arxiv: v1 [math.ca] 4 Aug 2017

Similar documents
arxiv: v1 [math.ca] 10 Dec 2012

Chapter 4. Inverse Function Theorem. 4.1 The Inverse Function Theorem

arxiv: v1 [math.cv] 17 Feb 2014

Continuity. Chapter 4

Continuity. Chapter 4

APPLICATIONS OF DIFFERENTIABILITY IN R n.

Notes on the Implicit Function Theorem

Differentiation. f(x + h) f(x) Lh = L.

Midterm 1 Solutions Thursday, February 26

1 Directional Derivatives and Differentiability

1. Bounded linear maps. A linear map T : E F of real Banach

THE INVERSE FUNCTION THEOREM FOR LIPSCHITZ MAPS

MA2AA1 (ODE s): The inverse and implicit function theorem

MATH 409 Advanced Calculus I Lecture 10: Continuity. Properties of continuous functions.

ARCS IN FINITE PROJECTIVE SPACES. Basic objects and definitions

Yunhi Cho and Young-One Kim

SMSTC (2017/18) Geometry and Topology 2.

Analysis II: The Implicit and Inverse Function Theorems

The Contraction Mapping Theorem and the Implicit and Inverse Function Theorems

Analysis Finite and Infinite Sets The Real Numbers The Cantor Set

Chapter 3. Differentiable Mappings. 1. Differentiable Mappings

Real Analysis Math 131AH Rudin, Chapter #1. Dominique Abdi

Course Summary Math 211

INVERSE FUNCTION THEOREM and SURFACES IN R n

Selected Solutions. where cl(ω) are the cluster points of Ω and Ω are the set of boundary points.

SPRING OF 2008 IM. IMPLICIT DIFFERENTIATION. F (x, y, z) = 0,

NOTES ON MULTIVARIABLE CALCULUS: DIFFERENTIAL CALCULUS

fy (X(g)) Y (f)x(g) gy (X(f)) Y (g)x(f)) = fx(y (g)) + gx(y (f)) fy (X(g)) gy (X(f))

DEVELOPMENT OF MORSE THEORY

Multivariable Calculus

van Rooij, Schikhof: A Second Course on Real Functions

van Rooij, Schikhof: A Second Course on Real Functions

Math 147, Homework 5 Solutions Due: May 15, 2012

Preliminary Exam 2016 Solutions to Morning Exam

Introduction to Real Analysis Alternative Chapter 1

MATHEMATICAL ECONOMICS: OPTIMIZATION. Contents

Analysis Qualifying Exam

The Inverse Function Theorem 1

Several variables. x 1 x 2. x n

3 (Due ). Let A X consist of points (x, y) such that either x or y is a rational number. Is A measurable? What is its Lebesgue measure?

9. Banach algebras and C -algebras

arxiv: v1 [math.ag] 25 Jun 2017

2. Function spaces and approximation

A NICE PROOF OF FARKAS LEMMA

Bounded Derivatives Which Are Not Riemann Integrable. Elliot M. Granath. A thesis submitted in partial fulfillment of the requirements

Review of Multi-Calculus (Study Guide for Spivak s CHAPTER ONE TO THREE)

Course 212: Academic Year Section 1: Metric Spaces

The Contraction Mapping Theorem and the Implicit Function Theorem

An introduction to some aspects of functional analysis

THE NEARLY ADDITIVE MAPS

Defn 3.1: An n-manifold, M, is a topological space with the following properties:

arxiv: v1 [math.ca] 12 Jul 2018

Static Problem Set 2 Solutions

Topological vectorspaces

Economics 204. The Transversality Theorem is a particularly convenient formulation of Sard s Theorem for our purposes: with r 1+max{0,n m}

A fixed point theorem for weakly Zamfirescu mappings

LECTURE NOTES ELEMENTARY NUMERICAL METHODS. Eusebius Doedel

Many of you got these steps reversed or otherwise out of order.

MATH 409 Advanced Calculus I Lecture 11: More on continuous functions.

A WHITNEY EXTENSION THEOREM FOR CONVEX FUNCTIONS OF CLASS C 1,ω.

Chapter 2 Functions from R p to R q

P-adic Functions - Part 1

Continuous Functions on Metric Spaces

THE INVERSE FUNCTION THEOREM

The Implicit Function Theorem for Lipschitz Functions and Applications. A Masters Thesis. presented to. the Faculty of the Graduate School

MAT 257, Handout 13: December 5-7, 2011.

1.1 Limits and Continuity. Precise definition of a limit and limit laws. Squeeze Theorem. Intermediate Value Theorem. Extreme Value Theorem.

Boston College, Department of Mathematics, Chestnut Hill, MA , May 25, 2004

Continuity. Matt Rosenzweig

CHAPTER 6. Limits of Functions. 1. Basic Definitions

Contents Ordered Fields... 2 Ordered sets and fields... 2 Construction of the Reals 1: Dedekind Cuts... 2 Metric Spaces... 3

Math 600 Day 1: Review of advanced Calculus

Commutative Banach algebras 79

x +3y 2t = 1 2x +y +z +t = 2 3x y +z t = 7 2x +6y +z +t = a

ANALYSIS QUALIFYING EXAM FALL 2017: SOLUTIONS. 1 cos(nx) lim. n 2 x 2. g n (x) = 1 cos(nx) n 2 x 2. x 2.

ON SPACE-FILLING CURVES AND THE HAHN-MAZURKIEWICZ THEOREM

The Heine-Borel and Arzela-Ascoli Theorems

MAT 570 REAL ANALYSIS LECTURE NOTES. Contents. 1. Sets Functions Countability Axiom of choice Equivalence relations 9

MATH 409 Advanced Calculus I Lecture 12: Uniform continuity. Exponential functions.

arxiv: v1 [math.fa] 1 Nov 2017

Implicit Functions, Curves and Surfaces

1.4 The Jacobian of a map

REVIEW FOR THIRD 3200 MIDTERM

Regularity for Poisson Equation

Multivariate Differentiation 1

Tangent Planes, Linear Approximations and Differentiability

TESTING HOLOMORPHY ON CURVES

2 Sequences, Continuity, and Limits

The method of stationary phase

THE DEGREE OF THE INVERSE OF A POLYNOMIAL AUTOMORPHISM

Solution to Set 7, Math 2568

Section 3.5 The Implicit Function Theorem

8 Further theory of function limits and continuity

Lecture 2 - Unconstrained Optimization Definition[Global Minimum and Maximum]Let f : S R be defined on a set S R n. Then

1 The Observability Canonical Form

Implications of the Constant Rank Constraint Qualification

Immerse Metric Space Homework

Math 421, Homework #9 Solutions

We denote the derivative at x by DF (x) = L. With respect to the standard bases of R n and R m, DF (x) is simply the matrix of partial derivatives,

Subdifferential representation of convex functions: refinements and applications

Transcription:

arxiv:170802065v1 [mathca] 4 Aug 2017 THE IMPLICIT FUNCTION THEOREM FOR MAPS THAT ARE ONLY DIFFERENTIABLE: AN ELEMENTARY PROOF Oswaldo Rio Branco de Oliveira Abstract This article shows a very elementary and straightforward proof of the Implicit Function Theorem for differentiable maps F(x, y) defined on a finite-dimensional Euclidean space There are no hypotheses on the continuity of the partial derivatives of F The proof employs determinants theory, the mean-value theorem, the intermediate-value theorem, and Darboux s property (the intermediate-value property for derivatives) The proof avoids compactness arguments, fixed-point theorems, and integration theory A stronger than the classical version of the Inverse Function Theorem is also shown An example is given Mathematics Subject Classification: 26B10, 26B12 Key words and phrases: Implicit Function Theorems, Jacobians, Transformations with Several Variables, Calculus of Vector Functions 1 Introduction The aim of this article is to present a very elementary and straightforward proof of a version of the Implicit Function Theorem that is fairly stronger than the classical version We prove the implicit function theorem for differentiable maps F(x, y), defined on a finite-dimensional Euclidean space, assuming that all the leading principal minors of the partial Jacobian matrix (x,y) are nowhere vanishing(these hypothesis are already enough to show the existence of an implicit solution) plus an additional non-degeneracy condition on the matrix to ensure the uniqueness of the implicit solution There are no hypotheses on the continuity of the partial derivatives of the map F The results in this article extend de Oliveira [1] and [2] In de Oliveira [1] are proven the classical versions (enunciated for maps of class C 1 on an open set) of the implicit and inverse function theorems In de Oliveira [2] is proven the implicit function theorem for maps F(x,y) such that the partial Jacobian matrix (x,y) is only continuous at the base point 1

The proof of the implicit function theorem shown in this article follows Dini s inductive approach (see [3]) Moreover, the proofs of the implicit and the inverse function theorems that we present avoid compactness arguments (ie, Weierstrass s theorem on maxima), fixed-point theorems (eg, Banach s fixed point theorem and Brouwer s fixed-point theorem), and Lebesgue s theories of measure and integration Instead of such tools, we give elementary proofs that are based on the intermediate-value and the mean-value theorems, both on the real line, the intermediate-value property for derivatives on R (also known as Darboux s property), and determinants theory As a corollary of the implicit function theorem shown in this article we obtain a version of the inverse function theorem that is stronger than the classical one proven in most textbooks An example is given Some remarks are appropriate regarding proofs of the classical implicit and inverse function theorems Most of these proofs start by showing the inverse function theorem and then derive the implicit function theorem as a trivial consequence In general, these proofs employ either a compactness argument or the contraction mapping principle (Banach s fixed point theorem), see Krantz and Parks [5, pp 41 52] and Rudin [6, pp 221 228] On the other hand, proofs of the classical implicit and inverse function theorems that do not use either of these two tools can be seen in de Oliveira [1] Taking into account maps that are everywhere differentiable (their differentials may be everywhere discontinuous), a proof of the implicit function theorem can be found in Hurwicz and Richter [4], whereas a proof of the inverse function theorem can be seen in Saint Raymond [7] While these two results are quite general, they also have proofs that are quite technical and not that easy to follow The first of these proofs employs Brouwer s fixed-point theorem while the second relies on Lebesgue s theories of measure and integration Henceforth, we shall freely assume that all the functions are defined on a subset of a finite-dimensional Euclidean space 2 Notations and Preliminaries Apart from the intermediate-value and the mean-value theorems, both on the real line, we assume the intermediate-value theorem for derivatives on R (also called Darboux s property) stated right below Lemma 1 (Darboux s Property) Given f : [a, b] R differentiable, the image of the derivative function is an interval Given a n n real matrix A, we denote its determinant by deta The determinant of the sub-matrix of A obtained by deleting the last n k rows and the last n k columns of A is the kth order leading principal minor of A 2

Given a nonempty subset X of R n and a nonempty subset Y of R m, it is well-known that the cartesian product X Y = {(x,y) : x X and y Y} is open in R n R m if and only if X and Y are open sets Let us consider n and m, both in N, and fix the canonical bases {e 1,,e n } and {f 1,,f m }, of R n and R m, respectively Given x = (x 1,,x n ) and y = (y 1,,y n ), bothinr n, wehavethe innerproduct x,y = x 1 y 1 ++x n y n and the norm x = x,x We denote the open ball centered at a point x in R n, with radius r > 0, by B(x;r) = {y in R n : y x < r} We identify a linear map T : R n R m with the m n matrix M = (a ij ), where T(e j ) = a 1j f 1 ++a mj f m for each j = 1,,n We also write Tv for T(v), where v R n Inthis section, ΩdenotesanonemptyopensubsetofR n, wheren 1 Given a map F : Ω R m and a point p in Ω, we write F(p) = ( F 1 (p),,f m (p) ) Let us suppose that F is differentiable at p The Jacobian matrix of F at p is JF(p) = ( ) i (p) = x j 1 i m 1 j n x 1 (p) m x 1 (p) x n (p) m x n (p) If F is a real function, then we have JF(p) = F(p), the gradient of F at p Given p and q, both in R n, we denote the linear segment with endpoints p and q by pq = {p + t(q p) : 0 t 1} The following result is a trivial corollaryof the mean-value theorem on the real line and thus we omit the proof Lemma 2 (The mean-value theorem in several variables) Let us consider a differentiable real function F : Ω R, with Ω open in R n Let p and q be points in Ω such that the segment pq is within Ω Then, there exists c in pq, with c p and c q, that satisfies F(p) F(q) = F(c),p q Given a real function F : Ω R, a short computation shows that the following definition of differentiability is equivalent to that which is most commonly employed We say that F is differentiable at p in Ω if there are an open ball B(p;r) within Ω, where r > 0, a vector v in R n, and a vector-valued map E : B(0;r) R n satisfying { F(p+h) = F(p)+ v,h + E(h),h, for all h B(0;r), where E(0) = 0 and E(h) 0 as h 0 3 Example and Motivation Right below we give an example of a function F : R 2 R 2 so that F is differentiable everywhere, the Jacobian matrix JF is not continuous at the origin, the leading principal minors of JF do not vanish near the origin, F is invertible near the origin (proven in the last section) 3

Example Let us consider the function { ( ) 8x+x F(x,y) = 3 1 cos x 2 +y,8y +y 3 1 sin 2 x 2 +y outside the origin 2 (0, 0) at the origin The Jacobian matrix of F outside the origin is given by 8+3x 2 1 cos x 2 +y + 2x4 1 2 (x 2 +y 2 ) sin 2 x 2 +y 2 2x 3 y (x 2 +y 2 ) 2 sin 1 x 2 +y 2 2xy3 (x 2 +y 2 ) 2 cos 1 x 2 +y 2 8+3y 2 sin 1 x 2 +y 2 2y4 (x 2 +y 2 ) 2 cos 1 x 2 +y 2 On the other hand, a short computation shows that ( ) 8 0 JF(0,0) = 0 8 Let us show that F is differentiable at the origin (and thus over the plane) Let T be the linear map associated to the matrix JF(0,0) Given a non-null vector v = (h,k) in the plane we have F(v) F(0) Tv v = ( 8h+h 3 1 cos h 2 +k,8k +k 3 1 sin 2 h 2 +k ) (8h,8k) 2 h2 +k 2 ( h 3 1 cos h = 2 +k,k 3 1 sin 2 h2 +k 2 h 2 +k 2 ) (h,k) (0,0) (0, 0) Thus, F is differentiable at the origin We claim that the four entries of JF(x,y) are discontinuous at the origin For instance, let us look the first entry, which has three terms The first two terms are continuous at the origin However, the third term is not In fact, by employing polar coordinates and writing (x,y) = (rcosθ,rsinθ) we find x 4 (x 2 +y 2 ) 2 sin 1 x 2 +y 2 = (cos4 θ)sin 1 r 2 Thus, the Jacobian matrix JF is not continuous at the origin At last, let us fix (x,y) with x 2 +y 2 1 There exist a,b,c,d,e and f, all in [ 1,1], such that the two principal minors of JF(x,y) respectively satisfy 8+3a+2b 3 and detjf(x,y) = 8+3a+2b 2c 2d 8+3e+2f 32 2 2 So, the principal minors of JF do not vanish in the unit disc centered at (0,0) In the last section we prove that F is invertible on a neighborhood of (0,0) 4

4 The Implicit Function Theorem The first implicit function result we prove concerns one equation, several real variables and a differentiable real function In its proof, we denote the variable in R n+1 = R n R by (x,y), where x = (x 1,,x n ) is in R n and y is in R In the next theorem, Ω denotes a nonempty open set within R n R Theorem 1 Let F : Ω R be differentiable, with nowhere vanishing, and (a,b) a point in Ω such that F(a,b) = 0 Then, there exists an open set X Y, within Ω and containing the point (a, b), that satisfies the following There exists a unique function g : X Y that satisfies F ( x,g(x) ) = 0, for all x in X We have g(a) = b The function g : X Y is differentiable and satisfies g x (x) = j (x,g(x)), for all x in X, where j = 1,,n x j (x,g(x)) Moreover, if F(x,y) is continuous at (a,b) then g(x) is continuous at x = a Proof By considering the function F(x+a, y c +b), with c = (a,b), we may assume that (a,b) = (0,0) and (0,0) = 1 Next, we split the proof into three parts: existence and uniqueness, continuity at the origin, and differentiability Existence and Uniqueness Let us choose a non-degenerate(n+1)-dimensional parallelepiped X [ r,r], centered at (0,0) and within Ω, whose edges are parallel to the coordinate axes and X is open Then, the function ϕ(y) = F(0,y), where y runs over [ r,r], is differentiable with ϕ nowhere vanishing and ϕ (0) = 1 Thus, by Darboux s property we have ϕ > 0 everywhere and we conclude that ϕ is strictly increasing Hence, by the continuity of F and shrinking X (if necessary) we may assume that F is strictly negative at the bottom of the parallelepiped and F is strictly positive at the top of the parallelepiped That is, F < 0 and F > 0 X { r} X {r} As a consequence, having fixed an arbitrary x in X, the function ψ(y) = F(x,y), where y [ r,r], satisfiesψ( r) < 0 < ψ(r) Hence, bythemean-valuetheoremthereexists a point η in the open interval Y = ( r,r) such that ψ (η) = (x,η) > 0 Therefore, by Darboux s property we have ψ (y) > 0 at every y in Y Thus, ψ is strictly increasing and the intermediate-value theorem yields the existence of a unique y, we then write y = g(x), in the open interval Y such that F(x,g(x)) = 0 5

Continuity at the origin Let δ satisfy 0 < δ < r From above, there exists an open set X, contained in X and containing 0, such that g(x) is in the interval ( δ,δ), for all x in X Thus, g is continuous at x = 0 Differentiability From the differentiability of the real function F at (0, 0), and writing F(0,0) = (v,1) R n R for the gradient of F at (0,0), it follows that there are functions E 1 : Ω R n and E 2 : Ω R satisfying F(h,k) = v,h +k + E 1 (h,k),h +E 2 (h,k)k, where lim E j(h,k) = 0 = E j (0,0), for j = 1,2 (h,k) (0,0) Hence, substituting [we already proved that g(h) h 0 g(0) = 0] { k = ( g(h), ) E j h,g(h) = ǫj (h), with lim ǫ j (h) = ǫ j (0) = 0 for j = 1,2, h 0 and noticing that we have F ( h,g(h) ) = 0, for all possible h, we obtain v,h +g(h) + ǫ 1 (h),h +ǫ 2 (h)g(h) = 0 Thus, [1+ǫ 2 (h)]g(h) = v,h ǫ 1 (h),h If h is small enough, then we have 1+ǫ 2 (h) 0 and we may write where g(h) = v,h + ǫ 3 (h),h, ǫ 3 (h) = ǫ 2(h) 1+ǫ 2 (h) v ǫ 1(h) 1+ǫ 2 (h) Therefore, g is differentiable at 0 and g(0) = v and lim h 0 ǫ 3(h) = 0 Now, given any a in X, we put b = g(a ) Then, g : X Y solves the problem F ( x,h(x) ) = 0, for all x in X, with the condition h(a ) = b From what we have just done it follows that g is differentiable at a Next, we prove the implicit function theorem for a finite number of equations Some notations are appropriate We denote the variable in R n R m = R n+m by (x,y), where x = (x 1,,x n ) is in R n and y = (y 1,,y m ) is in R m Given Ω an open subset of R n R m and a differentiable map F : Ω R m we write F = (F 1,,F m ), with F i the ith component of F and i = 1,,m, and = ( ) i j = 1 i m 1 j m m m m m Analogously, we define the matrix x = ( i x k ), where 1 i m and 1 k n 6

Theorem 2 (The Implicit Function Theorem) Let F : Ω R m be differentiable, with Ω a non-degenerate open ball within R n R m and centered at (a,b) Let us suppose that F(a,b) = 0 and that all the leading principal minors of the matrix are nowhere vanishing The following are true There exists an open set X Y, within Ω and containing (a,b), and a differentiable function g : X Y that satisfies We have Jg(x) = F ( x,g(x) ) = 0, for all x X, and g(a) = b [ ] 1 [ ] (x,g(x)) m m x (x,g(x)), for all x in X m n Let us suppose that we also have det ( i j (ξ ij ) ) 1 i,j m 0, for all ξ ij in Ω and 1 i,j m Then, the following is true If h : X Y satisfies F ( x,h(x) ) = 0 for all x in X, then we have h = g Proof Let us split the proof into three parts: existence and differentiability, differentiation formula, and uniqueness Existence and differentiability We claim that the system F 1 (x,y 1,,y m ) = 0, F 2 (x,y 1,,y m ) = 0, with the conditions F m (x,y 1,,y m ) = 0, y 1 (a) = b 1 y 2 (a) = b 2 y m (a) = b m, has a differentiable solution g(x) = ( g 1 (x),,g m (x) ) on some open set X containing a [ie, we have F ( x,g(x) ) = 0 for all x in X and g(a) = b] Let us employ induction on m The case m = 1 follows immediately from Theorem 1 Assuming that the claim holds for m 1, let us examine the case m Then, given a pair (x,y) = (x,y 1,,y m ) we introduce the helpful notations y = (y 2,,y m ), y = (y 1,y ), and (x,y) = (x,y 1,y ) As a first step, we consider the equation F 1 (x,y 1,y ) = 0, with the condition y 1 (a,b ) = b 1, where x and y are independent variables and y 1 is the dependent one Since 1 (x,y 1,y ) is nowhere vanishing, by Theorem 1 it follows that there exists a differentiable function ϕ(x,y ) on some open set [let us say, X Y ] containing (a,b ) that satisfies F 1 [x,ϕ(x,y ),y ] = 0 (on X Y ) and the condition ϕ(a,b ) = b 1 7

From Theorem 1 we see that ϕ(x,y ) also satisfies the m 1 equations ϕ j (x,y ) = j [x,ϕ(x,y ),y ], for all j = 2,,m (1) [x,ϕ(x,y ),y ] As a second step, we look at solving the system with m 1 equations F 2 [x,ϕ(x,y ),y ] = 0, with the condition y (a) = b F m [x,ϕ(x,y ),y ] = 0 Here, x is the independent variable while y is the dependent variable Let us define F i (x,y ) = F i [x,ϕ(x,y ),y ], with i = 2,,m, and write F = (F 2,,F m ) Evidently, the map F is differentiable In order to employ the induction hypothesis, let us show that all the leading principal minors of the partial Jacobian matrix are nowhere vanishing Thus, let us consider the leading principal minor (a general one) ϕ 2 + 2 ϕ 2 3 + 2 3 3 ϕ 2 + 3 3 ϕ 2 3 + 3 3 ϕ 2 + 2 ϕ 3 + 3 ϕ + 2 3 ϕ + 3 ϕ + Developing this determinant by the columns and then canceling the everywhere vanishing determinants we arrive at (a sum of k determinants) 2 3 2 ( ) i det = 3 3 j 2 3 3 2 i,j k 2 3 k ϕ 2 3 2 ϕ 2 3 4 2 + 3 ϕ 3 2 3 3 + 3 3 ϕ 3 2 3 4 3 ϕ 2 3 k ϕ 2 3 4 k 2 2 ϕ 1 ++ 3 2 3 3 ϕ 1 2 1 ϕ 8

Thus, we obtain (keeping track of ϕ j for j even and also for j odd) ( ) i det j 2 i,j k = 1 ϕ 2 ϕ 3 2 3 3 3 3 2 2 3 3 ϕ 3 The already remarked identity ϕ j = 1 j / 1 [see formula (1)] leads to ( ) i det j 2 i,j k = 1 2 3 2 3 3 3 3 2 2 3 3 3 Hence, all the leading principal minors of are nowhere vanishing Thus, by induction hypothesis there exists a differentiable function ψ defined on an open set X containing a [with ψ(x) within Y ] that satisfies { Fi [x,ϕ ( x,ψ(x) ),ψ(x) ] = 0, for all x in X, for all i = 2,,m, and the condition ψ(a) = b Clearly, we also have F 1 [ x,ϕ ( x,ψ(x) ),ψ(x) ] = 0, for all x in X Defining g(x) = ( ϕ(x,ψ(x)),ψ(x) ), where x X, we obtain F[x,g(x)] = 0, for every x in X, with g differentiable on X, and also the identity g(a) = ( ϕ(a,b ),b ) = (b 1,b ) = b Differentiation formula Differentiating F[x, g(x)] = 0 we find i x k + m j=1 i j g j x k = 0, with 1 i m and 1 k n ( ) ( ) In matricial form, we write x x,g(x) + x,g(x) Jg(x) = 0 Uniqueness If h : X Y and x in X satisfy F(x,h(x)) = 0, by Lemma 2 (the mean-value theorem in several variables) there exist c 1,,c m, all 9

in the open ball Ω (a convex set), satisfying 0 = F ( x,h(x) ) F ( x,g(x) ) (c 1 ) 1 m (c 1 ) = m (c m ) h 1 (x) g 1 (x) m m (c m ) h m (x) g m (x) From the hypothesis we have det ( i j (c i ) ) 0 Thus, h(x) = g(x) 5 The Inverse Function Theorem Theorem 3 (The Inverse Function Theorem) Let F : Ω R n be differentiable, with Ω a non-degenerate open ball within R n and centered at x 0 Let us suppose that all the leading principal minors of JF(x) are nowhere vanishing We also suppose det ( i x j (ξ ij ) ) 1 i,j n 0 for all ξ ij Ω, where i,j = 1,,n Under such conditions, there exist an open set X containing x 0, an open set Y containing y 0 = F(x 0 ), and a differentiable G : Y X satisfying F ( G(y) ) = y, for all y Y, and G ( F(x) ) = x, for all x X In addition, JG(y) = JF ( G(y) ) 1, for all y in Y Proof Let us split it into two parts: injectivity of F and existence of G Injectivity of F Let us suppose F(p) = F(q), with p in Ω and q in Ω By employing Lemma 2 we obtain c 1,,c n, all in the ball Ω, such that x 1 (c 1 ) 1 x n (c 1 ) p 1 q 1 0 = F(p) F(q) = n x 1 (c n ) n x n (c n ) p n q n The hypotheses imply det ( i x j (c i ) ) 0 Thus, p = q Existence of G The map Φ(y,x) = F(x) y, where (y,x) R n Ω, is differentiable and Φ(y 0,x 0 ) = 0 From the hypotheses it follows that all the leading principal minors of Φ x (y,x) = JF(x) are nowhere vanishing in R n Ω and ( ) ( ) Φ det x (η ij,ξ ij ) = det x (ξ ij) 0, 10

for all (η ij,ξ ij ) R n Ω The Implicit Function Theorem guarantees an open set Y containing y 0 and a differentiable map G : Y Ω satisfying F ( G(y) ) = y, for all y in Y Thus, G is bijective from Y to X = G(Y) and F is bijective from X to Y We also have X = F 1 (Y) Since F is continuous, the set X is open (and contains x 0 ) Putting F(x) = ( F 1 (x),,f n (x) ) and G(y) = ( G 1 (y),,g n (y) ) and differentiating ( F 1 (G(y)),,F n (G(y)) ) we find n k=1 i G k = { i 1, if i = j, = x k j j 0, if i j Remark 1 Is is clear that the function in section 3 (Example and Motivation) satisfy the conditions of the above inverse function theorem and is thus invertible, with differentiable inverse function, on a neighborhood of the origin Remark 2 It is not difficult to see that Theorem 2 implies the implicit function theorem for a differentiable function F : Ω R n R m R m, with F(a,b) = 0 and Ω an open set, whose partial Jacobian matrix (x,y) is continuous at the base point (a,b) and det (a,b) 0 In fact, by a linear change of coordinates in the y variable, we may assume (a,b) = I, with I the m m identity matrix Thus, on some open neighborhood of (a,b), we have det ( i j (ξ ij ) ) 0 1 i,j k for all ξ ij in this neighborhood, where 1 i,j k, for each k = 1,,m Remark 3 Similarly, Theorem 3 implies the inverse function theorem for a differentiable function F : Ω R n R n, with F(x 0 ) = y 0 and Ω an open set in R n, whose Jacobian matrix JF(x) is continuous at x 0 and detjf(x 0 ) 0 Acknowledgments The author is gratful to Professor P A Martin for discussions that lead to the example given in this article References [1] O R B de Oliveira, The implicit and the inverse function theorems: easy proofs, Real Anal Exchange, 39(1), 2013/2014, pp 207 218 [2] O R B de Oliveira, The implicit function theorem when the partial Jacobian matrix is only continuous at the base point, Real Anal Exchange, 41(2), 2016, pp 377 388 11

[3] U Dini, Lezione di Analisi Infinitesimale, volume 1, Pisa, 1907, 197 241 [4] L Hurwicz and M K Richter, Implicit functions and diffeomorphisms without C 1, Adv Math Econ, 5 (2003) 65 96 [5] S G Krantz and H R Parks, The Implicit Function Theorem - History, Theory, and Applications, Birkhäuser, Boston, 2002 [6] W Rudin, Principles of Mathematical Analysis, third edition, McGraw- Hill, Singapore, 1976 [7] J Saint Raymond, Local inversion for differentiable functions and the Darboux property, Mathematika, 49 (2002), 141 158 Departamento de Matemática, Universidade de São Paulo Rua do Matão 1010 - CEP 05508-090 São Paulo, SP - Brasil oliveira@imeuspbr 12