Sei sulla pagina 1di 3

THE IMPLICIT FUNCTION THEOREM

1. The contraction mapping fixed-point theorem


Let (X, ρ) be a non-empty complete metric space and K : X → X be a contrac-
tion mapping. That is, there exists a constant C < 1 such that for all x, y ∈ X,
ρ(K(x), K(y)) ≤ Cρ(K(x), K(y)). Then it is a classical result that K has a unique
fixed-point x. Uniqueness is obvious. To show existence, let x0 be any point in X
and consider the sequence {xk } defined inductively by xk = K(xk−1 ).
The following corollaries are immediate consequences.
Corollary 1. Let D(p, r) be a closed ball in X. If ρ(K(p), p) ≤ (1 − C)r, then K
has a fixed-point in D(p, r).
Corollary 2. Let B(p, r) be an open ball in X. If ρ(K(p), p) < (1 − C)r, then K
has a fixed-point in B(p, r).
Theorem 1. Let S be a non-empty topological space. Suppose K : S × X → X is a
contraction mapping for which there exists C < 1 such that for all s ∈ S, x, y ∈ X,
ρ(K(s, x), K(s, y)) ≤ Cρ(x, y). Further assume that for each x ∈ X, s 7→ K(s, x) is
continuous. Let ps be the fixed-point at s. Then the mapping s 7→ ps is continuous.
Proof. For all x, t ∈ S,
ρ(ps , pt ) = ρ(K(s, ps ), K(t, pt )) ≤ ρ(K(s, ps ), K(t, ps )) + Cρ(ps , pt ) .
Rearrange to finish the proof. 
Corollary 3. Let B(p, r) be an open ball in X and S and K : S × B(p, r) → X
be as in Theorem 1. Suppose that for any s ∈ S, ρ(K(s, p), p) < (1 − C)r. Then
for each s, there is a unique point ps ∈ B(p, r) such that K(s, ps ) = ps , and the
mapping s 7→ ps is continuous.

2. The implicit function theorem


Throughout this section, we denoted Banach spaces by V , W and X. Complete-
ness is not needed for every theorem, but we assume it anyway. For a function
F : V → W , we denote its derivative at α ∈ V by dFα , provided that it exists. If
QN
V = i=1 Vi , the i-th partial derivative of F at α is denoted by dFαi . In this case,
the sum of the norms on Vi is taken as the norm on V .
Suppose that F : V → W and G : V × W → X are differentiable. Define H by
H(x) = G(x, F (x)). Then H is differentiable, and
dHx = dG1(x,F (x)) + dG2(x,F (x)) ◦ dFx .
If F and G are such that dG2F (x) is invertible and G(x, F (x)) = 0 for all x, then we
can solve for dFx :
 −1
dFx = − dG2(x,F (x)) ◦ dG1(x,F (x)) .
The implicit function theorem addresses the converse case.
1
2 THE IMPLICIT FUNCTION THEOREM

Theorem 2. Let A × B be an open subset in V × W . Suppose that G : A × B → X


is differentiable at (α, β), G(α, β) = 0 and dG2(α,β) is invertible. Also, suppose that
F : A → B is a continuous function such that F (α) = β and G(x, F (x)) = 0 for all
x ∈ A. Then F is differentiable at α, and
 −1
dFα = − dG2(α,β) ◦ dG1(α,β) .

Proof. Set η(ξ) = F (α + ξ) − F (α). Since G(α + ξ, β + η(ξ)) − G(α, β) = 0, we have


dG(α,β) (ξ, η(ξ)) + u(ξ, η(ξ)) = dG1(α,β) (ξ) + dG2(α,β) (η(ξ)) + u(ξ, η(ξ)) = 0 ,
where u is such that ku(ζ)k/kζk → 0 as ζ → 0. Put
 −1
v = η + dG2(α,β) ◦ dG1(α,β) ,
 −1
and dG2(α,β) = T . Then

v(ξ) = −T (u(ξ, η(ξ))) .


By continuity of F , η(ξ) → 0 as ξ → 0. Hence for any  > 0, we can choose δ > 0
such that kξk < δ implies
ku(ξ, η(ξ))k < (kξk + kη(ξ)k) .
Then
kv(ξ)k < C1 (kξk + C2 (kξk + kv(ξ)k))
for some non-negative constants C1 , C2 independent of . Rearrange to get
kv(ξ)k C1 (1 + C2 )
< < C1 (1 + C2 )
kξk 1 − C1 C2 
for  small enough. Therefore, kv(ξ)k/kξk → 0 as ξ → 0, and F is differentiable at
α with derivative as desired. 
Completeness is needed in the following theorem so the contraction mapping
fixed-point theorem applies.
Theorem 3 (The implicit function theorem). Let G and (α, β) be as in Theorem 2.
In addition, let G be continuously differentiable in A×B. Then there is an open ball
M about α and a uniquely determined continuously differentiable function F : M →
B such that F (α) = β and G(x, F (x)) = 0 for all x ∈ M .
 −1
Proof. Set T = dG2(α,β) and define K : A × B → W by

K(x, y) = y − T (G(x, y)) .


2
Then K is continuously differentiable, and dK(α,β) = 0. Choose a product of open
2
balls M × N about (α, β) such that for any (x, y) ∈ M × N , kdK(x,y) k ≤ 1/2 and
2
dG(x,y) is invertible. Since K is continuous and K(α, β) = β, we can also improve
M so that kK(x, β) − βk < r/2, where r is the radius of N . Then by the mean
value theorem, for all x ∈ M and y1 , y2 ∈ N ,
1
kK(x, y1 ) − K(x, y2 )k ≤ ky1 − y2 k ,
2
THE IMPLICIT FUNCTION THEOREM 3

so K is a contraction mapping in y. Since kK(x, β) − βk < (1 − 1/2)r for each


x ∈ M , K(x, ·) has a unique fixed point px in N , and the mapping F : x 7→ px is
continuous. Since K(x, y) = y if and only if G(x, y) = 0, F is the desired function
on M . More specifically,
 −1
dFx = − dG2(x,F (x)) ◦ dG1(x,F (x))
implies that F is continuously differentiable.
Suppose that H : M → B is another function as desired. Then by continuity of
F and H, the set E = {x ∈ M : F (x) 6= H(x)} is open. Also, H −1 (N ) is an open
neighbourhood of α on which F and H must agree, because the fixed point of K
is unique. Hence α is an interior point of M − E. The same argument applies to
all x ∈ M − E, so M − E is open. But M is connected and M − E is non-empty.
Hence E = ∅, and F is the uniquely determined. 
The inverse function theorem is a special case of the implicit function theorem.
Theorem 4 (The inverse function theorem). Let B be an open subset of W and
H : B → V be a continuously differentiable function such that dHβ is invertible for
some β ∈ B. Then
(i) there exist an open ball M about α = H(β) and an open neighbourhood U
of β such that H is a homeomorphism from U to M ;
(ii) the inverse F of H on M is continuously differentiable, and
−1
dFx = dHF (x) .
Proof. Define G : V × B → V by G(x, y) = x − H(y). Then G is continuously
differentiable, and dG2(α,β) = −dHβ is invertible. Let K : V × B → W and F : M →
B be as in the implicit function theorem and put U = F (M ). Then F has the
desired derivatives, and H(F (x)) = x for all x ∈ M . Clearly the restriction of H
on U is onto M . H is also injective, because the fixed point of K in U for each
x ∈ M is unique. Therefore, H : U → M is a continuous bijection with a continuous
inverse F . 

Potrebbero piacerti anche