Harmonic Functions Kurt Bryan MA 436

Similar documents
MATH 425, PRACTICE FINAL EXAM SOLUTIONS.

3 Contour integrals and Cauchy s Theorem

Practice with Proofs

DERIVATIVES AS MATRICES; CHAIN RULE

1 Completeness of a Set of Eigenfunctions. Lecturer: Naoki Saito Scribe: Alexander Sheynis/Allen Xue. May 3, The Neumann Boundary Condition

88 CHAPTER 2. VECTOR FUNCTIONS. . First, we need to compute T (s). a By definition, r (s) T (s) = 1 a sin s a. sin s a, cos s a

1 if 1 x 0 1 if 0 x 1

No: Bilkent University. Monotonic Extension. Farhad Husseinov. Discussion Papers. Department of Economics

Reference: Introduction to Partial Differential Equations by G. Folland, 1995, Chap. 3.

Lecture 7: Continuous Random Variables

Linear Programming Notes V Problem Transformations

Dimensional Analysis

Class Meeting # 1: Introduction to PDEs

So let us begin our quest to find the holy grail of real analysis.

Solutions for Review Problems

An Introduction to Partial Differential Equations

BANACH AND HILBERT SPACE REVIEW

Notes on metric spaces

1 Error in Euler s Method

Fundamental Theorems of Vector Calculus

1 Norms and Vector Spaces

Second Order Linear Nonhomogeneous Differential Equations; Method of Undetermined Coefficients. y + p(t) y + q(t) y = g(t), g(t) 0.

1. Prove that the empty set is a subset of every set.

An Introduction to Partial Differential Equations in the Undergraduate Curriculum

The Heat Equation. Lectures INF2320 p. 1/88

Math 4310 Handout - Quotient Vector Spaces

1.7 Graphs of Functions

Inner Product Spaces

Elasticity Theory Basics

Scalar Valued Functions of Several Variables; the Gradient Vector

6. Define log(z) so that π < I log(z) π. Discuss the identities e log(z) = z and log(e w ) = w.

To give it a definition, an implicit function of x and y is simply any relationship that takes the form:

A QUICK GUIDE TO THE FORMULAS OF MULTIVARIABLE CALCULUS

DIFFERENTIABILITY OF COMPLEX FUNCTIONS. Contents

The Math Circle, Spring 2004

Polynomial Invariants

Real Roots of Univariate Polynomials with Real Coefficients

Metric Spaces. Chapter Metrics

Probability and Statistics Prof. Dr. Somesh Kumar Department of Mathematics Indian Institute of Technology, Kharagpur

correct-choice plot f(x) and draw an approximate tangent line at x = a and use geometry to estimate its slope comment The choices were:

Undergraduate Notes in Mathematics. Arkansas Tech University Department of Mathematics

MATH 4330/5330, Fourier Analysis Section 11, The Discrete Fourier Transform

The Mean Value Theorem

CBE 6333, R. Levicky 1 Differential Balance Equations

Method of Green s Functions

Mechanics 1: Conservation of Energy and Momentum

Applications to Data Smoothing and Image Processing I

Lecture 16. Newton s Second Law for Rotation. Moment of Inertia. Angular momentum. Cutnell+Johnson: 9.4, 9.6

The last three chapters introduced three major proof techniques: direct,

Integrals of Rational Functions

Linear Algebra Notes for Marsden and Tromba Vector Calculus

Limits and Continuity

TOPIC 4: DERIVATIVES

Notes on Factoring. MA 206 Kurt Bryan

Unified Lecture # 4 Vectors

MA4001 Engineering Mathematics 1 Lecture 10 Limits and Continuity

HOMEWORK 5 SOLUTIONS. n!f n (1) lim. ln x n! + xn x. 1 = G n 1 (x). (2) k + 1 n. (n 1)!

Constrained optimization.

3. Mathematical Induction

Math 120 Final Exam Practice Problems, Form: A

Descriptive Statistics and Measurement Scales

3. Reaction Diffusion Equations Consider the following ODE model for population growth

PUTNAM TRAINING POLYNOMIALS. Exercises 1. Find a polynomial with integral coefficients whose zeros include

Vector and Matrix Norms

The Method of Partial Fractions Math 121 Calculus II Spring 2015

Chapter 7. Potential flow theory. 7.1 Basic concepts Velocity potential

Solutions to Homework 5

I. GROUPS: BASIC DEFINITIONS AND EXAMPLES

MATH APPLIED MATRIX THEORY

c 2008 Je rey A. Miron We have described the constraints that a consumer faces, i.e., discussed the budget constraint.

10.2 Series and Convergence

1 The basic equations of fluid dynamics

Math 21a Curl and Divergence Spring, Define the operator (pronounced del ) by. = i

0 <β 1 let u(x) u(y) kuk u := sup u(x) and [u] β := sup

LS.6 Solution Matrices

Metric Spaces Joseph Muscat 2003 (Last revised May 2009)

Solutions to Homework 10

Solutions to Linear First Order ODE s

Calculus 1: Sample Questions, Final Exam, Solutions

Polynomial and Rational Functions

Differential Relations for Fluid Flow. Acceleration field of a fluid. The differential equation of mass conservation

Vectors. Objectives. Assessment. Assessment. Equations. Physics terms 5/15/14. State the definition and give examples of vector and scalar variables.

Correlation and Convolution Class Notes for CMSC 426, Fall 2005 David Jacobs

Introduction to the Finite Element Method

Scalars, Vectors and Tensors

Solutions to Practice Problems for Test 4

The small increase in x is. and the corresponding increase in y is. Therefore

Solution to Homework 2

(a) We have x = 3 + 2t, y = 2 t, z = 6 so solving for t we get the symmetric equations. x 3 2. = 2 y, z = 6. t 2 2t + 1 = 0,

THE BANACH CONTRACTION PRINCIPLE. Contents

Chapter 4 One Dimensional Kinematics

LEARNING OBJECTIVES FOR THIS CHAPTER

4. Complex integration: Cauchy integral theorem and Cauchy integral formulas. Definite integral of a complex-valued function of a real variable

FINAL EXAM SOLUTIONS Math 21a, Spring 03

Duality of linear conic problems

Pigeonhole Principle Solutions

MINIMIZATION OF ENTROPY FUNCTIONALS UNDER MOMENT CONSTRAINTS. denote the family of probability density functions g on X satisfying

Vector surface area Differentials in an OCS

Methods for Finding Bases

5 Homogeneous systems

Transcription:

1 Introduction Harmonic Functions Kurt Bryan MA 436 Let be a bounded region in lr n (we ll focus on lr 2 for now) and u(x, t) the temperature of, where x = (x 1,..., x n ) and t is time. We found that (up to rescaling) u must satisfy the heat equation u t u = 0 in, with appropriate boundary and initial conditions. For now we re going to study a special case of the heat equation, that in which the solution assumes a steadystate configuration, i.e., doesn t depend on time t. In this case u t 0; the function u becomes a function only of the spatial variables and we find that u = 0 in. This is called Laplace s Equation. The nonhomogeneous equation u = f is called Poisson s Equation. efinition: A function u which satisfies u = 0 is called harmonic. Harmonic functions are a very special and important class of functions, not only in PE, but also in complex analysis, electromagnetics, fluids, etc. But we can t solve u = 0 without more information. Specifically, the irichlet boundary condition from the heat equation is still needed, so we must specify u = h on for some given function h. Alternatively we can specify u = g on for a given function g, though it turns out that g must satisfy a certain condition (more on this later) if a solution is to exist. The initial condition from the heat equation is no longer relevant, since u doesn t change with time now. All in all then our problem is to find a function which satisfies u = 0 in, (1) u u = h on (2) OR = g on. (3) Before proceeding it might be instructive to consider a few examples of harmonic functions. First, what does Poisson s equation looks like in one 1

dimension? If is the interval (a, b) then Poisson s equation just becomes u (x) = f(x) in (a, b). If have irichlet boundary conditions then u(a) = h 1 and u(b) = h 2, where h 1 and h 2 are some given numbers. This is easy to solve, just integrate twice and choose the constants of integration to get the right boundary values; it can always be done. In the special case that f = 0 the solution is a straight line. Harmonic functions in one dimension are just linear functions. Exercise: Work out the solution to u = f with u(a) = h 1, u(b) = h 2 in terms of h 1, h 2, a, b, and F, where F = f is a second anti-derivative for f. Now consider solving u (x) = 0 on (a, b) with Neumann data (the special case f 0). The Neumann condition at x = 0 is u (a) = g 1 (we use u (a), since the unit outward normal at x = a points in the MINUS x direction) and u (b) = g 2 for some given g 1, g 2. Integrate u (x) = 0 once to find u (x) = m for some constant m. This immediately forces g 1 = m and g 2 = m for otherwise we have a contradiction. If we do indeed have g 1 = g 2 then we can take m = g 2, so u (x) = mx, and integrating again shows that u(x) = mx + c is harmonic and has the right Neumann data for ANY choice of c. In this case there is not a unique solution, but rather an entire family of solutions, all differing by an additive constant. Exercise: Consider solving u (x) = f(x) on (a, b) with Neumann data u (a) = g 1, u (b) = g 2. What relation must hold between g 1, g 2, and f in order for a solution to exist? Will it be unique? In two dimensions it s easy to check that any linear function of the form u(x, y) = ax+by +c for constants a, b, c, is harmonic. The function u(x, y) = x 2 y 2 is also harmonic. Not surprisingly, there are infinitely many others. 2 Basic Properties of Laplace s Equation Let s consider the big three questions for Laplace s equation: Existence, Uniqueness, and Stability. We ll also consider the allied question of what 2

interesting properties do harmonic functions have? We are not yet in a position to answer all questions. We ll deal with uniqueness and stability first, but consider only special cases for existence. For the moment, let s focus on uniqueness. 2.1 Uniqueness It s easy to show that there is only one function that can satisfy Poisson s equation with given irichlet boundary data. Suppose that both u 1 and u 2 are C 2 and satisfy Poisson s equation with forcing function f and boundary data h. Then the function v = u 2 u 1 satisfies v = 0 in and v = 0 on. We will show that v = 0 in, so u 1 = u 2 in. Take the equation v = 0 and multiply both sides by v to obtain v v = 0. By Green s first identity v v dv = v v da v 2 dv. But since v = 0 on and since v = 0 this becomes v 2 dv = 0. (4) But v 2 is clearly non-negative on ; if the integral is zero then we must have v 0 on, so that v is constant. But since v = 0 on we conclude that v 0 throughout. Thus u 1 = u 2 in. Now consider the same situation but with Neumann data. The same computations which led to equation (4) still work, and we conclude that v = u 2 u 1 is constant. But since we now have only v = 0 on, rather than v = 0, we can t conclude that v 0 on, only that v = c, so that u 2 = u 1 + c. And indeed, it s easy to see that if u 1 = f on with u 1 = g then u 2 = u 1 + c satisfies the same conditions for any choice of c. So when dealing with Neumann boundary conditions we obtain uniqueness only up to an additive constant. We can obtain uniqueness by specifying one additional condition, e.g., the value of the solution u at a given point in, or, for example, the condition that u(x) dx = 0. (5) 3

Thus, for example, if two solutions with the same Neumann data differ by a constant, say u 2 = u 1 + c, and both satisfy the additional condition (5) we can immediately deduce that c = 0, so u 1 u 2. 2.2 The Maximum Principle for Laplace s Equation Let u be a solution to Laplace s equation u = 0 which is C 2 on with u continuous up to and on the boundary of some region (meaning that for any point x 0 we have lim x x0 u(x) = u(x 0 ), where x limits to x 0 from inside ). The maximum principle states that sup u(x) = sup u(x). x x Note that the supremum might also be attained somewhere inside too (consider the case in which u is constant) the assertion is simply that it must be attained on the boundary. Proof: The is exactly like the proof of the maximum principle for the heat equation. In fact, we use the same trick. Let v(x) = u(x) + ɛ x 2, where x = x 2 1 + + x 2 n means the usual Pythagorean norm of x and ɛ is some small positive number. Then v has no maximum inside, for if it did have a maximum at x = p then v(p) = u + 4ɛ > 0 which is impossible at an interior maximum (from the second derivative test in Calculus III). We can conclude that the maximum value of v occurs on the boundary, so sup v(x) = sup v(x). x x Suppose that p is a point in at which v attains its maximum value. Then we have, for x, u(x) < v(x) v(p) = u(p) + ɛ p 2 sup u(x) + ɛd 2 x where d is the maximum distance from any point in to the origin. But since ɛ is arbitrary we conclude that u(x) sup u(x). x 4

The same reasoning applied to u(x) proves that the minimum value is attained on the boundary. Also, the same reasoning proves the Extended Maximum Principle: If f 0 on then a function u satisfying u = f on then u must attain its maximum value (but NOT necessarily minimum) value on. Similarly if f 0 then u must attain its minimum (but not necessarily maximum) value on. Actually, the maximum principle provides yet another proof of uniqueness. If u 1 and u 2 both satisfy Poisson s equation on with the same f and irichlet data h then v = u 2 u 1 is harmonic with zero boundary data. By the maximum principle the maximum and minimum value attained by v on is attained on, and hence is zero. Thus v must be identically zero, so u 1 = u 2 on. 2.3 Stability Suppose that u 1 and u 2 both satisfy Poisson s equation with the same forcing function, so u j = f on. Suppose that they have different boundary conditions, say u j = h j on, for j = 1, 2. Then we can prove a stability result, namely sup u 2 (x) u 1 (x) sup h 2 (x) h 1 (x). x x In other words, if the functions h 1 and h 2 are close in the supremum norm on then u 1 and u 2 are close in the supremum norm on all of. The proof is easy. Let v = u 2 u 1. Then v is harmonic and has boundary data h 2 h 1. efine M = sup h 2 (x) h 1 (x) x and note that sup x (h 2 (x) h 1 (x)) M and inf x (h 2 (x) h 1 (x)) M. By the maximum/minimum principle for harmonic functions sup x inf x v(x) sup (h 2 (x) h 1 (x)) M, (6) x v(x) inf x (h 2(x) h 1 (x)) M, (7) 5

Also note that sup x v equals either sup x v(x) or inf x v(x). Equations (6) and (7) then yield sup v M. x Thus if M (the supremum norm of h 2 h 1 on ) is small then v has a small supremum norm, i.e., u 1 is close to u 2 in supremum norm. 2.4 The Green s Function and Green s Third Identity Let x = (x 1, x 2 ) denote a point in two dimensions and let x = x 2 1 + x 2 2. The function G(x) = 1 ln x (8) 2π is called the Green s Function or Green s Kernel or Fundamental Solution for the Laplacian. In n dimensions with n 3 the Green s function is G(x) = 1 ω n x 2 n (9) where x = (x 1,..., x n ), x is the magnitude of x, and ω n is the surface area of the unit ball in n dimensions. The Green s kernel here plays a similar role as the Green s kernel for the heat equation. First, you can easily check that G(x) = 0 IF x 0. At x = 0 the function G has an asymptote, so G makes no sense there. Here s a picture of G in two dimensions: 6

Note that G(x) is radially symmetric about the singularity. Along any radial line from the singularity G drops off with distance r as 1 ln(r). 2π Given a fixed point y = (y 1,..., y n ) in n dimensional space we can also consider the function G(x y), which translates G so that the singularity is at x = y. To understand the significance and usefulness of G we need Green s Third Identity, which we derive below for the two-dimensional case, though it works in any dimension. Recall Green s second identity: Let be a bounded domain in lr 2 ; for C 2 functions u and v we have (u v v u) dx = (u v v u ) ds where n is an outward unit normal vector on (where dx = dx 1 dx 2 and ds is arc length) Let s take y to be some fixed point inside and then take v(x) = G(x y). Then G = 0 for x y, where it s understood that we apply in the x variable. We want to put v = G into Green s second identity and see what comes out, but there s one problem: G is NOT C 2 on, so Green s identity isn t valid. To get around this let us remove from a tiny ball B ɛ (y) of radius ɛ centered at y as illustrated below. 7

Let s use the notation ɛ for with the ball B ɛ (y) removed. On ɛ the function G(x y) is smooth and so Green s second identity is valid. If we put in v(x) = G(x y) (and u is still just some arbitrary C 2 function, not necessarily harmonic) then we obtain G(x y) u(x) dx = (u G ɛ ɛ G u ) ds (10) where all derivatives hitting G are with respect to the x variable. Now notice that ɛ really consists of two pieces: and B ɛ (y). On the vector n points outward, while on B ɛ (y) the vector n points INTO the ball (which is OUT of ɛ ). Let s take equation (10) and split the integrals over ɛ into integrals over and B ɛ (y) to obtain G(x y) u(x) dx = ɛ + u G ds G u ds B ɛ(y) u G ds B ɛ(y) G u Let s look at what happens to equation (11) as ɛ tends to zero. ds. (11) Claim: As ɛ approaches zero the integral on the left in equation (11) becomes an integral over. The last integral on the right vanishes, and the second to last integral approaches u(y). In summary, equation (11) becomes G(x y) u(x) dx = u G ds G u ds u(y). (12) 8

Proof: We ll examine the left side first, then the right side. First of all, I claim that the left side tends to G(x y) u(x) dx. To see this, look at the difference G(x y) u(x) dx G(x y) u(x) dx = G(x y) u(x) dx. ɛ B ɛ (y) (13) If u is C 2 then u(x) is bounded by some constant C near the point x = y, so the magnitude of the integral on the right above is bounded by C G(x y) dx. B ɛ (y) Also, G(x y) is just 1 ln x y, and this integral can be worked explicitly. 2π Just change variables let w = x y so dw = dx. This integral becomes C G(w) dw. B ɛ(0) Switch to polar coordinates and the integral becomes, explicitly C 2π ɛ 0 0 1 2π ln(r)r dr dθ = C ɛ2 2 ( ln(ɛ) 1 ) 2 which approaches zero as ɛ 0. Thus the right side of equation (13) approaches zero, and so the integral over ɛ approaches the integral over. Now let s examine the right side of equation (11). First, the terms involving integral over don t even involve ɛ, so there s nothing to do there. The very last term involving G u tends to zero as ɛ approaches zero. To see this, note that if u is C 2 on then u = u n is bounded on B ɛ(y). Thus the final integral in equation (11) can be bounded by C G(x y) ds B ɛ(y) where the integral is with respect to x. Parameterize the boundary of B ɛ (y) as x 1 = y 1 + ɛ cos(w), x 2 = y 2 + ɛ sin(w) for 0 w < 2π. Also, ds = ɛdv and then this integral becomes C 2π 0 1 ln(ɛ)ɛ dw = Cɛ ln(ɛ) 2π 9

which approaches zero as ɛ 0, as asserted. Finally, let s examine the most interesting term in equation (11), the second to last integral on the right. First of all let s again parameterize the boundary of B ɛ (y) as x 1 = y 1 + ɛ cos(w), x 2 = y 2 + ɛ sin(w) for 0 w < 2π. Then ds = ɛdw and an outward unit normal vector is just n = (cos(w), sin(w)). It s easy to check that ( x G(x, y) = x ) 1 y 1 2πr, x 2 y 2 2 2πr 2 where r 2 = (x 1 y 1 ) 2 + (x 2 y 2 ) 2 and x means compute the gradient with respect to the x variable. Now compute G = xg n, making sure to put everything in terms of w. We obtain on B ɛ(y). Thus G = 1 2πɛ u G 2π B ɛ(y) ds = 1 u(y + ɛ(cos(w), sin(w))) dw. (14) 0 2πɛ But since u is continuous, if ɛ is small then u(y +ɛ(cos(w), sin(w))) is close to u(y) (which is constant in x). In the limit that ɛ approaches zero the integral on the right in equation (14) looks like 1 2π 2πɛ u(y) 0 dw = u(y). This proves that the second to last integral on the right in equation (11) approaches u(y), as asserted, and finishes the proof of the claim. In summary, we can write equation (12) as u(y) = G(x y) u(x) dx+ u(x) G (x y) ds x G(x y) u (x) ds x (15) where ds x means integrate with respect to x. This is Green s Third Identity. As a special case, suppose we have a harmonic function u on, so u = 0 in, with u = h on and u = g on. Then Green s third identity tells us how to find u at any interior point y, as u(y) = G (x y)h(x) ds x 10 G(x y)g(x) ds x. (16)

But be warned if we simply make up functions g and h on and plug them into equation (16) we won t obtain a harmonic function u with irichlet data h and normal derivative g, for we already know that h alone determines u, or likewise, g alone determines u up to an additive constant. We can t pick BOTH of them and expect to find a solution. If you do choose both and plug into equation (16) you ll end up with a function that is harmonic on, but it probably won t have the correct irichlet OR Neumann data. Exactly the same argument given above works in higher dimensions too, though I won t go through it in detail. But equations (15) and (16) are true in lr n. 2.5 The Mean Value Property This is without doubt the most amazing property of harmonic functions. In fact, it s a property that ONLY harmonic functions have. Suppose that u = 0 on some region. Let B be a ball of radius r around some point y = (y 1, y 2 ) in (r not necessarily small, but small enough so B is contained in ). Claim: The value of u at the center of the ball equals the average value of u over the boundary of the ball, i.e., u(y) = 1 u ds = 1 2π u(y 1 + r cos(w), y 2 + r sin(w)) dw (17) B B 2π 0 where B = 2πr denotes the length of the boundary of B (or in n 3 dimensions the surface area of the boundary). To prove the Mean Value Property, start with Green s third identity in the form (16). We have u(y) = u(x) G B (x y) ds x. G(x y) u B (x) ds x. Since B is a ball of radius r we ll parameterize its boundary as x 1 = y 1 + r cos(w), x 2 = y 2 + r sin(w), so ds x = r dw. On B the function G(x y) becomes just a constant 1 G ln(r), while also becomes a constant, 1. The 2π 2πr above equation becomes u(y) = 1 2π u(y 1 + r cos(w), y 2 + r sin(w))r dw + ln(r) u 2πr 0 2π B ds x (18) 11

The second integral above is actually zero. To see this note that if u is harmonic then from Green s FIRST identity with v(x) 1 we have B u B ds = u dx = 0. Plugging this into equation (18) proves the claim. Again, the same argument works in lr n for n > 2. The mean value property even holds in one dimension, where it becomes the assertion that if u = 0 then u(x) = 1 (u(x r) + u(x + r)) 2 which is certainly true for harmonic (i.e., linear) functions. The mean value property also works if you use the interior of the ball, instead of the boundary. Exercise: Prove the last claim. Specifically, if u has the mean value property in the form (17) then u(y) = 1 u da B B where B = 4πr 2 is the area of the ball B and da means dx 1 dx 2. Then prove the converse; assume whatever continuity or differentiability you need from u. The mean value property provides another proof of the maximum principle. Informally, if x 0 is an interior point for a region then we can put a small ball of some radius around x 0. Now we can t have u(x 0 ) > u(x) for all x, because u(x 0 ) is an average of the values of u on the boundary of any ball surrounding x 0. But the average can never be larger than EVERY point out of which it is constructed. It turns out that harmonic functions are the ONLY ones which possess the mean value property. This is easy to prove, at least if you re willing to restrict your attention to C 2 functions. For the moment, let u be ANY C 2 function. Choose a fixed point y in and let B r denote a ball of radius r around y (r small enough so B r is contained in ). efine a function φ(r) as φ(r) = 1 u ds = 1 2π u(y 1 + r cos(t), y 2 + r sin(t)) dt. B r B r 2π 0 12

The function φ(r) is just the average value of u over the ball of radius r centered at y. Compute φ (r) = 1 ( 2π u (y 1 + r cos(t), y 2 + r sin(t)) cos(t) 2π 0 x 1 + u ) (y 1 + r cos(t), y 2 + r sin(t)) sin(t) dt x 2 = = 1 B r 1 B r B r u ds B r u da. (19) The last step is Green s first identity. Equation (19) holds for any C 2 function. Note also that lim r 0 + φ(r) = u(y) if u is continuous, that is, any continuous function has the mean value property on a ball of radius zero. Now suppose that u is some function which is NOT harmonic. Then we can find some ball B R contained in with either u > 0 or u < 0 on B R ; assume the former. For r < R equation (19) forces φ (r) > 0. Since lim r 0 + φ(r) = u(y) this shows that u(y) < φ(r) for any r > 0, that is, u does not have the mean value property. A similar conclusion holds if u < 0 anywhere. The proves that if u has the mean value property then u is harmonic. By the way, this also gives another proof of the mean value property for harmonic functions: If u = 0 then φ (r) = 0, i.e., φ is constant. Since lim r 0 + φ(r) = u(y), u must have the mean value property. 13