
Jpn. 14, 242–245 (1938)
11. Kakutani, S.: A generalization of Brouwer’s fixed point theorem. Duke Math. J. 8, 457–459
(1941)
12. Kitamura, Y., Kusano, T.: Oscillation of first-order nonlinear differential equations with devi-
ating arguments. Proc. Am. Math. Soc. 78, 64–68 (1980)
13. Kuczma, M., Choczewski, B., Ger, R.: Iterative Functional Equations. Encyclopedia of Math-
ematics and Its Application. Cambridge University Press, Cambridge (1990)
14. Markov, A.A.: Quelques theorems sur les ensembles abeliens. C.R. (Dokl.) Acad. Sci. URSS
(N.S) 1, 311–313 (1936)
15. Murugan, V., Subrahmanyam, P.V.: Existence of continuous solutions for an iterative functional
series equation with variable coefficients. Aequ. Math. 78, 167–176 (2009)
16. Rothe, E.: Zur theorie der topologischen Ordnung und der Vektorfelder in Banachschen Rau-
men. Compos. Math. 5, 177–196 (1937)
17. Rudin, W.: Functional Analysis. McGraw Hill International Edition (1991)
References 275
18. Ryll-Nardzewski, C.: On fixed points of semigroups of endomorphisms of linear spaces. In:
Proceedings of the Fifth Berkely Symposium on Statistics and Probability, Berkeley, vol. II
(1966)
19. Sadovski, B.N.: Asymptotically compact and densifying operators. Usp. Math. Nauk. 27, 81–
146 (1972)
20. Schauder, J.: Der Fixpunktsatz in funktional raumen. Stud. Math. 2, 171–180 (1930)
21. Terkelsen, F.: A short proof of Fan’s fixed point theorem. Proc. Am. Math. Soc. 42, 643–644
(1974)
22. Tychonoff, A.: Ein Fixpunktsatz. Math. Ann. 111, 767–776 (1935)
23. Werner, D.: A proof of the Markov-Kakutani fixed point theorem via the Hahn-Banach theorem.
Extr. Math. 8, 37–38 (1993)
24. Zhang, W., Baker, J.A.: Continuous solutions for a polynomial-like iterative equation with
variable coefficients. Ann. Pol. Math. 73, 29–36 (2000)
Chapter 12
Basic Analytic Degree Theory
of a Mapping
12.1 Introduction
The problem of finding the number of solutions of a given equation has engaged a
number of mathematicians. Brouwer, Bohl, Cauchy, Descartes, Gauss, Hadamard,
Hermite, Jacobi, Kronecker, Ostrowski, Picard, Sturm and Sylvester had contributed
to this topic. The Argument principle propounded by Cauchy on the zeros of a func-
tion inside a domain and Sturm’s theorem on the number of zeros of a real polynomial
in a closed bounded interval have evolved into Degree theory of mappings. Even as
the degree of a nonconstant polynomial gives the number of zeros of a polynomial,
the degree of a mapping provides the number of zeros of nonlinear mapping in a
domain. In this chapter, an elementary degree theory of mappings is described from
an analytic point of view proposed by Heinz [3]. For more elaborate treatment, Cronin
[1], Deimling [2], Lloyd [4] Outerelo and Ruiz [6] and Rothe [7] may be referred. It
should be mentioned that Ortega and Rheinboldt [5] had provided a more accessible
version of Heinz’s treatment.
Heinz [3] based his approach on some lemmata and relevant definitions. Throughout
we assume that is an open set in Rn , ∂ its boundary and its closure with respect
to the topology generated by the euclidean norm. We consider a map y : → Rn
with y = (y1 , . . . , yn ) where yi = yi (x1 , . . . , xn ) is the ith component function of
y, where i = 1, 2, . . . , n. For y ∈ C 1 (), the Jacobian J [y(x)] is well-defined for
each x ∈ . Let Ai j (x) be the cofactor of ai j (x) in the determinant of J [y(x)] =
det (ai j (x)).
∂bik ∂ 2 zi ∂ ∂z i ∂ y j n
= =
∂xi ∂xi ∂ yk j=1
∂ yk ∂ y j ∂xi
n
∂bi j
= a ji
j=1
∂ yk
and hence
∂e ∂bik n
=e (12.2.1)
∂ yk i=1
∂ yk
n ∂bik
As de = 1, d ∂∂eyk + e ∂∂dyk = 0, we get from (12.2.1), 0 = ∂∂dyk + d i=1 ∂xi
=
n ∂d ∂bik
i=1 ∂xi bik + e ∂xi = i=1 ∂∂yk (dbik ) = i=1 ∂∂yk Aki , k = 1, . . . n.
n n
If y (x0 ) is singular, then the mapping y = y(x) + x is invertible. So by the first
n
∂
part of the lemma (Aji ) = 0 where Aji is the cofactor of (i, j)th element in
i=1
∂x i
n
∂ n
∂
J [y (x)]. By continuity A ji = lim A = 0. Thus in this case also
i=1
∂xi >0
i=1
∂xi ji
the Lemma is true.
Lemma 12.2.2 Suppose
(i) y : → is a function with y ∈ C 1 (), being an open bounded subset of
Rn , y is continuous on and for some > 0 y(x) > for all x ∈ ∂;
(ii) ϕ : [0, ∞) ∞→ R is continuous, vanishing in [, ∞) and also in a neighbourhood
of 0 and 0 r n−1 ϕ(r )dr = 0.
12.2 Heinz’s Elementary Analytic Theory of Mapping … 279
Then ϕ(y(x))J [y(x)]d x = 0.
n
∂
n n
∂ fj
Ai j (x) f j (y(x)) = J [y(x)]
i=1
∂xi j=1 j=1
∂yj y=y(x)
= J [(y(x)](r ψ (r ) + nψ(r ))r =y(x)
= ϕ(y(x))J [y(x)]. (12.2.2)
div F(x)d x = 0
Rn
ϕ(y(x))J [y(x)]d x = 0
Rn
Proof Let D be the linear space of all continuous real-valued functions satisfying
(i). Define L , M, N : D → R by
∞
L = r n−1 (r )dr
0
M = (x)d x
Rn
Proof Clearly, d[yi (x); , z] is well-defined and equals d[yi (x) − z; , 0] for
each i = 1, 2. Thus without loss of generality it may be assumed that z = 0.
Let f : [0, ∞) → R be a continuously differentiable function which vanishes in
[3, ∞) and is 1 in [0, 2] with range in [0, 1]. Define y3 : → Rn by y3 (x) =
[1 − f (y1 (x))]y1 (x) + f (y1 (x))y2 (x). Since 0 ∈ / , y3 ∈ C 1 () as f, y1 , y2
are C -functions. Further y3 ∈ C(). Clearly, yi (x) − y j (x)| < for x ∈ , for
1
i, j ∈ {1, 2, 3}. From the definition of f , for y1 (x) > 3, y3 (x) = y1 (x) and for
||y1 (x)|| < 2, y3 (x) = y2 (x).
Let i : [0, ∞) → R be be two continuous functions that vanish in a neighbour-
hood of 0 satisfying
i (x)d x = 1 for i = 1, 2.
Rn
12.2 Heinz’s Elementary Analytic Theory of Mapping … 281
and
2 (y3 (x))J [y3 (x)] = 2 (y2 (x))J [y2 (x)].
and
d[y3 (x); , 0] = d[y2 (x); , 0]
Proof Be Lemma 12.2.5, {d(yk (x); , z)} is a Cauchy sequence of real num-
bers and hence is convergent. (Indeed it is constant after some stage.) Again
as sup{y(x) − yk (x) : x ∈ } can be made less than 10 1
sup{|y(x) − z : x ∈
∂}. d(y(x); , z) = d(yk (x); , z) after some stage. Thus d(y(x); , z) = lim
k→∞
d(yk (x); , z) for z ∈
/ {y(x); x ∈ ∂}.
Remark 12.2.7 For y ∈ C() where is a nonvoid bounded open set and y() ⊆
Rn , for z ∈
/ {y(x); x ∈ ∂}, d[y(x); , z] is well-defined. Indeed by Weierstrass
approximation theorem there is a sequence of polynomials (yk ) on such that
(yk ) converges uniformly on to y and yk (x) = z for all x ∈ ∂ for all k. So
d[y(x); , z) is well-defined for all y ∈ C() with z ∈ / y(∂) and is called the
degree of y.
Remark 12.2.8 For a linear operator A : Rn → Rn , = B(0, 1) the unit open ball.
⎧
⎪
⎨+1, if det of A > 0
d[A, = B(0, 1), 0] = −1, if det of A < 0
⎪
⎩
not defined if det of A = 0
282 12 Basic Analytic Degree Theory of a Mapping
we can choose (yk ) and ( ŷ j ) such that for k, j ≥ m 0 , inf{yk (x) − z : x ∈ ∂} and
inf{ ŷ j (x) − z : x ∈ ∂} > . So by Lemma 12.2.5, d[ ŷ j (x); , z] = d[yk (x);
, z] for all j, k ≥ m 0 . Proceeding to the limit we get d[ ŷ(x); , z] = d[y(x); , z]
by Lemma 12.2.6.
Remark 12.3.7 The above corollary implies that the degree of a mapping depends
only on its value on the boundary of .
The following theorem shows that the degree remains the same under certain
translates.
The following theorem is the first step in relating the degree of a mapping to the
solution of (non-linear) equations.
m
d[y(x); , z] = (y(x) − z)J [y(x)]d x
i=1 Wi
m
= φ(yi (x) − z)J [yi (x)]d x
i=1 yi−1 (K )
m
= sgn det J [yi (x)] φ(x)d x
i=1 K
m
= sgn det J [yi (x)]
i=1
as φ(x)d x = Rn φ(x)d x = 1.
K
Thus in this case degree of the mapping y at z with respect to is an integer.
Naturally one would like to know if the degree of the mapping y(x) at a point
remains an integer even if the Jacobian of the mapping in singular at some of the roots
of y(x) = z. Thanks to Sard’s theorem detailed below the answer to this question is
in the affirmative.
y(u) − (y(u) + y (x − y))∞ ≤ βx − u∞ ≤ β .
m
Proof By Sard’s Theorem 12.3.13 y(S (G)) has measure zero. So there exists a
sequence z k ∈
/ y(S (G)) ∪ y(∂G) with lim z k = z. By assumption there exists >
k→∞
0 such that z k ∈ B(z, ) for all k ≥ k0 and the line segments joining y to yk for k ≥ k0
lie in B(z, ). So these paths do not meet y(∂G). So by Theorem 12.3.9 for k ≥ k0 .
d[y(x); G, z] = d[y(x); G, z k ].
Proof We can find a sequence of functions yk ∈ C 1 () such that yk converges uni-
formly in to y, with z ∈/ yk (∂). Clearly, d[yk (x); , z] is an integer for each k by
Theorem 12.3.15. Also we can find k0 such that d[yk (x); , z] = d[yk0 (x); , z] for
all k ≥ k0 . Proceeding to the limit as k tends to infinity and noting that d[yk0 (x); , z)
is an integer and lim d[yk (x); , z] = d[y(x); , z] it follows that d[y(x); , z]
k→∞
is an integer.
Proof Let be the unit open ball {x ∈ Rn : x < 1}. Suppose f (x) = x for
all x ∈ = B(0, 1). For each t ∈ [0, 1] and x ∈ define f (x, t) = x − t f (x).
Now f (x, t) = x − t f (x) ≥ x − t f (x) = 1 − t f (x) ≥ 1 − t > 0 for
x ∈ ∂ and t ∈ [0, 1). By assumption x = f (x) for x ∈ ∂. Thus f (x, t) = 0 for
all x ∈ ∂ and t ∈ [0, 1]. So by Theorem 12.3.4, d[ f (x, t); , 0] is a constant. Thus
d[ f (x, 1); , 0] = d[ f (x0 ); , 0] = 1 by definition of the degree. So by Theorem
12.4.1, the equation f (x) − x = 0 has a solution in , a contradiction. So f has a
fixed poin in .
It may be added by way of conclusion that Heinz had also presented Leray’s
product theorem for the degree of a mapping. This result can be used to deduce
important theorems of topology such as Brouwer’s domain invariance theorem. The
reader is referred to Lloyd [4] for further details.
References
1. Cronin, J.: Fixed Points and Topological Degree in Nonlinear Analysis. Surveys, vol. II. Amer-
ican Mathematical Society, Providence (1964)
2. Deimling, K.: Nichtlineare Gleichungen und Abbildungs grade. Springer, Berlin (1974)
3. Heinz, E.: An elementary analytic theory of the degree of mapping in n-dimensional space. J.
Math. Mech. 8, 231–247 (1959)
4. Lloyd, N.G.: Degree Theory. Cambridge University Press, Cambridge (1978)
5. Ortega, J.M., Rheinboldt, W.C.: Iterative Solution of Nonlinear Equations in Several Variables.
Academic, New York (1970)
6. Outerelo, E., Ruiz, J.M.: Mapping Degree Theory. Graduate Studies in Mathematics, vol. 108.
American Mathematical Society, Providence (2009)
7. Rothe, E.: Introduction to Various Aspects of Degree Theory in Banach Spaces. Mathematical
Surveys and Monographs, vol. 23. American Mathematical Society, Providence (1986)
8. Sard, A.: The measure of the critical values of differentiable maps. Bull. Am. Math. Soc. 48,
883–897 (1942)
Appendix A
A Counterexample on Common Fixed Points
Since every continuous self-map on a bounded closed interval of real numbers has a
fixed point, a natural question is whether two commuting continuous self-maps on a
closed bounded interval of real numbers have a common fixed point. This has been
answered in the negative, independently by Boyce [1] and Huneke [2]. In the sequel,
we highlight the solution provided by Huneke in Part II of his paper [2].
Let h : S ⊆ R → R be a map where S is a subset of the set of real numbers
R. Define h ∗ by h ∗ (x) = 1 − h(1 − x) for each x for which 1 − x ∈ S. Let b ∈
1 1
0, 2 and let s = 3−2b+(6−4b)
2
1−2b
. Clearly, s is well-defined and positive. Define three
continuous functions h 1 , h 2 and h 3 with linear graphs as follows:
(1 − b + sb)
h 1 : I1 = b, → [b, 1] by h 1 (x) = s(x − b) + b
s
(1 − b + sb) (2 − b + sb)
h 2 : I2 = , → [0, 1] by h 2 (x) = 2 − sx + sb − b
s s
(2 − b + sb) (3 − 2b + sb)
h 3 : I3 = , → [0, 1 − b] by h 3 (x) = −2 + sx − sb + b
s s
Define h : b, (3−2b+sb)
s
→ [0, 1] by h(x) = {h j (x), x ∈ I j , j = 1, 2, 3. Clearly, h
is continuous and has a linear graph and h j is invertible for each j = 1, 2, 3.
Definition A.1.1 Let Cb be the set of all continuous self-maps on [0, b] with b as a
fixed point. For each g ∈ Cb define g : [0, 1] → R the unique continuous extension
of g defined by
(1) g(x) = g(x) on [0, b];
(2) g(x) = h(x) on [b, h −1 3 (1 − b)];
(3) g(x) = (h ∗1 )−1 g(h ∗ (x)) on [h −1 ∗ −1 −1
3 (1 − b), (h 2 ) h 2 (0)];
∗ −1 ∗ ∗ −1 −1
(4) g(x) = (h 2 ) g(h (x)) on [(h 2 ) (h 2 (0)), 1 − b];
(5) g(x) = the fixed point of h ∗2 on [1 − b, 1].
Remark A.1.2 The verification that g defined above is a unique extension of g is left
as an exercise.
Remark A.1.3 If g ∈ Cb satisfies Lipschitz condition with Lipschitz constant s, then
g too satisfies Lipschitz condition with Lipschitz constant s.
Proof (attributed to David Boyd in [2]) Let g ∈ Cb satisfy Lipschitz condition with
Lipschitz constants and L be the set of all s-Lipschitz self-maps on [0, 1] satisfying
(1), (2) and (5) for g in Definition A.1.1.
Define T : L → L by
⎧
⎪
⎨ f (x) if x ∈ [0, h −13 (1 − b)] ∪ [1 − b, 1]
T f (x) = (h 1 ) f (h (x)), if x ∈ [h 3 (1 − b), (h ∗2 )−1 h −1
∗ −1 ∗ −1
2 (0)]
⎪
⎩ ∗ −1
(h 2 ) f (h ∗ (x)), if x ∈ [(h ∗2 )(h −1
2 (0)), 1 − b]
Since (h ∗1 )−1 and (h ∗2 )−1 have linear graphs and 1s Lipschitzian, (h ∗1 )−1 (0) = (h ∗2 )−1
(s−1)(1−b)
(0), h ∗ is s-Lipschitzian and the fixed points of h ∗ are h −1 3 (1 − b), s+1
and
1 − b. Clearly, L is a complete metric space with the supremum metric on which T
is a contraction with Lipschitz constant 1s < 1. This follows from the fact that (h ∗1 )−1
and (h ∗2 )−1 are contractions with Lipschitz constant 1s . So, there is unique function
g in L with T (g) = g and it can be seen that g satisfies the conditions of Definition
A.1.1.
1
Lemma A.1.4 Let f, g ∈ Cb and x ∈ 2 , 1 . Then
(i) ( f )∗ x = x implies g(x) = x and
(ii) ( f )∗ (g(x)) = g(( f )∗ (x)).
Proof (i) The domain of h ∗ is [(h ∗3 )−1 (b), 1 − b] and (h ∗3 )−1 (b) = − 3+2b+s−sb
s
=
(1−2b)(2b−3) ∗
1 − b + 3−2b+(6−4b)1/2 < 2 as 0 ≤ b < 2 . So, the fixed points of ( f ) are in [1 − b, 1]
1 1
Thus, we have
Proposition A.1.5 For any f and g in Cb , f and (g)∗ are commuting functions
without a common fixed point.
References
1. Boyce, W.M.: Commuting functions with no common fixed point. Trans. Am. Math. Soc. 137,
77–92 (1969)
2. Huneke, J.P.: On common fixed points of commuting continuous functions on an interval. Trans.
Am. Math. Soc. 139, 371–381 (1969)
Appendix B
A Compact Contractible Space Without Fixed
Point Property
B.1 Introduction
A natural question is whether a compact contractible space has the fixed point prop-
erty for continuous functions. This question was raised in 1932 by Borsuk [1] and
was settled in the negative by Kinoshita [2] in 1953. In the following, Kinoshita’s
counterexample is described.
Recall that a topological space X is called contractible if the identity map on X
is null-homotopic (i.e. homotopic to a constant).
f maps (0, 0, 0) onto (0, 0, 1), the circle r = π2 arctan(π ) to (0, 0, 0) and the interior
of this circle into the segment r = 0, z ∈ [0, 1]. The annulus of the circle is rotated
π radians while the inner boundary is contracted to (0, 0, 0). f has no fixed points
in A1 .
For (r, θ, z) ∈ A2 , since r = 1
(1, (θ − π + 2π z) mod 2π, z + 2z , z ∈ 0, 21
f (1, θ, z) =
(1, (θ − π + 2π z) mod 2π, 21 + 2z , z ∈ 21 , 1 .
References
1. Borsuk, K.: Uber eine Klasse von lokal zusammen hängenden Räumen. Fund. Math. 19, 220–242
(1932)
2. Kinoshita, S.: On some contractible continua without fixed point property. Fund. Math. 40,
96–98 (1953)
Appendix C
Fractals via Fixed Points
C.1 Introduction
Fractals may intuitively be understood as highly irregular non-smooth sets with ‘non-
integral dimension’ arising often from a recursive process of construction displaying
self-similarity. Some of these fractals can be realized as fixed points of set-functions.
(see Hutchinson [2]).
Mandelbrot has pointed out how fractals can be used to model several physical
phenomena. Falconer [1] treats interesting aspects of fractals from a geometric point
of view.
∞ ∞
Hδs (E) = inf (diam Ui ) : E ⊆
s
Ui and 0 < diam Ui < δ .
i=1 i=1
Remark C.2.2 One can show that H s (E) = sup Hδs (E) for any E ⊆ Rv . Further
δ>0
when H s in restricted to the σ -algebra of H s -measurable subsets of Rn , it is a
measure called s-dimensional Hausdorff measure.
Remark C.2.4 For each s ≥ 0 H s is a regular outer measure. Thus all Borel subsets
of Rn are H s -measurable.
Definition C.2.5 Let E ⊆ Rn . The unique real number, dim H E is called the Haus-
dorff dimension of E provided
∞ for 0 ≤ s < dim H (E)
H s (E) =
0 for dim H (E) < s < ∞.
Using the concept of iterated function systems (IFS), a large class of fractal sets
can be constructed with explicit computation of their Hausdorff dimension.
The theorem below stated without proof is used in the existence theorem.
Theorem C.2.6 Let (X, d) be a complete metric space. K (X ) the set of all nonempty
compact subsets of X is a complete metric space under the Hausdorff metric (vide
Definition 5.3.1).
Remark C.2.7 If (An ) ∈ K (X ) is a Cauchy sequence with the Hausdorff metric then
it converges to lim sup(An ) in K (X ).
≤ max ci H (A, B)
i=1,...,m
m
m
Proof Clearly for the function t → cit , there is a unique s with cis = 1. For
i=1 i=1
Ik = {(i 1 , . . . , i k ) : i k ∈ {1, . . . , m}} and any set B and a given element (i
1 , . . . , ik )
of Ik , define Bi1 ...ik = Si1 ◦ · · · ◦ Sik (B). As F is fixed under the IFS, F = Fi1 ...ik .
Ik
So {Fi1 ...ik } cover F. Since Si ◦ · · · ◦ Sk is a contraction similarity with constant
ci1 . . . cik ,
(diam Fi1 ...ik )s = (ci1 . . . cik )s (diam F)s
Ik
= cis1 . . . cisk (diam F)s
m
= (diam F)s since cis = 1.
i=1
For δ > 0, we can choose k such that (max ci )k diam F ≤ δ. Since diam Fi1 ...ik ≤
(maxi ci )k diam F the sets {F i1 ...ik } cover F with diameter ≤ δ, for sequences
(i 1 , . . . , i k ) in Ik . So H s (F) ≤ Ik (diam(Fi1 . . . Fik ))s .
Consider I the set of all sequences (i n ) such that i k ∈ {1, . . . , m} for all k and let
Ii1 ...ik be the subset of I consisting of only those sequences starting with i 1 , . . . , i k .
∞
Define μ(Ii1 ,...,ik ) = (ci1 . . . cik )s . Define xi1 ,i2 ,... = Fi1 ...ik . For a subset A of F
k=1
define μ(A) = μ{Ii1 ...ik : xi1 ...,ik , · · · ∈ A}. Now μ(F) = μ(I ) = 1 as F contains
every point xi1 ,i2 ,... for each (i 1 , i 2 , . . . ) ∈ I .
Since the IFS satisfies an open set condition, there is an open set V with V ⊃
m m
Si (V ), where Si (V ) ∩ S j (V ) = φ for i = j. So S(V ) = Si (V ) and {S k (V )}
i=1 i=1
converges to F. So F ⊆ V i1 ,...,ik for any sequence {i 1 , . . . , i k }. For a ball B of radius
0 < r < 1, each sequence in I is truncated at the first i k for which
min ci r ≤ ci1 . . . cik ≤ r
1≤i≤m
radius (mini ci r ) α and is contained in a ball of radius rβ. Let J be be the subset
of I consisting of only those sequences (i 1 , . . . , i k ) such that V i1 ...ik intersects B.
By Proposition C.3.5 there exist at most γ = (1 + 2β)n (mini ci α)−n elements in J .
Now
μ(B) = μ(F ∩ B)
= μ({Ii1 ,...,ik : xi1 ,...,ik ... ∈ F ∩ B})
= μ({Ii1 ,...,ik : xi1 ,...,ik ... ∈ ( V i1 ...ik ) ∩ B}).
J
So
μ(B) ≤ μ({Ii1 ,...,ik : xi1 ,...,ik ... ∈ V i1 ...ik })
J
= μ( Ii1 ,...ik )
J
< μ(Ii1 . . . Iik )
J
≤ (ci1 . . . cik )s ≤ r s = qr s .
J J
References
1. Falconer, K.J.: The geometry of fractal sets, Cambridge University Press (1962)
2. Hutchinson, J.E.: Fractals and self-similarity, Indiana University of Mathematics Jl., 30,
713–747 (1981)
302 Appendix C: Fractals via Fixed Points
Postscript
Apart from the references cited in each chapter, the following books may be referred for further
study.
1. Aksoy, A., Khamsi, M.A.: Nonstandard Methods in Fixed Point Theory. Springer, Berlin (1990).
2. Agarwal, R.P., Meehan, M., Oregan, D.: Fixed Point Theory and Applications. Cambridge
University Press, Cambridge (2001).
3. Bonsall, F.F.: Lectures on some fixed point theorems of functional analysis. Tata Institute,
Bombay (1962).
4. Brown, R.F.: The Letschetz Fixed Point Theorem. Scott Foreman & Co, London (1971).
5. Cronin, J.: Fixed Points and Topological Degree in Nonlinear Analysis. Survey, vol. 11. Amer-
ican Mathematical Society, New York (1964).
6. Deimling, K.: Nichtlineare Gleichungen und Abbildungsgrade. Hochschult Text. Springer,
Berlin (1974).
7. Deimling, K.: Nonlinear Functional Analysis. Springer, Berlin (1985).
8. Dugundji, J., Granas, A.: Fixed Point Theory. PWN-Polish Scientific Publishers, Warsaw
(1982).
9. Goebel, K., Kirk, W.A.: Topics in Metric Fixed Point Theory. Cambridge University Press,
Cambridge (1990).
10. Hadzic, O.: Fixed point theory in topological vector spaces. University of Novisad (1964).
11. Istratescu, V.I.: Fixed point theory. Reidel of Dordrecht (1981).
12. Khamsi, M.A., Kirk, W.A.: An Introduction to Metric Spaces and Fixed Point Theory. Wiley,
New York (2001).
13. Kirk, W., Shahzad, N.: Fixed Point Theory in Distance Spaces. Springer, Berlin (2014).
14. Kirk, W., Sims, B. (eds.) Handbook of Metric Fixed Point Theory. Springer, Netherlands (2001).
15. Rothe, E.H.: Introduction to Various Aspects of Degree Theory in Banach Spaces. Mathematical
Surveys and Monographs, vol. 23. American Mathematical Society, Providence (1986).
16. Smart, D.R.: Fixed Point Theorems. Cambridge University Press, Cambridge (1994).
17. Van der Walt, T.: Fixed and Almost Fixed Points. Mathematical Centrum, Amsterdam (1963).
18. Zeidler, E.: Nonlinear Functional Analysis and Its Applications I: Fixed Point Theorems.
Springer, Berlin (1986).