Introduction to Partial Differential Equations, Math 463/513, Spring 2015

Similar documents
MATH 220: MIDTERM OCTOBER 29, 2015

Partial Differential Equations

Chapter 3 Second Order Linear Equations

z x = f x (x, y, a, b), z y = f y (x, y, a, b). F(x, y, z, z x, z y ) = 0. This is a PDE for the unknown function of two independent variables.

The first order quasi-linear PDEs

MATH 220: Problem Set 3 Solutions

Math 126 Final Exam Solutions

MATH 425, FINAL EXAM SOLUTIONS

THE WAVE EQUATION. d = 1: D Alembert s formula We begin with the initial value problem in 1 space dimension { u = utt u xx = 0, in (0, ) R, (2)

INTRODUCTION TO PDEs

CLASSIFICATION AND PRINCIPLE OF SUPERPOSITION FOR SECOND ORDER LINEAR PDE

Partial Differential Equations for Engineering Math 312, Fall 2012

PARTIAL DIFFERENTIAL EQUATIONS. Lecturer: D.M.A. Stuart MT 2007

Partial Differential Equations

Differential equations, comprehensive exam topics and sample questions

Final Exam May 4, 2016

Class Meeting # 1: Introduction to PDEs

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.

UNIVERSITY OF MANITOBA

Green s Functions and Distributions

(1) u (t) = f(t, u(t)), 0 t a.

Introduction of Partial Differential Equations and Boundary Value Problems

MATH 173: PRACTICE MIDTERM SOLUTIONS

Lecture No 1 Introduction to Diffusion equations The heat equat

Final: Solutions Math 118A, Fall 2013

Mathematical Methods - Lecture 9

Partial differential equation for temperature u(x, t) in a heat conducting insulated rod along the x-axis is given by the Heat equation:

MATH-UA 263 Partial Differential Equations Recitation Summary

Ordinary Differential Equations (ODEs)

Lecture Introduction

In this chapter we study elliptical PDEs. That is, PDEs of the form. 2 u = lots,

A proof for the full Fourier series on [ π, π] is given here.

Numerical Solutions to Partial Differential Equations

First order Partial Differential equations

Hyperbolic PDEs. Chapter 6

Lucio Demeio Dipartimento di Ingegneria Industriale e delle Scienze Matematiche

u xx + u yy = 0. (5.1)

Math 46, Applied Math (Spring 2008): Final

Partial Differential Equations

1 Assignment 1: Nonlinear dynamics (due September

Mathematical Tripos Part IA Lent Term Example Sheet 1. Calculate its tangent vector dr/du at each point and hence find its total length.

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

Various lecture notes for

Lecture Notes on PDEs

4 Divergence theorem and its consequences

MATH COURSE NOTES - CLASS MEETING # Introduction to PDEs, Spring 2018 Professor: Jared Speck

13 PDEs on spatially bounded domains: initial boundary value problems (IBVPs)

UNIVERSITY of LIMERICK OLLSCOIL LUIMNIGH

Laplace s Equation. Chapter Mean Value Formulas

LEGENDRE POLYNOMIALS AND APPLICATIONS. We construct Legendre polynomials and apply them to solve Dirichlet problems in spherical coordinates.

Salmon: Lectures on partial differential equations

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

Mathematics for Engineers II. lectures. Differential Equations

swapneel/207

The method of lines (MOL) for the diffusion equation

Various lecture notes for

4 Introduction to First-Order Partial Differential

Implicit Functions, Curves and Surfaces

1 Basic Second-Order PDEs

where F denoting the force acting on V through V, ν is the unit outnormal on V. Newton s law says (assume the mass is 1) that

Modeling Rarefaction and Shock waves

MATH COURSE NOTES - CLASS MEETING # Introduction to PDEs, Fall 2011 Professor: Jared Speck

Waves in a Shock Tube

THE WAVE EQUATION. F = T (x, t) j + T (x + x, t) j = T (sin(θ(x, t)) + sin(θ(x + x, t)))

Chapter 8: Taylor s theorem and L Hospital s rule

Math 220A - Fall 2002 Homework 5 Solutions

i=1 α i. Given an m-times continuously

MATH 32A: MIDTERM 2 REVIEW. sin 2 u du z(t) = sin 2 t + cos 2 2

Suggested Solution to Assignment 7

Ordinary Differential Equation Theory

Math 4263 Homework Set 1

Lecture 4: Numerical solution of ordinary differential equations

Definition of differential equations and their classification. Methods of solution of first-order differential equations

1 Functions of Several Variables Some Examples Level Curves / Contours Functions of More Variables... 6

R. Courant and D. Hilbert METHODS OF MATHEMATICAL PHYSICS Volume II Partial Differential Equations by R. Courant

Math Partial Differential Equations 1

MATH 220: Problem Set 3 Solutions

Math The Laplacian. 1 Green s Identities, Fundamental Solution

Vibrating-string problem

MATH FALL 2014

Numerical Approximation of Phase Field Models

Introduction and some preliminaries

Contents. MATH 32B-2 (18W) (L) G. Liu / (TA) A. Zhou Calculus of Several Variables. 1 Multiple Integrals 3. 2 Vector Fields 9

MATH 126 FINAL EXAM. Name:

An introduction to Birkhoff normal form

Mathematics of Physics and Engineering II: Homework problems

PDEs, part 1: Introduction and elliptic PDEs

Starting from Heat Equation

Math 342 Partial Differential Equations «Viktor Grigoryan

α u x α1 1 xα2 2 x αn n

for all subintervals I J. If the same is true for the dyadic subintervals I D J only, we will write ϕ BMO d (J). In fact, the following is true

LMI Methods in Optimal and Robust Control

Multivariable Calculus

1 Review of di erential calculus

2. Method of Characteristics

Introduction - Motivation. Many phenomena (physical, chemical, biological, etc.) are model by differential equations. f f(x + h) f(x) (x) = lim

Partial Differential Equations Example sheet 1

h(x) lim H(x) = lim Since h is nondecreasing then h(x) 0 for all x, and if h is discontinuous at a point x then H(x) > 0. Denote

Travelling waves. Chapter 8. 1 Introduction

Numerical Analysis of Differential Equations Numerical Solution of Parabolic Equations

Transcription:

Introduction to Partial Differential Equations, Math 463/513, Spring 215 Jens Lorenz April 1, 215 Contents Department of Mathematics and Statistics, UNM, Albuquerque, NM 87131 1 First Order Scalar PDEs 4 1.1 Linear Problems.................................... 4 1.2 Semi Linear Problems................................. 12 1.3 Burgers Equation: A First Study of Shocks..................... 13 1.4 Quasi Linear Problems................................ 18 1.5 Characteristics for General Quasilinear Equations.................. 21 1.6 General Nonlinear Scalar PDEs............................ 23 1.7 The Eikonal Equation and the Wave Equation................... 28 1.7.1 Derivation of the Eikonal Equation...................... 28 1.7.2 The Approximate Solution of the Wave Equation.............. 29 1.7.3 Terminology for Waves............................. 31 1.8 Generalizations..................................... 33 1.9 Supplement: The Implicit Function Theorem.................... 34 1.1 Notes.......................................... 35 2 Laplace s Equation and Poisson s Equation 36 2.1 Terminology....................................... 36 2.2 Volumes and Surface Areas of Balls in R n...................... 36 2.3 Integration of x λ Over Circular Regions in R n................... 39 2.4 Rotational Invariance of ; Radial Harmonic Functions.............. 4 2.5 Physical Interpretation of Poisson s Equation in R 3................. 42 2.6 Poisson s Equation in R 3................................ 44 2.6.1 The Newtonian Potential........................... 44 2.6.2 Uniqueness of Decaying Solutions....................... 47 2.6.3 Remarks on the Relation Φ = δ..................... 49 1

2.6.4 Remarks on Coulomb s Law for Point Charges............... 5 2.7 The Fundamental Solution Φ(x) for the Laplace Operator on R n......... 51 2.8 The Dirichlet Problem for Poisson s Equation in a Bounded Domain....... 52 2.9 The Green s Function for a Half Space........................ 56 2.1 The Green s Function for the Unit Ball........................ 63 2.11 Symmetry of the Green s Function.......................... 66 3 The Heat Equation and Other Evolution Equations With Constant Coefficients 69 3.1 The Cauchy Problem for the Heat Equation..................... 69 3.1.1 Solution Using the Heat Kernel........................ 69 3.1.2 Decay of the Heat Kernel Solution as x............... 74 3.1.3 Uniqueness of Decaying Solutions Via Maximum Principle......... 75 3.1.4 Derivation of the Heat Kernel Using Fourier Transformation........ 76 3.2 Evolution Equations with Constant Coefficients: Well Posed and Ill Posed Problems........................................... 79 3.2.1 Solution via Fourier Expansion........................ 79 3.2.2 The Operator Norm of the Solution Operator................ 81 3.3 Well Posed and Ill Posed Initial Value Problems for Constant Coefficient Operators 83 3.4 First Order Hyperbolic Systems in One Space Dimension............. 85 3.5 The 1D Wave Equation Written as a Symmetric Hyperbolic System....... 86 3.5.1 Derivation of d Alembert s Formula..................... 87 3.6 The Euler Equations for 1D Compressible Flow................... 88 3.7 General Constant Coefficient Systems, First Order in Time............ 9 3.8 Symmetric Hyperbolic Systems: Maxwell s Equations as an Example....... 93 3.9 An Ill Posed Problem................................. 95 3.1 Strongly Parabolic Systems in 1D........................... 96 4 The Wave Equation and First Order Hyperbolic Systems 97 4.1 The Wave Equation in 1D: d Alembert s Formula.................. 97 4.2 An Initial Boundary Value Problem for the 1D Wave Equation.......... 97 4.3 The Wave Equation in 3D: Kirchhoff s Formula................... 98 4.4 The Wave Equation in 2D: The Method of Descent................. 13 4.5 The Inhomogeneous Wave Equation: Duhamel s Principle............. 16 4.6 Energy Conservation for Symmetric Hyperbolic Systems and for the Wave Equation18 4.7 Initial Boundary Value Problems for Strongly Hyperbolic Systems in 1D..... 19 5 Parabolic Systems with Variable Coefficients: Solution by Iteration 115 5.1 Uniform Smoothness of the Functions u n (x, t).................... 116 5.2 Convergence of u n in the L 2 Norm.......................... 12 5.3 Existence of a Solution of (5.1)............................ 121 5.4 Results on Fourier Expansion............................. 123 5.4.1 Application to the System (5.3)........................ 125 2

6 Exact Boundary Conditions and Their Approximations 127 6.1 A Model Problem: Derivation of Exact Boundary Conditions........... 127 6.2 Approximation of the Exact Boundary Conditions................. 13 6.3 Introduction to Normal Mode Analysis........................ 131 6.4 Exact Solution of the Model Problem......................... 133 7 Strongly Hyperbolic Systems in 1D: The Cauchy problem and IBVPs 134 8 Extensions 137 8.1 Weak Solutions of Burgers Equation......................... 137 8.2 Piecewise Smooth Weak Solutions of Burgers Equation.............. 137 8.3 Examples........................................ 14 9 Existence of Nonlinear PDEs via Iteration 142 9.1 Setup.......................................... 142 9.2 Solution of Linear Problems via Fourier Expansion................. 143 9.3 Auxiliary Results.................................... 143 9.3.1 A Sobolev Inequality.............................. 143 9.3.2 Picard s Lemma................................ 144 9.4 A Priori Estimates................................... 146 9.5 Uniform Smoothness of the Iterates.......................... 151 9.6 Application of Arzela Ascoli.............................. 152 9.7 Convergence of the Whole Sequence......................... 154 1 Appendix 1: Fourier Expansion 156 11 Appendix 2: Fourier Transformation 157 11.1 Fourier Transform on the Schwartz Space...................... 157 11.2 Tempered Distributions................................ 16 12 Appendix 3: Fundamental Solution of Poisson s Equation via Fourier Transform 161 13 Appendix 4: The Ray Equation 163 13.1 Derivation........................................ 163 13.2 Snell s Law....................................... 164 13.3 Electromagnetic Waves in the Atmosphere...................... 165 14 Appendix 5: The Wave Equation via Fourier Transform 166 14.1 General Remarks on the Cauchy Problem...................... 166 14.2 The 1D Case via Fourier Transform......................... 167 14.3 The 3D Case via Fourier Transform......................... 17 3

1 First Order Scalar PDEs First order scalar PDEs can be solved by solving families of ODEs. This is not true anymore for higher order PDEs or for systems of first order PDEs. The general quasilinear first order equation for an unknown function u of two independent variables x, y is a(x, y, u)u x + b(x, y, u)u y = c(x, y, u). In applications, y is often the time variable, y = t, and b(x, y, u) for all arguments of interest. Then we can divide by b and achieve the standard form u t + a(x, t, u)u x = c(x, t, u). We start with some simpler special forms and discuss the following types of linear equations first: u t + au x =, a R u t + au x = F (x, t) u t + au x = c(x, t)u + F (x, t) u t + a(x, t)u x = u t + a(x, t)u x = c(x, t)u + F (x, t) It is not difficult to generalize to more than two independent variables, e.g., to pdes for u(x, y, t). 1.1 Linear Problems 1. The simplest problem. Consider the Cauchy problem 1 u t + au x =, x R, t, (1.1) u(x, ) = f(x), x R, (1.2) where f C 1 is a given function and a is a real constant. It is not difficult to check that the problem is solved by u(x, t) = f(x at) (1.3) since u t (x, t) = af (x at) and u x (x, t) = f (x at). The solution describes the propagation of the initial data f(x) at speed a. Consider a so called projected characteristic, parameterized by t, { } Γ x = (x + at, t) : t. (1.4) If we evaluate the solution u(x, t) = f(x at) on Γ x we obtain that u(x + at, t) = f(x ), 4

t Γ x x x Figure 1: Projected characteristic Γ x for a > i.e., the solution carries the initial value f(x ) along the projected characteristic Γ x. We want to show that the above Cauchy problem does not have another solution. To this end, let v(x, t) denote any C 1 solution. Fix any x R and consider the function h(t) = v(x + at, t), i.e., consider v along the projected characteristic Γ x. Obtain h (t) = (v t + av x )(x + at, t) =. Since h() = v(x, ) = f(x ) we obtain that v equals f(x ) along the projected characteristic Γ x, and therefore v agrees with the solution (1.3). 2. Addition of a forcing. Consider the Cauchy problem u t + au x = F (x, t), x R, t, (1.5) u(x, ) = f(x), x R. (1.6) As above, let f C 1, a R; we also assume F, F x C. Assume that u C 1 is a solution. Define h(t) = u(x + at, t), i.e., consider u along the projected characteristic Γ x. Obtain Also, h() = f(x ). Therefore, h (t) = (u t + au x )(x + at, t) = F (x + at, t). h(t) = f(x ) + t F (x + as, s)ds. 1 A PDE together with an initial condition is called a Cauchy problem. 5

u(x, t) t u(x + at, t ) f(x) (x + at, t ) t = t Γ x x x Figure 2: Solution of u t + au x =, u(x, ) = f(x), a > Since u(x + at, t) = h(t) we can obtain u(x, t) as follows: Given (x, t), first determine x with x + at = x, i.e., x = x at. (This determines the projected characteristic Γ x u(x, t) = h(t) = f(x at) + t which passes through x at time t.) Second, F (x at + as, s)ds. (1.7) So far, we have assumed that u solves the problem (1.5), (1.6) and have derived the above formula (1.7). Exercise: Show that the function given by (1.7) solves the problem (1.5), (1.6). Example 1.1: Solve The formula (1.7) yields u t + au x = 1, u(x, ) = sin x. u(x, t) = sin(x at) + t. We note that the forcing term 1 on the right hand side of the pde u t + au x = 1 leads to the growing term t in the solution. 3. Addition of a zero order term. Consider the Cauchy problem u t + au x = c(x, t)u + F (x, t), x R, t, (1.8) u(x, ) = f(x), x R. (1.9) Assume f C 1, c, c x, F, F x C, a R. As before, we first assume that u is a C 1 solution. This will lead us to a formula for u. We can then use the formula to show that the problem actually has a solution. 6

Consider the function Obtain h() = f(x ) and h(t) = u(x + at, t). h (t) = (u t + au x )(x + at, t) Thus h(t) satisfies a 1st order linear ODE. Recall that the ODE IVP is solved by h(t) = h exp( = c(x + at, t)h(t) + F (x + at, t). h (t) = γ(t)h(t) + g(t), h() = h, t γ(τ)dτ) + t exp( t s γ(τ)dτ)g(s)ds. Thus we can obtain the solution u(x, t) in explicit form by solving linear ODE initial value problems. Exercise: Derive the resulting formula for u(x, t) and show that it defines a solution of the Cauchy problem (1.8), (1.9). Example 1.2: Obtain thus u t + au x = u + 1, u(x, ) = sin x. h(t) = f(x )e t + t e (t s) ds = sin(x )e t + e t (e t 1), u(x, t) = sin(x at)e t + 1 e t. We note that the zero order term u on the right hand side of the pde leads to the exponetially decaying factor e t. 4. Variable signal speed a(x, t). Consider the Cauchy problem u t + a(x, t)u x =, u(x, ) = f(x), (1.1) where a, f C 1. Consider a line parameterized by t of the form Γ x = { (ξ(t), t), t }, ξ C 1, (1.11) where ξ : [, ) R is a C 1 function with ξ() = x. We first assume that ξ(t) is an arbitrary C 1 function with ξ() = x, but we will determine ξ(t) below. 7

Assume that u(x, t) solves the Cauchy problem (1.1) and consider the function h(t) = u(ξ(t), t), which is the solution u along the line Γ x. Then we obtain h() = f(x ) and Therefore, h (t) if This motivates the following definition. h (t) = (u t + ξ (t)u x )(ξ(t), t). ξ (t) = a(ξ(t), t), t. (1.12) Definition 1.1 The parameterized line (1.11) is called a projected characteristic for the equation u t + a(x, t)u x = if the function ξ(t) satisfies (1.12). Our considerations suggest to solve the problem (1.1) as follows: Let (x, t) be a given point. Determine the real number x = x (x, t) so that the solution ξ(t) = ξ(t; x ) of the IVP satisfies Then set Example 1.3: We consider the IVP The solution is Let (x, t) be a given point. We solve ξ (t) = a(ξ(t), t), ξ() = x, (1.13) ξ(t; x ) = x. u(x, t) = f(x (x, t)). u t + xu x =, u(x, ) = f(x). ξ (t) = ξ(t), ξ() = x. ξ(t; x ) = x e t. x e t = x for x = x (x, t). This yields Then we obtain Thus, our solution formula is x (x, t) = xe t. u(x, t) = f(x (x, t)) = f(xe t ). u(x, t) = f(xe t ). 8

(ξ(t;x ),t) t x x Figure 3: Characteristics ξ(t; x ) = e t x 5. Add a zero order term plus forcing. Consider the Cauchy problem u t + a(x, t)u x = c(x, t)u + F (x, t), u(x, ) = f(x). To solve the problem, first determine the projected characteristics of the problem u t +a(x, t)u x =, i.e., solve the IVPs (1.13). Denote the solution of (1.13) by ξ(t) = ξ(t; x ). Assuming that u(x, t) solves the given problem, we set Then h(t) sayisfies h() = f(x ) and h(t) = u(ξ(t; x ), t). h (t) = c(ξ(t; x ), t)h(t) + F (ξ(t; x ), t). This allows us to compute h(t) = h(t; x ). If (x, t) is a given point, we determine x = x (x, t) with ξ(t; x ) = x. Then we obtain u(x, t) = h(t; x (x, t)). Though the given Cauchy problem is linear in u, there is no guarantee that it is solvable in some time interval t T with T >. The reason is that the equation ξ (t) = a(ξ(t), t) for ξ(t) (which determines the projected characteristics) maybe nonlinear nonlinear in ξ. Example 1.4: In this example we let (x, t) R 2, i.e., we allow t to be negative. Consider the Cauchy problem u t + x 2 u x =, u(x, ) = f(x), (1.14) where f C 1. The projected characteristics } Γ x = {(ξ(t; x ), t) : t I x 9

t t = 1 x t x x x x t = 1 x (a) (b) Figure 4: Time interval I x for positive (a) and negative (b) values of x. Recall that ξ(t; x ) = x 1 x t are determined by thus ξ = ξ 2, ξ() = x, (1.15) ξ(t; x ) = x 1 x t. Here I x is the maximal interval of existence of the solution ξ(t) = ξ(t; x ) of the IVP (1.15). Clearly, ξ(t; x ) becomes singular for t = 1/x. Therefore, ξ(t; x ) solves the initial value problem (1.15) in the following time intervals: If x > we require < t < 1 x ; if x < we require 1 x < t < ; if x = then t R is not restricted, i.e., I = R. The solution u(x, t) is determined in the region R of the (x, t) plane which is covered by the projected characteristics. This is the region consisting of all points ( ) ξ(t; x ), t (1.16) where x R, t I x with I x = (, 1 x ) for x > I x = ( 1 x, ) for x < I x = R for x = Exercise: Show that the region R covered by the projected characteristics equals the region which lies strictly between the two branches of the hyperbola 1

t I x t = 1 x x Figure 5: Region of points (x, t) with x R and t I x Thus R is the shaded region in Figure 6. 1 + xt =, i.e., t = 1 x. To determine the solution u(x, t), fix any point (x, t) R. Then determine x R with x 1 x t = x. Obtain that x = x 1 + xt and note that 1 + xt for (x, t) R. This suggests that the solution of the Cauchy problem (1.14) is ( x ) u(x, t) = f for (x, t) R. 1 + xt Exercise: Show that the above function u(x, t) solves the Cauchy problem. One considers the function u(x, t) = f(x/(1 + xt)) as a solution of the initial value problem (1.14) only in the region R covered by the projected characteristics (1.16), i.e., in the region between the two branches of the hyperbola 1 + xt = or t = 1 x. In the region R + above the left branch of the hyperbola and in the region R below the right branch of the hyperbola, the formula u(x, t) = f(x/(1 + xt)) still defines a solution of the PDE u t + x 2 u x =. However, one does not consider u in R + or R as a solution of the initial value problem (1.14) since the projected characteristics of the equation which start at points (x, ) do not enter the regions R +, R. Thus, in R + and R, any solution u(x, t) of the PDE u t + x 2 u x = is completely unrelated to the initial condition u(x, ) = f(x). 11

t R + x R Figure 6: Regions R + and R 1.2 Semi Linear Problems The process described in Section 1.1 can be generalized to solve semi linear problems u t + a(x, t)u x = c(x, t, u), u(x, ) = f(x). The projected characteristics are the same as for u t + a(x, t)u x =. The IVP for h(t) = u(ξ(t; x ), t) becomes Example 1.5: We let and obtain Using separation of variables one obtains Therefore, h (t) = c(ξ(t; x ), t, h(t)), h() = f(x ). u t + au x = u 2, u(x, ) = f(x) = sin x. h(t) = u(x + at, t) h (t) = h 2 (t), h() = f(x ). h(t) = f(x ) 1 f(x )t. u(x, t) = sin(x at) 1 t sin(x at). The solution is C for t < 1, but blows up at certain x values when t 1. 12

1.3 Burgers Equation: A First Study of Shocks The equation u t + uu x =, called the inviscid Burgers equation, is a simple model for shock formation. Consider the Cauchy problem u t + uu x =, u(x, ) = sin x. Assume that u(x, t) is a solution. The projected characteristic Γ x is determined by ξ (t) = u(ξ(t), t), ξ() = x, and u carries the value sin x along Γ x. Therefore, ξ (t) = sin x, ξ() = x. One obtains that the projected characteristics Γ x are straight lines (ξ(t), t) given by ξ(t) = x + t sin x. The slope of the line Γ x depends on the initial value at x. Given a point (x, t) we must solve for x. x + t sin x = x (1.17) Lemma 1.1 If t < 1 then for each x R the equation (1.17) has a unique solution x = x (x, t). The function x (x, t) is C for x R, t < 1. Proof: Let H(y) = y + t sin y. Then H(y) as y and H(y) as y. Existence of a solution x of (1.17) follows from the intermediate value theorem. Furthermore, H (y) 1 t >. Therefore, the solution x is unique. Smoothness of the function x (x, t) follows from the implicit function theorem. Let us determine the first derivatives x x and x t in terms of the function x (x, t) through implicit differentiation: Differentiating (1.17) w.r.t x we obtain x x + t cos(x )x x = 1, thus Similarly, x x (x, t) = 1 1 + t cos x (x, t). (1.18) x t (x, t) = sin x (x, t) 1 + t cos x (x, t). (1.19) 13

Our considerations suggest that the solution of Example 1.6 is u(x, t) = sin x (x, t), where x (x, t) is the solution of (1.17). We can check this: Clearly, x (x, ) = x for all x, thus u(x, ) = sin x (x, ) = sin x. Furthermore, u t = (cos x )x t and u x = (cos x )x x. The formulas (1.18) and (1.19) then imply that u t + uu x =. Remarks on shock formation and weak solutions: Consider the Cauchy problem u t + uu x =, u(x, ) = sin x. We have shown that the problem has a C solution u(x, t) defined for x R, t < 1. However, if t > 1, then the projected characteristics { } Γ x = (x + t sin x, t) : t intersect. For example, consider Γ π = { } (π, t) : t and Γ π+ε = { } (π + ε + t sin(π + ε), t) : t for < ε << 1. By Taylor expansion, sin(π + ε) = ε + ε3 6 + O(ε5 ). The two projected characteristics Γ π and Γ π+ε intersect at time t > 1 if i.e., if t = π = π + ε + t sin(π + ε), ε sin(π + ε) = 1 1 ε2 6 + O(ε4 ). We obtain that intersection occurs for times t > 1, where t is arbitrarily close to the time 1. The constructed solution u(x, t) carries the value u(π, t) = sin π = along Γ π and carries the different value sin(π + ε) along Γ π+ε. This shows that a smooth solution u(x, t) of the above Cauchy problem does not exist in R [, T ) if T > 1. A shock forms at t = 1. It is interesting, however, to broaden the solution concept and consider so called weak solutions. The idea is the following: Write Burgers equation as u t + 1 2 (u2 ) x =. 14

Let φ C (R2 ) denote a test function. First assume that u(x, t) is a smooth solution of Burgers equation with u(x, ) = f(x). Multiply Burgers equation by φ(x, t) and integrate to obtain = ( u t φ + 1 ) 2 (u2 ) x φ dxdt. Now move the derivative operators from u to φ through integration by parts and obtain (uφ t + 1 2 (u2 )φ x ) dxdt + f(x)φ(x, ) dx =. (1.2) One calls u a weak solution of Burgers equation with initial condition u(x, ) = f(x) if the above equality (1.2) holds for all test functions φ C (R2 ). The integrals in (1.2) all make sense if, for example, u L (R [, )). Thus, weak solutions do not necessarily have classical derivatives. Of special interest are piecewise smooth weak solutions. These satisfy the differential equation classically in regions where they are smooth and satisfy the so called Rankine Hugoniot jump condition along shocks, which are lines of discontinuity. Auxiliary Results: Green Gauss Theorem Let U denote a bounded open subset of R k with smooth boundary U. Denote the unit outword normal on U by n(x) for x U. Let g : U R denote a smooth function, which can be smoothly extended to the boundary U. We denote the extended function again by g. Then the following formula holds: U x j g(x) dx = U g(x)n j (x) ds(x) for j = 1,..., k. Here ds(x) denotes the surface measure on the boundary U. Let us specialize the result to a region U R 2 and let us assume that the boundary curve Γ of U has the parametrization (x 1 (s), x 2 (s)), s L, by arclength s. We assume that this parametrization is counterclockwise. Then the unit outward normal is We obtain the formulas n(x 1 (s), x 2 (s)) = (x 2(s), x 1(s)). U x 1 g(x) dx = = Γ L g(x)n 1 (x) ds(x) g(x 1 (s), x 2 (s)) dx 2(s) ds ds 15

and = Γ g(x 1, x 2 ) dx 2 If we use the notation U x 2 g(x) dx = g(x)n 2 (x) ds(x) Γ L = = Γ g(x 1 (s), x 2 (s)) dx 1(s) ds g(x 1, x 2 ) dx 1 ds then the resulting formulas become (x 1, x 2 ) = (x, t) U U g x (x, t) dxdt = g(x, t) dt (1.21) Γ g t (x, t) dxdt = g(x, t) dx (1.22) Derivation of the Rankine Hugoniot Jump Condition Let F : R R denote a smooth function, a so called flux function. In the case of Burgers equation, we have F (u) = 1 2 u2. Let u(x, t) denote a weak solution of This means that u t + F (u) x =, u(x, ) = f(x). Γ ) (uφ t + F (u)φ x dxdt + f(x)φ(x, t) dx = (1.23) for all φ C (R2 ). Let U R (, ) denote an open bounded set and let Γ denote a curve crossing U, which has a parametrization Write (x(t), t), t t t 1. U = U l U r where U l lies to the left and U r lies to the right of Γ. 16

Let φ C (U). We have = = (uφ t + F (u)φ x ) dxdt U... +... U l U r Assume that the weak solution u(x, t) is a smooth classical solution of u t + F (u) x = in U l and in U r which has a jump discontinuity along Γ. Using the Green Gauss formulas (1.21) and (1.22) we have (F (u)φ) x (x, t) dxdt = (F (u l )φ)(x, t) dt U l Γ (F (u)φ) t (x, t) dxdt = (F (u l )φ)(x, t) dx U l Γ Since u t + F (u) x = in U l we obtain (uφ t + F (u)φ x )(x, t) dxdt = (F (u l ) dt u l dx)φ. U l Γ If we integrate over U r instead of U l we obtain in the same way: (uφ t + F (u)φ x )(x, t) dxdt = (F (u r ) dt u r dx)φ. U r Γ The reason for the minus sign is that the direction of Γ agrees with the orientation of U l, but is opposite to the orientation U r. Adding the last two equations and recalling that u is a weak solution, we obtain that (F (u l ) dt u l dx)φ = (F (u r ) dt u r dx)φ. Since φ is an arbitray function in C Γ Γ (U) the above equation implies that along Γ. Therefore, F (u l ) dt u l dx = F (u r ) dt u r dx (u r u l ) dx dt = F (u r) F (u l ) along Γ. This is the Ranking Hugoniot jump condition. It couples the shock speed dx dt jump of u and the jump of F (u) along a shock curve Γ. We have shown: 17 to the

Theorem 1.1 Let u denote a piecewise smooth weak solution of u t + F (u) x = which has a jump discontinuity along a curve Γ with parametrization (x(t), t), t t t 1. Then we have (u r u l )(x(t), t) dx(t) dt = (F (u r ) F (u l ))(x(t), t), t t t 1. 1.4 Quasi Linear Problems We now consider equations where the signal speed a also depends on u. The general form of the equation is u t + a(x, t, u)u x = c(x, t, u). (1.24) Example 1.7: u t + uu x = u, u(x, ) = sin x. Assume that u(x, t) is a solution. The projected characteristic Γ x is determined by ξ (t) = u(ξ(t), t), ξ() = x. Set h(t) = u(ξ(t), t). Then we have h() = u(x, ) = sin x and Thus we must solve the system h (t) = u t + ξ (t)u x = (u t + uu x )(ξ(t), t) = h(t). ξ (t) = h(t), h (t) = h(t) with initial condition Obtain and ξ() = x, h() = sin x. h(t; x ) = e t sin x ξ(t; x ) = x + (1 e t ) sin x. 18

Given a point (x, t) we must solve for x. Denote the solution by x (x, t). Then x + (1 e t ) sin x = x (1.25) u(x, t) = h(t; x ) solves the problem. Note: If H(y) = y + (1 e t ) sin y then = e t sin x (x, t) H (y) 1 (1 e t ) = e t > for all t. Therefore, (1.25) is uniquely solvable for x for any given t and all x R. The solution x (x, t) is C by the implicit function theorem, for all t. Therefore, u(x, t) is C for all t. Exercise: Use implicit differentiation to compute the derivatives x t and x x in terms of x (x, t). Then show that the function u(x, t) = e t sin x (x, t) solves the problem. In principle, the general equation (1.24) with initial condition u(x, ) = f(x) can be solved in the same way: We consider the system of ODEs with initial condition Denote the solution by Given a point (x, t) we solve ξ (t) = a(ξ(t), t, h(t)), h (t) = c(ξ(t), t, h(t)) (1.26) ξ() = x, h() = f(x ). ξ(t; x ), h(t; x ). ξ(t; x ) = x for x. Assume that there is a unique solution x whenever t T and x R. Denote the solution by x (x, t). Then we expect that solves the problem. We have ξ(; x ) = x for all x. Therefore, if we solve u(x, t) = h(t; x (x, t)) (1.27) ξ(; x ) = x for x, then the solution is x = x. This ensures that x (x, ) = x 19

for all x. Obtain u(x, ) = h(; x (x, )) = h(; x) = f(x). Thus the function (1.27) satisfies the initial condition. We want to discuss now under what assumptions u(x, t) (given by (1.27)) satisfies the PDE u t + a(x, t, u)u x = c(x, t, u). To this end, let us assume that the implicitly defined function x (x, t) is C 1. We make the following formal computations: thus u x = h x x x, u t = h t + h x x t, u t + au x = h t + h x x t + ah x x x = c + h x (x t + ax x ). The function x satisfies thus and This yields and Since ξ t = a it follows that ξ(t; x (x, t)) = x, ξ x x x = 1 ξ t + ξ x x t =. x x = 1 ξ x x t = ξ t ξ x. x t + ax x =, which implies u t + au x = c. These computations can be justified as long as for t T, x R. ξ x (t; x ) Remark: Reconsider Example 1.4, u t + x 2 u x =, where ξ(t; x ) = x 1 x t 2

and ξ x (t; x ) = 1 (1 x t) 2. Though the condition ξ x is always satisfied, there is no time interval of existence for u(x, t) since there is no time interval t T, T >, where ξ(t; x ) is defined for all x R. The trouble is due to the strong growth of the coefficient a = x 2. 1.5 Characteristics for General Quasilinear Equations Consider the equation a(x, t, u)u x + b(x, t, u)u t = c(x, t, u) (1.28) where a, b, c C 1. A parametrized C 1 curve (in R 3 ) ( ) x(s), t(s), ū(s), a s b, (1.29) is called a characteristic of equation (1.28) if d x ds d t ds dū ds = a( x, t, ū) (1.3) = b( x, t, ū) (1.31) = c( x, t, ū) (1.32) The corresponding curve ( ) x(s), t(s), a s b, (1.33) in the (x, t) plane is called a projected characteristic. Characteristics are important because of the following result. Theorem 1.2 Suppose that u(x, t) is a solution of (1.28) defined in a neighborhood of the projected characteristic (1.33). Also, assume that u( x(a), t(a)) = ū(a), i.e., the starting point of the characteristic (1.29) lies in the solution surface S R 3 determined by u. Under these assumptions the whole characteristic (1.29) lies in S. Proof: In the following the argument s of the functions x etc. notation. Set U(s) = ū(s) u( x(s), t(s)). is often suppressed in the 21

u(x, t) ū( x(s), t(s)) t ū(a) s = b s = a ( x(s), t(s)) x Figure 7: Theorem 1.1 By assumption, U(a) =. Also, du ds = ū u x ( x, t) x u y ( x, t) t = c( x, t, ū) u x ( x, t)a( x, t, ū) u y ( x, t)b( x, t, ū) = c( x, t, U + u( x, t)) u x ( x, t)a( x, t, U + u( x, t)) u y ( x, t)b( x, t, U + u( x, t)) Given the solution u(x, t) and the curve ( x(s), t(s)), we may consider the above equation as an ODE for the unknown function U = U(s). Note that U solves this ODE since u solves (1.28). Since U(a) =, uniqueness of solutions of ordinary initial value problems implies that U, thus u( x(s), t(s)) ū(s). The theorem says that all solution surfaces { } S = (x, t, u(x, t)) of the PDE (1.28) are made up of characteristic curves. Therefore, the solutions of the PDE (1.28) can be constructed by solving the ODE system (1.3) to (1.32). Previously we have considered the special case b 1. If b 1, then equation (1.31) reads d t ds = 1 and we can choose the variable t as parameter of any characteristic; we do not need the abstract parameter s. With b 1 and t = s the characteristic system (1.3) to (1.32) becomes d x dt = a( x, t, ū), dū dt = c( x, t, ū). 22

This is exactly the same system as in our previous notation; see (1.26). 1.6 General Nonlinear Scalar PDEs ξ = a(ξ, t, h), h = c(ξ, t, h) Let F (x, y, u, p, q) be a smooth function of five variables. We denote the partial derivatives of F by F x etc. Consider the PDE F (x, y, u, u x, u y ) =. (1.34) Suppose that u(x, y) is a solution and set p(x, y) = u x (x, y), q(x, y) = u y (x, y). For later reference we derive the equations (1.36) and (1.37) below. If one differentiates the identity w.r.t. x, then one obtains thus Similarly, by differentiating (1.35) w.r.t. y, F (x, y, u(x, y), p(x, y), q(x, y)) = (1.35) F x + F u u x + F p p x + F q q x =, p x F p q x F q = F x + pf u. (1.36) p y F p q y F q = F y + qf u. (1.37) We want to show that, in principle, the PDE (1.34) can again be solved by solving a system of ODEs. The characteristic system for (1.34) consists of the following five coupled ODEs: d x ds dȳ ds dū ds d p ds d q ds = F p (1.38) = F q (1.39) = pf p + qf q (1.4) = F x pf u (1.41) = F y qf u (1.42) 23

where F p = F p ( x, ȳ, ū, p, q), etc. An instructive example is given by the so called Eikonal equation of geometrical optics: Example 1.8: u 2 x + u 2 y = 1 Here thus The characteristic system becomes F (p, q) = 1 2 (p2 + q 2 1), F x = F y = F u =, F p = p, F q = q. d x ds = p dȳ ds = q dū ds = p 2 + q 2 d p ds = d q ds = We discuss this system below. Let us first continue the discussion of the general case. Suppose that ( ) x(s), ȳ(s), ū(s), p(s), q(s), a s b, (1.43) is a solution of the characteristic system (1.38) to (1.42). Such a solution is called a characteristic for (1.34). Let u(x, y) be a solution of (1.34) and recall the settings p = u x, q = u y. Suppose that u(x, y) is defined in a neighborhood of the projected characteristic ( ) x(s), ȳ(s), a s b. Theorem 1.3 Under the above assumptions, suppose that u( x(a), ȳ(a)) = ū(a) p( x(a), ȳ(a)) = p(a) q( x(a), ȳ(a)) = q(a) Then the whole characteristic (1.43) lies in the solution surface corresponding to u, and we have for a s b, u( x(s), ȳ(s)) = ū(s) p( x(s), ȳ(s)) = p(s) q( x(s), ȳ(s)) = q(s) 24

Proof: The proof is similar to the proof of Theorem 1.2. Set U(s) = ū(s) u( x(s), ȳ(s)) P (s) = p(s) p( x(s), ȳ(s)) Q(s) = q(s) q( x(s), ȳ(s)) By assumption, U(a) = P (a) = Q(a) =. We calculate (note that p y = u xy = u yx = q x ) U = ū u x x u y ȳ = pf p + qf q pf p qf q P = p p x x p y ȳ = F x pf u p x F p p y F q = F x pf u p x F p q x F q Q = q q x x q y ȳ = F y qf u q x F p q y F q = F y qf u p y F p q y F q Here all functions F x etc. are evaluated at ( ) x(s), ȳ(s), ū(s), p(s), q(s). Recall that, by (1.36) and (1.37), p x F p q x F q = F x + pf u and p y F p q y F q = F y + qf u. The left hand side p x F p q x F q appears on the right hand side of the equation for P if one ignores the difference between variables with and without bars. Similarly, p y F p q y F q appears on the right hand side for Q. If one could ignore the difference between variables with and without bars, then the above equations would read U = P = Q =. Using the same argument as in the proof of Theorem 1.2 (based on the unique solvability of ordinary initial value problems), it follows that Example 1.8 (continued) U P Q =. 25

The characteristic equations are solved by x(s) = x() + p()s ȳ(s) = ȳ() + q()s ū(s) = ū() + ( p 2 () + q 2 ())s p(s) = p() q(s) = q() Suppose that Γ : (f(σ), g(σ)), σ L, is a smooth curve in the (x, y) plane parametrized (for simplicity) by arclength σ, thus ( ) 2 ( 2 f (σ) + g (σ)) = 1. We want to determine a solution u(x, y) of the eikonal equation satisfying u = on Γ, i.e., u(f(σ), g(σ)) =, σ L. (1.44) Suppose first that u(x, y) is a solution of the PDE u 2 x + u 2 y = 1 satisfying the side condition (1.44); set p = u x, q = u y. Furthermore, set Differentiating (1.44) w.r.t. σ we obtain Also, the PDE u 2 x + u 2 y = 1 requires φ(σ) = p(f(σ), g(σ)), ψ(σ) = q(f(σ), g(σ)). f (σ)φ(σ) + g (σ)ψ(σ) =. (1.45) φ 2 (σ) + ψ 2 (σ) = 1. (1.46) We consider φ(σ) and ψ(σ) as unknowns in (1.45), (1.46), and note that there are two solutions of (1.45), (1.46), namely (φ, ψ) = ( g, f ) and (φ, ψ) = (g, f ). For definiteness, let us assume that the solution u(x, y) under consideration corresponds to the first solution, φ(σ) = g (σ), ψ(σ) = f (σ). The vector (φ(σ), ψ(σ)) is a unit normal to the curve Γ in the point P σ = (f(σ), g(σ)). 26

Fix the point P σ on the curve Γ. We want to start a projected characteristic at this point and want the whole characteristic to lie in the solution surface. Then the corresponding initial condition for the characteristic system is x() = f(σ) ȳ() = g(σ) ū() = p() = φ(σ) q() = ψ(σ) The solution of the characteristic system with this initial condition is x(s) = f(σ) + φ(σ)s ȳ(s) = g(σ) + ψ(σ)s ū(s) = s p(s) = φ(σ) q(s) = ψ(σ) For the solution u(x, y) under consideration, this implies that ( ) u f(σ) + φ(σ)s, g(σ) + ψ(σ)s = s. (1.47) This says that, along any straight line, ( ) f(σ) + φ(σ)s, g(σ) + ψ(σ)s, s R, (1.48) the solution u equals s, as long as the solution u exists. The straight line (1.48) is a projected characteristic. As in the quasilinear case, projected characteristics can intersect, which generally leads to breakdown of the solution u(x, y). In regions where the projected characteristics do not intersect, they can be used to determine a solution u(x, y). If, in the above example, we fix s and vary σ, we obtain the curve ( ) Γ s : f(σ) + φ(σ)s, g(σ) + ψ(σ)s, σ L, (1.49) which is roughly parallel to the given curve Γ. On Γ s the solution u carries the value s, i.e., Γ s is a level curve of u. If we think of s as time, then the mapping Γ Γ s describes a motion of the given initial curve Γ in the (x, y) plane. A Particular Case: Let Γ : (f(σ), g(σ)) = (cos σ, sin σ), σ 2π, 27

be the parametrized unit circle. Then Γ s is the circle (1 s)(cos σ, sin σ), σ 2π, of radius 1 s. On this circle, u equals s. Setting r = x 2 + y 2 and noting that 1 s = r implies s = 1 r, one obtains u(x, y) = 1 x 2 + y 2. If instead of (φ, ψ) = ( g, f) the solution u(x, y) corresponds to (φ, ψ) = (g, f), then one obtains u(x, y) = x 2 + y 2 1. This example shows that the solution of a nonlinear problem with a side condition F (x, y, u, u x, u y ) = u(f(σ), g(σ)) = h(σ) is generally not unique, not even locally near the curve Γ : (f(σ), g(σ)). 1.7 The Eikonal Equation and the Wave Equation 1.7.1 Derivation of the Eikonal Equation Consider the 2D wave equation where c = c(x, y). Let v tt = c 2 (v xx + v yy ) v(x, y, t) = V (x, y)e iωt, i.e., we consider a solution of the wave equation which oscillates in time at the frequency ω. One obtains Helmholtz equation for the amplitude function V (x, y): V xx + V yy + ω2 c 2 V =. Let c denote a constant reference value for the wave speed. For example, c may be the speed of light in vacuum and c(x, y) the speed of light in a material. The number is called the index of refraction. We have n(x, y) = c c(x, y) ω ωn(x, y) = = kn(x, y) with k = ω. c(x, y) c c 28

Here k = ω c is a wave number. The equation for V (x, y) becomes V xx + V yy + k 2 n 2 (x, y)v =. (1.5) The following values are approximate values for visible light: c = 3 1 8 m sec λ = 4 1 7 m (wavelength) ω = c λ = 7.5 114 sec 1 k = ω c = 1 λ = 2.5 16 m 1 Thus, if we choose 1m as the unit of length, then the wave number k is very large. We try to find solutions of (1.5) of the form V (x, y) = A(x, y)e iku(x,y) where u(x, y) is the phase function and A(x, y) the amplitude. Setting E = e iku(x,y) we have V x = A x E + AEiku x V xx = A xx E + 2A x Eiku x + AE( k 2 )(u x ) 2 + AEiku xx and similar equations hold for V y, V yy. After dividing by E equation (1.5) becomes ( ) ( ) A + ik 2A x u x + 2A y u y + A u k 2 A (u x ) 2 + u y ) 2 n 2 (x, y) =. If k >> 1 it is reasonable to neglect the term A and to solve the two equations (u x ) 2 + (u y ) 2 = n 2 (x, y) 2A x u x + 2A y u y + A u = For the phase function u(x, y) we have obtained the eikonal equation. If u(x, y) is known, then the second equation is a transport equation for A(x, y). If one does not neglect the term A, then one obtains a singular perturbation problem for A. 1.7.2 The Approximate Solution of the Wave Equation For simplicity, let be constant. The eikonal equation is c(x, y) = c = c 29

(u x ) 2 + (u y ) 2 = 1. Let u(x, y) denote a solution with u = on Γ where Γ has the arclength parametrization Let Γ s denote the curve with parametrizaton We have obtained that (Case 1): or (Case 2): Γ : (f(σ), g(σ)), σ L. Γ s : (f(σ), g(σ)) + s( g (σ), f (σ)), σ L. u(f(σ), g(σ)) + s( g (σ), f (σ)) = s u(f(σ), g(σ)) + s( g (σ), f (σ)) = s. For definiteness, assume Case 1. The corresponding approximated solution of the wave equation is v(x, y, t) = A(x, y)e ik(u(x,y) ct). Here A(x, y) does not depend on k and we may expect that A(x, y) varies slowly on the length scale 1/k. Consider the real part of the exponential, Fix a straight line a(x, y, t) = cos(ku(x, y) ct)). (x(s), y(s)) = (f(σ), g(σ)) + s( g (σ), f (σ)), s s s, (1.51) where σ is fixed. On this line we have thus u(x(s), y(s)) = s, φ(s, t) := a(x(s), y(s), t) = cos(k(s ct)). Since k >> 1, the function φ(s, t) is highly oscillator. It moves at speed c in positive s direction. Therefore, the function a(x, y, t) moves at speed c along the straight line (1.51). The lines (1.51) are the light rays. The lines Γ s are the wave fronts. 3

1.7.3 Terminology for Waves Plane Waves: Consider the wave equation u tt = c 2 u in n space dimensions. Let F : R C denote a smooth function and fix It is easy to check that the function k R n with k = 1. u(x, t) = F (k x ct), x R n, t R, (1.52) solves the wave equation. We want to explain that the above function u(x, t) describes a plane wave which moves at speed c in direction k. Let H = {ξ R n : k ξ = } denote the hyperplane through the origin orthogonal to k and for any real positive α let H α = {x R n : k ξ = α} = {x = ξ + αk : ξ H } denote the hyperplane parallel to H with distance α from H. At time t = we have and at time t > we have u(x, ) = F (k x) = F (α) for x H α u(x, t) = F (k x ct) = F (β ct) for x H β. If β = α + ct then we have F (β ct) = F (α). We can interprete this as follows: At time t = the function (1.52) has the value F (α) on the plane H α and at time t > it has the same value F (α) on the plane H α+ct. The value F (α) has moved from H α to H α+ct, i.e., it has moved at speed c in direction k. Since the function u(x, t) = F (k x ct) is constant on each plane H α (at fixed t), it is called a plane wave. Plane Waves With Periodic F : Consider the function where, as above, k R n, k = 1, and let κ >. Choose κ = 5, for example. Then we have u(x, t) = cos(κ(k x ct)) (1.53) 31

u(x, ) = cos(5α) for x H α. For α 2π the function cos(5α) has 5 waves with wave length λ = 2π 5. In general, the number κ in (1.53) is called the wave number. The wave length is λ = 2π κ. We can also write the function (1.53) in the form u(x, t) = cos(κk x ωt)) with ω = κc. (1.54) Here ω = κc is the circular frequency of the wave. The time period of the oscillation (1.54) is T = 2π ω. The frequency is and we can write f = 1 T = ω 2π u(x, t) = cos(κk x 2πft)). Data For Visible Light: The following approximate values hold for visible light in vacuum: Speed: c = 3 1 8 meter sec Frequency: f = 6 1 14 1 sec In fact, visible light has (approximately) frequencies in the rather narrow band 4 1 14 1 sec f 8 114 1 sec. Time Period: T = 1 f = 1 6 1 14 sec 32

Wave Number: κ = ω c = 2πf c The number of waves in meter length is 2 1 6. = 4π 1 6 1 meter Wave Length: λ = 2π κ =.5 1 6 meter = 5 nanometer 1.8 Generalizations More than two independent variables. There are no difficulties to generalize the approach via characteristics to problems where the unknown function u depends on more than two independent variables. Example 1.9: Consider u t + a(x, y, t)u x + b(x, y, t)u y = with initial condition u(x, y, ) = f(x, y). Projected characteristics are now parametrized lines of the form where (ξ 1 (t), ξ 2 (t), t), t, ξ 1 = a(ξ 1, ξ 2, t), ξ 2 = b(ξ 1, ξ 2, t). Let u(x, y, t) solve the differential equation and consider u along a projected characteristic: Clearly, h(t) = u(ξ 1 (t), ξ 2 (t), t). h (t) = u t + ξ 1(t)u x + ξ 2(t)u y =. Thus u carries the initial data along the projected characteristics. This can again be used conversely to construct solutions. Data not given at t =. If the data are not given at t = it is often good (and sometimes necessary) to parameterize the projected characteristics by some parameter s, not by t, even if the PDE can be solved for u t. Example 1.1: Consider u t + u x = u 2 for u = u(x, t) with side condition u(x, x) = x for all x R. The projected characteristics are the straight lines (x + s, x + s), s R. 33

Assume that u is a solution and consider u along a projected characteristic, One finds that h = h 2 and h(s) = u(x + s, x + s). h() = u(x, x ) = x. Therefore, h(s) = x. 1 sx If (x, t) is a given point we determine x and s with This yields and therefore x + s = x, x + s = t. x = 1 2 (x t), s = 1 (x + t), 2 2(x t) u(x, t) = 4 + t 2 x 2. Note that u blows up at all points of the hyperbola t 2 = x 2 4. 1.9 Supplement: The Implicit Function Theorem Let U R n, V R m denote nonempty open sets and let Φ : U V R m be a C p function, p 1. For every fixed u U the system Φ(u, v) = (1.55) consists of m equations for m variables, the m components of v. It is reasonable to expect that under certain assumptions for any fixed u there is a unique solution v, and that v = v(u) changes smoothly as a function of u. Then the equation (1.55) defines v implicitly as a function of u. The implicit function theorem gives a condition ensuring this to be correct locally near a solution (u, v ) U V of the equation Φ(u, v) =. Theorem 1.4 (Implicit Function Theorem) Under the above assumptions, let (u, v ) U V and assume the following two conditions: Φ(u, v ) =, det(φ v (u, v )). (1.56) (Note that Φ v (u, v ) is an m m matrix.) Then there are open sets U, V with u U U, v V V so that the following holds: For all u U the equation (1.55) has a solution v V, which is unique in V. In addition, the function u v, which assigns to each u U the corresponding solution v V, is C p. 34

Application: Consider the equation x + t sin x = x for x R and t < 1. We have already shown that there is a unique solution x = x (x, t) and want to argue that the assignment (x, t) x (x, t) is C. This follows from the implicit function theorem with U = R ( 1, 1), V = R, and Note that Φ(x, t, x ) = x + t sin x x. Φ x = 1 + t cos x. 1.1 Notes The PDE F (x, u, u) where x R n is treated in Evans, Partial Differential Equations, section 3.2. A side condition u = h on Γ is assumed given. Here Γ is a submanifold of R n of dimensions n 1. Using a transformation, the side condition and PDE are transformed so that the new Γ becomes a subset of the hyperplane x n =. Then, under the assumption that the new boundary Γ is noncharacteristic, a local solution is constructed via the implicit function theorem. For details, see Evans. A geometric interpretation of the PDE F (x, y, u, u x, u y ) using Monge cones is in John, Partial Differential Equations. 35

2 Laplace s Equation and Poisson s Equation 2.1 Terminology Let D j = / x j. Then = n j=1 is the Laplace operator acting on functions u(x), x Ω R n. Here, typically, Ω is an open subset of R n. The equation u = in Ω (2.1) is called Laplace s equation; its solutions are called harmonic functions in Ω. The inhomogeneous equation u = f(x) in Ω (2.2) is called Poisson s equation. Typically, the equations (2.1) and (2.2) have to be solved together with boundary conditions for u, or with decay conditions if Ω is unbounded. For example, the problem u = f(x) in Ω, u = g(x) on Ω, (2.3) is called a Dirichlet problem for Poisson s equation. Here Ω is the boundary of the set Ω. 2.2 Volumes and Surface Areas of Balls in R n The fundamental solution Φ(x) for the Laplace operator in R n will be discussed below; Φ(x) depends on a constant ω n, the surface area of the unit sphere in R n. In this section we compute ω n and the volume of the unit ball in R n. Notations: For x, y R n let the Euclidean inner product be denoted by D 2 j x, y = x y = j x j y j. The Euclidean norm is x = (x x) 1/2. Let B r = B r (R n ) = {x R n : x < r} denote the open ball of radius r centered at. Its surface is B r = B r (R n ) = {x R n : x = r}. We will use Euler s Γ function, which may be defined by Γ(s) = t s 1 e t dt, Re s >. 36

(We will need the function only for real s >.) Let ω n = area( B 1 (R n )), i.e., ω n is the surface area of the unit ball in R n. For example, ω 2 = 2π, vol(b 1 (R 2 )) = π = ω 2 /2 and ω 3 = 4π, vol(b 1 (R 3 )) = 4π 3 = ω 3/3. Theorem 2.1 The following formulas are valid: ω n = 2πn/2 Γ(n/2) (2.4) vol(b 1 (R n )) = ω n (2.5) n area( B r (R n )) = ω n r n 1 (2.6) vol(b r (R n )) = ω n n rn (2.7) Formula (2.6) says that the area of the (n 1) dimensional surface B r scales like r n 1. Similarly, (2.7) says that the n dimensional volume of B r scales like r n. These are results of integration theory. We will prove (2.4) and (2.5) below. First we evaluate an important integral and show some simple properties of the Γ function. Lemma 2.1 I 1 := e x2 dx = π (2.8) Γ(s + 1) = sγ(s), Re s > (2.9) Γ(m + 1) = m!, m =, 1, 2,... (2.1) Proof: The proof of (2.8) uses a trick and polar coordinates: I1 2 = dxdy Γ( 1 2 ) = π (2.11) = = π = π = π x2 y2 e R 2 2π 37 re r2 drdφ 2re r2 dr e ρ dρ

Note that we have used the result that the circumference of the unit circle is 2π, which may be taken as the definition of π. Equation (2.9), which is the fundamental functional equation of the Γ function, follows through integrating by parts: This proves (2.9). Also, Γ(s + 1) = t s e t dt = t s ( e t ) + s = sγ(s) Γ(1) = e t dt = 1, t s 1 e t dt and then (2.1) follows from (2.9) by induction. Finally, (2.11) follows from (2.8) by substituting t = x 2, dt = 2xdx, This completes the proof of Lemma 2.1. Γ( 1 2 ) = t 1/2 e t dt = 2 = π 1 x xe x2 dx We now calculate ω n by considering the integral of e x 2 over R n. By Fubini s theorem and (2.8) we know that the result is π n/2 : π n/2 = dx = e x 2 R n area( B r )e r2 dr = ω n r n 1 e r2 dr = ω n 2 = ω n 2 = ω n 2 Γ(n/2) r n 2 2re r2 dr t n 2 1 e t dt 38

This proves (2.4). Also, vol(b 1 (R n )) = 1 area( B r ) dr 1 = ω n r n 1 dr = ω n /n Note that one can use the functional equation Γ(s + 1) = sγ(s) and the two equations Γ(1/2) = π, Γ(1) = 1 to evaluate Γ(n/2) for every integer n. For example, ( 3 Γ 2) ( 1 ) = Γ 2 + 1 = 1 ( 1 ) 2 Γ 2 etc. ( 5 Γ 2) ( 7 Γ 2) = 1 π 2 ( 3 ) = Γ 2 + 1 = 3 ( 3 ) 2 Γ 2 = 1 3 π 2 2 ( 5 ) = Γ 2 + 1 = 1 3 5 π 2 2 2 2.3 Integration of x λ Over Circular Regions in R n For λ < the function x λ, x R n, x, has a singularity at x =. It is important to understand when the singularity is locally integrable. It is also important to know when the function x λ decays fast enough as x to make x λ integrable over an outer region x a >. Let < a < b < and consider the n dimensional integral I(a, b, λ, n) := x λ dx a< x <b b = ω n r λ+n 1 dr a { b ω λ+n a λ+n = n λ+n if λ + n ω n log b a if λ + n = 39

Consider the limit a with b > fixed. One obtains { x λ b ω λ+n dx = n λ+n if λ > n if λ n x <b For > λ > n the singularity of the function x λ at x = is integrable; for λ n the singularity is so strong that x λ is not integrable over B b. Now consider the limit of I(a, b, λ, n) as b with a > fixed. One obtains x >a x λ dx = { a ω λ+n n λ+n if λ < n if λ n If λ < n, then, as x, the function x λ decays fast enough to be integrable over the domain R n \ B a ; for λ n the decay is too slow to make the function integrable over R n \ B a. Note that there is no value of λ for which the function x λ is integrable over R n : Either the singularity at x = is too strong (for λ n) or the decay as x is too slow (for λ n). If λ = n then the singularity is too strong and the decay is too slow. 2.4 Rotational Invariance of ; Radial Harmonic Functions A remarkable property of the Laplace operator = n j=1 D2 j is its invariance under an orthogonal change of coordinates. We make precise what this means and prove it. This invariance motivates to look for radial solutions of Laplace s equation, u =. Here u(x) is called a radial function, if there is a function v(r), depending only on a single variable < r <, with u(x) = v( x ) for all x R n, x. The radial solutions of Laplace s equation will lead to the fundamental solution. The Laplacian is invariant under orthogonal transformations of the coordinate system. A precise statement is the following. Theorem 2.2 Let u C 2 (R n ) and let S denote an orthogonal n n matrix, SS T = I. Let v(y) = u(sy), y R n, i.e., v(y) is obtained from u(x) by changing from x coordinates to y coordinates. Then we have ( v)(y) = ( u)(sy) for all y R n. The theorem says that the following two operations lead to the same result: A: First change from x coordinates to y coordinates (where x = Sy, SS T = I) and then take the Laplacian; B: First take the Laplacian and then change from x coordinates to y coordinates. Proof: First note that (Sy) j = s jl y l, l 4

thus Since v(y) = u(sy) we have by the chain rule: D k (Sy) j = s jk where D k = / y k D k v(y) = j (D j u)(sy)d k (Sy) j = j (D j u)(sy)s jk Therefore, Summation over k yields Here and the claim follows. D 2 k v(y) = j v(y) = k (D i D j u)(sy)s ik s jk. i (D i D j u)(sy)s ik s jk. j i s ik s jk = δ ij k The Laplacian applied to radial functions. Let v(r), r >, be a C 2 function and let u(x) = v( x ), x R n. Then Proof: We have u(x) = v (r) + n 1 r v (r) with r = x. D j u(x) = v (r) x j r D 2 j u(x) = v (r) x2 j r 2 + v (r) Summation over j from 1 to n yields the result. ( 1 r x2 ) j r 3 We can now determine all solutions of u = in R n \ {} which are radially symmetric. We must solve v + n 1 v =, r thus v = w + const where w + n 1 w =. r 41