P435 Lect 07
P435 Lect 07
Steven Errede
LECTURE NOTES 7
LAPLACE’S EQUATION
As we have seen in previous lectures, very often the primary task in an electrostatics problem
is e.g. to determine the electric field E ( r ) of a given stationary/static charge distribution
– e.g. via Coulomb’s Law:
Charge density
ρ ( r ′) ẑ
Field Point
Source r = r − r′ P
Point(s)
S r
v′
Volume r′
element dτ ′ Ο ŷ
in volume v′
x̂
1 rˆ r r − r′
E (r ) = ∫ r ρ ( r ′) dτ ′ r = r − r′ rˆ = =
4πε o v′
2
r r − r′
(x − xs ) + ( y p − ys ) + ( z p − zs )
2 2 2
r = r − r′ = p
Oftentimes it is easier to first calculate the potential V ( r ) , and then use E ( r ) = −∇V ( r )
1 1
Here: V ( r ) = ∫ r ρ ( r ′) dτ ′
4πε o v′
But even doing this integral analytically often can be very challenging. . .
Furthermore, often in problems involving conductors, ρ ( r ′ ) may not apriori (i.e. beforehand)
be known! Charge is free to move around, and often only the total free charge Qfree is controlled
/ known in the problem.
In such cases, it is usually better to recast the problem in DIFFERENTIAL form, using Poisson’s
equation:
ρ (r )
∇i E ( r ) = −∇i∇V ( r ) = −∇ 2V ( r ) =
εo
ρ (r )
Or: ∇ 2V ( r ) = − ⇐ Poisson’s Equation
εo
Poisson’s equation, together with the boundary conditions associated with the value(s) allowed
for V ( r ) e.g. on various conducting surfaces, or at r = ∞, etc. enables one to uniquely determine
V ( r ) (we’ll see how / why shortly. . .).
∇ 2V ( r ) = 0 ⇐ Laplace’s Equation
commensurate with the boundary conditions for the specific problem at hand.
Very often, in fact, we are interested in finding the potential V ( r ) in a charge-free region,
containing no electric charge, i.e. where ρ ( r ′ ) = 0.
Now, before we go any further on this discussion, let’s back up a bit and take a (very) broad
generalized MATHEMATICAL view (or approach) to find V ( r ) .
d 2V ( x ) d ⎛ dV ( x ) ⎞ ⎛ dV ⎞ dV ( x )
∫ 2
d x
dx = ∫ ⎜
dx ⎝ dx ⎠
⎟ dx = ∫ d ⎜
⎝ dx
⎟=
⎠ dx
= ∫ dx = m = 1st constant of integration
dV ( x )
Then: ∫ dx
dx = ∫ mdx = m ∫ dx
∫ dV ( x ) = V ( x ) = mx + b ← 2
nd
Or: constant of integration
d 2V ( x )
So: V ( x) = b + mx (equation for a straight line) is the general solution for = 0.
dx 2
y-intercept slope
Depending on the boundary conditions for the problem, e.g. suppose V ( x = 5 ) = 0 Volts and
V ( x = 1) = 4 Volts, then together, these two boundary conditions uniquely specify what b and m
must be – we have two equations, and two unknowns (m & b) – solve simultaneously:
V ( x = 5 ) = 0 = b + 5m → b = −5m
V ( x = 1) = 4 = b + 1m → 4 = −5m + 1m = −4m
∴ V ( x ) = 5 − 1x or: m = −1 and b = 5.
V ( x ) = 5 − 1x is the equation of a straight line for this problem.
5 b = 5 is y-intercept
4
V(x)
3
2 slope m = −1
0 1 2 3 4 5 x
x−a x x+a
The solutions of V ( x ) are as “boring” as possible, but fit the endpoints (boundary conditions)
properly.
This may be “obvious” in one-dimension, but it is also true / also holds in 2-D and 3-D cases of
∇ 2V ( r ) = 0 .
∂ 2V ∂ 2V
If V = V ( x, y ) then ∇ 2V = 0 ⇒ + =0
∂x 2 ∂y 2
Nevertheless, V ( x, y ) will still wind up being the average value of V around a point ( x, y )
within a circle of radius R centered on the point ( x, y ) .
1
V ( x, y ) = V ( r ) dl
2π R circle∫C of
radius R
centered
on (x,y)
iteration n iteration n −1
ΔV ( x, y ) = Vn ( x, y ) − Vn −1 ( x, y ) ≤ tolerance
V ( x, y ) again will have no local maxima or minima – all extrema will occur on boundaries.
∇ 2V ( x, y ) = 0 has solution V ( x, y ) which is the most featureless function – as smooth as
possible.
1
i.e. V (r ) = ∫ Vda
4π R 2 sphere at r
of radius R
Uniqueness Theorem(s):
Suppose we have two solutions of Laplace’s equation, V1 ( r ) and V2 ( r ) , each satisfying the
same boundary condition(s), i.e. the potentials V1 ( r ) and V2 ( r ) are specified on the boundaries.
We assert that the two solutions can at most differ by a constant. (n.b. Only differences in the
scalar potential V ( r ) are important / physically meaningful!)
Proof: Consider a closed region of space with volume v which is exterior to n charged
conducting surfaces S1, S2, S3. . . Sn that are responsible for generating the potential V.
The volume v is bounded (outside) by the surface S.
S4 v, S
Suppose we have two solutions V1(r) and V2(r) both satisfying ∇ 2V ( r ) = 0 i.e. ∇ 2V1 ( r ) = 0
and ∇ 2V2 ( r ) = 0 in the charge-free region(s) of the volume v.
V1(r) and V2(r) satisfy either Dirichlet boundary conditions or satisfy Neumann boundary
conditions ∇V ( r )inˆ on the surfaces S1, S2, S3. . . Sn. We also demand that V(r) be finite at r = ∞.
The potentials Vi =1,2 are uniquely specified on charged (equipotential) surfaces S1, S2, S3, . . . Sn
in the volume v .
( )
Now apply the divergence theorem to the quantity VΔ ∇VΔ ; we also define: EΔ ( r ) ≡ −∇VΔ ( r )
∫ ∇i(V ∇V ) dτ =
v
Δ Δ
S+
∫ (V ∇V )idA =
Δ Δ
S+
∫ ( )
VΔ − EΔ idA
S1 + S2 + S3 +... Sn S1 + S2 + S3 +... Sn
Volume integral
over enclosing Surface integral over
volume v ALL surfaces in v
Then:
− ∫ ( ) ( )
VΔ EΔ idA = − ∫ VΔ EΔ idAS − ∫ VΔ EΔ idAS1 − ∫ VΔ EΔ idAS2 − ... − ∫ VΔ EΔ idASn ( ) ( ) ( )
S+ S S1 S2 Sn
S1 + S2 + S3 +... Sn
Recognizing that:
1. The conducting surfaces S1, S2, S3, . . . Sn, are equipotentials.
Thus: VΔ ( r ) = V1 ( r ) − V2 ( r ) (= a constant on surfaces S1, S2, S3, . . . Sn) must = 0 at/on those surfaces!!!
2. The volume v is arbitrary, so let’s choose volume v → ∞, and thus surface area S → ∞ as well.
3. ∫ EΔ idAi = Φ Ei = electric flux through ith surface.
Si
∴ ∫ (
∇i VΔ ∇VΔ dτ = − VΔS ) ∫ EΔ idAS − VΔS1 ∫ EΔ idAS1 − VΔS2 ∫E
Δ idAS2 − ... − VΔSn ∫E
Δ idASn
v =0 S = 0 S1 = 0 S2 = 0 Sn
all space all space
= Φ SE1 = Φ SE2 = Φ En
S
= Φ SE
Thus: ∫ (
∇i VΔ ∇VΔ dτ = 0 )
v
all space
=∇VΔ i ∇VΔ
( ) VΔ ( ∇ 2VΔ ) dτ + ∫ ( ∇V )
2
Then:
v
∫ ∇i VΔ ∇VΔ dτ = ∫
v v
Δ dτ = 0
all space all space =0 all space
∫ ( ∇V ) ∫ ( ∇V i∇V ) dτ = 0
2
= Δ dτ = Δ Δ
v v
all space mathematically all space
≥0
can be
∫ ( ∇V ) ( ) = ( ∇V ( r )i∇VΔ ( r ) ) = 0 .
2 2
The only way Δ dτ = 0 is iff (i.e. if and only if) the integrand ∇VΔ ( r ) Δ
V
mathematically
≥0
i.e. the two solutions V1(r) and V2(r) for ∇ 2V ( r ) = 0 are identical – there is only one unique solution.
If ∇V1 inˆ = − E1⊥ and ∇V2 inˆ = − E2⊥ are specified on the surfaces S1, S2, S3, . . ., Sn in the volume v
enclosed by surface S (Neumann boundary conditions), then ∇VΔ ( r ) = ∇V1 ( r ) − ∇V2 ( r ) = 0 at
all points in volume v and ∇VΔ inˆ = 0.
Here, solutions V1(r) and V2(r) can differ, but only by a constant Vo.
e.g. V1 ( r ) = V2 ( r ) + Vo ⇒ problem is NOT over-determined for V ( r ) .
( E ( r ) is over-determined / unique, but not V ( r ) ).
Physical Example:
If we instead specify the charge densit(ies) ρ ( r ) within the volume v (see figure below), then we
also have a uniqueness theorem for the electric field associated with Poisson’s equation
(∇i E ( r ) = −∇ 2V ( r ) = ρ ( r ) ε o . )
Q4 S4 v, S
ρ ( r ) specified
Suppose there are two electric fields E1 ( r ) and E2 ( r ) , both satisfying all of the boundary
conditions of this problem. Both obey Gauss’ law in differential and integral form everywhere
within the volume v:
∇i E1 ( r ) = ρ ( r ) ε o and: ∇i E2 ( r ) = ρ ( r ) ε o
1 1
∫ E1 ida =
εo
Qiencl and: ∫ E2 ida =
εo
Qiencl
i th conducting i th conducting
surface , Si surface , Si
Even though we do not know how the charge Qi on the ith conducing surface Si is distributed,
we do know that each surface Si is an equipotential, hence the scalar potential VΔ ≡ V1 − V2 on
each surface is at least a constant on each surface Si (n.b. VΔ may not necessarily be = 0, since in
general V2 may not in general be equal to V1 on each/every surface Si).
( )
Using Griffith’s product rule # 5: ∇i fA = f ∇i A + Ai ∇f , then: ( ) ( )
( )
∇i VΔ EΔ = VΔ ∇i EΔ + EΔ i ∇VΔ( ) ( )
However, in the region between conductors, we have shown (above) that ∇i EΔ ( r ) = 0 ,
( ) (
and EΔ ≡ −∇VΔ , hence: ∇i VΔ EΔ = EΔ i ∇VΔ = − EΔ i EΔ = − EΔ2 . )
If we integrate this relation over the entire volume v (with associated enclosing surface S):
∫ ∇i(V E ) dτ = ∫
v Δ Δ all S
VΔ EΔ ida = − ∫ EΔ2 dτ
v
Note that the surface integral covers all boundaries of the region in question – the enclosing outer
surface S and all of the Si inner surfaces associated with the i conductors. Since VΔ is a constant
on each surface, it can be pulled outside of the surface integral (n.b. if the outer surface S is at
infinity, then for localized sources of charge, VΔ ( r = ∞ ) = 0 ). Thus:
VΔ ∫ EΔ ida = − ∫ EΔ2 dτ
all S v
Hence, in general, the only way that this integral can vanish is if EΔ ( r ) ≡ E1 ( r ) − E2 ( r ) = 0
everywhere, thus, we must have E1 ( r ) = E2 ( r ) .
In general, when solving the potential V ( r ) problems in 3 (or less) dimensions, first note the
symmetries associated with the problem. Then, if you have:
In 2-D and 3-D problems, the general solutions to ∇ 2V ( r ) = 0 are the harmonic functions (an ∞-
series solution, in principle) e.g. of sines and cosines, Bessel functions, or Legendre
Polynomials and/or Spherical Harmonics.
⎛ ∂2 ∂2 ∂2 ⎞ ∂ 2V ∂ 2V ∂ 2V
∇ 2V ( x, y, z ) = ⎜ 2 + 2 + 2 ⎟ V ( x, y, z ) = 2 + 2 + 2 = 0
⎝ ∂x ∂y ∂z ⎠ ∂x ∂y ∂z
The solutions of ∇ 2V = 0 in rectangular coordinates are known as harmonic functions (i.e. sines
and cosines) (→ Fourier Series Solutions).
It is usually (but not always) possible to find a solution to the Laplace Equation, ∇ 2V = 0 which
also satisfies the boundary conditions, via separation of variables technique, i.e. try a product
solution of the form:
V ( x, y , z ) = X ( x ) Y ( y ) Z ( z )
where: X (x) x
Y (y) are functions only of y respectively.
Z (z) z
∂ 2V ( x, y, z ) ∂ 2V ( x, y, z ) ∂ 2V ( x, y, z )
Then: ∇ 2V ( x, y, z ) = 0 ⇒ + + =0
∂x 2 ∂y 2 ∂z 2
But: V ( x, y, z ) = X ( x ) Y ( y ) Z ( z )
∂2 X ( x)Y ( y ) Z ( z ) ∂2 X ( x)Y ( y ) Z ( z ) ∂2 X ( x )Y ( y ) Z ( z )
Thus: + + =0
∂x 2 ∂y 2 ∂z 2
∂2 X ( x) ∂ 2Y ( y ) ∂2Z ( z )
= Y ( y) Z ( z) + X ( x) Z ( z ) + X ( x)Y ( y ) =0
∂x 2 ∂y 2 ∂z 2
1 ∂2 X ( x) 1 ∂ Y ( y)
2
1 ∂ Z ( z)
2
Then: = + + =0
X ( x ) ∂x 2 Y ( y ) ∂y 2 Z ( z ) ∂z 2
1 ∂2 X ( x) 1 ∂ Y ( y)
2
1 ∂ Z ( z)
2
But: = + + = 0 i.e. C1 + C2 + C3 = 0
X ( x ) ∂x 2 Y ( y ) ∂y 2 Z ( z ) ∂z 2
fcn(x) only fcn(y) only fcn(z) only
independent of y, z independent of x, z independent of x, y
= C1 = C2 = C3
1 ∂2 X ( x) d 2 X ( x)
= constant C1 ⇒ − C1 X ( x ) = 0 #1
X ( x) ∂x 2 dx 2
1 ∂ Y ( y) d 2Y ( y )
2
= constant C 2 ⇒ − C2Y ( y ) = 0 #2
Y ( y ) ∂y 2 dy 2
1 ∂ Z ( z) d 2Z ( z )
2
= constant C3 ⇒ − C3 Z ( z ) = 0 #3
Z ( z ) ∂z 2 dz 2
Note total derivatives now!!!
Subject to the constraint: C1 + C2 + C3 = 0
Can now solve 3 ORDINARY 1-D differential equations, #1–3, which are subject to C1 + C2 + C3 = 0,
PLUS the specific Dirichlet / Neumann boundary conditions for the problem on either V (x, y, z)
or ∇V ( x, y, z )inˆ at surfaces for this 3-D problem.
Essentially, we have replaced the 3-D problem with three 1-D problems, and the constraint:
C1 + C2 + C3 = 0.
* If one has a 2-D rectangular coordinate problem ( ∇ 2V ( x, y ) = 0 ) , then: V (x, y) = X(x)Y(y) (only).
1 d 2 X ( x) d 2 X ( x)
= C1 ⇒ − C1 X ( x ) = 0
X ( x ) dx 2 dx 2
1 d Y ( y) d 2Y ( y )
2
= C2 ⇒ − C2Y ( y ) = 0
Y ( y ) dy 2 dy 2
* If one has a 1-D rectangular coordinate problem ( ∇ 2V ( x ) = 0 ) , then: V(x) = X(x) (only).
d 2V ( x ) d 2 X ( x) 1 d X ( x)
2
= 0 ⇒ = 0 ⇒ = 0 = C1
dx 2 dx 2 X ( x) dx 2
d 2 X ( x)
= 0 ⇒ X ( x) = V ( x) = ax + b is the 1-D general solution.
dx 2
For 1-D problem ( ∇ 2V ( x ) = 0 ) , only need to solve one ordinary differential equation subject to
dV ( x)
the constraint C1 = 0 and BC’s on either V(x) or .
dx
Since we have the constraint C1 + C2 + C3 = 0, at least one of the Ci’s (i = 1, 2 or 3) must be less
than zero.
Let us “choose” C1 = −α 2 , C2 = − β 2 , C3 = γ 2
Then: C1 + C2 + C3 = 0
−α 2 − β 2 + γ 2 = 0 or: α2 + β2 = γ 2
The boundary conditions on the surfaces will define α and β , and hence define γ .
IMPORTANT NOTE:
The geometry (x – y – z) of the problem and the boundary conditions dictate whether:
C1 > 0 or C1 < 0
C2 > 0 or C2 < 0
C3 > 0 or C3 < 0
i.e. have sine / cosine type solutions vs. sinh / cosh (or e x , e− x ) type solutions for x, y, z.
∞
+ ∑ Bmn cos (α n x ) cos ( β m y ) sinh ( γ mn z )
m,n=0
= α n2 + β m2
∞
+ ∑ Cmn sin (α n x ) sin ( β m y ) cosh ( γ mn z )
m,n=0
= α n2 + β m2
∞
+ ∑ Dmn cos (α n x ) cos ( β m y ) cosh ( γ mn z )
m,n=0
= α n2 + β m2
1 x −x 1 x −x
n.b. cosh( x) =
2
(e + e ) sinh( x) =
2
(e − e )
1 ix − ix 1
n.b. sin( x) =
2i
(e − e ) cos( x) = ( eix + e − ix )
2
i ≡ −1
− ix
n.b. eix = cos( x) + i sin( x) e = cos( x) − i sin( x)
n.b. e x = cosh( x) + sinh( x) e − x = cosh( x) − sinh( x)
The BC’s and symmetries will determine which of the coefficients Amn, Bmn, Cmn, Dmn = 0.
We solve for the non-zero coefficients Apq, Bpq, Cpq and Dpq by taking inner products.
i.e. we multiply V ( x, y, z ) = ∑ ( stuff ) by e.g. sin(α p x) sin( β q y ) to project out the p-qth
component (i.e. we use the orthogonality properties of the individual terms in sin( ) and cos( )
Fourier Series.) and then integrate over the relevant intervals in x and y:
e.g.
xo yo
∫ ∫
0 0
V ( x, y ) sin(α p x) sin( β q y )dxdy
⎧⎪⎛ ∞ ⎞
⎨⎜ ∑ Amn sin(α n x) sin( β m y ) sinh(γ mn z ) *sin(α p x) sin( β q y ) ⎟
xo yo
=∫ ∫
⎪⎩⎜⎝ m, n =0 ⎟
0 0
= constant here ⎠
⎛ ∞ ⎞
+ ⎜ ∑ Bmn cos(α n x) cos( β m y ) sinh(γ mn z ) *sin(α p x) sin( β q y ) ⎟
⎜ m, n =0 ⎟
⎝ = constant here ⎠
⎛ ∞ ⎞
+ ⎜ ∑ Cmn sin(α n x) sin( β m y ) cosh(γ mn z ) *sin(α p x) sin( β q y ) ⎟
⎜ m, n =0 ⎟
⎝ = constant here ⎠
⎛ ∞ ⎞ ⎫⎪
+ ⎜ ∑ Dmn cos(α n x) cos( β m y ) cosh(γ mn z ) *sin(α p x) sin( β q y ) ⎟ ⎬ dxdy
⎜ m,n =0 ⎟⎪
⎝ = constant here ⎠⎭
cos (α n x ) sin (α p x ) dx = 0
xo
∫ 0
So all terms in above Σ’s vanish, except for a single term (in each sum) – that for the Apq /Bpq
/Cpq /Dpq coefficient!!! The BC’s will e.g. kill off 3 out of remaining 4 non-zero terms, thus only
one term survives…
Suppose only the Apq coefficient survives. Its analytic form is now known for all integers p and q.
Then the analytic form of 3-D potential V ( x, y, z ) is now known – it is an infinite series solution
of the form:
∞
V ( x, y , z ) = ∑A
m,n=0
mn sin (α n x ) sin ( β m y ) sinh ( γ mn z )
= α n2 + β m2
Laplace’s Equation ∇ 2V ( ρ , ϕ , z ) = 0
And Potential Problems with Cylindrical Symmetry (Cylindrical Coordinates)
ẑ r r = ρ + z = ρρˆ + zzˆ r = ρ 2 + z2
Ο ŷ ∇ 2V ( ρ , ϕ , z ) = 0
1 ∂ ⎛ ∂V
⎞ 1 ∂ 2V ∂ 2V
ϕ ρ = ρ
⎟+ 2 + 2 =0
ρ ∂ρ ⎜⎝ ∂ρ
⎠ ρ ∂ϕ
2
∂z
∂ V 1 ∂V 1 ∂ V ∂ 2V
2 2
ϕ̂ = 2+ + + + =0
∂ρ ρ ∂ρ ρ 2 ∂ϕ 2 ∂z 2
x̂ ρ̂
d 2Z ( z )
2
− k 2 Z ( z ) = 0 ⇒ Z ( z ) = e ± kz
dz
d Q (ϕ )
2
+ν 2Q (ϕ ) = 0 ⇒ Q (ϕ ) = e ± iνϕ
dϕ 2
d 2R ( ρ ) 1 dR ( ρ ) ⎛ 2 ν 2 ⎞
+ + ⎜k − 2 ⎟ R(ρ) = 0
dρ2 ρ dρ ⎝ ρ ⎠
Note(s):
1.) k is arbitrary without imposing boundary conditions.
2.) k appears in both Z(z) and R(ρ) equations.
3.) In order for Q(φ) to be single-valued (i.e. Q (ϕ ) = Q (ϕ + 2π ) ), ν must be an integer!
d 2R ( x) 1 dR ( x ) ⎛ ν 2 ⎞
Let x ≡ kp Then: + + ⎜1 − 2 ⎟ R ( x ) = 0 ⇐ Bessel’s Equation
dx 2 x dx ⎝ x ⎠
∞
R ( x ) = xα ∑ a j x j ⇐ Power Series Solution α = ±ν
j =0
1
a2 j ≡ − a2 j − 2 for j = 0, 1, 2, 3, . . . .
4 j ( j +α )
⎡ ( −1) j Γ (α + 1) ⎤ ( −1)
j
a2 j = ⎢ 2 j ⎥ a0 = 2( j +α )
⎢⎣ 2 j !Γ ( j + α + 1) ⎥⎦ 2 j !Γ ( j + α + 1)
1
where a0 = α Γ ( x ) = Gamma Function
2 Γ (α + 1)
There exist TWO solutions of the Radial Equation (i.e. Bessel’s Equation):
They are:
Bessel Functions of 1st kind, of order ±ν :
ν
( −1)
j 2j
∞
⎛x⎞ ⎛x⎞
J +ν ( x ) = ⎜ ⎟
⎝2⎠
∑ ⎜ ⎟
j = 0 j !Γ ( j + ν + 1) ⎝ 2 ⎠
These series converge for
−ν
( −1)
j 2j
⎛x⎞ ∞ ⎛x⎞
J −ν ( x ) = ⎜ ⎟ ∑ ⎜ ⎟ all values of x.
⎝ 2 ⎠ j =0 j !Γ ( j −ν + 1) ⎝ 2 ⎠
If ν is not an integer (which is not the case here), then the J ±ν ( x ) form a pair of
n.b. linearly independent solutions to the 2nd order Bessel’s Equation:
R ( x ) = Aν Jν ( x ) + A−ν J −ν ( x ) for ν ≠ integer
However, note that if ν = integer (which is the case for us here) then the Bessel functions
Jν ( x ) and J −ν ( x ) are NOT linearly independent!!
∴ ⇒ We must find another linearly independent solution for R(x) when ν = m = integer
Hankel Functions are complex linear combinations of Jν ( x ) and Nν ( x ) (Bessel Functions of 1st
and 2nd kind respectively). They are defined as follows:
Hν(1) ( x ) ≡ Jν ( x ) + iNν ( x ) The Hankel Functions Hν(1) ( x ) and Hν( 2) ( x ) also form a
Hν( 2) ( x ) ≡ Jν ( x ) − iNν ( x ) fundamental set/basis of solutions to the Bessel equation.
V ( ρ , ϕ , z ) = R ( ρ ) Q (ϕ ) Z ( z )
Note that sometimes we want V ( r ) only inside some finite region of space, e.g. coaxial
capacitor – if so, then don’t have to worry about r = ∞ solutions being finite – an example – the
Coaxial Capacitor:
End View of a Coaxial Capacitor
If the r = 0 region is an excluded region in the problem, then must include (i.e. allow) the Nν ( x )
solutions (singular at x = k ρ = 0 )!!!
If r = 0 region is included in problem, then ALL coefficients Cmn = Dmn ≡ 0 (for all m, n),
if V ( ρ , ϕ , z ) is finite @ r = 0 .
ρ
ϕ x
∇ 2V ( ρ , ϕ ) = 0 V ( ρ , ϕ ) = R ( ρ ) Q (ϕ )
Again try product solution
Potential, V ( ρ , ϕ ) is independent of z
(e.g. infinitely long coaxial cable)
1 ∂ ⎛ ∂V ⎞ 1 ∂ 2V
∇ V ( ρ ,ϕ ) =
2
ρ ⎟+ 2 =0
ρ ∂ρ ⎜⎝ ∂ρ ⎠ ρ ∂ϕ
2
Get:
ρ d ⎛ dR ( ρ ) ⎞ 1 d Q (ϕ )
2
⎜ρ ⎟ = C1 = − Let C1 = k2
R(ρ) dρ ⎝ dρ ⎠ Q (ϕ ) d ϕ 2
d ⎛ dR ( ρ ) ⎞ 2 d 2 Q (ϕ )
Then: ρ ⎜ρ ⎟ − k R ( ρ ) = 0 and + k 2 Q (ϕ ) = 0
dρ ⎝ dρ ⎠ dϕ 2
singular @ ρ → ∞ singular @ ρ = 0
d ⎛ dR ( ρ ) ⎞ 2
ρ ⎜ρ ⎟ − n R ( ρ ) = 0 ⇒ Rn ( ρ ) = Cn ρ + Dn ρ
−n
n
for n ≥ 1 (i.e. n = 1, 2, 3, . . .)
dρ ⎝ dρ ⎠
Ro ( ρ ) = Co + Do ln ( ρ ) for n = 0 only
Again, apply BC’s on all relevant surfaces, impose V ( r → ∞ ) = finite, etc. – these will dictate /
determine all coefficients, V0, V1, an, bn, cn and dn.
i.e. Solve for V0, V1, an, bn, cn and dn by applying all boundary conditions, V ( r → ∞ ) = finite,
and using orthogonality conditions / properties:
2π ρo
an = " A " ∫ dϕ ∫ d ρ ρ V ( ρ , ϕ )ρ n cos ( nϕ ) dA = ρ d ρ dϕ
0 0
2π ρo
bn = " B " ∫ dϕ ∫ d ρ ρ V ( ρ , ϕ )ρ − n cos ( nϕ )
0 0
2π ρo
cn = " C " ∫ dϕ ∫ d ρ ρ V ( ρ , ϕ )ρ n sin ( nϕ )
0 0
2π ρo
d n = " D " ∫ dϕ ∫ d ρ ρ V ( ρ , ϕ )ρ − n sin ( nϕ )
0 0
“A”, “B”, “C”, “D” are appropriate normalization factors (we will discuss later).
ẑ V = V ( r ,ϑ , ϕ )
ϕ̂
r̂
r
θ θ
Ο ŷ
ϕ ϕ̂
x̂
∇ 2V ( r , ϑ , ϕ ) = 0
1 ∂ ⎛ 2 ∂V ⎞ 1 ∂ ⎛ ∂V ⎞ 1 ∂ 2V
= ⎜ r ⎟ + ⎜ sin θ ⎟ + =0
r 2 ∂r ⎝ ∂r ⎠ r 2 sin θ ∂θ ⎝ ∂θ ⎠ r 2 sin 2 θ ∂ϕ 2
d 2U ( r ) U ( r ) Q (ϕ ) d ⎛ dP (θ ) ⎞ U ( r ) P (θ ) d 2Q (ϕ )
P (θ ) Q (ϕ ) + ⎜ sin θ ⎟+ 2 2 =0
dr 2 r 2 sin θ dθ ⎝ dθ ⎠ r sin θ dϕ 2
Multiply by r 2 sin 2 θ / U ( r ) P (θ ) Q (ϕ ) :
⎡ 1 d 2U ( r ) 1 1 d ⎛ dP (θ ) ⎞ ⎤ 1 d Q (ϕ )
2
r 2 sin 2 θ ⎢ + 2 ⎜ sin θ ⎟⎥ + =0
⎢⎣U ( r ) dr r sin θ P (θ ) dθ dθ ⎠ ⎥⎦ Q (ϕ ) dϕ 2
2
⎝
function of r +θ only function of ϕ only
1 d 2 Q (ϕ ) d 2 Q (ϕ )
Now: = −m 2
⇒ + m 2 Q (ϕ ) = 0
Q (ϕ ) dϕ 2
dϕ 2
⎢⎣U ( r ) dr r sin θ P (θ ) dθ ⎝ dθ ⎠ ⎥⎦
2
1 d U (r ) 1 d ⎛ dP (θ ) ⎞
2
1 m2
= − ⎜ sin θ ⎟ +
U ( r ) dr 2 r 2 sin θ P (θ ) dθ ⎝ dθ ⎠ r 2 sin 2 θ
d 2U ( r ) α 1 d ⎛ dP (θ ) ⎞ ⎡ m2 ⎤
∴ − 2 U (r ) = 0 and ⎜ sin θ +
⎟ ⎢α − ⎥ P (θ ) = 0
dr 2 r sin θ dθ ⎝ dθ ⎠ ⎣ sin 2 θ ⎦
d 2U ( r ) ( + 1) 1 d ⎛ dP (θ ) ⎞ ⎡ m2 ⎤
∴ − U ( r ) = 0 and ⎜ sin θ ⎟+⎢ ( + 1) − ⎥ P (θ ) = 0
dr 2 r2 sin θ dθ ⎝ dθ ⎠ ⎣ sin 2 θ ⎦
1 d ⎛ sin 2 θ dP (θ ) ⎞ ⎡ m2 ⎤
Then: ⎜ ⎟ + ( + 1) − P (θ ) = 0
sin θ dθ ⎝ sin θ dθ ⎠ ⎢⎣ sin 2 θ ⎥⎦
⎡ m2 ⎤
d ⎛ 2 dP ( x ) ⎞
Becomes: ⎜(1 − x ) dx ⎟ ( ) 1 − x 2 ⎥ P ( x ) = 0 ⇐
+ ⎢ + 1 − Generalized Legendre' Equation
dx ⎝ ⎠ ⎢⎣ ( ) ⎥⎦
d 2U ( r ) ( + 1) U r = 0 are of the form:
General Solutions of the radial equation, 2
− ( )
dr r2
U ( r ) = Ar + Br − ( +1) (l + A + B) are determined by boundary conditions…
⎛ ( 2 + 1) ⎞ π
A =⎜ ⎟ ∫0 V ( r = a, θ ) P ( cos θ ) sin θ dθ
⎝ 2a ⎠
normalization
factor
n.b. The P ( cos θ ) functions form a complete orthonormal basis set on the unit circle (r = 1)
for −1 ≤ cos θ ≤ 1 or: 0 ≤ θ ≤ π
1 d ⎛ dP (θ ) ⎞ ⎡ m2 ⎤
⎜ sin θ ⎟+⎢ ( + 1) − ⎥ P (θ ) = 0 x = cos θ
sin θ dθ ⎝ dθ ⎠ ⎣ sin 2 θ ⎦
dx m "ordinary "
Legendre '
Polynomial
m = ± integer ≠ 0
i.e. m = ±1, ±2, ±3,... but have a constraint on m !!! − ≤m≤+
Also: P − m ( x ) = ( −1)
m ( − m )! m
P ( x)
( + m )!
1 2 ( + m )!
∫ P m′ ( x ) P m ( x ) dx = δ′
−1
( 2 + 1) ( − m )! Kroenecker
δ -function
Y m (θ , ϕ ) ≡
(2 + 1) ( − m ) ! m
P ( cos θ ) eimϕ
4π ( + m ) ! = Qm (ϕ )
The Spherical Harmonics Y m (θ , ϕ ) form a complete orthonormal set of basis “vectors” on the
surface of the unit sphere (r = 1)
i.e. i → −i where i ≡ −1
2π π
∫ dϕ ∫ sin θ dθ Y *′m′ (θ , ϕ ) Y m (θ , ϕ ) = δ ′ δ m′m
0 0
Ω= 4π
i.e. ∫ d ΩY *′m′ (θ , ϕ ) Y m (θ , ϕ ) = δ ′ δ m′m d Ω = sin θ dθ dϕ
Ω= 0
∞
Completeness’ Relation: ∑ ∑ Y (θ ,ϕ ) Y (θ ,ϕ ) = δ ( cos θ − cos θ ') δ (ϕ − ϕ ')
=0 m =−
*
m m
DIRAC δ -functions
Y m (θ , ϕ ) Spherical Harmonics
1
Use Y − m (θ , ϕ ) = ( −1) Y *m (θ , ϕ )
m
=0 Y00 =
4π
in order to obtain Y2− 2 , Y2−1 , Y3−3 , Y3− 2 , Y3−1 etc.
3
=1 Y11 = − sin θ eiϕ
8π
3
Y10 = − cos θ Note:
4π
Y m (θ , ϕ ) ≡
(2 + 1) ( − m ) ! m
P ( cos θ ) eimϕ
4π ( + m ) !
Y22 =
1 15
sin 2 θ e 2iϕ Y 0 (θ , ϕ ) =
(2 + 1)
P ( cos θ )
4 2π 4π
5
=2 Y21 = − sin θ cos θ eiϕ
8π
5 ⎛3 1⎞
Y20 = ⎜ cos θ − ⎟
2
4π ⎝ 2 2⎠
1 35
Y33 = − sin 3 θ e3iϕ
4 4π
1 105 2
=3 Y32 = sin θ cos θ e 2iϕ
4 2π
1 21
Y31 = − sin θ ( 5cos 2 θ − 1) eiϕ
4 4π
7 ⎛5 3 ⎞
Y30 = ⎜ cos θ − cos θ ⎟
3
4π ⎝ 2 2 ⎠
…. …. etc.
∞ +
V ( r ,θ , ϕ ) = ∑ ∑ ⎡⎣ A m r + B m r −( +1)
⎤Y m (θ , ϕ )
⎦
=0 m =−
If V = V (θ , ϕ ) on surface (e.g. at r = a)
(i.e. no charge at r = 0 in problem → B m = 0 ∀ ,m )
∞ +
Then: V (θ , ϕ ) = ∑ ∑A m Y m (θ , ϕ ) on surface (r = a).
=0 m =−
4π
And: A m = ∫ d ΩY *m (θ , ϕ ) V (θ , ϕ ) on surface (r = a).
0
Note: V (θ = 0, ϕ ) = ∑
∞
(2 + 1)
A0
" north " =0 4π
pole
(2 + 1) 4π
d ΩP ( cos θ ) V (θ , ϕ )
A0 =
4π ∫ 0
∇ ψ (r,t ) + 2
2
= 0 and in Schrödinger’s wave equation Hψ = Eψ in Quantum
c ∂t 2
Mechanics problems. These equations will appear again and again, in one form or another for
E&M, Classical Mechanics, Quantum Mechanics courses as well as for Classical / Newtonian
Gravity problems…
For more detailed information e.g. on separation of variables and solutions to 3-D Wave
1 ∂ 2ψ
Equation ∇ 2ψ = − 2 2 in rectangular, cylindrical and spherical coordinates see Prof. S.
c ∂t
Errede lecture notes (Lecture IV – parts 1 & 2) on (sound) waves in 1-D, 2-D, 3D Physics 406
Acoustical Physics of Music website:
http://online.physics.uiuc.edu/courses/phys406/406_lectures.html
and also see/read his Fourier Analysis Lectures on this website, if interested.