Notes On The Implicit Function Theorem: KC Border
Notes On The Implicit Function Theorem: KC Border
1
KC Border Notes on the Implicit Function Theorem 2
ȳ = f (x̄, p̄).
c. ξ is continuous on W .
The next result, also due to Halkin [13, Theorem E] takes the second ap-
proach. It concludes that ξ is differentiable at a single point. Related results
may be found in Hurwicz and Richter [14, Theorem 1], Leach [16, 17], Nijen-
huis [21], and Nikaidô [22, Theorem 5.6, p. 81].
v. 2013.08.27::14.08
KC Border Notes on the Implicit Function Theorem 3
Dξ(p̄) = −T −1 ◦ S.
The following result is Theorem 9.4 in Loomis and Sternberg [19, p. 231].
It strengthens the hypotheses of both Theorems 3 and 4. In return we get
differentiability of ξ on W .
c. ξ is differentiable on W , and
−1
∂ξ 1 ∂ξ 1 ∂f 1 ∂f 1 ∂f 1 ∂f 1
··· ··· ···
∂p1 ∂pm ∂x1 ∂xn ∂p1 ∂pm
.. .. = − .. .. .. .. .
. . . . . .
∂ξ n ∂ξ n ∂f n ∂f n ∂f n ∂f n
··· ··· ···
∂p1 ∂pm ∂x1 ∂xn ∂p1 ∂pm
The classical version maybe found, for instance, in Apostol [3, Theorem 7-6,
p. 146], Rudin [24, Theorem 9.28, p. 224], or Spivak [26, Theorem 2-12, p. 41].
Some of these have the weaker statement that there is a unique function ξ within
the class of continuous functions satisfying both ξ(p̄) = x̄ and f (ξ(p); p) = 0
v. 2013.08.27::14.08
KC Border Notes on the Implicit Function Theorem 4
for all p. Dieudonné [8, Theorem 10.2.3, p. 272] points out that the C k case
follows from the formula for Dξ and the fact that the mapping from invertible
linear transformations to their inverses, A 7→ A−1 , is C ∞ . (See Marsden [20,
Lemma 2, p. 231].)
c. ξ is C k on W , and
−1
∂ξ 1 ∂ξ 1 ∂f 1 ∂f 1 ∂f 1 ∂f 1
··· ··· ···
∂p1 ∂pm ∂x1 ∂xn ∂p1 ∂pm
.. .. = − .. .. .. .. .
. . . . . .
∂ξ n ∂ξ n ∂f n ∂f n ∂f n ∂f n
··· ··· ···
∂p1 ∂pm ∂x1 ∂xn ∂p1 ∂pm
f (x, p) − f (y, p)
m⩽ ⩽ M.
x−y
( )
Then there is a unique function ξ : P → R satisfying f ξ(p), p = 0. Moreover,
ξ is continuous.
v. 2013.08.27::14.08
KC Border Notes on the Implicit Function Theorem 5
T = Dx f (x̄, p̄).
( )
Define φ : X × P → Rn by φ = πX − T −1 f − ȳ . That is,
( )
φ(x, p) = x − T −1 f (x, p) − ȳ . (2)
( )
Note that φ(x, p) = x if and only if T −1 f (x, p)− ȳ = 0. But the invertibility of
T −1 guarantees that this happens if and only if f (x, p) = ȳ. Thus the problem
of finding a zero of f (·, p) − ȳ is equivalent to that of finding a fixed point of
φ(·, p). Note also that
φ(x̄, p̄) = x̄. (3)
Observe that φ is continuous and also has a derivative Dx φ with respect to x
whenever f does. In fact,
v. 2013.08.27::14.08
KC Border Notes on the Implicit Function Theorem 6
That is, φ(x, p) ∈ B̄r (x̄). Since φ is continuous and B̄r (x̄) is compact and
convex, by the Brouwer Fixed Point Theorem (e.g., [5, Corollary 6.6, p. 29]),
there is some x ∈ B̄r (x̄) satisfying φ(x, p) = x, or in other words f (x, p) = 0.
We have just proven parts (a) and (b) of Theorem 2. That is, for every
neighborhood X of x̄, there is a neighborhood W = B̄ε(r) (p̄) of p̄ and a function
( )
ξ from B̄ε(r) (p̄) into B̄r (x̄) ⊂ X satisfying ξ(p̄) = x̄ and f ξ(p), p = 0 for all
p ∈ W . (Halkin actually breaks this part out as Theorem A.)
We can use the above result to construct a ξ that is continuous at p̄. Start
with a given neighborhood U of x̄. Construct a sequence of r1 > r2 > · · · > 0
satisfying lim rn = 0 and for each n consider the neighborhood Un = U ∩Brn (x̄).
From the argument above there is a neighborhood ( Wn of) p̄ and a function ξn
from Wn into Un ⊂ U satisfying ξn (p̄) = x̄ and f ξn (p), p = 0 for all p ∈ Wn .
Without loss of generality we may assume Wn ⊃ Wn+1 (otherwise replace Wn+1
with Wn ∩ Wn+1 ), so set W = W1 . Define ξ : W → U by ξ(p) (= ξn (p) ) for
p ∈ Wn \ Wn+1 . Then ξ is continuous at p̄, satisfies ξ(p̄) = x̄, and f ξ(p), p = 0
for all p ∈ W .
Note that the above proof used in an essential way the compactness of B̄r (x̄),
which relies on the finite dimensionality of Rn . The compactness was used first
to show that m(p) is finite, and second to apply the Brouwer fixed point theorem.
Theorem 3 adds to the hypotheses of Theorem 2. It assumes that Dx f exists
everywhere on X × P and is continuous. The conclusion is that there are some
neighborhoods U of x̄ and W of p̄ (and a continuous
) function ξ : W → U such
that ξ(p) is the unique solution to f ξ(p), p = 0 lying in U . It is the uniqueness
of ξ(p) that puts a restriction on U . If U is too large, say U = X, then the
solution need not be unique. (On the other hand, it is easy to show, as does
Dieudonné [8, pp. 270–271], there is at most one continuous ξ, provided U is
connected.) The argument we use here, which resembles that of Loomis and
v. 2013.08.27::14.08
KC Border Notes on the Implicit Function Theorem 7
for some z lying on the segment between x and y. If x and y lie in B̄r (x̄), then
z too must lie in B̄r (x̄), so ∥Dx φ(z, p)∥ < 12 . It follows that
1
φ(x, p) − φ(y, p) < |x − y| for all x, y ∈ B̄r (x̄), p ∈ Bε (p̄), (5)
2
so φ(·, p) is a contraction on B̄r (x̄) with contraction constant 21 .
To see that B̄r (x̄) is mapped into itself, let (x, p) belong to B̄r (x̄) × W and
observe that
v. 2013.08.27::14.08
KC Border Notes on the Implicit Function Theorem 8
Note that the above proof nowhere uses the finite dimensionality of Rm , so
the theorem actually applies to a general Banach space.
Since Df exists at (x̄, p̄), there exists r > 0 such that Br (x̄) × Br (p̄) ⊂ X × W
and if |x − x̄| < r and |p − p̄| < r, then
∆(x, p) 1
< ,
|x − x̄| + |p − p̄| 2 ∥T −1 ∥
which in turn implies
1 1
T −1 ∆(x, p) < |x − x̄| + |p − p̄|.
2 2
Since ξ is continuous at p̄ and ξ(p̄) = x̄, there is some r ⩾ δ > 0 such that
|p − p̄| < δ implies |ξ(p) − x̄| < r. Thus
( ) 1 1
T −1 ∆ ξ(p), p < ξ(p) − x̄ + |p − p̄| for all p ∈ Bδ (p̄). (6)
2 2
( )
But f ξ(p), p − f (x̄, p̄) = 0 implies
( ) ( )
T −1 ∆ ξ(p), p = ξ(p) − x̄ + T −1 S(p − p̄) . (7)
Therefore, from the facts that |a + b| < c implies |a| < |b| + c, and ξ(p̄) = x̄,
equations (6) and (7) imply
1 1
ξ(p) − ξ(p̄) < T −1 S(p − p̄) + ξ(p) − ξ(p̄) + |p − p̄| for all p ∈ Bδ (p̄)
2 2
or, ( )
ξ(p) − ξ(p̄) < 2∥T −1 S∥ + 1 |p − p̄| for all p ∈ Bδ (p̄).
That is, ξ satisfies a local Lipschitz condition at p̄. For future use set M =
2∥T −1 S∥ + 1.
Now we are in a position to prove that −T −1 S is the differential of ξ at p̄.
Let ε > 0 be given. Choose 0 < r < δ so that |x − x̄| < r and |p − p̄| < r implies
∆(x, p) ε
< ,
|x − x̄| + |p − p̄| (M + 1) ∥T −1 ∥
v. 2013.08.27::14.08
KC Border Notes on the Implicit Function Theorem 9
so
( ) ( ) ε ( )
ξ(p)−ξ(p̄) +T −1 S(p−p̄) = T −1 ∆ ξ(p), p < |ξ(p)−ξ(p̄)|+|p−p̄| ⩽ ε|p−p̄|,
(M + 1)
for |p − p̄| < r, which shows that indeed −T −1 S is the differential of ξ at p̄.
Proof of Theorem 6: Let f satisfy the hypotheses of the theorem. Let C(P )
denote the set of continuous real functions on P . Then C(P ) is complete under
the uniform norm metric, ∥f − g∥ = supp |f (p) − g(p)| [2, Lemma 3.97, p. 124].
For each p define the function ψp : R → R by
1
ψp (x) = (x) −
f (x, p).
M
Note that ψp (x) = x if and only if f (x, p) = 0. If ψp has a unique fixed
point,
( then
) we shall have shown that there is a unique function ξ satisfying
f ξ(p), p = 0. It suffices to show that ψp is a contraction.
To see this, write
f (x, p) − f (y, p)
ψp (x) − ψp (y) = x−y−
( M )
1 f (x, p) − f (y, p)
= 1− (x − y).
M x−y
By hypothesis
f (x, p) − f (y, p)
0<m⩽ ⩽ M,
x−y
so ( m)
|ψp (x) − ψp (y)| ⩽ 1 − |x − y|.
M
This shows that ψp is a contraction with constant 1 − M m
< 1.
To see that ξ is actually continuous, define the function ψ : C(P ) → C(P )
via
1 ( )
ψg(p) = g(p) − f g(p), p .
M
(Since f is continuous, ψg is continuous whenever g is continuous.) The point-
(wise argument
) above is independent of p, so it also shows that |ψg(p)−ψh(p)| ⩽
1− M m
|g(p) − h(p)| for any functions g and h. Thus
( m)
∥ψg − ψh∥ ⩽ 1 − ∥g − h∥.
M
In other words ψ is a contraction on C(P ), so( it has )a unique fixed point ḡ in
C(P ), so ḡ is continuous. But ḡ also satisfies f ḡ(p), p , but since ξ(p) is unique
we have ξ = ḡ is continuous.
v. 2013.08.27::14.08
KC Border Notes on the Implicit Function Theorem 10
x
f (x, p) > 0
(x1 , p1 ) f (x, p) = 0
f′ (x3 , p3 )
(x2 , p2 )
f (x, p) < 0 p
Proof : Fix a point p in P and let ε > 0 be given. Let ρ be a compatible metric
on P and using the continuity of φ(x, ·) on P , choose δ > 0 so that ρ(p, q) < δ
implies that ( ( ) ( ))
d φ ξ(p), p , φ ξ(p), q < (1 − α)ε.
So if ρ(p, q) < δ, then
( ) ( ( ) ( ))
d ξ(p), ξ(q) = d φ ξ(p), p , φ ξ(q), q
( ( ) ( )) ( ( ) ( ))
⩽ d φ ξ(p), p , φ ξ(p), q + d φ ξ(p), q , φ ξ(q), q
( )
< (1 − α)ε + αd ξ(p), ξ(q)
so ( )
(1 − α)d ξ(p), ξ(q) < (1 − α)ε
or ( )
d ξ(p), ξ(q) < ε,
which proves that ξ is continuous at p.
1.2 Examples
Figure 1 illustrates the Implicit Function Theorem for the special case n = m =
1, which is the only one I can draw. The figure is drawn sideways since we
are looking for x as a function of p. In this case, the requirement that the
differential with respect to x be invertible reduces to ∂f ∂x ̸= 0. That is, in the
diagram the gradient of f may not be horizontal. In the figure, you can see
that the points, (x1 , p1 ), (x2 , p2 ), and (x3 , p3 ), the differentials Dx f are zero.
At (x1 , p1 ) and (x2 , p2 ) there is no way to define x as a continuous function of
v. 2013.08.27::14.08
KC Border Notes on the Implicit Function Theorem 11
Example 10 Define f : R2 → R2 by
f 1 (x, y) = x
and
y − x2 0 ⩽ x2 ⩽ y
f 2 (x, y) = y 2 − x2 y
0 ⩽ y ⩽ x2
x2
−f 2 (x, −y) y ⩽ 0.
v. 2013.08.27::14.08
KC Border Notes on the Implicit Function Theorem 12
Then f is everywhere differentiable on R2 , and Df (0, 0) is the identity mapping, Work out the details.
f (x, p) = g(x) − p.
Then f (x, p) = 0 if and only if p = g(x). Thus if( there )is a unique implicitly
defined function ξ : P → X implicitly defined by f ξ(p), p = 0, it follows that g
is invertible and ξ = g −1 . Now compare the Jacobian matrix of f with respect
to x and observe that it is just the Jacobian matrix of g. Thus each of the
implicit function theorems has a corresponding inverse function theorem.
We could also proceed in the other direction, as is usually the case in text-
books. Let X × P be a subset of Rn × Rm , and let f : X × P → Rn , and
suppose f (x̄, p̄) = 0. Define a function g : X × P → Rn × P by
( )
g(x, p) = f (x, p), p .
for all p ∈ P . Now let’s compare hypotheses. The standard Inverse Function
Theorem, e.g. [20, Theorem 7.1.1, p. 206], says that if g is continuously dif-
ferentiable and has a nonsingular Jacobian matrix at some point, then there
is a neighborhood of the point where g is invertible. The Jacobian matrix for
v. 2013.08.27::14.08
KC Border Notes on the Implicit Function Theorem 13
( )
g(x, p) = f (x, p), p above is
∂f 1 ∂f 1 ∂f 1 ∂f 1
∂x ··· ···
1 ∂xn ∂p1 ∂pm
.. .. .. ..
. . . .
∂f n ∂f n ∂f n ∂f n
··· ···
∂x1 ∂xn ∂p1 ∂pm .
0 ··· 0 1 0
.. .. ..
. . .
0 ··· 0 0 1
Since this is block diagonal, it is easy to see that this Jacobian matrix is non-
singular at (x̄, p̄) if and only if the derivative Dx f (x̄, p̄), is invertible.
which has determinant e2x (cos2 y + sin2 y) = e2x > 0 everywhere. Nonetheless,
f is not invertible since f (x, y) = f (x, y + 2π) for every x and y. □
v. 2013.08.27::14.08
KC Border Notes on the Implicit Function Theorem 14
Proof : Since the gi ′ (x∗ )s are linearly independent, n ⩾ m, and without loss of
generality, we may assume the coordinates are numbered so that the m × m
matrix
∂g 1 ∂g 1
···
∂x1 ∂xm
. ..
.
. .
∂g m ∂g m
···
∂x1 ∂xm
is invertible at x∗ .
Fix v satisfying gi ′ (x∗ ) · v = 0 for all i = 1, . . . , m. Rearranging terms we
have
∑m
∂g i (x∗ ) ∑ n
∂g i (x∗ )
· vj = − · vj i = 1, . . . , m,
j=1
∂xj j=m+1
∂xj
or in matrix terms
∂g 1 ∂g 1 ∂g 1 ∂g 1
··· ... vm+1
∂x1 ∂xm v1 ∂xm+1 ∂xn
. .
. .. . = −
.. .. ..
. ,
. . . . .
∂g m ∂g m ∂g m ∂g m vn
vm
··· ...
∂x1 ∂xm ∂xm+1 ∂xn
so
−1 ∂g 1 ∂g 1
∂g 1 ∂g 1
··· ...
v1 vm+1
∂x1 ∂xm ∂xm+1 ∂xn
..
. = − .. .. .. .. ..
. .
. . . .
∂g m ∂g m ∂g m ∂g m vn
vm
··· ...
∂x1 ∂xm ∂xm+1 ∂xn
Observe that these conditions completely characterize v. That is, for any y ∈
Rn ,
( ′ ∗ )
gi (x ) · y = 0, i = 1, . . . , m, and yj = vj , j = m+1, . . . , n =⇒ y = v.
(8)
Define the C ∞ function f : Rm × R → Rn by
v. 2013.08.27::14.08
KC Border Notes on the Implicit Function Theorem 15
Then x̂(0) = x∗ , ( )
g x̂(α) = 0 for all α ∈ (−δ, δ),
and x̂ is differentiable at 0, and if g is C k , then x̂ is C k . So by the Chain Rule,
Now by construction (9), x̂′j (0) = vj , for j = m+1, . . . , n. Thus (8) implies
x̂′ (0) = v.
v. 2013.08.27::14.08
KC Border Notes on the Implicit Function Theorem 16
Typically we can write the equilibrium conditions of our model as the zero of a
system of equations in the endogenous variables and the exogenous parameters:
F 1 (x1 , . . . , xn ; p1 , . . . , pm ) = 0
.. (10)
.
n
F (x1 , . . . , xn ; p1 , . . . , pm ) = 0
for at least a rectangle of values of p. The Implicit Function Theorem tells that
such an explicit function exists whenever it is possible to solve for all its partial
derivatives. ( )
Setting G(p) = F ξ(p); p , and differentiating Gi with respect to pj , yields,
by equation (11),
∑ ∂F i ∂ξ k ∂F i
+ =0 (12)
∂xk ∂pj ∂pj
k
v. 2013.08.27::14.08
KC Border Notes on the Implicit Function Theorem 17
The old-fashioned derivation (see, e.g., Samuelson [25, pp. 10–14]) of this
same result runs like this: “Totally differentiate” the ith row of equation (10)
to get
∑ ∂F i ∑ ∂F i
dxk + dpℓ = 0 (15)
∂xk ∂pℓ
k ℓ
for all i. Now set all dpℓ s equal to zero except pj , and divide by dpj to get
∑ ∂F i dxk ∂F i
+ =0 (16)
∂xk dpj ∂pj
k
for all i and j, which is equivalent to equation (12). For further information on
total differentials and how to manipulate them, see [3, Chapter 6].
Using Cramer’s Rule (e.g. [4, pp. 93–94]), we see then that
∂F 1 ∂F 1 ∂F 1 ∂F 1 ∂F 1
··· ···
∂x1 ∂xi−1 ∂pj ∂xi+1 ∂xn
.. .. .. .. ..
. . . . .
∂F n ∂F n ∂F n ∂F n ∂F n
··· ···
dxi ∂ξ i ∂x1 ∂xi−1 ∂pj ∂xi+1 ∂xn
= =− . (17)
dpj ∂pj ∂F 1 ∂F 1
···
∂x1 ∂xn
.. ..
. .
∂F n ∂F n
···
∂x1 ∂xn
∂F 1 ∂F 1
···
∂x1 ∂xn
. ..
Or, letting ∆ denote the determinant of ..
. , and letting ∆i,j
∂F n ∂F n
···
∂x1 ∂xn
denote the determinant of the matrix formed by deleting its i-th row and j-th
column, we have
∂ξ i ∑
n
∂F k ∆k,i
= − (−1)i+k . (18)
∂pj ∂pj ∆
k=1
v. 2013.08.27::14.08
KC Border Notes on the Implicit Function Theorem 18
f >0 f >0
f <0
f <0
v. 2013.08.27::14.08
KC Border Notes on the Implicit Function Theorem 19
References
[1] S. N. Afriat. 1971. Theory of maxima and the method of Lagrange. SIAM
Journal of Applied Mathematics 20:343–357.
http://www.jstor.org/stable/2099955
[2] C. D. Aliprantis and K. C. Border. 2006. Infinite dimensional analysis: A
hitchhiker’s guide, 3d. ed. Berlin: SpringerVerlag.
[3] T. M. Apostol. 1957. Mathematical analysis: A modern approach to ad-
vanced calculus. Addison-Wesley series in mathematics. Reading, Mas-
sachusetts: Addison Wesley.
[10] D. Gale and H. Nikaidô. 1965. Jacobian matrix and global univalence of
mappings. Mathematische Annalen 159(2):81–93.
DOI: 10.1007/BF01360282
[11] W. Ginsberg. 1973. Concavity and quasiconcavity in economics. Journal
of Economic Theory 6(6):596–605. DOI: 10.1016/0022-0531(73)90080-X
[12] H. Halkin. 1972. Necessary conditions for optimal control problems with
infinite horizons. Discussion Paper 7210, CORE.
[13] . 1974. Implicit functions and optimization problems without con-
tinuous differentiability of the data. SIAM Journal on Control 12(2):229–
236. DOI: 10.1137/0312017
v. 2013.08.27::14.08
KC Border Notes on the Implicit Function Theorem 20
[18] A. Leroux. 1984. Other determinantal conditions for concavity and quasi-
concavity. Journal of Mathematical Economics 13(1):43–49.
DOI: 10.1016/0304-4068(84)90023-5
v. 2013.08.27::14.08
KC Border Notes on the Implicit Function Theorem 21
v. 2013.08.27::14.08