Math 201 Handout 1. Rich Schwartz. September 6, 2006

Similar documents
Implicit Functions, Curves and Surfaces

MATH 51H Section 4. October 16, Recall what it means for a function between metric spaces to be continuous:

Math 328 Course Notes

Richard S. Palais Department of Mathematics Brandeis University Waltham, MA The Magic of Iteration

Math 600 Day 1: Review of advanced Calculus

Nonlinear equations. Norms for R n. Convergence orders for iterative methods

VANDERBILT UNIVERSITY. MATH 2300 MULTIVARIABLE CALCULUS Practice Test 1 Solutions

Advanced Calculus I Chapter 2 & 3 Homework Solutions October 30, Prove that f has a limit at 2 and x + 2 find it. f(x) = 2x2 + 3x 2 x + 2

Partial Derivatives October 2013

Example: Limit definition. Geometric meaning. Geometric meaning, y. Notes. Notes. Notes. f (x, y) = x 2 y 3 :

d(x n, x) d(x n, x nk ) + d(x nk, x) where we chose any fixed k > N

Analysis II: The Implicit and Inverse Function Theorems

MATH The Chain Rule Fall 2016 A vector function of a vector variable is a function F: R n R m. In practice, if x 1, x n is the input,

b) The system of ODE s d x = v(x) in U. (2) dt

Introduction to Algebraic and Geometric Topology Week 3

Non-linear wave equations. Hans Ringström. Department of Mathematics, KTH, Stockholm, Sweden

Math 127: Course Summary

Math 361: Homework 1 Solutions

Lecture 10. (2) Functions of two variables. Partial derivatives. Dan Nichols February 27, 2018

Metric Spaces and Topology

D-MATH Algebra I HS18 Prof. Rahul Pandharipande. Solution 1. Arithmetic, Zorn s Lemma.

Problem List MATH 5143 Fall, 2013

SMSTC (2017/18) Geometry and Topology 2.

Math 111, Introduction to the Calculus, Fall 2011 Midterm I Practice Exam 1 Solutions

MATH20411 PDEs and Vector Calculus B

APPLICATIONS OF DIFFERENTIABILITY IN R n.

THE INVERSE FUNCTION THEOREM FOR LIPSCHITZ MAPS

SOLVING DIFFERENTIAL EQUATIONS

How to Use Calculus Like a Physicist

Math 126: Course Summary

1 The Local-to-Global Lemma

Lecture Notes on Metric Spaces

Writing proofs for MATH 51H Section 2: Set theory, proofs of existential statements, proofs of uniqueness statements, proof by cases

converges as well if x < 1. 1 x n x n 1 1 = 2 a nx n

Math 212-Lecture 8. The chain rule with one independent variable

The Heine-Borel and Arzela-Ascoli Theorems

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

MATH 819 FALL We considered solutions of this equation on the domain Ū, where

LECTURE 10: THE PARALLEL TRANSPORT

Functional Analysis HW #3

COMPLETE METRIC SPACES AND THE CONTRACTION MAPPING THEOREM

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

The Inverse Function Theorem 1

APPLICATIONS OF DIFFERENTIATION

The Contraction Mapping Theorem and the Implicit and Inverse Function Theorems

The Liapunov Method for Determining Stability (DRAFT)

The Symmetric Space for SL n (R)

DEPARTMENT OF MATHEMATICS AND STATISTICS UNIVERSITY OF MASSACHUSETTS. MATH 233 SOME SOLUTIONS TO EXAM 2 Fall 2018

Theory of Ordinary Differential Equations

TANGENT VECTORS. THREE OR FOUR DEFINITIONS.

1.4 The Jacobian of a map

Lecture 3. Econ August 12

Section 2.8: The Existence and Uniqueness Theorem

FIXED POINT ITERATIONS

Introduction to Real Analysis Alternative Chapter 1

Midterm 1 Solutions Thursday, February 26

MINIMAL GRAPHS PART I: EXISTENCE OF LIPSCHITZ WEAK SOLUTIONS TO THE DIRICHLET PROBLEM WITH C 2 BOUNDARY DATA

DIFFERENTIAL GEOMETRY Multivariable Calculus, a refresher

UNIQUENESS OF nth ORDER DIFFERENTIAL SYSTEMS WITH STRONG SINGULARITY

1. Continuous Functions between Euclidean spaces

Midterm 1. Every element of the set of functions is continuous

X. Linearization and Newton s Method

Lecture 10. Riemann-Stieltjes Integration

Chain Rule. MATH 311, Calculus III. J. Robert Buchanan. Spring Department of Mathematics

Math 114: Course Summary

Economics 204 Fall 2011 Problem Set 2 Suggested Solutions

Pogorelov Klingenberg theorem for manifolds homeomorphic to R n (translated from [4])

INVERSE FUNCTION THEOREM and SURFACES IN R n

Coordinate systems and vectors in three spatial dimensions

1.1. The category of Banach spaces. A Banach space E over F is a complete normed vector space where F = R or C. This means that

1 The Existence and Uniqueness Theorem for First-Order Differential Equations

Math 118B Solutions. Charles Martin. March 6, d i (x i, y i ) + d i (y i, z i ) = d(x, y) + d(y, z). i=1

Existence and Uniqueness

Solutions to Problem Set 5 for , Fall 2007

THE INVERSE FUNCTION THEOREM

V (v i + W i ) (v i + W i ) is path-connected and hence is connected.

2tdt 1 y = t2 + C y = which implies C = 1 and the solution is y = 1

Applied Math Qualifying Exam 11 October Instructions: Work 2 out of 3 problems in each of the 3 parts for a total of 6 problems.

Math 117: Continuity of Functions

Locally convex spaces, the hyperplane separation theorem, and the Krein-Milman theorem

Fixed point iteration Numerical Analysis Math 465/565

Lecture 10. Econ August 21

Metric Spaces Math 413 Honors Project

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

Math 101: Course Summary

Math 131 Exam 2 Spring 2016

Math Advanced Calculus II

Handout 5, Summer 2014 Math May Consider the following table of values: x f(x) g(x) f (x) g (x)

Riemannian geometry of surfaces

Problem List MATH 5173 Spring, 2014

arxiv:math/ v3 [math.co] 15 Oct 2006

Asymptotic Stability by Linearization

Infinite series, improper integrals, and Taylor series

1 Lyapunov theory of stability

Review for the First Midterm Exam

Metric spaces and metrizability

The Contraction Mapping Theorem and the Implicit Function Theorem

Contents. 1. Introduction

2015 Math Camp Calculus Exam Solution

M311 Functions of Several Variables. CHAPTER 1. Continuity CHAPTER 2. The Bolzano Weierstrass Theorem and Compact Sets CHAPTER 3.

Transcription:

Math 21 Handout 1 Rich Schwartz September 6, 26 The purpose of this handout is to give a proof of the basic existence and uniqueness result for ordinary differential equations. This result includes the statement that the solutions depend smoothly on the initial conditions if the data for the differential equation is smooth. (See below for a precise formulation.) There are plenty of accounts of this theorem in the literature. However, I found the various written accounts somehow complicated and opaque. I hope to give a streamlined and clear proof here. For ease of exposition, I will give the proof in R 2. This case shares all the features of the general case of R n. Mainly, I m doing R 2 instead of R n so that I can write (x, y) in place of (x 1,..., x n ). 1 The Result We will always coordinatize R 3 as (x, y, t). We think of t as time. A map φ : R 3 R is smooth if all orders of partial derivatices of φ exist and are continuous. (Just the existence of all orders of derivatives forces their continuity.) I ll denote partial derivatives by d t, d x, d y. For instance d t φ = φ/ t. Suppose that V : U R 2 is a smooth function defined in a neighborhood U of (, ). We can write V = (V 1, V 2 ). Given a map F : U R 2 we will write F = (F 1, F 2 ). Here F 1 and F 2 are both maps from R 2 R. The basic ordinary differential equation is given by d t F = V F. Geometrically, V defines a vector field in a neighborhood of the origin and the curve γ(t) = F(x, y, t) (for x and y held fixed) is always tangent to V. To write out the above equation in detail, we have 1

d t F 1 (x, y) = V 1 (F 1 (x, y), F 1 (x, y)). d t F 2 (x, y) = V 2 (F 2 (x, y), F 2 (x, y)). Assuming that F exists, we define f(x, y) = F(x, y, ). We say that the function F is a V -extension of f in this case. In particular, F should be defined both for small positive values of t and small negative values. Here is the basic result: Theorem 1.1 Suppose that U is a small neighborhood of the origin in R 2 on which V is defined and smooth. Suppose that f : U R 2 is a smooth function. Then f has a smooth V -extension F, defined in perhaps a smaller neighborhood about the origin. 2 Existence and Uniqueness Here we will hold the point (x, y) fixed. We show the existence and uniqueness of a smooth map φ(t) = F(x, y, t) which satisfies the basic equation, at least in some interval ( ǫ, ǫ). The interval ( ǫ, ǫ) only depends on some properties of V which are the same for all choices (x, y). Thus, when we run our argument at each point (x, y) in a neighborhood of (, ), we can use the same value of ǫ for all points. This gives us our V -extension F of f. All we know about F from this construction is that the partial derivatives d (n) t F exist for all n. We don t know about the other partial derivatives. Dealing with these other partial derivatives, which we do in later sections, is the most painful part of the argument. So, suppose now that the closed interval I = [ ǫ, ǫ] is fixed. We will specify the choice of ǫ later on. We let X denote the set of all continuous functions ψ from I to R 2 such that ψ() = F(,, ). Our desired solution φ is going to be an element of X. There is a natural metric on X, namely d(g 1, g 2 ) = sup g 1 (t) g 2 (t). t I Here is just the Euclidean norm on R 2. It is not hard to show that X is a complete metric space. That is, any Cauchy sequence in X converges to a 2

unique point in X. This result is sometimes stated in terms of equicontinuous families of functions in a basic analysis course. A map T : X X is called a contraction if there is some c (, 1) such that d(t(g 1 ), T(g 2 )) cd(g 1, g 2 ) If T is a contraction, then T has a unique fixed point. The uniqueness is immediate. For existence, you just look at the sequence {T n (x)}, where x X is any starting point, and note that this is a Cauchy sequence. Now for our proof. Consider the following map T : X X. Given any function ψ : [ ǫ, ǫ] R we define Tψ(s) = ψ() + V (ψ(t)) dt. Lemma 2.1 If ǫ is small enough then T is a contraction. Proof: Since V is smooth in a neighborhood of the origin, we have a bound K on all the first order directional derivatives of V in a possibly smaller neighborhood of the origin. This gives us a bound V (p 1 ) V (p 2 ) K p 1 p 2, for any two points p 1 and p 2 sufficiently close to the origin. (In the literature, this condition is summarized by saying the V is Lipschitz near the origin.) Taking ǫ = K/2 we have Tψ 1 (s) Tψ 2 (s) K ψ 1 (t) ψ 2 (t) dt K s d(ψ 1, ψ 2 ) Kǫd(ψ 1, ψ 2 ) d(ψ 1, ψ 2 )/2. Since this inequality holds for all s I, we have Hence T is a contraction. d(tψ 1, Tψ 2 ) d(ψ 1, ψ 2 )/2. So, we can start with any function, say the constant function ψ. Then the limit φ = lim T n ψ exists and is the unique fixed point of T. Hence φ(s) = Tφ(s) = φ() + 3 V (φ(t)) dt. (1)

Now for the fun: We know that φ is continuous because it is a member of X. Hence V φ is continuous. Hence φ is the integral of a continuous function. Hence φ is differentiable. Hence V φ is differentiable. But now φ is the integral of a differentiable function. Hence φ is twice differentiable. And so on. This shows inductively that φ is smooth. When we differentiate both sides of the last equation and apply the Fundamental Theorem of Calculus, we arrive at d t φ = V φ. This proves the existence and uniqueness of solutions to the basic differential equation. 2.1 A General Estimate In the last section we used the Lipschitz condition on V. That is V (p 1 ) V (p 2 ) K p 1 p 2, for any two points p 1 and p 2 sufficiently close to the origin. Here we will prove a general estimate about two different solutions to two different equations. Suppose that V is some other smooth and K-Lipschitz function defined in a neighborhood of the origin. We want to consider solutions to the two equations d t F = V F; d t F = V F, where F is a V -extension of f and F is a V -extension of another smooth function f. Intuitively, we want to prove the result that the extensions F and F are close provided that f and f are close and V and V are close. Lemma 2.2 Suppose that t [ t, t ] where exp((k + 1)t ) = 2. If ( x, ŷ) (x, y) < ǫ, f( x,ŷ) f(x, y) < ǫ, V (p) V (p) < ǫ for all points p in a neighborhood of the origin. then F( x, ŷ, s) F(x, y, s) 2ǫ. 4

Proof: Using the integral form for the basic equation (Equation 1), and the chain rule, we have F( x, ŷ, ŝ) F(x, y, s) f(x, y) f( x, ŷ) + f(x, y) f( x, ŷ) + + ǫ + ǫs + V (F(x, y, t)) V ( F( x,ŷ, t)) dt V (F(x, y, t)) V (F(x, y, t)) dt V (F(x, y, t)) V (F( x,ŷ, t)) dt K F(x, y, t) F(x, y, t) dt. The last line just comes from putting in all the hypotheses of the lemma. It is convenient to write The calculation above shows that A(t) = F(x, y, t) F( x, ŷ, t). A(s) ǫ + ǫs + KA(t)dt. Differentiating both sides and remembering the initial conditions, we see that A() ǫ; d t A ǫ + KA. Suppose, for the sake of contradiction that there is some t [ t, t ] such that A(t) > 2ǫ. Without loss of generality assume t >. There is some maximal interval [t 1, t 2 ] [, t ] such that t [t 1, t 2 ] and A ǫ on all of [t 1, t 2 ]. For this interval we have the bound But then A(t 1 ) = ǫ; which is a contradiction. d t A (K + 1)A. A(t) exp((k + 1)(t t )) ǫ 2ǫ, As an immediate application of our result we can take V = V and F = F. The above result shows that F(x, y, t) F( x,ŷ, t) 2 f(x, y) f( x, ŷ). as long as t is sufficiently small and ( x,ŷ) is sufficiently close to (x, y). This shows that the function F(x, y, t) is continuous in all variables, sufficiently close to the origin. 5

3 Smoothness We fix some small value of h and consider the new function F h (x, y, t) = F(x + h, y, t) F(x, y, t). h Of course, we might need to work in a smaller neighborhood so that everything in this last equation is defined. We also define f h (x, y) = F h (x, y, t). Our goal is to show that lim h F h (x, y, t) exists in a neighborhood of the origin in R 3. Below we will prove Lemma 3.1 There is a smooth function V h (x, y) such that F h satisfies the differential equation t F h = V h F h. Furthermore, V h is K-Lipschitz in a neighborhood of the origin, where K and the neighborhood are independent of h. Finally, the limit W = lim h V ( x, y, t) exists at all points in a neighborhood of the origin in R 3. Given Lemma 3.1 we compute the proof as follows. For any ǫ > of interest to us, we choose h small enough such that f h (x, y) d x f(x, y) < ǫ V h (p) W(p) < ǫ for all p in a neighborhood U of the origin. From the basic existence theorem we know that d x f has a continuous W- extension G. From the estimate in the previous section we know that F h (x, y, t) G(x, y, t) 2ǫ throughout our neighborhood. We have ǫ as h. Hence F h G. Hence d x F exists in a neighborhood of the origin. The same argument shows that d y F exists in a neighborhood of the origin. 6

Before we prove Lemma 3.1 we deal with the higher derivatives of F. The argument above works for a sufficiently small neighborhood of the origin. However, we can simply relabel the origin and re-run the argument based at some other point in order to show that d x F and d y F exist in a small neighborhood of any point where F itself exists and satisfies the basic differential equation. This shows that d x F and d y F exist on exactly the same neighborhood that F exists on. Now we can repeat the argument. Lemma 3.1 says that d x F is a W-extension of d x f, where f and W are both smooth. Hence d x F is continuous, and both d xx F exist and d yx F exist. Likewise d xy F and d yy exist. And so on. This shows that F is smooth in the neighborhood on which it exists and satisfies the differential equation. 4 Proof of Lemma 3.1 This is a painful exercise in the chain rule. Using Equation 1 we have F h (x, y, s) = F(x + h, y, s) F(x, y, s) h F(x + h, y, ) F(x, y, ) + h V (F(x + h, y, t)) V (F(x, y, t)) dt = h Now let s do something different. Holding s fixed, let Ψ(u) = V (u(f(x, y, s)) + (1 u)f(x + h, y, s)). From the Fundamental Theorem of Calculus, we have Ψ(1) Ψ() = d u (Ψ)du = d x V (F(x + uh, y, s))h F h,1 (x, y, s)+ = R 1 (s)ds. (2) d y V (F(x + uh, y, s))h F h,2 (x, y, s). (3) This last line comes from the chain rule. we have set F h = ( F h,1, F h,2 ). 7

Note that Ψ(1) = V (F(x, y, t)) and Ψ() = V (F((x + h, y, t)). Plugging this into the last equation and dividing by h we have Now we define and R(s) = d x V (F(x + uh, y, s)) F h,1 (x, y, s)du+ d y V (F(x + uh, y, s)) F h,2 (x, y, s)du. A 11 (F(x, y), s) = A 12 (F(x, y), s) = With these definitions, we have d x V (F(x + uh, y, s))du, d y V (F(x + uh, y, s))du. R 1 (s) = A 11 (F(x, y), s) F h,1 (x, y, s) + A 12 (F(x, y), s) F h,1 (x, y, s). (4) Combining Equations 2 and 4 we see that F h,1 (x, y, s) = F h,1 (x, y, )+ A 11 (F(x, y), t) F h,1 (x, y, t) + A 12 (F(x, y), t) F h,1 (x, y, t)dt. We have done all this with respect to the x-coordinate. Re-doing everything with respect to the y-coordinate, we have F h,2 (x, y, s) = F h,1 (x, y, )+ A 21 (F(x, y), t) F h,1 (x, y, t) + A 22 (F(x, y), t) F h,1 (x, y, t)dt. Combining these last two equations into a matrix equation, we have F h (x, y, s) = F h (x, y, ) + A(F(x, y), t) F h (x, y, t)dt. This last equation is a matrix valued equation. This is an integral form of the basic differential equation, a special case of Equation 1. This completes the proof of Lemma 3.1. 8