Fixed Point Constructive Mmath
Fixed Point Constructive Mmath
1
ISSN 1759-9008
Abstract: This paper gives the beginnings of a development of the theory of fixed
point theorems within Bishop’s constructive analysis. We begin with a construc-
tive proof of a result, due to Borwein, which characterises when some sets have the
contraction mapping property. A review of the constructive content of Brouwer’s
fixed point theorem follows, before we turn our attention to Schauder’s general-
isation of Brouwer’s fixed point theorem. As an application of our constructive
Schauder’s fixed point theorem we give an approximate version of Peano’s theorem
on the existence of solutions of differential equations. Other fixed point theorems
are mentioned in passing.
1 Introduction
Fixed-point theorems are a major tool in both functional analysis and mathematical
economics1 and are used to prove the existence of solutions to differential equations
and the existence of Nash equilibria among other things. Despite this, the constructive
literature on fixed point theorems has been scant2 . There are (at least) two reasons for
this:
(i) The standard proof of the simplest and most useful of the well known fixed point
theorems, the Banach fixed point theorem, is essentially constructive.
(ii) The non-constructive nature of Brouwer’s fixed point theorem, and the subse-
quent rejection of this theorem by Brouwer, is well known, and a constructive
approximate version for simplices (via Sperner’s lemma) is part of the folklore.
1
The main fixed point theorem used in mathematical economics, Kakutani’s fixed point
theorem, is given a constructive treatment in [16].
2
Although [9] gives a Bishop-style constructive treatment of Edelstein’s fixed point theorem;
and Kohlenbach [18, Chapter 18] examines contractive and nonexpansive fixed point theorems
for computational content using tools from proof theory.
Only recently has a fully constructive proof of the approximate version of Brouwer’s
fixed point theorem, for simplices, been presented [23].
This paper is broken up into three sections. In the first we consider the fixed point
theorems of Banach and Brouwer; we prove that spaces with a strong connectedness
property are complete if and only if every contraction mapping has a fixed point, and
we give an approximate version of Brouwer’s fixed point theorem for uniformly se-
quentially continuous functions on totally bounded subsets of Rn with convex closures.
The second section presents a constructive treatment of Schauder’s fixed point theorem
and its extension due to Rothe. In the final section we give an application of our con-
structive Schuader’s fixed point theorem: we prove an approximate version of Peano’s
theorem on the existence of solutions of differential equations.
The simplest and most useful of the standard fixed point theorems is the Banach fixed
point theorem: every contraction mapping of a complete metric space into itself has a
fixed point. The standard proof of this theorem (see [22] or the first part of the proof
of Theorem 1) is fully constructive.
Conversely, suppose that S has the contraction mapping property, let (xn )n>1 be a
Cauchy sequence in S, and let x̂ be the limit of (xn )n>1 in the completion (Ŝ, ρ̂) of
(S, ρ). Without loss of generality we may take
for all s, t ∈ [0, 1]. Using the gluing lemma [8], define a mapping
[
g : {0} ∪ [2−(k+1) , 2−k ] ∪ (1, ∞) → S ∪ {x̂}
k>1
by
x̂ if t = 0
g(t) ≡ gk (2k+1 t − 1) if t ∈ [2−(k+1) , 2−k ]
x1 if t > 1.
Suppose that there exist t1 , t2 with t2 < t1 and ρ(g(t1 ), g(t2 )) > L|t1 −t2 |. By continuity
we may assume, without loss of generality that ti ∈ [2−ni , 2−ni +1 ] (i = 1, 2); set
s0 = t1 , sn1 −n2 = t2 , and sk = 2−n2 −1+k for 1 6 k 6 n1 − n2 − 1. Then
1 −n2
nX
ρ(g(sk−1 ), g(sk )) > ρ(g(t1 ), g(t2 ))
k=1
> L|t1 − t2 |
1 −n2
nX
= |sk−1 − sk |,
k=1
—a contradiction. Hence
for all s, t in the domain of g. It follows that g is uniformly continuous on its domain,
and so extends to a uniformly continuous mapping g : [0, ∞) → S ∪ {b x} such that (1)
holds for all s, t ∈ [0, ∞).
For completeness, we give a constructive proof of the approximate Brouwer fixed point
theorem, extended to uniformly sequentially continuous functions; for novelty we give
a proof based on David Gale’s proof from [15]. Before we do this we require a few
more definitions.
Let X be a metric space and let f be a function from X into X . If ρ(x, f (x)) < ε, then x
is called an ε-fixed point. A function f : X → X has approximate fixed points if for each
ε > 0 there exists an ε-fixed point of f in X . If every uniformly continuous function
from X into X has approximate fixed points, then X is said to have the approximate
fixed point property.
Gale’s proof of Brouwer’s fixed point theorem uses a generalisation of the game of
Hex. An n-dimensional Hex board of size k consists of vertices V = {1, . . . , k}n with
edges between two vertices x, y ∈ V if5 kx − yk = 1 and either xi 6 yi for each i or
yi 6 xi for each i. Then n-dimensional Hex is an n-player game where players take
turns to pick unclaimed vertices. A player gains an edge of the hex board if she owns
the nodes at either end; player i wins the game by connecting the two i-banks
i-bank 1 = {(v1 , . . . , vn ) ∈ V : vi = 0},
i-bank 2 = {(v1 , . . . , vn ) ∈ V : vi = k},
with her edges. The ‘Hex Theorem’ of [15] (which, being finitely combinatorial, is
fully constructive) says that any colouring of an n-dimensional Hex board with at most
n colours has a winner (for n > 2 there may be more than one).
5
We use k · k throughout to represent either the maximum norm or the supremum norm.
Lemma 2 Let f be a function from the unit hypercube [0, 1]n into itself. Then for
n
all ε, δ > 0 either there exists x ∈ [0, 1] such that ρ(x, f (x)) < ε or there exist
0
x, x ∈ [0, 1] such that ρ x, x < δ and ρ(f (x), f (x0 )) > ε.
n 0
Proof Write
f (x) = (f1 (x), . . . , fn (x)) ;
Fixing ε, δ > 0, without loss of generality take δ < ε/3. Pick N > 0 such that
1/N < δ , and subdivide [0, 1]n into an n-dimensional Hex board of size N . We
partition the set V of vertices of this Hex board into sets C1+ , C1− , . . . , Cn+ , Cn− , and B
such that
2ε
x ∈ C1+ ⇒ f1 (x) − x1 > ;
3
2ε
x ∈ C1− ⇒ x1 − f1 (x) > ;
3
..
.
+ 2ε
x ∈ Cn ⇒ fn (x) − xn > ;
3
2ε
x ∈ Cn− ⇒ xn − fn (x) > ;
3
x ∈ B ⇒ kf (x) − xk < ε.
By the Hex theorem, either B is inhabited, and there exists x ∈ [0, 1] such that
ρ(x, f (x)) < ε, or, as we may assume, there exists an i-path from i-bank 1 to i-bank
2 for some 1 6 i 6 n. Since no vertex of Ci+ has i-th coordinate 1 and no vertex of
Ci− has i-th coordinate 0, such a path contains points from each set. Hence there exist
adjacent vertices x, x0 such that x ∈ Ci+ and x0 ∈ Ci− . Then kx − x0 k < δ < ε/3,
fi (x) > fi (x0 ), and
fi (x) − fi (x0 ) = (fi (x) − x) + x − x0 + x0 − fi (x0 )
2ε ε 2ε
> − + = ε.
3 3 3
Therefore ρ(f (x), f (x0 )) > |fi (x) − fi (x0 )| > ε.
With this lemma at hand we can weaken the standard hypothesis of the approximate
Brouwer fixed point theorem; we only require that f : [0, 1]n → [0, 1]n be uniformly
sequentially continuous6 : for all sequences (xn )n>1 , (yn )n>1 in [0, 1]n , if ρ (xn , yn )
6
Throughout this paper, uniformly sequentially continuous can be substituted for uniformly
continuous in the definition of the approximate fixed point property.
Proof We construct, using countable choice, sequences (xn )n>1 , xn0 n>1 as follows.
For each n ∈ N, apply Lemma 2 to construct either x ∈ [0, 1]n such that ρ(x, f (x)) < ε
or x, x0 ∈ [0, 1]n such that ρ(x, x0 ) < 1/n and ρ(f (x), f (x0 )) > ε. In the lattercase
we set xn = x and xn0 = x0 ; in the former we set xn = xn0 = x. Then ρ xn , xn0 n>1
converges to zero. Sincef is uniformly sequentially continuous, there exists N ∈ N
such that ρ f (xn ) , f xn0 < ε for all n > N . It follows that ρ (xN , f (xN )) < ε.
Next we extend the approximate Brouwer fixed point theorem, for uniformly continuous
functions, to compact convex subsets of Rn . A subset S of a normed space X is strictly
convex if for each ε > 0there exists δ > 0 such that for all x, y in the boundary ∂S
of S, if7 ρ 12 (x − y), ∂S < δ , then kx − yk < ε. A normed space X is uniformly
convex if its unit ball is strictly convex: for all ε > 0 there exists δ > 0 such that for
all x, y ∈ X with kxk = kyk = 1, if 21 (x − y) > 1 − δ , then kx − yk < ε. Any inner
product space is uniformly convex [12, Page 93], and the Lp spaces for 1 < p < ∞
are uniformly convex [4, Chapter 7, (3.22)].
Proof The proof of Theorem 6 in [11] establishes that for each x ∈ X and each ε > 0
the set
Sεx = {y ∈ S : ρ(x, y) < ρ(x, S) + ε}
7
We do not require ∂S to be located here: for an arbitrary subset S of a metric space X we
use ‘ρ(x, S) < ε’ as a shorthand for ‘there exists s ∈ S with ρ(x, s) < ε. If S is located, then
this coincides with the standard meaning.
has diameter no greater than ε, and hence that x has a unique best approximation in S.
y
To see that Q is uniformly continuous, observe that if kx − yk < ε/2, then Sε/2 ⊂ Sεx .
x
Hence Q(x), Q(y) ∈ Sε , so kQ(x) − Q(y)k 6 ε.
We call the mapping Q from the preceding theorem the projection onto S.
Theorem 5 Every totally bounded set S of Rn with convex closure has the approxi-
mate fixed point property.
Proof Let S be a subset of Rn satisfying the conditions of the theorem; without loss
of generality we may assume that S is both closed and a subset of the unit cube [0, 1]n .
Fix ε > 0 and let Q be the projection mapping from [0, 1]n onto S (which exists, by
the preceding theorem). Applying Theorem 3 to the mapping f ◦ Q : [0, 1]n → [0, 1]n ,
construct x ∈ [0, 1]n such that kx − f ◦ Q(x)k < ε/2. Then
kx − Q(x)k = ρ(x, S)
ε
6 kx − f ◦ Q(x)k < ,
2
so
kQ(x) − f (Q(x))k 6 kQ(x) − xk + kx − f ◦ Q(x)k
ε ε
< + = ε.
2 2
Hence Q(x) is an ε-fixed point of f .
Proof Let f be a uniformly continuous function from T into itself, and fix ε > 0. Let
δ > 0 be such that for all x, y ∈ T , if kx − yk < δ , then
fε/2 (x) − fε/2 (y) < ε/2,
where fε/2 is as in the statement of the proposition. Let Q be the projection onto
fε/2 (S) restricted to T , and let I : fε/2 (S) → T be the inclusion mapping; note that
kQ(t) − tk < ε/2 for all t ∈ T .
−1
Then fε/2 ◦ Q ◦ f ◦ I ◦ fε/2 is a uniformly continuous function from S into S. Hence
there exists x ∈ S such that
−1
kfε/2 ◦ Q ◦ f ◦ fε/2 (x) − xk < δ.
Then
kQ ◦ f ◦ fε/2 (x) − fε/2 (x)k < ε/2,
and so
f fε/2 (x) − fε/2 (x) 6 f ◦ fε/2 (x) − Q ◦ f ◦ fε/2 (x)
+ Q ◦ f ◦ fε/2 (x) − fε/2 (x)
ε ε
6 + = ε.
2 2
−1
Thus Q fε/2 (x) is an ε-fixed point of f .
The Brouwer fixed point theorem for uniformly continuous functions is equivalent to
the essentially nonconstructive lesser limited principle of omniscience,
LLPO: For each binary sequence (an )n>1 , either an = 0 for all even n or
an = 0 for all odd n.
To obtain the usual conclusion of the intermediate value theorem, it suffices to assume
that f : R → R is locally nonzero: for all x ∈ R and all ε > 0 there exists
y ∈ (x − ε, x + ε) such that f (y) 6= 0. The equivalent notion for Brouwer’s fixed point
theorem, that f : X → X is locally John Doe—for all x ∈ X and all ε > 0 there
exists y ∈ B(x, ε) such that f (y) 6= y—is not sufficient to prove Brouwer’s fixed point
theorem in higher dimensions: Orevkov [20] has given an example of a continuous
function from [0, 1]2 into itself with no fixed point. Further, Veldman [23] has shown
that Brouwer’s fixed point theorem for continuous functions with at most one fixed
point is equivalent to Brouwer’s fan theorem for detachable bars (see also [3]). (We
say that f has at most one fixed point if
max{ρ(x, f (x)), ρ(y, f (y))} > 0
whenever x 6= y.)
The standard Brouwerian example showing that Brouwer’s fixed point theorem implies
LLPO also shows that we cannot prove constructively that every nonexpansive mapping
on [0, 1] has the fixed point property.9 Using a standard classical argument (see for
example [22]), we can, however, show that such a mapping has approximate fixed
points.
Proof Since we are only interested in approximate fixed points we may replace X by
its completion X̂ , and S by its closure in X̂ ; we may also assume that 0 ∈ S. Fix ε > 0
and let N > 0 be such that S is contained in the ball of radius N centered on 0. Pick
9
Classically, every continuous nonexpansive mapping on a bounded closed subset of a
uniformly convex Banach space has the fixed point property. The standard classical proof for
Hilbert spaces (see [22]) requires the statement that ‘every convex bounded closed subset of
a Hilbert space is weakly compact’, which implies LPO. As a consequence it seems likely
that the most general formulation of the nonexpansive fixed point theorem is not equivalent to
LLPO; however, if we restrict ourselves to weakly compact subsets of a Hilbert space, then
the nonexpansive fixed point theorem follows from MIN.
r ∈ (1 − ε/N, 1), and let x be the unique fixed point of the contraction mapping rf .
Then
kf (x) − xk = kf (x) − rf (x)k = (1 − r)kf (x)k 6 (1 − r)N < ε.
In this section we extend Brouwer’s fixed point theorem by considering compact convex
subsets of arbitrary Banach spaces; this gives us Schauder’s fixed point theorem.
Lemma 8 Let S be a totally bounded subset of a metric space X , fix β > α > 0,
and let S0 be a convex set such that for each x ∈ S there exists x0 ∈ S0 such that
ρ(x, x0 ) < α/2. Then there exists a uniformly continuous function P from S into S0
such that kP(x) − xk < β for all x ∈ S.
fi (x) > γ − α;
whence Pn
fi (x)xi0
P(x) ≡ Pi=1
n
i=1 fi (x)
Let r > 0 and write {1, . . . , n} as the disjoint union of two sets P, Q such that
i ∈ P ⇒ kx − x0 k < γ + r;
i ∈ Q ⇒ kx − x0 k > γ.
for all i ∈ {1, . . . , n}. For each such i pick xi0 ∈ V such that kxi − xi0 k < ε/8. Then
for each x ∈ S, there exists i ∈ {1, . . . , n} such that kx − xi0 k < ε/4. Let S be the
closed convex hull of {x10 , . . . , xn0 }, and let Q : S0 → S be the restriction to S0 of the
projection onto S. If ni=1 λi = 1 and each λi ≥ 0, then
P
n
! n n
X X X
Q λi xi0 − λi xi0 ≤ λi Qxi0 − xi0
i=1 i=1 i=1
Xn
λi ρ xi0 , S
=
i=1
n
X ε
≤ λi xi0 − xi < ;
8
i=1
Proof Since we are interested in approximate fixed points, replacing X with its
completion X̂ and S with its closure in X̂ , we may assume that S is compact and
convex. The result then follows from Theorems 4 and 9.
Strictly convex sets are also projective; the proof is similar to that of Theorem 4 (which,
in turn, is based on the proof of Theorem 6 of [11]).
Proof Let x be a point of X , fix ε > 0, and let δ ∈ (0, ε/2) be such that for all
x, y ∈ ∂S, if ρ 21 (x − y), ∂S < 2δ , then kx − yk < ε/2. Set
Hence the diameter of Sεx is no greater than ε. The proof then proceeds as in Theorem
4.
Theorem 13 Every inhabited, open, totally bounded, convex subset of a normed space
has the approximate fixed points property.
Proof The proof is similar to that of the preceding theorem. Let S be an inhabited,
open, totally bounded, convex subset of a normed space X , and let f : S → S be a
uniformly continuous function. Let {x1 , . . . , xn } be an ε/6-approximation to S. We
construct, as follows, a finite-dimensional subspace V of X such that V contains an
ε/3-approximation {x10 , . . . , xn0 } to S. Let V1 = span{x1 } and x10 = x1 . Suppose that
we have constructed Vk−1 and x10 , . . . , xk0 and let r ∈ (0, ε/6) be such that B(xk , r) ⊂ S.
Either ρ(xk , Vk−1 ) > 0 or ρ(xk , Vk−1 ) < r. In the first case we set
Vk = span{Vk−1 , xk }
and xk = xk0 . In the second case, pick xk0 ∈ V such that kxk − xk0 k < r and set
Vk = Vk−1 . Set V = Vn ; it is easy to see that {x10 , . . . , xn0 } is an ε/3-approximation to
S.
Let S0 be the convex hull of {x10 , . . . , xn0 }. Then S0 ⊂ S and, by Lemma 8, there exists
a uniformly continuous function P : S → S0 such that kP(x) − xk < ε/2 for all x ∈ S.
Using Brouwer’s fixed point theorem, applied to P ◦ f |S0 : S0 → S0 , construct x ∈ S0
such that kP ◦ f (x) − xk < ε/2. Then
kf (x) − xk 6 kf (x) − P ◦ f (x)k + kP ◦ f (x) − xk
< ε/2 + ε/2 = ε.
Hence x is an ε-fixed point of f .
We can extend Theorem 9 to give an approximate version of Rothe’s theorem [21, 22]
for projective sets.
Proof Fix ε > 0, let Q be the projection onto S, and let δ ∈ (0, ε/4) be such that if
kx − yk < δ , then kf (x) − f (y)k < ε. Since Q ◦ f is a uniformly continuous function
from S into S, it follows from Schauder’s fixed point theorem for projective sets that
there exists x ∈ S such that kQ ◦ f (x) − xk < δ . Suppose that ρ(f (x), S) > ε/4.
Thenf (x) ∈
/ S and ρ(x, y) > δ for all y ∈ ∂S—this contradicts that kQ ◦ f (x) − xk < δ .
Since Schauder’s fixed point theorem implies Brouwer’s fixed point theorem and fol-
lows from MIN, it is equivalent to LLPO.
4 An application
We give an application of the approximate Schauder fixed point theorem for uniformly
convex spaces (Corollary 10). A standard application of Schauder’s fixed point the-
orem is in proving Peano’s Theorem asserting the existence of solutions to particular
differential equations:
Peano’s Theorem Let A be a closed subset of R, let (x0 , y0 ) ∈ A, and let
r > 0 be such that if |x − x0 | 6 r and |y − y0 | 6 r, then (x, y) ∈ A. Let
f : A → R be continuous, let
M > sup {|f (x, y)| : |x − x0 | 6 r, |y − y0 | 6 r} ,
and set h = min {r, r/M}. Then the differential equation
(2) y0 = f (x, y), y (x0 ) = y0
has a solution y on the interval [x0 − h, x0 + h].
However, since the exact version of Peano’s Theorem is constructively equivalent to
LLPO (see [6], which also gives an alternative constructive proof of an approximate
Peano’s Theorem, [2] gives a proof that Peono’s Theorem implies LLPO), we can only
hope to prove an approximate version of Peano’s Theorem.
There is another, more pressing, problem: Peano’s Theorem asserts the existence
of solutions to particular differential equations in the normed space C(I), for some
interval I , with the supremum norm, but this normed space is not uniformly convex.
To overcome this difficulty, we first approximate the sup norm with a uniformly convex
norm—this relies on being able to restrict the possible solutions of (2) to a sufficiently
friendly subset of C(I).
Lemma 15 If S is a bounded Lipschitz subset of C(I), then for each ε > 0 there
exists p > 1 such that | kyk − kykp | < ε for all y ∈ S.
Proof Fix ε > 0 and let N be a bound for S and M be a Lipschitz constant for S. It
suffices to choose p > 1 such that |kyk − kykp | < ε, where y is given by
4N 2 a+b
y(x) = max M, 1− x− − N.
b−a b−a 2
This is possible because, in C([−1, 1]),
Z 1 1/p
p
k1 − |x|kp = |1 − |x|| dt
−1
Z 1 1/p
p
= 2 |1 − x| dt
0
1/p
2
= −→ 1,
p+1
as p → ∞.
References
[17] H Ishihara, An omniscience principle, the König lemma and the Hahn-
Banach theorem, Z. Math. Logik Grundlagen Math. 36(1990), 237 – 240; doi:
10.1002/malq.19900360307.
[18] U Kohlenbach, Applied Proof Theory: Proof Interpretations and their Use in Mathe-
matics, Springer Monographs in Mathematics, 2008.
[19] J Myhill, Constructive set theory, J. Symbolic Logic 40(3) (1975), 347–382; doi:
10.2307/2272159.
[20] V P Orevkov, A constructive map of the square into itself, which moves every con-
structive point, Dokl. Akad. Nauk SSSR 152(1963), 55–58.
[21] E Rothe, Zur Theorie der topologischen Ordnung und der Vektorfelder in Banachschen
Räumen, Compos. Math. 5(1937), 177-196.
[22] D R Smart, Fixed Point Theorems, Cambridge Univ.Press, 1974.
[23] W Veldman, Brouwer’s approximate fixed-point theorem is equivalent to Brouwer’s fan
theorem, in Logicism, Intuitionism, and Formalism, Part II, Synth. Libr., 341, Springer,
Dordrecht, 2009, 277–299,; doi: 10.1007/978-1-4020-8926-8 14.