23
3 Laplace’s Equation We now turn to studying Laplace’s equation Δu =0 and its inhomogeneous version, Poisson’s equation, -Δu = f. We say a function u satisfying Laplace’s equation is a harmonic function. 3.1 The Fundamental Solution Consider Laplace’s equation in R n , Δu =0 x R n . Clearly, there are a lot of functions u which satisfy this equation. In particular, any constant function is harmonic. In addition, any function of the form u(x)= a 1 x 1 + ... + a n x n for constants a i is also a solution. Of course, we can list a number of others. Here, however, we are interested in finding a particular solution of Laplace’s equation which will allow us to solve Poisson’s equation. Given the symmetric nature of Laplace’s equation, we look for a radial solution. That is, we look for a harmonic function u on R n such that u(x)= v(|x|). In addition, to being a natural choice due to the symmetry of Laplace’s equation, radial solutions are natural to look for because they reduce a PDE to an ODE, which is generally easier to solve. Therefore, we look for a radial solution. If u(x)= v(|x|), then u x i = x i |x| v 0 (|x|) |x|6 =0, which implies u x i x i = 1 |x| v 0 (|x|) - x 2 i |x| 3 v 0 (|x|)+ x 2 i |x| 2 v 00 (|x|) |x|6 =0. Therefore, Δu = n - 1 |x| v 0 (|x|)+ v 00 (|x|). Letting r = |x|, we see that u(x)= v(|x|) is a radial solution of Laplace’s equation implies v satisfies n - 1 r v 0 (r)+ v 00 (r)=0. Therefore, v 00 = 1 - n r v 0 = v 00 v 0 = 1 - n r = ln v 0 = (1 - n) ln r + C = v 0 (r)= C r n-1 , 1

3 Laplace’s Equation - Stanford Universityweb.stanford.edu/class/math220b/handouts/laplace.pdf3 Laplace’s Equation We now turn to studying Laplace’s equation ∆u = 0 and its

  • Upload
    lelien

  • View
    224

  • Download
    2

Embed Size (px)

Citation preview

Page 1: 3 Laplace’s Equation - Stanford Universityweb.stanford.edu/class/math220b/handouts/laplace.pdf3 Laplace’s Equation We now turn to studying Laplace’s equation ∆u = 0 and its

3 Laplace’s Equation

We now turn to studying Laplace’s equation

∆u = 0

and its inhomogeneous version, Poisson’s equation,

−∆u = f.

We say a function u satisfying Laplace’s equation is a harmonic function.

3.1 The Fundamental Solution

Consider Laplace’s equation in Rn,

∆u = 0 x ∈ Rn.

Clearly, there are a lot of functions u which satisfy this equation. In particular, anyconstant function is harmonic. In addition, any function of the form u(x) = a1x1 + . . .+anxn

for constants ai is also a solution. Of course, we can list a number of others. Here, however,we are interested in finding a particular solution of Laplace’s equation which will allow usto solve Poisson’s equation.

Given the symmetric nature of Laplace’s equation, we look for a radial solution. Thatis, we look for a harmonic function u on Rn such that u(x) = v(|x|). In addition, to beinga natural choice due to the symmetry of Laplace’s equation, radial solutions are natural tolook for because they reduce a PDE to an ODE, which is generally easier to solve. Therefore,we look for a radial solution.

If u(x) = v(|x|), then

uxi=

xi

|x|v′(|x|) |x| 6= 0,

which implies

uxixi=

1

|x|v′(|x|)− x2

i

|x|3v′(|x|) +x2

i

|x|2v′′(|x|) |x| 6= 0.

Therefore,

∆u =n− 1

|x| v′(|x|) + v′′(|x|).Letting r = |x|, we see that u(x) = v(|x|) is a radial solution of Laplace’s equation impliesv satisfies

n− 1

rv′(r) + v′′(r) = 0.

Therefore,

v′′ =1− n

rv′

=⇒ v′′

v′=

1− n

r=⇒ ln v′ = (1− n) ln r + C

=⇒ v′(r) =C

rn−1,

1

Page 2: 3 Laplace’s Equation - Stanford Universityweb.stanford.edu/class/math220b/handouts/laplace.pdf3 Laplace’s Equation We now turn to studying Laplace’s equation ∆u = 0 and its

which implies

v(r) =

c1 ln r + c2 n = 2

c1(2−n)rn−2 + c2 n ≥ 3.

From these calculations, we see that for any constants c1, c2, the function

u(x) ≡

c1 ln |x|+ c2 n = 2c1

(2−n)|x|n−2 + c2 n ≥ 3.(3.1)

for x ∈ Rn, |x| 6= 0 is a solution of Laplace’s equation in Rn − 0. We notice that thefunction u defined in (3.1) satisfies ∆u(x) = 0 for x 6= 0, but at x = 0, ∆u(0) is undefined.We claim that we can choose constants c1 and c2 appropriately so that

−∆xu = δ0

in the sense of distributions. Recall that δ0 is the distribution which is defined as follows.For all φ ∈ D,

(δ0, φ) = φ(0).

Below, we will prove this claim. For now, though, assume we can prove this. That is, assumewe can find constants c1, c2 such that u defined in (3.1) satisfies

−∆xu = δ0. (3.2)

Let Φ denote the solution of (3.2). Then, define

v(x) =

Rn

Φ(x− y)f(y) dy.

Formally, we compute the Laplacian of v as follows,

−∆xv = −∫

Rn

∆xΦ(x− y)f(y) dy

= −∫

Rn

∆yΦ(x− y)f(y) dy

=

Rn

δxf(y) dy = f(x).

That is, v is a solution of Poisson’s equation! Of course, this set of equalities above is entirelyformal. We have not proven anything yet. However, we have motivated a solution formulafor Poisson’s equation from a solution to (3.2). We now return to using the radial solution(3.1) to find a solution of (3.2).

Define the function Φ as follows. For |x| 6= 0, let

Φ(x) =

− 12π

ln |x| n = 21

n(n−2)α(n)1

|x|n−2 n ≥ 3,(3.3)

where α(n) is the volume of the unit ball in Rn. We see that Φ satisfies Laplace’s equationon Rn−0. As we will show in the following claim, Φ satisfies −∆xΦ = δ0. For this reason,we call Φ the fundamental solution of Laplace’s equation.

2

Page 3: 3 Laplace’s Equation - Stanford Universityweb.stanford.edu/class/math220b/handouts/laplace.pdf3 Laplace’s Equation We now turn to studying Laplace’s equation ∆u = 0 and its

Claim 1. For Φ defined in (3.3), Φ satisfies

−∆xΦ = δ0

in the sense of distributions. That is, for all g ∈ D,

−∫

Rn

Φ(x)∆xg(x) dx = g(0).

Proof. Let FΦ be the distribution associated with the fundamental solution Φ. That is, letFΦ : D → R be defined such that

(FΦ, g) =

Rn

Φ(x)g(x) dx

for all g ∈ D. Recall that the derivative of a distribution F is defined as the distribution Gsuch that

(G, g) = −(F, g′)

for all g ∈ D. Therefore, the distributional Laplacian of Φ is defined as the distribution F∆Φ

such that(F∆Φ, g) = (FΦ, ∆g)

for all g ∈ D. We will show that

(FΦ, ∆g) = −(δ0, g) = −g(0),

and, therefore,(F∆Φ, g) = −g(0),

which means −∆xΦ = δ0 in the sense of distributions.By definition,

(FΦ, ∆g) =

Rn

Φ(x)∆g(x) dx.

Now, we would like to apply the divergence theorem, but Φ has a singularity at x = 0. Weget around this, by breaking up the integral into two pieces: one piece consisting of the ballof radius δ about the origin, B(0, δ) and the other piece consisting of the complement of thisball in Rn. Therefore, we have

(FΦ, ∆g) =

Rn

Φ(x)∆g(x) dx

=

B(0,δ)

Φ(x)∆g(x) dx +

Rn−B(0,δ)

Φ(x)∆g(x) dx

= I + J.

3

Page 4: 3 Laplace’s Equation - Stanford Universityweb.stanford.edu/class/math220b/handouts/laplace.pdf3 Laplace’s Equation We now turn to studying Laplace’s equation ∆u = 0 and its

We look first at term I. For n = 2, term I is bounded as follows,∣∣∣∣−

B(0,δ)

1

2πln |x|∆g(x) dx

∣∣∣∣ ≤ C|∆g|L∞∣∣∣∣∫

B(0,δ)

ln |x| dx

∣∣∣∣

≤ C

∣∣∣∣∫ 2π

0

∫ δ

0

ln |r|r dr dθ

∣∣∣∣

≤ C

∣∣∣∣∫ δ

0

ln |r|r dr

∣∣∣∣≤ C ln |δ|δ2.

For n ≥ 3, term I is bounded as follows,∣∣∣∣∫

B(0,δ)

1

n(n− 2)α(n)

1

|x|n−2∆g(x) dx

∣∣∣∣ ≤ C|∆g|L∞∫

B(0,δ)

1

|x|n−2dx

≤ C

∫ δ

0

(∫

∂B(0,r)

1

|y|n−2dS(y)

)dr

=

∫ δ

0

1

rn−2

(∫

∂B(0,r)

dS(y)

)dr

=

∫ δ

0

1

rn−2nα(n)rn−1 dr

= nα(n)

∫ δ

0

r dr =nα(n)

2δ2.

Therefore, as δ → 0+, |I| → 0.Next, we look at term J . Applying the divergence theorem, we have∫

Rn−B(0,δ)

Φ(x)∆xg(x) dx =

Rn−B(0,δ)

∆xΦ(x)g(x) dx−∫

∂(Rn−B(0,δ))

∂Φ

∂νg(x) dS(x)

+

∂(Rn−B(0,δ))

Φ(x)∂g

∂νdS(x)

= −∫

∂(Rn−B(0,δ))

∂Φ

∂νg(x) dS(x) +

∂(Rn−B(0,δ))

Φ(x)∂g

∂νdS(x)

≡ J1 + J2.

using the fact that ∆xΦ(x) = 0 for x ∈ Rn −B(0, δ).We first look at term J1. Now, by assumption, g ∈ D, and, therefore, g vanishes at

∞. Consequently, we only need to calculate the integral over ∂B(0, ε) where the normalderivative ν is the outer normal to Rn − B(0, δ). By a straightforward calculation, we seethat

∇xΦ(x) = − x

nα(n)|x|n .

The outer unit normal to Rn −B(0, δ) on B(0, δ) is given by

ν = − x

|x| .

4

Page 5: 3 Laplace’s Equation - Stanford Universityweb.stanford.edu/class/math220b/handouts/laplace.pdf3 Laplace’s Equation We now turn to studying Laplace’s equation ∆u = 0 and its

Therefore, the normal derivative of Φ on B(0, δ) is given by

∂Φ

∂ν=

(− x

nα(n)|x|n)·(− x

|x|)

=1

nα(n)|x|n−1.

Therefore, J1 can be written as

−∫

∂B(0,δ)

1

nα(n)|x|n−1g(x) dS(x) = − 1

nα(n)δn−1

∂B(0,δ)

g(x) dS(x) = −−∫

∂B(0,δ)

g(x) dS(x).

Now if g is a continuous function, then

−−∫

g(x) dS(x) → −g(0) as δ → 0.

Lastly, we look at term J2. Now using the fact that g vanishes as |x| → +∞, we only needto integrate over ∂B(0, δ). Using the fact that g ∈ D, and, therefore, infinitely differentiable,we have

∣∣∣∣∫

∂B(0,δ)

Φ(x)∂g

∂νdS(x)

∣∣∣∣ ≤∣∣∣∣∂g

∂ν

∣∣∣∣L∞(∂B(0,δ))

∂B(0,δ)

|Φ(x)| dS(x)

≤ C

∂B(0,δ)

|Φ(x)| dS(x).

Now first, for n = 2,

∂B(0,δ)

|Φ(x)| dS(x) = C

∂B(0,δ)

| ln |x|| dS(x)

≤ C| ln |δ||∫

∂B(0,δ)

dS(x)

= C| ln |δ||(2πδ) ≤ Cδ| ln |δ||.

Next, for n ≥ 3,

∂B(0,δ)

|Φ(x)| dS(x) = C

∂B(0,δ)

1

|x|n−2dS(x)

≤ C

δn−2

∂B(0,δ)

dS(x)

=C

δn−2nα(n)δn−1 ≤ Cδ.

Therefore, we conclude that term J2 is bounded in absolute value by

Cδ| ln δ| n = 2Cδ n ≥ 3.

Therefore, |J2| → 0 as δ → 0+.

5

Page 6: 3 Laplace’s Equation - Stanford Universityweb.stanford.edu/class/math220b/handouts/laplace.pdf3 Laplace’s Equation We now turn to studying Laplace’s equation ∆u = 0 and its

Combining these estimates, we see that

Rn

Φ(x)∆xg(x) dx = limδ→0+

I + J1 + J2 = −g(0).

Therefore, our claim is proved.

Solving Poisson’s Equation. We now return to solving Poisson’s equation

−∆u = f x ∈ Rn.

From our discussion before the above claim, we expect the function

v(x) ≡∫

Rn

Φ(x− y)f(y) dy

to give us a solution of Poisson’s equation. We now prove that this is in fact true. First, wemake a remark.

Remark. If we hope that the function v defined above solves Poisson’s equation, we mustfirst verify that this integral actually converges. If we assume f has compact support onsome bounded set K in Rn, then we see that

Rn

Φ(x− y)f(y) dy ≤ |f |L∞∫

K

|Φ(x− y)| dy.

If we additionally assume that f is bounded, then |f |L∞ ≤ C. It is left as an exercise toverify that ∫

K

|Φ(x− y)| dy < +∞

on any compact set K.

Theorem 2. Assume f ∈ C2(Rn) and has compact support. Let

u(x) ≡∫

Rn

Φ(x− y)f(y) dy

where Φ is the fundamental solution of Laplace’s equation (3.3). Then

1. u ∈ C2(Rn)

2. −∆u = f in Rn.

Ref: Evans, p. 23.

Proof. 1. By a change of variables, we write

u(x) =

Rn

Φ(x− y)f(y) dy =

Rn

Φ(y)f(x− y) dy.

6

Page 7: 3 Laplace’s Equation - Stanford Universityweb.stanford.edu/class/math220b/handouts/laplace.pdf3 Laplace’s Equation We now turn to studying Laplace’s equation ∆u = 0 and its

Letei = (. . . , 0, 1, 0, . . .)

be the unit vector in Rn with a 1 in the ith slot. Then

u(x + hei)− u(x)

h=

Rn

Φ(y)

[f(x + hei − y)− f(x− y)

h

]dy.

Now f ∈ C2 implies

f(x + hei − y)− f(x− y)

h→ ∂f

∂xi

(x− y) as h → 0

uniformly on Rn. Therefore,

∂u

∂xi

(x) =

Rn

Φ(y)∂f

∂xi

(x− y) dy.

Similarly,∂2u

∂xixj

(x) =

Rn

Φ(y)∂2f

∂xixj

(x− y) dy.

This function is continuous because the right-hand side is continuous.

2. By the above calculations and Claim 1, we see that

∆xu(x) =

Rn

Φ(y)∆xf(x− y) dy

=

Rn

Φ(y)∆yf(x− y) dy

= −f(x).

3.2 Properties of Harmonic Functions

3.2.1 Mean Value Property

In this section, we prove a mean value property which all harmonic functions satisfy. First,we give some definitions. Let

B(x, r) = ball of radius r about x in Rn

∂B(x, r) = boundary of ball of radius r about x in Rn

α(n) = volume of unit ball in Rn

nα(n) = surface area of unit ball in Rn.

For a function u defined on B(x, r), the average of u on B(x, r) is given by

−∫

B(x,r)

u(y) dy =1

α(n)rn

B(x,r)

u(y) dy.

7

Page 8: 3 Laplace’s Equation - Stanford Universityweb.stanford.edu/class/math220b/handouts/laplace.pdf3 Laplace’s Equation We now turn to studying Laplace’s equation ∆u = 0 and its

For a function u defined on ∂B(x, r), the average of u on ∂B(x, r) is given by

−∫

∂B(x,r)

u(y) dS(y) =1

nα(n)rn−1

∂B(x,r)

u(y) dS(y).

Theorem 3. (Mean-Value Formulas) Let Ω ⊂ Rn. If u ∈ C2(Ω) is harmonic, then

u(x) = −∫

∂B(x,r)

u(y) dS(y) = −∫

B(x,r)

u(y) dy

for every ball B(x, r) ⊂ Ω.

Proof. Assume u ∈ C2(Ω) is harmonic. For r > 0, define

φ(r) = −∫

∂B(x,r)

u(y) dS(y).

For r = 0, define φ(r) = u(x). Notice that if u is a smooth function, then limr→0+ φ(r) =u(x), and, therefore, φ is a continuous function. Therefore, if we can show that φ′(r) = 0,then we can conclude that φ is a constant function, and, therefore,

u(x) = −∫

∂B(x,r)

u(y) dS(y).

We prove φ′(r) = 0 as follows. First, making a change of variables, we have

φ(r) = −∫

∂B(x,r)

u(y) dS(y)

= −∫

∂B(0,1)

u(x + rz) dS(z).

Therefore,

φ′(r) = −∫

∂B(0,1)

∇u(x + rz) · z dS(z)

= −∫

∂B(x,r)

∇u(y) · y − x

rdS(y)

= −∫

∂B(x,r)

∂u

∂ν(y) dS(y)

=1

nα(n)rn−1

∂B(x,r)

∂u

∂ν(y) dS(y)

=1

nα(n)rn−1

B(x,r)

∇ · (∇u) dy (by the Divergence Theorem)

=1

nα(n)rn−1

B(x,r)

∆u(y) dy = 0,

8

Page 9: 3 Laplace’s Equation - Stanford Universityweb.stanford.edu/class/math220b/handouts/laplace.pdf3 Laplace’s Equation We now turn to studying Laplace’s equation ∆u = 0 and its

using the fact that u is harmonic. Therefore, we have proven the first part of the theorem.It remains to prove that

u(x) = −∫

B(x,r)

u(y) dy.

We do so as follows, using the first result,

B(x,r)

u(y) dy =

∫ r

0

(∫

∂B(x,s)

u(y) dS(y)

)ds

=

∫ r

0

(nα(n)sn−1 −

∂B(x,s)

u(y) dS(y)

)ds

=

∫ r

0

nα(n)sn−1u(x) ds

= nα(n)u(x)

∫ r

0

sn−1 ds

= α(n)u(x) sn|s=rs=0

= α(n)u(x)rn.

Therefore, ∫

B(x,r)

u(y) dy = α(n)rnu(x),

which implies

u(x) =1

α(n)rn

B(x,r)

u(y) dy = −∫

B(x,r)

u(y) dy,

as claimed.

3.2.2 Converse to Mean Value Property

In this section, we prove that if a smooth function u satisfies the mean value propertydescribed above, then u must be harmonic.

Theorem 4. If u ∈ C2(Ω) satisfies

u(x) = −∫

∂B(x,r)

u(y) dS(y)

for all B(x, r) ⊂ Ω, then u is harmonic.

Proof. Let

φ(r) = −∫

∂B(x,r)

u(y) dS(y).

If

u(x) = −∫

∂B(x,r)

u(y) dS(y)

9

Page 10: 3 Laplace’s Equation - Stanford Universityweb.stanford.edu/class/math220b/handouts/laplace.pdf3 Laplace’s Equation We now turn to studying Laplace’s equation ∆u = 0 and its

for all B(x, r) ⊂ Ω, then φ′(r) = 0. As described in the previous theorem,

φ′(r) =r

n−∫

B(x,r)

∆u(y) dy.

Suppose u is not harmonic. Then there exists some ball B(x, r) ⊂ Ω such that ∆u > 0 or∆u < 0. Without loss of generality, we assume there is some ball B(x, r) such that ∆u > 0.Therefore,

φ′(r) =r

n−∫

B(x,r)

∆u(y) dy > 0,

which contradicts the fact that φ′(r) = 0. Therefore, u must be harmonic.

3.2.3 Maximum Principle

In this section, we prove that if u is a harmonic function on a bounded domain Ω in Rn,then u attains its maximum value on the boundary of Ω.

Theorem 5. Suppose Ω ⊂ Rn is open and bounded. Suppose u ∈ C2(Ω)∩C(Ω) is harmonic.Then

1. (Maximum principle)max

Ωu(x) = max

∂Ωu(x).

2. (Strong maximum principle) If Ω is connected and there exists a point x0 ∈ Ω suchthat

u(x0) = maxΩ

u(x),

then u is constant within Ω.

Proof. We prove the second assertion. The first follows from the second. Suppose thereexists a point x0 in Ω such that

u(x0) = M = maxΩ

u(x).

Then for 0 < r < dist(x0, ∂Ω), the mean value property says

M = u(x0) = −∫

B(x0,r)

u(y) dy ≤ M.

But, therefore,

−∫

B(x0,r)

u(y) dy = M,

and M = maxΩ u(x). Therefore, u(y) ≡ M for y ∈ B(x0, r). To prove u ≡ M throughoutΩ, you continue with this argument, filling Ω with balls.

10

Page 11: 3 Laplace’s Equation - Stanford Universityweb.stanford.edu/class/math220b/handouts/laplace.pdf3 Laplace’s Equation We now turn to studying Laplace’s equation ∆u = 0 and its

Remark. By replacing u by −u above, we can prove the Minimum Principle.Next, we use the maximum principle to prove uniqueness of solutions to Poisson’s equa-

tion on bounded domains Ω in Rn.

Theorem 6. (Uniqueness) There exists at most one solution u ∈ C2(Ω) ∩ C(Ω) of theboundary-value problem, −∆u = f x ∈ Ω

u = g x ∈ ∂Ω.

Proof. Suppose there are two solutions u and v. Let w = u− v and let w = v − u. Then wand w satisfy

∆w = 0 x ∈ Ωw = 0 x ∈ ∂Ω.

Therefore, using the maximum principle, we conclude

maxΩ|u− v| = max

∂Ω|u− v| = 0.

3.2.4 Smoothness of Harmonic Functions

In this section, we prove that harmonic functions are C∞.

Theorem 7. Let Ω be an open, bounded subset of Rn. If u ∈ C(Ω) and u satisfies the meanvalue property,

u(x) = −∫

∂B(x,r)

u(y) dS(y)

for every ball B(x, r) ⊂ Ω, then u ∈ C∞(Ω).

Remarks.

1. As proven earlier, if u ∈ C2(Ω) ∩ C(Ω) and u is harmonic, then u satisfies the meanvalue property, and, therefore, u ∈ C∞(Ω).

2. In fact, if u satisfies the hypothesis of the above theorem, then u is analytic, but wewill not prove that here. (See Evans.)

Proof. First, we introduce the function η such that

η(x) ≡

Ce1

|x|2−1 |x| < 10 |x| ≥ 1

where the constant C is chosen such that∫Rn η(x) dx = 1. Notice that η ∈ C∞(Rn) and η

has compact support. Now define the function ηε(x) such that

ηε(x) ≡ 1

εnη

(x

ε

).

11

Page 12: 3 Laplace’s Equation - Stanford Universityweb.stanford.edu/class/math220b/handouts/laplace.pdf3 Laplace’s Equation We now turn to studying Laplace’s equation ∆u = 0 and its

Therefore, ηε ∈ C∞(Rn) and supp(ηε) ⊂ x : |x| < ε. Further,

Rn

ηε(x) dx = 1.

Now choose ε such that ε < dist(x, ∂Ω). Define

uε(x) =

Ω

ηε(x− y)u(y) dy.

Now we claim

1. uε ∈ C∞

2. uε(x) = u(x).

First, for (1), uε ∈ C∞ because ηε ∈ C∞. We prove (2) as follows. Using the fact thatsuppηε(x− y) ⊂ y : |x− y| < ε. Therefore,

uε(x) =

B(x,ε)

ηε(x− y)u(y) dy

=1

εn

B(x,ε)

η

( |x− y|ε

)u(y) dy

=1

εn

∫ ε

0

(∫

∂B(x,r)

η

( |x− y|ε

)u(y) dS(y)

)dr

=1

εn

∫ ε

0

(∫

∂B(x,r)

η(r

ε

)u(y) dS(y)

)dr

=1

εn

∫ ε

0

η(r

ε

) ∫

∂B(x,r)

u(y) dS(y) dr

=1

εn

∫ ε

0

η(r

ε

)nα(n)rn−1 −

∂B(x,r)

u(y) dS(y) dr

=1

εn

∫ ε

0

η(r

ε

)nα(n)rn−1u(x) dr

= u(x)1

εn

∫ ε

0

η(r

ε

) ∫

∂B(0,r)

dS(y) dr

= u(x)1

εn

B(0,ε)

η

( |y|ε

)dy

= u(x)

B(0,ε)

ηε(y) dy

= u(x).

12

Page 13: 3 Laplace’s Equation - Stanford Universityweb.stanford.edu/class/math220b/handouts/laplace.pdf3 Laplace’s Equation We now turn to studying Laplace’s equation ∆u = 0 and its

3.2.5 Liouville’s Theorem

In this section, we show that the only functions which are bounded and harmonic on Rn areconstant functions.

Theorem 8. Suppose u : Rn → R is harmonic and bounded. Then u is constant.

Proof. Let x0 ∈ Rn. By the mean value property,

u(x0) = −∫

B(x0,r)

u(y) dy

for all B(x0, r). Now by the previous theorem, we know that if u ∈ C2(Ω) ∩ C(Ω) and u isharmonic, then u is C∞. Therefore,

∆u = 0 =⇒ ∆uxi= 0

for i = 1, . . . , n. Therefore, uxiis harmonic and satisfies the mean value property. Therefore,

uxi(x0) = −

B(x0,r)

uxi(y) dy

=1

α(n)rn

B(x0,r)

uxi(y) dy

=1

α(n)rn

∂B(x0,r)

uνi dS(y),

by the Divergence theorem, where ν = (ν1, . . . , νn) is the outward unit normal to B(x0, r).Therefore,

|uxi(x0)| ≤

∣∣∣∣1

α(n)rn

∂B(x0,r)

uνi dS(y)

∣∣∣∣

≤ |u|L∞(∂B(x0,r))|νi|L∞∣∣∣∣

1

α(n)rn

∂B(x0,r)

dS(y)

∣∣∣∣

≤ |u|L∞(Rn)

∣∣∣∣nα(n)rn−1

α(n)rn

∣∣∣∣≤ n

r|u|L∞(Rn).

Therefore,

|uxi(x0)| ≤ n

r|u|L∞(Rn)

≤ Cn

r,

by the assumption that u is bounded. Now this is true for all r. Taking the limit as r → +∞,we see that |uxi

(x0)| = 0. Therefore, uxi(x0) = 0. This is true for i = 1, . . . , n and for all

x0 ∈ Rn. Therefore, we conclude that u ≡ constant.

13

Page 14: 3 Laplace’s Equation - Stanford Universityweb.stanford.edu/class/math220b/handouts/laplace.pdf3 Laplace’s Equation We now turn to studying Laplace’s equation ∆u = 0 and its

As a corollary of Liouville’s Theorem, we have the following representation formula forall bounded solutions of Poisson’s equation on Rn, n ≥ 3.

Theorem 9. (Representation Formula) Let f ∈ C2(Rn) with compact support. Let n ≥ 3.Then every bounded solution of

−∆u = f x ∈ Rn (3.4)

has the form

u(x) =

Rn

Φ(x− y)f(y) dy + C

for some constant C, where Φ(x) is the fundamental solution of Laplace’s equation in Rn.

Proof. Recall that the fundamental solution of Laplace’s equation in Rn, n ≥ 3 is given by

Φ(x) =K

|x|n−2

where K = 1/n(n− 2)α(n). As shown earlier,

u(x) =

Rn

Φ(x− y)f(y) dy

is a solution of (3.4). Here we show this is a bounded solution for n ≥ 3. Fix ε > 0. Then,we have

|u(x)| =∣∣∣∣∫

Rn

Φ(x− y)f(y) dy

∣∣∣∣

=

∣∣∣∣K∫

Rn

1

|x− y|n−2f(y) dy

∣∣∣∣

=

∣∣∣∣K∫

B(x,ε)

1

|x− y|n−2f(y) dy

∣∣∣∣ +

∣∣∣∣K∫

Rn−B(x,ε)

1

|x− y|n−2f(y) dy

∣∣∣∣

≤ |f(y)|L∞∫

B(x,ε)

1

|x− y|n−2dy + C

Rn−B(x,ε)

|f(y)| dy.

It is easy to see that the first term on the right-hand side is bounded. The second termon the right-hand side is bounded, using the assumption that f ∈ C2(Rn) with compactsupport. Therefore, we conclude that

u(x) =

Rn

Φ(x− y)f(y) dy

is a bounded solution of (3.4). Now suppose there is another bounded solution of (3.4). Letu be such a solution. Let

w(x) = u(x)− u(x).

14

Page 15: 3 Laplace’s Equation - Stanford Universityweb.stanford.edu/class/math220b/handouts/laplace.pdf3 Laplace’s Equation We now turn to studying Laplace’s equation ∆u = 0 and its

Then w is a bounded, harmonic function on Rn. Then, by Liouville’s Theorem, w must beconstant. Therefore, we conclude that

u(x) = u(x) + C

=

Rn

Φ(x− y)f(y) dy + C,

as claimed.

3.3 Solving Laplace’s Equation on Bounded Domains

3.3.1 Laplace’s Equation on a Rectangle

In this section, we will solve Laplace’s equation on a rectangle in R2. First, we consider thecase of Dirichlet boundary conditions. That is, we consider the following boundary valueproblem. Let Ω = (x, y) ∈ R2 : 0 < x < a, 0 < y < b. We want to look for a solution ofthe following,

uxx + uyy = 0 (x, y) ∈ Ωu(0, y) = g1(y), u(a, y) = g2(y) 0 < y < bu(x, 0) = g3(x), u(x, b) = g4(y) 0 < x < a.

(3.5)

In order to do so, we consider the following simpler example. From this, we will show howto solve the more general problem above.

Example 10. Let Ω = (x, y) ∈ R2 : 0 < x < a, 0 < y < b. Consider

uxx + uyy = 0 (x, y) ∈ Ωu(0, y) = g1(y), u(a, y) = 0 0 < y < bu(x, 0) = 0, u(x, b) = 0 0 < x < a.

(3.6)

We use separation of variables. We look for a solution of the form

u(x, y) = X(x)Y (y).

Plugging this into our equation, we get

X ′′Y + XY ′′ = 0.

Now dividing by XY , we arrive at

X ′′

X+

Y ′′

Y= 0,

which impliesY ′′

Y= −X ′′

X= −λ

15

Page 16: 3 Laplace’s Equation - Stanford Universityweb.stanford.edu/class/math220b/handouts/laplace.pdf3 Laplace’s Equation We now turn to studying Laplace’s equation ∆u = 0 and its

for some constant λ. By our boundary conditions, we want Y (0) = 0 = Y (b). Therefore, webegin by solving the eigenvalue problem,

Y ′′ = −λY 0 < y < b

Y (0) = 0 = Y (b).

As we know, the solutions of this eigenvalue problem are given by

Yn(y) = sin(nπ

by)

, λn =(nπ

b

)2

.

We now turn to solving

X ′′ =(nπ

b

)2

X

with the boundary condition X(a) = 0. The solutions of this ODE are given by

Xn(x) = An cosh(nπ

bx)

+ Bn sinh(nπ

bx)

.

Now the boundary condition X(a) = 0 implies

An cosh(nπ

ba)

+ Bn sinh(nπ

ba)

= 0.

Therefore,

un(x, y) = Xn(x)Yn(y) =[An cosh

(nπ

bx)

+ Bn sinh(nπ

bx)]

sin(nπ

by)

where An, Bn satisfy the condition

An cosh(nπ

ba)

+ Bn sinh(nπ

ba)

= 0.

is a solution of Laplace’s equation on Ω which satisfies the boundary conditions u(x, 0) = 0,u(x, b) = 0, and u(a, y) = 0. As we know, Laplace’s equation is linear. Therefore, wecan take any combination of solutions un and get a solution of Laplace’s equation whichsatisfies these three boundary conditions. Therefore, we look for a solution of the form

u(x, y) =∞∑

n=1

un(x, y) =∞∑

n=1

[An cosh

(nπ

bx)

+ Bn sinh(nπ

bx)]

sin(nπ

by)

where An, Bn satisfy

An cosh(nπ

ba)

+ Bn sinh(nπ

ba)

= 0. (3.7)

To solve our boundary-value problem (3.6), it remains to find coefficients An, Bn which notonly satisfy (3.7), but also satisfy the condition u(0, y) = g1(y). That is, we need

u(0, y) =∞∑

n=1

An sin(nπ

by)

= g1(y).

16

Page 17: 3 Laplace’s Equation - Stanford Universityweb.stanford.edu/class/math220b/handouts/laplace.pdf3 Laplace’s Equation We now turn to studying Laplace’s equation ∆u = 0 and its

That is, we want to be able to express g1 in terms of its Fourier sine series on the interval[0, b]. Assuming g1 is a “nice” function, we can do this. From our earlier discussion of Fourierseries, we know that the Fourier sine series of a function g1 is given by

g1(y) ∼∞∑

n=1

An sin(nπ

by)

where the coefficients An are given by

An =〈g1, sin

(nπb

y)〉

〈sin (nπb

y), sin

(nπb

y)

where the L2-inner product is taken over the interval [0, b].Therefore, to summarize, we have found a solution of (3.6) given by

u(x, y) =∞∑

n=1

un(x, y) =∞∑

n=1

[An cosh

(nπ

bx)

+ Bn sinh(nπ

bx)]

sin(nπ

by)

where

An =〈g1, sin

(nπb

y)〉

〈sin (nπb

y), sin

(nπb

y)

and

Bn = − coth(nπ

ba)

An.

¦Now we return to considering (3.5). For the general boundary value problem on a rect-

angle with Dirichlet boundary conditions, we can find a solution by finding four separatesolutions ui for i = 1, . . . , 4 such that each ui is identically zero on three of the sides andsatisfies the boundary condition on the fourth side. For example, for the boundary valueproblem (3.5), we use the procedure in the above example to find a function u1(x, y) whichis harmonic on Ω and such that u1(0, y) = g1(y) and u1(a, y) = 0 for 0 < y < b, andu1(x, 0) = 0 = u1(x, b) for 0 < x < a. Similar we find functions u2, u3 and u4 which vanishon three of the sides but satisfy the fourth boundary condition.

We now consider an example where we have a mixed boundary condition on one side.

Example 11. Let Ω = (x, y) ∈ R2, 0 < x < L, 0 < y < H. Consider the followingboundary value problem,

uxx + uyy = 0 (x, y) ∈ Ωu(0, y) = 0, u(L, y) = 0 0 < y < Hu(x, 0)− uy(x, 0) = 0, u(x,H) = f(x) 0 < x < L.

(3.8)

Using separation of variables, we have

X ′′

X= −Y ′′

Y= −λ.

17

Page 18: 3 Laplace’s Equation - Stanford Universityweb.stanford.edu/class/math220b/handouts/laplace.pdf3 Laplace’s Equation We now turn to studying Laplace’s equation ∆u = 0 and its

We first look to solve X ′′ = −λX 0 < x < LX(0) = 0 = X(L).

As we know, the solutions of this eigenvalue problem are given by

Xn(x) = sin(nπ

Lx)

, λn =(nπ

L

)2

.

Now we need to solve

−Y ′′ = −(nπ

L

)2

Y

with the boundary condition Y (0)− Y ′(0) = 0. The solutions of this ODE are given by

Yn(y) = An cosh(nπ

Ly)

+ Bn sinh(nπ

Ly)

.

The boundary condition Y (0)− Y ′(0) = 0 implies

An −Bnnπ

L= 0.

Therefore,

Yn(y) = Bnnπ

Lcosh

(nπ

Ly)

+ Bn sinh(nπ

Ly)

.

Therefore, we look for a solution of (3.8) of the form

u(x, y) =∞∑

n=1

Bn sin(nπ

Lx) [nπ

Lcosh

(nπ

Ly)

+ sinh(nπ

Ly)]

.

Substituting in the condition u(x,H) = f(x), we have

u(x,H) =∞∑

n=1

Bn sin(nπ

Lx) [nπ

Lcosh

(nπ

LH

)+ sinh

(nπ

LH

)]= f(x).

Recall the Fourier sine series of f on [0, L] is given by

f ∼∞∑

n=1

An sin(nπ

Lx)

where

An =〈f, sin

(nπL

x)〉

〈sin (nπL

x), sin

(nπL

x)〉

where the L2-inner product is taken over (0, L). Therefore, in order for our boundary con-dition u(x,H) = f(x) to be satisfied, we need Bn to satisfy

Bn

[nπ

Lcosh

(nπ

LH

)+ sinh

(nπ

LH

)]=

〈f, sin(

nπL

x)〉

〈sin (nπL

x), sin

(nπL

x)〉 .

18

Page 19: 3 Laplace’s Equation - Stanford Universityweb.stanford.edu/class/math220b/handouts/laplace.pdf3 Laplace’s Equation We now turn to studying Laplace’s equation ∆u = 0 and its

Using the fact that

〈sin(nπ

Lx)

, sin(nπ

Lx)〉 =

∫ L

0

sin2(nπ

Lx)

dx =L

2,

the solution of (3.8) is given by

u(x, y) =∞∑

n=1

Bn sin(nπ

Lx) [nπ

Lcosh

(nπ

Ly)

+ sinh(nπ

Ly)]

where

Bn =2

L

[nπ

Lcosh

(nπ

LH

)+ sinh

(nπ

LH

)]−1∫ L

0

f(x) sin(nπ

Lx)

dx.

¦

3.3.2 Laplace’s Equation on a Disk

In this section, we consider Laplace’s Equation on a disk in R2. That is, let Ω = (x, y) ∈R2 : x2 + y2 < a2. Consider

uxx + uyy = 0 (x, y) ∈ Ωu = h(θ) (x, y) ∈ ∂Ω.

(3.9)

To solve, we write this equation in polar coordinates as follows. To transform our equationin to polar coordinates, we will write the operators ∂x and ∂y in polar coordinates. We willuse the fact that

x2 + y2 = r2

y

x= tan θ.

Consider a function u such that u = u(r, θ), where r = r(x, y) and θ = θ(x, y). That is,

u = u(r(x, y), θ(x, y)).

Then

∂xu(r(x, y), θ(x, y)) = urrx + uθθx

= urx

(x2 + y2)1/2− uθ

y

x2 sec2 θ

= ur cos θ − sin θ

ruθ.

Therefore, the operator ∂∂x

can be written in polar coordinates as

∂x= cos θ

∂r− sin θ

r

∂θ.

19

Page 20: 3 Laplace’s Equation - Stanford Universityweb.stanford.edu/class/math220b/handouts/laplace.pdf3 Laplace’s Equation We now turn to studying Laplace’s equation ∆u = 0 and its

Similarly, the operator ∂∂y

can be written in polar coordinates as

∂y= sin θ

∂r+

cos θ

r

∂θ.

Now squaring these operators we have

∂2

∂x2=

[cos θ

∂r− sin θ

r

∂θ

]2

= cos2 θ∂2

∂r2+ 2

sin θ cos θ

r2

∂θ− 2

sin θ cos θ

r

∂2

∂r∂θ+

sin2 θ

r

∂r+

sin2 θ

r2

∂2

∂θ2.

Similarly,

∂2

∂y2=

[sin θ

∂r+

cos θ

r

∂θ

]2

= sin2 θ∂2

∂r2− 2

sin θ cos θ

r2

∂θ+ 2

sin θ cos θ

r

∂2

∂r∂θ+

cos2 θ

r

∂r+

cos2 θ

r2

∂2

∂θ2.

Combining the above terms, we can write the operator ∂2x + ∂2

y in polar coordinates asfollows,

∂2

∂x2+

∂2

∂y2=

∂2

∂r2+

1

r

∂r+

1

r2

∂2

∂θ2.

Therefore, in polar coordinates, Laplace’s equation is written as

urr +1

rur +

1

r2uθθ = 0. (3.10)

Now we will solve it using separation of variables. In particular, we look for a solution ofthe form u(r, θ) = R(r)Θ(θ). Then letting u(x, y) = u(r(x, y), θ(x, y)), we will arrive at asolution of Laplace’s equation on the disk.

Substituting a function of the form u(r, θ) = R(r)Θ(θ), into (3.10), our equation is writtenas

R′′Θ +1

rR′Θ +

1

r2RΘ′′ = 0.

Dividing by RΘ,R′′

R+

R′

rR+

Θ′′

r2Θ= 0.

Multiplying by r2, we are led to the equations

Θ′′

Θ= −r2R′′

R− rR′

R= −λ

for some scalar λ. The boundary condition for this problem is u = h(θ) for (x, y) ∈ ∂Ω.Therefore, we are led to the following eigenvalue problem with periodic boundary conditions,

Θ′′ = −λΘ 0 < θ < 2πΘ(0) = Θ(2π), Θ′(0) = Θ′(2π).

20

Page 21: 3 Laplace’s Equation - Stanford Universityweb.stanford.edu/class/math220b/handouts/laplace.pdf3 Laplace’s Equation We now turn to studying Laplace’s equation ∆u = 0 and its

Recall from our earlier work that periodic boundary conditions imply our eigenfunctions andeigenvalues are

Θn(θ) = An cos(nθ) + Bn sin(nθ), λn = n2 n = 0, 1, 2, . . .

For each λn, we need to solver2R′′

n + rR′n = λnRn.

That is, we need to solve the second-order ODE,

r2R′′n + rR′

n − n2Rn = 0

for n = 0, 1, 2, . . .. Recall that a second-order ODE will have two linearly independentsolutions. We look for a solution of the form R(r) = rα for some α. Doing so, our ODEbecomes

(α2 − n2)rα = 0.

Therefore, for n ≥ 1, we have found two linearly independent solutions, Rn(r) = rn andRn(r) = r−n. Now for n = 0, we have only found one linearly independent solution so far,R0(r) = 1. We look for another linearly independent solution. If n = 0, our equation can bewritten as

r2R′′ + rR′ = 0.

Dividing by r, our equation becomes

rR′′ + R′ = 0.

A linearly independent solution of this equation is R0(r) = ln r. Therefore, for each n ≥ 0,we have found a solution of (3.10) of the form

un(r, θ) = Rn(r)Θn(θ) =

[Cnrn +

Dn

rn

][An cos(nθ) + Bn sin(nθ)]

A0 [C0 + D0 ln r] .

But, we don’t want a solution which blows up as r → 0+. Therefore, we reject the solutions1rn and ln r. Therefore, we consider a solution of (3.10) of the form

u(r, θ) =∞∑

n=0

rn [An cos(nθ) + Bn sin(nθ)] .

Now in order to solve (3.9), we need u(a, θ) = h(θ). That is, we need

∞∑n=0

an [An cos(nθ) + Bn sin(nθ)] = h(θ).

Using the fact that our eigenfunctions are orthogonal on [0, 2π], we can solve for our coef-ficients An and Bn as follows. Multiplying the above equation by cos(nθ) and integratingover [0, 2π], we have

An =1

an

〈h(θ), cos(nθ)〉〈cos(nθ), cos(nθ)〉 =

1

πan

∫ 2π

0

h(θ) cos(nθ) dθ for n = 1, 2, . . .

21

Page 22: 3 Laplace’s Equation - Stanford Universityweb.stanford.edu/class/math220b/handouts/laplace.pdf3 Laplace’s Equation We now turn to studying Laplace’s equation ∆u = 0 and its

A0 =〈h(θ), 1〉〈1, 1〉 =

1

∫ 2π

0

h(θ) dθ.

Similarly, multiplying by sin(nθ) and integrating over [0, 2π], we have

Bn =1

an

〈h(θ), sin(nθ)〉〈sin(nθ), sin(nθ)〉 =

1

πan

∫ 2π

0

h(θ) sin(nθ) dθ.

To summarize, we have found a solution of Laplace’s equation on the disk in polarcoordinates, given by

u(r, θ) =∞∑

n=0

rn [An cos(nθ) + Bn sin(nθ)]

where

A0 =1

∫ 2π

0

h(θ) dθ

An =1

πan

∫ 2π

0

h(θ) cos(nθ) dθ

Bn =1

πan

∫ 2π

0

h(θ) sin(nθ) dθ.

Now we will rewrite this solution in terms of a single integral by substituting An and Bn

into the series solution above. Doing so, we have

u(r, θ) =1

∫ 2π

0

h(φ) dφ

+∞∑

n=1

rn

[1

πan

∫ 2π

0

h(φ) cos(nφ) dφ

]cos(nθ) +

[1

πan

∫ 2π

0

h(φ) sin(nφ) dφ)

]sin(nθ)

=1

∫ 2π

0

h(φ)

1 + 2

∞∑n=1

[rn

an[cos(nφ) cos(nθ) + sin(nφ) sin(nθ)]

]dφ

=1

∫ 2π

0

h(φ)

1 + 2

∞∑n=1

[rn

ancos(n(θ − φ))

]dφ.

Now

1 + 2∞∑

n=1

rn

ancos(n(θ − φ))

=

1 + 2

∞∑n=1

rn

an

[ein(θ−φ) + e−in(θ−φ)

2

]

= 1 +

∞∑n=1

(rei(θ−φ)

a

)n

+∞∑

n=1

(re−i(θ−φ)

a

)n

= 1 +rei(θ−φ)

a− rei(θ−φ)+

re−i(θ−φ)

a− re−i(θ−φ)

=a2 − r2

a2 − 2ar cos(θ − φ) + r2.

22

Page 23: 3 Laplace’s Equation - Stanford Universityweb.stanford.edu/class/math220b/handouts/laplace.pdf3 Laplace’s Equation We now turn to studying Laplace’s equation ∆u = 0 and its

Therefore,

u(r, θ) =1

∫ 2π

0

h(φ)a2 − r2

a2 − 2ar cos(θ − φ) + r2dφ.

We can write this in rectangular coordinates as follows. Let x be a point in the disk Ωwith polar coordinates (r, θ). Let x′ be a point on the boundary of the disk Ω with polarcoordinates (a, φ). Therefore, |x − x′|2 = a2 + r2 − 2ar cos(θ − φ) by the law of cosines.Therefore,

u(x) =1

|x′|=a

u(x′)(a2 − |x′|2)|x− x′|2

ds

a,

using the fact that ds = a dφ is the arc length of the curve. Rewriting this, we have

u(x) =a2 − |x′|2

2πa

|x′|=a

u(x′)|x− x′|2 ds.

This is known as Poisson’s formula for the solution of Laplace’s equation on the disk.

23