1. Introduction
    1. Overview
    2. Finding V V
      1. Type 1 Problems
      2. Type 2 Problems
      3. Deriving Poisson’s Equation
      4. Problem
  2. 1-D
    1. Cartesian
      1. Charge Method
      2. Laplace’s Equation Method
      3. Problems
        1. Computing Οƒ \sigma
        2. Checking Solution
    2. Cylindrical
      1. Computing Οƒ \sigma
      2. Computing E \mathbf{E} using Charge Method
    3. Spherical
      1. Example
  3. 2–D Cartesian
    1. General Solution
    2. Problem – Relationship Between Forms 1.–4.
    3. Solution Step 1.
      1. Example 3.3 of Griffiths
      2. Example 3.4 of Griffiths
      3. Problem – Long Square Tube
      4. Problem – Long Rectangular Tube
    4. Step 2 preparation
      1. Problem – Integrals
      2. Problem – Superposition
    5. Solution Step 2.
      1. Example 3.3
      2. Problem
      3. Plotting Results
  4. 2-D Cartesian Problems
    1. Long Rectangular Duct
    2. Long Rectangular Duct

See also 3.1 of Griffiths.

Many laws of physics are expressed as either ordinary differential equations or partial differential equations.

Recall from mechanics that the equation for the position of a particle in a uniform gravitational field is given by the ordinary differential equation

m d 2 x d t 2 = βˆ’ m g m\frac{d^2x}{dt^2}=-mg

This equation does not explicitly give the position of a mass versus time. To determine, one must first obtain a solution to this equation.

A general solution for x ( t ) x(t) has the form

x ( t ) = C 1 + C 2 t βˆ’ g t 2 / 2 x(t)=C_1+C_2t-gt^2/2

where C 1 C_1 and C 2 C_2 are two constants.

To use this equation to find x ( t ) x(t) for a mass, the values of the two constants must be known. Usually these two constants are determined based on the initial values of x x and d x / d t dx/dt . For example, if x ( t = 0 ) = 0 x(t=0)=0 and v ( t = 0 ) = 0 v(t=0)=0 , then C 1 = C 2 = 0 C_1=C_2=0 and the trajectory of the mass is given by x ( t ) = βˆ’ g t 2 / 2 x(t)=-gt^2/2 .

This type of problem is called an initial value problem – given a differential equation for the position of the mass and it’s initial values of x x and v v , we arrive at a solution for x x for all t t . Another example of an inital value problem is for the position x x of a mass on a spring. The differential equation for x x is d 2 x / d t 2 βˆ’ Ο‰ o 2 x = 0 d^2x/dt^2-\omega_o^2x=0 . A general solution is x ( t ) = C 1 sin ⁑ ( Ο‰ o t ) + C 2 cos ⁑ ( Ο‰ o t ) x(t)=C_1\sin(\omega_o t)+C_2\cos(\omega_o t) .

A boundary value problem involves first finding a general solution to a partial differential equation with respect to the position (instead of time). The solution will have constants whose values are determined by the conditions at the boundary. For example, in electrostatics

βˆ‚ 2 V βˆ‚ x 2 = 0 \frac{\partial^2 V}{\partial x^2}=0

is a partial differential equation for V V for a system where the symmetry is such that the potential only depends on x x . If we want to know V ( x ) V(x) , we need to know a general solution to this equation. It is

V ( x ) = C 1 + C 2 x V(x)=C_1 + C_2x

The values for C 1 C_1 and C 2 C_2 are determined based on boundary conditions. For example, V ( x = 0 ) = 0 V(x=0)=0 and V ( x = L ) = V o V(x=L)=V_o .

In this course, you will only be required to know how to derive the general solution for partial differential equations in 1–D for cartesian, cylindrical, and spherical coordinates. Derivations are given in the 1–D section.

For 2– and 2–D partial differential equations, you will be given the general solution and asked to find the unknown constants given boundary value information.

There are generally two types of problems involving V V in electrostatics:

  1. given the locations of charges, compute the electric potential and electric field at all locations in space; and

  2. given the electric potential at certain locations in space, compute the electric potential and electric field at all locations in space.

For problems of type 1., we can compute V V using

V ( r ) = βˆ‘ i = 1 N k q i ∣ r βˆ’ r i β€² ∣ V(\mathbf{r})=\sum_{i=1}^{N} \frac{kq_{i}}{|\mathbf{r}-\mathbf{r}_{i}'|}

for N N discrete charges, or the continuous approximation

V ( r ) = ∫ k d q ∣ r βˆ’ r β€² ∣ V(\mathbf{r})=\int \frac{kdq}{|\mathbf{r}-\mathbf{r}'|}

along with a given line, surface, or volume charge density ( Ξ» , Οƒ \lambda, \sigma , or ρ \rho , respectively).

In many problems, we don’t know the locations of the charges in the system. Consider the scenario shown in the following figure in which a net charge of Q Q is placed on a conductor.

The charges on the conductor will distribute themselves so that the electric field in the conductor is zero. In general, we don’t know what the distribution will be – we only know that whatever it is, it is such that the electric field is zero inside.

If the top cap is centered on the origin and in the x x – y y plane, to compute the potential due to the top cap outside of the conductor, we may be tempted to start with

V ( r ) = ∫ 0 2 Ο€ ∫ 0 R k Οƒ ( s β€² ) ∣ r βˆ’ s β€² s ^ ∣ s β€² d s β€² d Ο• β€² V(\mathbf{r})= \int_0^{2\pi}\int_0^R \frac{k\sigma(s')}{|\mathbf{r}-s'\hat{\mathbf{s}}|}s'ds'd\phi'

which was the approach used to find the potential due to charges distributed on a disk with a given Οƒ ( s ) \sigma(s) . However, in this case, we don’t know Οƒ ( s ) \sigma(s) and so we are stuck. A brute-force method of computing Οƒ \sigma uses the fact that the V V must be constant on the conductor: guess Οƒ \sigma the top, bottom, and side surfaces, compute V V using integration, and then repeat until a V V is found that is constant on the conductor.

In summary, we can only use the integral formula for V V if the charge densities are known. In general, if we place a net charge on a conductor, we do not know how it will distribute on its surface.

Problems of type 2. are called boundary value problems and are generally solved by finding solutions to a partial differential equation called Poisson’s equation:

βˆ‡ 2 V = βˆ’ ρ Ο΅ 0 \nabla^2 V= -\frac{\rho}{\epsilon_0}

βˆ‡ 2 \nabla^2 is a new form of operator. It involves operation on a scalar function f f and results in a scalar function. In cartesian coordinates,

βˆ‡ 2 f = βˆ‚ 2 f βˆ‚ x 2 + βˆ‚ 2 f βˆ‚ y 2 + βˆ‚ 2 f βˆ‚ z 2 \displaystyle \nabla^2f=\frac{\partial^2 f}{\partial x^2}+\frac{\partial^2 f}{\partial y^2}+\frac{\partial^2 f}{\partial z^2}

Therefore, Poisson’s equation in cartesian coordinates is

βˆ‚ 2 V βˆ‚ x 2 + βˆ‚ 2 V βˆ‚ y 2 + βˆ‚ 2 V βˆ‚ z 2 = βˆ’ ρ Ο΅ 0 \frac{\partial^2 V}{\partial x^2}+\frac{\partial^2 V}{\partial y^2}+\frac{\partial^2 V}{\partial z^2}=-\frac{\rho}{\epsilon_0}

Similar to the divergence and gradient operators, the βˆ‡ 2 \nabla^2 operator (also called the Laplacian or β€œdel squared”) has different forms for each coordinate system. These are given on the second–to last page of Griffiths.

Laplace’s equation is Poisson’s equation with ρ = 0 \rho=0 :

βˆ‡ 2 V = 0 \nabla^2 V = 0

In cartesian coordinates, it is

βˆ‚ 2 V βˆ‚ x 2 + βˆ‚ 2 V βˆ‚ y 2 + βˆ‚ 2 V βˆ‚ z 2 = 0 \frac{\partial^2 V}{\partial x^2}+\frac{\partial^2 V}{\partial y^2}+\frac{\partial^2 V}{\partial z^2}=0

Poisson’s Equation, βˆ‡ 2 V = βˆ’ ρ / Ο΅ 0 \nabla^2 V= -{\rho}/{\epsilon_0} , follows directly from inserting the definition of the electric potential E = βˆ’ βˆ‡ V \mathbf{E}=-\nabla V into Gauss’ Law in differential form βˆ‡ β‹… E = ρ / Ο΅ 0 \nabla\boldsymbol{\cdot}\mathbf{E}=\rho/\epsilon_0 .

Gauss’s law in differential form is

βˆ‡ β‹… E = ρ Ο΅ 0 \nabla\boldsymbol{\cdot}\mathbf{E}=\frac{\rho}{\epsilon_0}

Using E = βˆ’ βˆ‡ V \mathbf{E}=-\boldsymbol{\nabla}V , this is

βˆ‡ β‹… ( βˆ’ βˆ‡ V ) = ρ Ο΅ 0 \boldsymbol{\nabla}\boldsymbol{\cdot}(-\boldsymbol{\nabla}V)=\frac{\rho}{\epsilon_0}

To evaluate βˆ‡ β‹… ( βˆ’ βˆ‡ V ) \boldsymbol{\nabla}\boldsymbol{\cdot}(-\boldsymbol{\nabla}V) , first write βˆ‡ V \boldsymbol{\nabla}V in cartesian

βˆ‡ V = x ^ βˆ‚ V βˆ‚ x + y ^ βˆ‚ V βˆ‚ y + z ^ βˆ‚ V βˆ‚ z \boldsymbol{\nabla}V=\mathbf{\hat{x}}\frac{\partial V}{\partial x}+\mathbf{\hat{y}}\frac{\partial V}{\partial y}+\mathbf{\hat{z}}\frac{\partial V}{\partial z}

Then

βˆ‡ β‹… ( βˆ’ βˆ‡ V ) = βˆ‡ β‹… ( βˆ’ x ^ βˆ‚ V βˆ‚ x βˆ’ y ^ βˆ‚ V βˆ‚ y βˆ’ z ^ βˆ‚ V βˆ‚ z ) \boldsymbol{\nabla}\boldsymbol{\cdot}(-\boldsymbol{\nabla}V)=\boldsymbol{\nabla}\boldsymbol{\cdot}\left(-\mathbf{\hat{x}}\frac{\partial V}{\partial x}-\mathbf{\hat{y}}\frac{\partial V}{\partial y}-\mathbf{\hat{z}}\frac{\partial V}{\partial z}\right)

Using the definition of divergence of a vector function U \mathbf{U} , which is

βˆ‡ β‹… U = x ^ βˆ‚ U x βˆ‚ x + y ^ βˆ‚ U y βˆ‚ y + z ^ βˆ‚ U z βˆ‚ z \boldsymbol{\nabla}\boldsymbol{\cdot}\mathbf{U}=\mathbf{\hat{x}}\frac{\partial U_x}{\partial x}+\mathbf{\hat{y}}\frac{\partial U_y}{\partial y}+\mathbf{\hat{z}}\frac{\partial U_z}{\partial z}

with U x = βˆ’ βˆ‚ V / βˆ‚ x U_x=-\partial V/\partial x , U y = βˆ’ βˆ‚ V / βˆ‚ y U_y=-\partial V/\partial y , U z = βˆ’ βˆ‚ V / βˆ‚ z U_z=-\partial V/\partial z gives

βˆ‡ β‹… ( βˆ’ x ^ βˆ‚ V βˆ‚ x βˆ’ y ^ βˆ‚ V βˆ‚ y βˆ’ z ^ βˆ‚ V βˆ‚ z ) = βˆ’ βˆ‚ 2 V βˆ‚ x 2 βˆ’ βˆ‚ 2 V βˆ‚ y 2 βˆ’ βˆ‚ 2 V βˆ‚ z 2 \boldsymbol{\nabla}\boldsymbol{\cdot}\left(-\mathbf{\hat{x}}\frac{\partial V}{\partial x}-\mathbf{\hat{y}}\frac{\partial V}{\partial y}-\mathbf{\hat{z}}\frac{\partial V}{\partial z}\right)=-\frac{\partial^2 V}{\partial x^2}-\frac{\partial^2 V}{\partial y^2}-\frac{\partial^2 V}{\partial z^2}

and so

βˆ‡ β‹… ( βˆ’ βˆ‡ V ) = βˆ’ βˆ‚ 2 V βˆ‚ x 2 βˆ’ βˆ‚ 2 V βˆ‚ y 2 βˆ’ βˆ‚ 2 V βˆ‚ z 2 \boldsymbol{\nabla}\boldsymbol{\cdot}(-\boldsymbol{\nabla}V)=-\frac{\partial^2 V}{\partial x^2}-\frac{\partial^2 V}{\partial y^2}-\frac{\partial^2 V}{\partial z^2}

and

βˆ’ βˆ‚ 2 V βˆ‚ x 2 βˆ’ βˆ‚ 2 V βˆ‚ y 2 βˆ’ βˆ‚ 2 V βˆ‚ z 2 = ρ Ο΅ 0 -\frac{\partial^2 V}{\partial x^2}-\frac{\partial^2 V}{\partial y^2}-\frac{\partial^2 V}{\partial z^2}=\frac{\rho}{\epsilon_0}

Moving the negative sign and using the definition of the Laplacian βˆ‡ 2 \nabla^2 , gives Poisson’s equation

βˆ‡ 2 V = βˆ’ ρ Ο΅ 0 \nabla^2V=-\frac{\rho}{\epsilon_0}

Outside of a solid sphere of radius R R with uniformly distributed charge Q Q , the field is

V ( r ) = k Q 1 r \displaystyle V(r)=kQ\frac{1}{r}

inside, it is

V ( r ) = k Q 2 R [ 1 βˆ’ r 2 R 2 ] \displaystyle V(r)=\frac{kQ}{2R}\left[1-\frac{r^2}{R^2}\right]

Compute βˆ‡ 2 V \nabla^2 V and verify that βˆ‡ 2 V = βˆ’ ρ Ο΅ 0 \nabla^2V=-\frac{\rho}{\epsilon_0} inside and outside of the sphere. You may use the formula for βˆ‡ 2 \nabla^2 in any coordinate system.

See also 3.2 of Griffiths.

Suppose that we connect a battery with a potential difference V o V_o to two large conducting plates. We want to know how V V varies between the plates.

This problem can be solved using two approaches.

  1. By assuming the potential causes equal and opposite amount of charge to be uniformly distributed on the plates. This approach can only be used because we know (or assume) how the charges will distribute on the surfaces. If the plates were not large, we could not use this method.

  2. By using the boundary value method.

Assume charges of Β± Q \pm Q appear on the plates when the battery is connected so that the surface charge densities are Β± Q / A \pm Q/A and use the method demonstrated in the capacitance notes notes to find V ( x ) V(x) .

The electric field between the plates is E = βˆ’ Οƒ / Ο΅ o x ^ \mathbf{E}=-\sigma/\epsilon_o\mathbf{\hat{x}} . Using V ( x ) = V ( a ) βˆ’ ∫ a x E β‹… d l V(x)=V(a)-\int_a^x\mathbf{E}\boldsymbol{\cdot} d\mathbf{l} with a = 0 a=0 and d l = d x β€² x ^ d\mathbf{l}=dx'\mathbf{\hat{x}}

V ( x ) = V ( 0 ) βˆ’ ∫ 0 x βˆ’ Οƒ Ο΅ o d x β€² \displaystyle V(x)=V(0)-\int_0^x -\frac{\sigma}{\epsilon_o}dx'

Integration gives

V ( x ) = V ( 0 ) + Οƒ Ο΅ o x \displaystyle V(x) = V(0)+\frac{\sigma}{\epsilon_o}x

We are not done because we need to write Οƒ \sigma in terms of V o V_o . To compute Οƒ \sigma , plug in x = d x=d to get

V ( d ) βˆ’ V ( 0 ) = Οƒ Ο΅ o d \displaystyle V(d)-V(0)=\frac{\sigma}{\epsilon_o}d

The difference in potential V ( d ) βˆ’ V ( 0 ) V(d)-V(0) was given as V o V_o . Substitution gives

V o = Οƒ Ο΅ o d β‡’ Οƒ = Ο΅ o V o d \displaystyle V_o=\frac{\sigma}{\epsilon_o}d\quad\Rightarrow\quad \sigma=\epsilon_o\frac{V_o}{d}

Plugging this Οƒ \sigma into V ( x ) = V ( 0 ) + Οƒ Ο΅ o x \displaystyle V(x) = V(0)+\frac{\sigma}{\epsilon_o}x gives

V ( x ) = V ( 0 ) + V o x d \displaystyle V(x)=V(0)+V_o\frac{x}{d}

This problem may also be solved using Laplace’s equation.

Between the plates, there are no charges, so ρ = 0 \rho=0 and Laplace’s equation applies. If the plates are large, then the potential will only vary in the x x –direction, in which case Laplace’s equation in cartesian components simplifies to:

βˆ‡ 2 V = βˆ‚ 2 V βˆ‚ x 2 = 0 \displaystyle\nabla^2V = \frac{\partial^2 V}{\partial x^2} = 0

This can be written as

βˆ‡ 2 V = d 2 V d x 2 = 0 \displaystyle\nabla^2V = \frac{d^2 V}{dx^2} = 0

because V V only depends on x x . This equation can be integrated twice to give

V ( x ) = a x + b \displaystyle V(x)=ax+b

where a a and b b are constants. You can verify that V = a x + b V = ax+b satisfies βˆ‚ 2 V / βˆ‚ x 2 = 0 \partial^2 V/\partial x^2 = 0 by differentiating it twice with respect to x x .

The interpretation of this equation is that in a configuration where it can be argued that potential only depends on x x , the potential must increase or decrease linearly or be constant. In this example, the two boundary conditions are

  1. V ( x = 0 ) = 0 V(x=0)=0

  2. V ( x = d ) = V o V(x=d)=V_o

These two conditions give two equations that can be solved to find the unknown constants a a and b b :

V ( x = 0 ) = 0 = a β‹… 0 + b β‡’ b = 0 \displaystyle V(x=0)=0=a\cdot 0+b \Rightarrow b = 0

V ( x = d ) = V o = a d + b = a d + 0 β‡’ a = V o / d \displaystyle V(x=d)=V_o=ad+b=ad+0 \Rightarrow a=V_o/d

so the solution is

V ( x ) = V o d x \displaystyle V(x) = \frac{V_o}{d}x

This is the same result found using the charge method, V ( x ) = V ( 0 ) + V o x d V(x)=V(0)+V_o\frac{x}{d} , if we set V ( 0 ) = 0 V(0)=0 to be consistent with what was used in Laplace’s equation method.

After developing a solution to Laplace’s equation, one should always verify that the solution matches the boundary conditions, in this case by plugging in the coordinates of the boundary:

  1. V ( x ) = V o d x β‡’ V ( 0 ) = 0 \displaystyle V(x) = \frac{V_o}{d}x \Rightarrow V(0)=0 so the solution matches the x = 0 x=0 boundary condition

  1. V ( x ) = V o d x β‡’ V ( d ) = V o \displaystyle V(x) = \frac{V_o}{d}x \Rightarrow V(d)=V_o so the solution matches the x = d x=d boundary condition

1-D examples are simple. As will be seen when 2- and 3-D problems are considered, typically, one cannot find a single equation that satisfies all boundary conditions by following this method; a function can be found that satisfies some but not all of the boundary conditions. To fully solve the problem, one has to rely on the so-called β€œFourier Trick” to find a solution. This method is described in 2-D Cartesian.

In section 2.5.3 of Griffiths, the equation for the electric field immediately outside of a conductor is stated to be E = ( Οƒ / Ο΅ o ) n ^ \mathbf{E}=(\sigma/\epsilon_o)\hat{\mathbf{n}} .

Find the surface charge densities on each plate by computing E \mathbf{E} using E = βˆ’ βˆ‡ V \mathbf{E}=-\boldsymbol{\nabla} V .

A student came up with the following equation for the potential between the plates: V ( x ) = V o ( tan ⁑ βˆ’ 1 ( x / d + 1 / 2 ) + ( x / d ) 3 ) V(x) = V_o(\tan^{-1}(x/d+1/2) + (x/d)^3) . Without doing a calculation, we know this is wrong or the expression simplifies to V ( x ) = V o x / d V(x)=V_ox/d . Why?

Two long and concentric conducting cylinders of radius a a and b b are configured as shown in the following figure.

Assume the potential on the inner cylinder is V o V_o and that on the outer cylinder is 0 0 .

The cylinders are conductors, and if they are long relative to their radius, the charges that appear on them will be approximately uniformly distributed on their surfaces. If the charge distribution is independent of Ο• \phi and z z , the potential will not depend on Ο• \phi and z z . So V = V ( s ) V=V(s) .

In cylindrical coordinates, the Laplacian is

βˆ‡ 2 V = 1 s βˆ‚ βˆ‚ s ( s βˆ‚ V βˆ‚ s ) + 1 s 2 βˆ‚ 2 V βˆ‚ Ο• 2 + βˆ‚ 2 V βˆ‚ z 2 \nabla^2V={1 \over s}{\partial \over \partial s}\left(s {\partial V \over \partial s}\right) + {1 \over s^2}{\partial^2 V \over \partial \phi^2} + {\partial^2 V \over \partial z^2}

if V = V ( s ) V=V(s) , the derivatives in the second and third terms are zero, and so

βˆ‡ 2 V = 1 s βˆ‚ βˆ‚ s ( s βˆ‚ V βˆ‚ s ) \nabla^2 V={1 \over s}{\partial \over \partial s}\left(s {\partial V \over \partial s}\right)

Because V V depends only on s s , we can replace the partial derivative with the total derivative to obtain the ODE

βˆ‡ 2 V = 1 s d d s ( s d V d s ) = 0 \nabla^2 V={1 \over s}{d \over d s}\left(s {d V \over d s}\right)=0

The boundary conditions are

  1. V ( a ) = V o V(a)=V_o

  2. V ( b ) = 0 V(b)=0

To solve the ODE, note that because 1 / s 1/s is not zero, the only way for the left-hand side of

1 s d d s ( s d V d s ) = 0 {1 \over s}{d \over d s}\left(s {d V \over d s}\right)=0

to be zero is if

s d V d s = C 1 s {d V \over d s}=C_1

where C 1 C_1 is a constant. Direct integration of this equation gives

V ( s ) = C 1 ln ⁑ s + C 2 V(s) = {C_1 \ln s} + C_2

The two unknowns are solved for by using the boundary conditions

V ( b ) = 0 = C 1 ln ⁑ b + C 2 β‡’ C 2 = βˆ’ C 1 ln ⁑ b V(b) = 0 = {C_1 \ln b} + C_2 \Rightarrow C_2=-C_1\ln b

V ( a ) = V o = C 1 ln ⁑ a + C 2 β‡’ V o = C 1 ( ln ⁑ a βˆ’ ln ⁑ b ) = C 1 ln ⁑ ( a / b ) V(a) = V_o = {C_1 \ln a} + C_2\Rightarrow V_o=C_1(\ln a-\ln b)=C_1\ln(a/b)

Solving for C 1 C_1 and C 2 C_2 gives

C 1 = V o ln ⁑ ( a / b ) C_1 = {V_o \over \ln(a/b)} C 2 = βˆ’ V o / ln ⁑ b ln ⁑ ( a / b ) C_2 = -{{V_o / \ln b} \over \ln(a/b)}

Subsitution of these constants into V(s)=C1ln⁑s+C2V(s) = {C_1 \ln s} + C_2 gives

V(s)=Voln⁑(a/b)(ln⁑sβˆ’ln⁑b)V(s) = {V_o \over \ln(a/b)}\left(\ln s - \ln b \right)

as a check of the algebraic steps, plug in s=as=a and s=bs=b into this equation and verify that the boundary conditions used, V(a)=VoV(a)=V_o and V(b)=0V(b)=0 are satisfied.

The electric field can be found using E=βˆ’βˆ‡V\mathbf{E}=-\nabla V. In cylindrical coordinates, when VV depends only on ss, the negative of the gradient of VV is given by

βˆ’βˆ‡V=βˆ’βˆ‚Vβˆ‚ss^-\nabla V=-{\partial V \over \partial s} \hat{\mathbf{s}}

Evaluation of the derivative gives

E = βˆ’ V o ln ⁑ ( a / b ) 1 s s ^ \mathbf{E}=-{V_o \over \ln(a/b)}{1 \over s}\hat{\mathbf{s}}

Using βˆ’ ln ⁑ ( a / b ) = ln ⁑ ( b / a ) -\ln(a/b)=\ln(b/a) , this can also be written as

E = V o ln ⁑ ( b / a ) 1 s s ^ \mathbf{E}={V_o \over \ln(b/a)}{1 \over s}\hat{\mathbf{s}}

This field points from a a to b b (in + s ^ +\hat{\mathbf{s}} direction) as expected because the potential on the inner surface is higher than that on the outer surface.

Find the surface charge densities on the cylinders.

For the example given, find E ( s ) \mathbf{E}(s) between the cylinders using the charge method.

In spherical coordinates, the Laplacian is

βˆ‡ 2 V = 1 r 2 βˆ‚ βˆ‚ r ( r 2 βˆ‚ V βˆ‚ r ) + 1 r 2 sin ⁑ ΞΈ βˆ‚ βˆ‚ ΞΈ ( sin ⁑ ΞΈ βˆ‚ V βˆ‚ ΞΈ ) + 1 r 2 sin ⁑ 2 ΞΈ βˆ‚ 2 V βˆ‚ Ο• 2 \nabla^2V = { {1 \over r^{2}}{\partial \over \partial r}\left(r^{2}{\partial V \over \partial r}\right)+{1 \over r^{2}\sin \theta }{\partial \over \partial \theta }\left(\sin \theta {\partial V \over \partial \theta }\right)+{1 \over r^{2}\sin ^{2}\theta }{\partial ^{2}V \over \partial \phi ^{2}}}

If V = V ( r ) V=V(r) , then this reduces to

βˆ‡ 2 V = 1 r 2 βˆ‚ βˆ‚ r ( r 2 βˆ‚ V βˆ‚ r ) \nabla^2V = { {1 \over r^{2}}{\partial \over \partial r}\left(r^{2}{\partial V \over \partial r}\right)}

1. Find V ( r ) V(r) that satisfies 1 r 2 βˆ‚ βˆ‚ r ( r 2 βˆ‚ V βˆ‚ r ) = 0 \displaystyle { {1 \over r^{2}}{\partial \over \partial r}\left(r^{2}{\partial V \over \partial r}\right)}=0 in terms of r r and two unknown constants.

2. Two concentric spherical conducting shells are connected to a battery such that the inner shell is at a potential of 0 0 and the outer shell is at a potential of V o V_o . The inner shell has an outer radius of a a . The outer shell has an inner radius of b b . Use your equation from part 1. and these boundary conditions to find V ( r ) V(r) between the conductors in terms of V o V_o , a a , and b b .

3. In section 2.5.3 of Griffiths, the equation for the electric field immediately outside of a conductor is stated to be E = ( Οƒ / Ο΅ o ) n ^ \mathbf{E}=(\sigma/\epsilon_o)\hat{\mathbf{n}} . Use the potential V ( r ) V(r) found in part 2. to find the electric field E ( r ) \mathbf{E}(r) between a a and b b and then evaluate this electric field at a a and b b to find the surface charge densities at a a and b b .

4. Use equation for capacitance for this configuration (see capacitance) to eliminate V o V_o from the surface charge densities. You should find that the surface charge densities on the inner and outer conductors are βˆ’ Q / 4 Ο€ a 2 -Q/4\pi a^2 and Q / 4 Ο€ b 2 Q/4\pi b^2 , respectively.

Answer:

1. In spherical coordinates, the Laplacian is

βˆ‡ 2 V = 1 r 2 βˆ‚ βˆ‚ r ( r 2 βˆ‚ V βˆ‚ r ) + 1 r 2 sin ⁑ ΞΈ βˆ‚ βˆ‚ ΞΈ ( sin ⁑ ΞΈ βˆ‚ V βˆ‚ ΞΈ ) + 1 r 2 sin ⁑ 2 ΞΈ βˆ‚ 2 V βˆ‚ Ο† 2 \displaystyle \nabla^2V={ {1 \over r^{2}}{\partial \over \partial r} \left(r^{2}{\partial V \over \partial r}\right) + {1 \over r^{2} \sin \theta }{\partial \over \partial \theta } \left(\sin \theta {\partial V \over \partial \theta }\right) + {1 \over r^{2} \sin ^{2}\theta }{\partial ^{2}V \over \partial \varphi ^{2}}}

Because the system is invariant with rotation by Ο• \phi and ΞΈ \theta , the potential must be independent of ΞΈ \theta and Ο• \phi ; as a result, the second two terms are zero. Therefore, we need to solve

βˆ‡ 2 V = 1 r 2 βˆ‚ βˆ‚ r ( r 2 βˆ‚ V βˆ‚ r ) = 0 \displaystyle \nabla^2V={1 \over r^{2}}{\partial \over \partial r} \left(r^{2}{\partial V \over \partial r}\right)=0

Becuase V V depends only on r r , we can replace the partial derivative with the total derivative

βˆ‡ 2 V = 1 r 2 d d r ( r 2 d V d r ) = 0 \displaystyle \nabla^2V={1 \over r^{2}}{d \over d r} \left(r^{2}{d V \over d r}\right)=0

To solve the ODE, note that the following must hold

r 2 d V d r = c 1 \displaystyle r^2{dV\over dr} = c_1

where C 1 C_1 is a constant. Direct integration gives

V ( r ) = c 1 r + c 2 \displaystyle V(r) = {c_1 \over r} + c_2

2. The two unknowns are solved for by using the boundary conditions V ( a ) = V o V(a)=V_o and V ( b ) = 0 V(b)=0 :

V ( a ) = 0 = c 1 a + c 2 \displaystyle V(a) = 0 = {c_1 \over a} + c_2

V ( b ) = V o = c 1 b + c 2 \displaystyle V(b) = V_o = {c_1 \over b} + c_2

Solving for c 1 c_1 and c 2 c_2 gives

c 1 = V o ( 1 b βˆ’ 1 a ) \displaystyle c_1 = {V_o \over \left({1\over b}-{1\over a}\right)}

c 2 = βˆ’ V o / a ( 1 b βˆ’ 1 a ) \displaystyle c_2 = -{{V_o/a}\over{\left({1\over b}-{1\over a}\right)}}

and subsitution of these constants into V ( r ) = c 1 / r + c 2 V(r) = {c_1/r} + c_2 gives

V ( r ) = V o ( 1 a βˆ’ 1 b ) ( 1 a βˆ’ 1 r ) \displaystyle V(r) = {V_o \over \left({1\over a}-{1\over b}\right)}{\left({1\over a}-{1\over r}\right)}

As a check of the algebraic steps, plug in r = a r=a and r = b r=b into this equation and verify that the boundary conditions used, V ( a ) = V o V(a)=V_o and V ( b ) = 0 V(b)=0 , are satisfied. Also note that this is the same result obtained in problem 1.4.

3. The electric field can be found using E = βˆ’ βˆ‡ V \mathbf{E}=-\nabla V . In spherical coordinates, when V V depends only on r r ,

βˆ‡ V = βˆ‚ V βˆ‚ r r ^ \displaystyle \boldsymbol{\nabla} V={\partial V \over \partial r} \hat{\mathbf{r}}

giving

E = βˆ’ V o ( 1 a βˆ’ 1 b ) 1 r 2 r ^ \displaystyle \mathbf{E}=-{V_o \over \left({1\over a}-{1\over b}\right)}{1\over r^2}\hat{\mathbf{r}}

This field points outward radially inward as expected given the potential on the outer surface is higher than that on the inner surface.

Evaluated at a a ,

E ( a ) = βˆ’ V o ( 1 a βˆ’ 1 b ) 1 a 2 r ^ \displaystyle \mathbf{E}(a)=-{V_o \over \left({1\over a}-{1\over b}\right)}{1\over a^2}\hat{\mathbf{r}}

At r = a r=a , n ^ = r ^ \hat{\mathbf{n}}=\hat{\mathbf{r}} , so Οƒ a = βˆ’ V o ( 1 a βˆ’ 1 b ) 1 a 2 \displaystyle \sigma_a = -{V_o \over \left({1\over a}-{1\over b}\right)}{1\over a^2}

Evaluated at b b ,

E ( b ) = βˆ’ V o ( 1 a βˆ’ 1 b ) 1 b 2 r ^ \displaystyle \mathbf{E}(b)=-{V_o \over \left({1\over a}-{1\over b}\right)}{1\over b^2}\hat{\mathbf{r}}

At r = b r=b , n ^ = βˆ’ r ^ \hat{\mathbf{n}}=-\hat{\mathbf{r}} , so Οƒ b = + V o ( 1 a βˆ’ 1 b ) 1 b 2 \displaystyle \sigma_b = +{V_o \over \left({1\over a}-{1\over b}\right)}{1\over b^2}

4.

Using V o = Q / C V_o=Q/C and the equation for capacitance for this configuration derived in capacitance:

C = 4 Ο€ Ο΅ o 1 a βˆ’ 1 b \displaystyle C=\frac{4\pi\epsilon_o}{\frac{1}{a}-\frac{1}{b}}

Substitution of V o = Q / C V_o=Q/C into

Οƒ a = βˆ’ V o ( 1 a βˆ’ 1 b ) 1 a 2 \displaystyle \sigma_a = -{V_o \over \left({1\over a}-{1\over b}\right)}{1\over a^2} and Οƒ b = V o ( 1 a βˆ’ 1 b ) 1 b 2 \displaystyle \sigma_b ={V_o \over \left({1\over a}-{1\over b}\right)}{1\over b^2}

gives

Οƒ a = βˆ’ Q 4 Ο€ a 2 \displaystyle \sigma_a =-\frac{Q}{4\pi a^2} and Οƒ b = + Q 4 Ο€ b 2 \displaystyle \sigma_b =+\frac{Q}{4\pi b^2}

See also 3.3.1 of Griffiths.

Laplace’s equation in 2-D cartesian coordinates is

βˆ‡ 2 V = βˆ‚ 2 V βˆ‚ x 2 + βˆ‚ 2 V βˆ‚ y 2 = 0 \nabla^2V = \frac{\partial^2 V}{\partial x^2} + \frac{\partial^2 V}{\partial y^2} = 0

For arbitrary constants A , B , C , D , A,B,C,D, and m m the following equations satisfy it

  1. V ( x , y ) = ( A cosh ⁑ m x + B sinh ⁑ m x ) ( C cos ⁑ m y + D sin ⁑ m y ) V(x,y) = \big(A\cosh mx+B\sinh mx\big)\big(C\cos my+D\sin my\big)

  2. V ( x , y ) = ( A cos ⁑ m x + B sin ⁑ m x ) ( C cosh ⁑ m y + D sinh ⁑ m y ) V(x,y) = \big(A\cos mx+B\sin mx)(C\cosh my+D\sinh my\big)

  3. V ( x , y ) = ( A e m x + B e βˆ’ m x ) ( C cos ⁑ m y + D sin ⁑ m y ) V(x,y) = \big(Ae^{mx}+Be^{-mx})(C\cos my+D\sin my\big)

  4. V ( x , y ) = ( A cos ⁑ m x + B sin ⁑ m x ) ( C e m y + D e βˆ’ m y ) V(x,y) = \big(A\cos mx+B\sin mx\big)\big(Ce^{my}+De^{-my}\big)

  5. V ( x , y ) = ( A cos ⁑ ( m x + B ) ) ( C cosh ⁑ ( m x + B ) ) V(x,y) = \big(A\cos(mx+B)\big)\big(C\cosh(mx+B)\big)

  6. V ( x , y ) = ( A cosh ⁑ ( m x + B ) ) ( C cosh ⁑ ( m x + B ) ) V(x,y) = \big(A\cosh(mx+B)\big)\big(C\cosh(mx+B)\big)

(This can be shown using the method of Separation of Variables covered in Chapter 3 of Griffiths. I use m m on this page instead of k k as is done in the text because I have been using k k for 1 / ( 4 Ο€ Ο΅ 0 ) 1/(4\pi\epsilon_0) .)

All forms are equivalent in the sense that the constants in one equation can be written in terms of constants in any of the other equations. For example, if we write form 3. as V ( x , y ) = ( A β€² e m β€² x + B e βˆ’ m β€² x ) ( C β€² cos ⁑ m β€² y + D β€² sin ⁑ m β€² y ) V(x,y) = \big(A'e^{m'x}+Be^{-m'x})(C'\cos m'y+D'\sin m'y\big) , then one can show that A = ( A β€² + B β€² ) / 2 A=(A'+B')/2 , B = ( A β€² βˆ’ B β€² ) / 2 B=(A'-B')/2 , C = C β€² C=C' , D = D β€² D=D' , and m = m β€² m=m' .

The reason that all four forms are listed is that for certain problems, a certain choice of the form to start with leads to less algebra.

The general solution steps are

  1. Start with one of the forms 1.–4. and use the boundary conditions to find an equation that satisfies some of the four boundary conditions. There is no general rule about which form to start with beyond considering a few boundary conditions and looking at which equation will give coefficients that are zero immediately.

  2. Use superposition and Fourier’s trick to find an equation that satisfies all four boundary conditions.

Laplace’s equation in cartesian coordinates when V = V ( x , y ) V=V(x,y) is

βˆ‡ 2 V = βˆ‚ 2 V βˆ‚ x 2 + βˆ‚ 2 V βˆ‚ y 2 = 0 \nabla^2V = \frac{\partial^2 V}{\partial x^2} + \frac{\partial^2 V}{\partial y^2} = 0

For arbitrary constants A , B , C , D , A,B,C,D, and m m the following four equations satisfy it

  1. V ( x , y ) = ( A cosh ⁑ m x + B sinh ⁑ m x ) ( C cos ⁑ m y + D sin ⁑ m y ) V(x,y) = \big(A\cosh mx+B\sinh mx\big)\big(C\cos my+D\sin my\big)

  2. V ( x , y ) = ( A cos ⁑ m x + B sin ⁑ m x ) ( C cosh ⁑ m y + D sinh ⁑ m y ) V(x,y) = \big(A\cos mx+B\sin mx)(C\cosh my+D\sinh my\big)

  3. V ( x , y ) = ( A e m x + B e βˆ’ m x ) ( C cos ⁑ m y + D sin ⁑ m y ) V(x,y) = \big(Ae^{mx}+Be^{-mx})(C\cos my+D\sin my\big)

  4. V ( x , y ) = ( A cos ⁑ m x + B sin ⁑ m x ) ( C e m y + D e βˆ’ m y ) V(x,y) = \big(A\cos mx+B\sin mx\big)\big(Ce^{my}+De^{-my}\big)

1. Show that equation 1. satisfies βˆ‡ 2 V = 0 \nabla^2V=0 . Recall that the definitions of the hyperbolic sin ⁑ \sin and cos ⁑ \cos are sinh ⁑ z = ( e m z βˆ’ e βˆ’ m z ) / 2 \sinh z = (e^{mz}-e^{-mz})/2 and cosh ⁑ z = ( e m z + e βˆ’ m z ) / 2 \cosh z = (e^{mz}+e^{-mz})/2 .

2. Show that equation 1. is related to equation 3. Do this by labeling the constants in equation 3. with primes and finding the constants in equation 1. in terms of the primed constants.

3. Show that equation 2. is related to equation 4. Do this by labeling the constants in equation 4. with primes and finding the constants in equation 2. in terms of the primed constants.

4. Show that equation 1. can be derived from equation 2. by using Euler’s identity e i z = cos ⁑ z + i sin ⁑ z e^{iz}=\cos z+i\sin z and the definitions of the hyperbolic sin ⁑ \sin and cos ⁑ \cos . Do this by labeling the constants in equation 2. with primes and finding the constants in equation 1. in terms of primed constants.

Answer:

2. Form 3. with primed constants is

V ( x , y ) = ( A β€² e m β€² x + B β€² e βˆ’ m β€² x ) ( C β€² cos ⁑ m y + D β€² sin ⁑ m β€² y ) V(x,y) = \big(A'e^{m'x}+B'e^{-m'x})(C'\cos my+D'\sin m'y\big)

Using

cosh ⁑ m x = ( 1 / 2 ) ( e m x + e βˆ’ m x ) \cosh mx=(1/2)(e^{mx}+e^{-mx}) and sinh ⁑ m x = ( 1 / 2 ) ( e m x βˆ’ e βˆ’ m x ) \sinh mx=(1/2)(e^{mx}-e^{-mx})

we can write

e m β€² x = cosh ⁑ m β€² x + sinh ⁑ m β€² x e^{m'x}=\cosh m'x+\sinh m'x and e βˆ’ m β€² x = cosh ⁑ m β€² x βˆ’ sinh ⁑ m β€² x e^{-m'x}=\cosh m'x-\sinh m'x

Inserting these into the equation for V ( x , y ) V(x,y) gives

V ( x , y ) = ( ( A β€² + B β€² ) cosh ⁑ m β€² x + ( A β€² βˆ’ B β€² ) sinh ⁑ m β€² x ) ( C β€² cos ⁑ m y + D β€² sin ⁑ m β€² y ) \displaystyle V(x,y) = \Big((A'+B')\cosh m'x+(A'-B')\sinh m'x\Big)(C'\cos my+D'\sin m'y\big)

and based on comparison with

V ( x , y ) = ( A cosh ⁑ m x + B sinh ⁑ m x ) ( C cos ⁑ m y + D sin ⁑ m y ) V(x,y) = \big(A\cosh mx+B\sinh mx\big)\big(C\cos my+D\sin my\big)

we conclude

A = A β€² + B β€² A=A'+B' , B = A β€² βˆ’ B β€² B=A'-B' , C = C β€² C=C' , D = D β€² D=D' , and m = m β€² m=m' .

As a check, if A β€² = B β€² = 1 A'=B'=1 and m β€² = 1 m'=1 , form 3. is

V ( x , y ) = ( e x + e βˆ’ x ) ( C β€² cos ⁑ y + D β€² sin ⁑ y ) = ( 2 cosh ⁑ x ) ( C β€² cos ⁑ y + D β€² sin ⁑ y ) V(x,y) = \big(e^{x}+e^{-x})(C'\cos y+D'\sin y\big)=(2\cosh x)(C'\cos y+D'\sin y)

With A β€² = B β€² = 1 A'=B'=1 , A = 2 A=2 , and B = 0 B=0 , so form 1. is

V ( x , y ) = ( 2 cosh ⁑ m x ) ( C cos ⁑ m y + D sin ⁑ m y ) V(x,y) = (2\cosh mx)(C\cos my+D\sin my)

3.

A = A β€² A=A' , B = B β€² B=B' , C = C β€² + D β€² C=C'+D' , D = C β€² βˆ’ D β€² D=C'-D' , and m = m β€² m=m' .

4.

Euler’s identity is

e i z = cos ⁑ z + i sin ⁑ z e^{iz}=\cos z+i\sin z

Replacing z z with βˆ’ z -z gives

e βˆ’ i z = cos ⁑ ( βˆ’ z ) + i sin ⁑ ( βˆ’ z ) = cos ⁑ z βˆ’ i sin ⁑ z e^{-iz}=\cos (-z)+i\sin (-z) = \cos z - i\sin z

Adding the last two equations gives cos ⁑ z = e i z + e βˆ’ i z 2 \displaystyle \cos z = \frac{e^{iz}+e^{-iz}}{2} . Subtracting gives sin ⁑ z = e i z βˆ’ e βˆ’ i z 2 i \displaystyle \sin z = \frac{e^{iz}-e^{-iz}}{2i} .

Comparison of

cos ⁑ z = e i z + e βˆ’ i z 2 \displaystyle \cos z = \frac{e^{iz}+e^{-iz}}{2} with cosh ⁑ z = e z + e βˆ’ z 2 \displaystyle \cosh z=\frac{e^{z}+e^{-z}}{2} gives

cos ⁑ z = cosh ⁑ i z \cos z=\cosh iz and cosh ⁑ z = cos ⁑ ( i z ) \cosh z=\cos (iz)

Comparison of sin ⁑ z = e i z βˆ’ e βˆ’ i z 2 i \displaystyle \sin z = \frac{e^{iz}-e^{-iz}}{2i} with sinh ⁑ z = e z βˆ’ e βˆ’ z 2 \displaystyle \sinh z=\frac{e^{z}-e^{-z}}{2} gives

sin ⁑ z = βˆ’ i sinh ⁑ i z \sin z=-i\sinh i z and sinh ⁑ z = βˆ’ i sin ⁑ ( i z ) \sinh z=-i\sin (iz) .

Form 2. with primed constants is

V ( x , y ) = ( A β€² cos ⁑ m β€² x + B β€² sin ⁑ m β€² x ) ( C β€² cosh ⁑ m β€² y + D β€² sinh ⁑ m β€² y ) V(x,y) = \big(A'\cos m'x+B'\sin m'x)(C'\cosh m'y+D'\sinh m'y\big)

Using the above identities, this can be written as

V ( x , y ) = ( A β€² cosh ⁑ m β€² i x βˆ’ i B β€² sinh ⁑ i m β€² x ) ( C β€² cos ⁑ ( i m β€² y ) βˆ’ i D β€² sin ⁑ ( i m β€² y ) ) V(x,y) = \big(A'\cosh m'ix-iB'\sinh im'x\big)\big(C'\cos (im'y)-iD'\sin (im'y)\big)

From comparison with form 1.

V ( x , y ) = ( A cosh ⁑ m x + B sinh ⁑ m x ) ( C cos ⁑ m y + D sin ⁑ m y ) V(x,y) = \big(A\cosh mx+B\sinh mx\big)\big(C\cos my+D\sin my\big)

we conclude

A = A β€² A=A' , B = βˆ’ i B β€² B=-iB' , C = C β€² C=C' , D = βˆ’ i D β€² D=-iD' , and m = i m β€² m=im' .

or A = A β€² A =A' , B = + i B β€² B=+iB' , C = C β€² C=C' , D = + i D β€² D=+iD' , and m = βˆ’ i m β€² m=-im' .

Start with one of the forms 1.–4. and use the boundary conditions to find an equation that satisfies some of the boundary conditions.

In Examples 3.3 and 3.4, Griffiths happens to choose

  1. The best form 1.–4. to start with

  2. The best order of boundary conditions to address

The problem students have with solving other problems is when they don’t happen to choose the β€œbest” way to get to the answer. In the following, I go through examples 3.3 and 3.4 by addressing the boundary conditions in a different order than Griffiths did. In doing so, I highlight some of the actual complications one will encounter if the β€œbest” order of addressing the boundary conditions is not selected.

Prior to reading the following two examples, read Griffiths’ solutions to Example 3.3 and 3.4.

The boundary conditions are

  1. V = 0 V = 0 when y = 0 y = 0

  2. V = 0 V = 0 when y = a y = a

  3. V = V o ( y ) V = V_o(y) when x = 0 x = 0 – (Griffiths starts with V o ( y ) V_o(y) and then gives the solution for V = V o = c o n s t V=V_o=const )

  4. V β†’ 0 V\rightarrow 0 as x β†’ ∞ x\rightarrow\infty

Griffiths starts with form 3.

V ( x , y ) = ( A e m x + B e βˆ’ m x ) ( C cos ⁑ m y + D sin ⁑ m y ) V(x,y) = \big(Ae^{mx}+Be^{-mx})(C\cos my+D\sin my\big)

and boundary condition 4. Here I am going to start with form 3. but boundary condition 1.

BC 1.: V = 0 V = 0 when y = 0 y = 0

Plugging V = 0 V = 0 and y = 0 y = 0 into

V ( x , y ) = ( A e m x + B e βˆ’ m x ) ( C cos ⁑ m y + D sin ⁑ m y ) V(x,y) = \big(Ae^{mx}+Be^{-mx})(C\cos my+D\sin my\big)

gives

0 = ( A e m x + B e βˆ’ m x ) ( C cos ⁑ m 0 + D sin ⁑ m 0 ) 0 = \big(Ae^{mx}+Be^{-mx})(C\cos m0+D\sin m0\big)

or

0 = ( A e m x + B e βˆ’ m x ) C 0 = \big(Ae^{mx}+Be^{-mx}\big)C

There are two ways for this equation to be true for any x β‰₯ 0 x \ge 0

  1. Both A A and B B are zero. We reject this option because then form 3. reduces to V ( x , y ) = 0 V(x,y)=0 , which satisfies Laplace’s equation, but can’t satisfy all of the boundary conditions.

  2. C = 0 C=0 .

If C = 0 C=0 , we are left with

V ( x , y ) = ( A e m x + B e βˆ’ m x ) D sin ⁑ m y V(x,y) = \big(Ae^{mx}+Be^{-mx}\big)D\sin my

or, defining A β€² = A D A'=AD and B β€² = B D B'=BD ,

V ( x , y ) = ( A β€² e m x + B β€² e βˆ’ m x ) sin ⁑ m y V(x,y) = \big(A'e^{mx}+B'e^{-mx}\big)\sin my

BC 2.: V = 0 V = 0 when y = a y = a

Plugging these values into the equation we left off with after addressing boundary condition 1. gives

0 = ( A β€² e m x + B β€² e βˆ’ m x ) sin ⁑ m a 0 = \big(A'e^{mx}+B'e^{-mx}\big)\sin ma

There are two ways for this equation to be true for any x β‰₯ 0 x \ge 0 :

  1. Both A β€² A' and B β€² B' are zero. We reject this option because then we are left with V ( x , y ) = 0 V(x,y)=0 , which satisfies Laplace’s equation, but can’t satisfy all of the boundary conditions.

  2. sin ⁑ m a = 0 \sin ma=0 .

The only way for sin ⁑ m a = 0 \sin ma=0 to be true generally is if m a = 0 , Β± Ο€ , Β± 2 Ο€ , … ma=0,\pm \pi,\pm 2\pi,… .

From this we conclude

m = n Ο€ a m=\frac{n\pi}{a} with n n constrained to be 0 , Β± 1 , Β± 2 , … 0, \pm 1, \pm 2, …

This leaves

V ( x , y ) = ( A β€² e n Ο€ x / a + B β€² e βˆ’ n Ο€ x / a ) sin ⁑ n Ο€ y / a V(x,y) = \big(A'e^{n\pi x/a}+B'e^{-n\pi x/a}\big)\sin n\pi y/a

Inspection of this equation allows us to reject the possibility that n = 0 n=0 ; if n = 0 n=0 , this equation reduces to V ( x , y ) = 0 V(x,y)=0 , which cannot satisfy all of the boundary conditions.

BC 3.: V = V o V = V_o when x = 0 x = 0

Plugging these values into the equation we left off with after addressing boundary condition 2. gives

V o = ( A β€² e n Ο€ 0 / a + B β€² e βˆ’ n Ο€ 0 / a ) sin ⁑ n Ο€ y / a V_o = \big(A'e^{n\pi 0/a}+B'e^{-n\pi 0/a}\big)\sin n\pi y/a

or, because e 0 = 1 e^0=1 ,

V o = ( A β€² + B β€² ) sin ⁑ n Ο€ y / a V_o = \big(A'+B'\big)\sin n\pi y/a

This equation is true but is not immediately useful for simplifying our equation for V ( x , y ) V(x,y) . So we move on to the next boundary condition.

BC 4.: V = 0 V = 0 when x β†’ ∞ x \rightarrow \infty

Plugging these values into the equation we left off with after addressing boundary condition 2. gives

0 = ( A β€² e n Ο€ ∞ / a + B β€² e βˆ’ n Ο€ ∞ / a ) sin ⁑ n Ο€ y / a 0 = \big(A'e^{n\pi \infty/a}+B'e^{-n\pi \infty/a}\big)\sin n\pi y/a

or, because e βˆ’ ∞ = 0 e^{-\infty}=0 ,

0 = ( A β€² e n Ο€ ∞ / a ) sin ⁑ n Ο€ y / a 0 = \big(A'e^{n\pi \infty/a}\big)\sin n\pi y/a

There is only one way for this equation to be true: A β€² = 0 A'=0 . (There is technically a mathematical issue that I am ignoring because zero times infinity is indeterminate; there is a way around this issue, but for now, it is not important and I’ll use this dubious notation.)

This leaves

V ( x , y ) = B β€² e βˆ’ n Ο€ x / a sin ⁑ n Ο€ y / a V(x,y) = B'e^{-n\pi x/a}\sin n\pi y/a

which the same as equations 3.28 and 3.29 with one exception. In the above I concluded n n could be Β± 1 , Β± 2 , … \pm 1, \pm 2, … where Griffiths concludes n = 1 , 2 , 3 , … n=1, 2, 3, … . The equivalence of these two conclusions will be shown when B β€² B' is computed.

The final step is determining B β€² B' . This would seem to be impossible because boundary condition 3., V = V o V = V_o when x = 0 x = 0 , gives

V o = B β€² e βˆ’ n Ο€ 0 / a sin ⁑ n Ο€ y / a = B β€² sin ⁑ ( n Ο€ y / a ) V_o = B'e^{-n\pi 0/a}\sin n\pi y/a=B'\sin(n\pi y/a)

and this equation needs to hold true for any 0 ≀ y ≀ a 0\le y \le a . Another way to show that finding a B β€² B' that satisfies boundary condition 3. is impossible, suppose y = a / 2 y=a/2 and n = 1 n=1 , then we have

V o = B β€² e βˆ’ n Ο€ 0 / a sin ⁑ n Ο€ y / a = B β€² sin ⁑ ( Ο€ / 2 ) V_o = B'e^{-n\pi 0/a}\sin n\pi y/a=B'\sin(\pi/2)

Giving B β€² = V o B'=V_o . Now suppose y = a / 4 y=a/4 and n = 1 n=1 , then

V o = B β€² e βˆ’ n Ο€ 0 / a sin ⁑ n Ο€ y / a = B β€² sin ⁑ ( Ο€ / 4 ) V_o = B'e^{-n\pi 0/a}\sin n\pi y/a=B'\sin(\pi/4)

Giving B β€² = 2 V o B'=\sqrt{2}V_o . One can repeat this for all allowed values and n n and you will come to the conclusion that there is no single value of B β€² B' that can satisfy this equation for any 0 ≀ y ≀ a 0\le y \le a .

To satisfy this last boundary condition, superposition and β€œFourier’s Trick” is needed, which is covered in step 2.

The boundary conditions are

  1. V = 0 V = 0 when y = 0 y = 0

  2. V = 0 V = 0 when y = a y = a

  3. V = V o V = V_o when x = b x = b

  4. V = V o V = V_o when x = βˆ’ b x = -b

Griffiths starts with form 3.

V ( x , y ) = ( A e m x + B e βˆ’ m x ) ( C cos ⁑ m y + D sin ⁑ m y ) V(x,y) = \big(Ae^{mx}+Be^{-mx})(C\cos my+D\sin my\big)

and a symmetry argument to conclude A = B A=B . Here I am going to start with form 3. but boundary condition 1. In addition, I won’t make use of the symmetry argument.

BC 1.: V = 0 V = 0 when y = 0 y = 0

BC 2.: V = 0 V = 0 when y = a y = a

The starting form 3. and boundary conditions 1. and 2. are identical to that from example 3.3, so we can re-use the result concluded from addressing them:

V ( x , y ) = ( A β€² e n Ο€ x / a + B β€² e βˆ’ n Ο€ x / a ) sin ⁑ n Ο€ y / a V(x,y) = \big(A'e^{n\pi x/a}+B'e^{-n\pi x/a}\big)\sin n\pi y/a

with n n constrained to be 0 , Β± 1 , Β± 2 , … 0, \pm 1, \pm 2, … .

BC 3.: V = V o V = V_o when x = b x = b

Plugging the boundary values into the equation we left off with after addressing boundary condition 2. gives

V o = ( A β€² e n Ο€ b / a + B β€² e βˆ’ n Ο€ b / a ) sin ⁑ n Ο€ y / a V_o = \big(A'e^{n\pi b/a}+B'e^{-n\pi b/a}\big)\sin n\pi y/a

Now we seem stuck. How does one constrain A β€² A' and B β€² B' for this equation to hold? We have one other boundary condition to address, so perhaps it will provide some insight.

BC 4.: V = V o V = V_o when x = βˆ’ b x = -b

V o = ( A β€² e βˆ’ n Ο€ b / a + B β€² e n Ο€ b / a ) sin ⁑ n Ο€ y / a V_o = \big(A'e^{-n\pi b/a}+B'e^{n\pi b/a}\big)\sin n\pi y/a

Here we would seem to be stuck again. However, notice that the equation here can be combined with that from BC 3:

( A β€² e n Ο€ b / a + B β€² e βˆ’ n Ο€ b / a ) sin ⁑ n Ο€ y / a = ( A β€² e βˆ’ n Ο€ b / a + B β€² e n Ο€ b / a ) sin ⁑ n Ο€ y / a \big(A'e^{n\pi b/a}+B'e^{-n\pi b/a}\big)\sin n\pi y/a=\big(A'e^{-n\pi b/a}+B'e^{n\pi b/a}\big)\sin n\pi y/a

The sin ⁑ n Ο€ y / a \sin n\pi y/a term appears on both sides and can be dropped, leaving

( A β€² e n Ο€ b / a + B β€² e βˆ’ n Ο€ b / a ) = ( A β€² e βˆ’ n Ο€ b / a + B β€² e n Ο€ b / a ) \big(A'e^{n\pi b/a}+B'e^{-n\pi b/a}\big)=\big(A'e^{-n\pi b/a}+B'e^{n\pi b/a}\big)

which simplifies to

A β€² ( e n Ο€ b / a βˆ’ e βˆ’ n Ο€ b / a ) = B β€² ( e n Ο€ b / a βˆ’ e βˆ’ n Ο€ b / a ) A'\big(e^{n\pi b/a}-e^{-n\pi b/a}\big)=B'\big(e^{n\pi b/a}-e^{-n\pi b/a}\big)

The terms in parentheses are identical, so we conclude

A β€² = B β€² A'=B'

Therefore, the equation we ended with when addressing BC 2.,

V ( x , y ) = ( A β€² e βˆ’ n Ο€ b / a + B β€² e n Ο€ b / a ) sin ⁑ n Ο€ y / a V(x,y) = \big(A'e^{-n\pi b/a}+B'e^{n\pi b/a}\big)\sin n\pi y/a

simplifies to

V ( x , y ) = A β€² ( e βˆ’ n Ο€ x / a + e n Ο€ x / a ) sin ⁑ n Ο€ y / a V(x,y) = A'\big(e^{-n\pi x/a}+e^{n\pi x/a}\big)\sin n\pi y/a

This can be futher simplified using the definition cosh ⁑ z = ( e z + e βˆ’ z ) / 2 \cosh z=(e^z+e^{-z})/2

V ( x , y ) = 2 A β€² cosh ⁑ ( n Ο€ x / a ) sin ⁑ ( n Ο€ y / a ) V(x,y) = 2A'\cosh(n\pi x/a)\sin(n\pi y/a)

This equation is equivalent to Griffiths’ 3.41 if 2 A β€² 2A' is replaced with a constant labeled C C . As with example 3.3, we ended up with one constant that is still unknown.

Note that n = 0 n=0 would give V ( x , y ) = 0 V(x,y)=0 , which does not satisfy all of the boundary conditions. As a result, the allowed values of n n are Β± 1 , Β± 2 , … \pm 1, \pm 2, … .

As was the case in my solution to example 3.3, above I concluded n n could be Β± 1 , Β± 2 , … \pm 1, \pm 2, … where Griffiths concludes n = 1 , 2 , 3 , … n=1, 2, 3, … . The equivalence of these two conclusions will be shown when A β€² A' is computed in step 2..

  1. Show that ∫ 0 Ο€ sin ⁑ 2 x sin ⁑ x d x \displaystyle\int_0^\pi \sin 2x \sin x dx = 0. (Note that without any calculation, this is expected from a plot of sin ⁑ x \sin x and sin ⁑ 2 x \sin 2x and thinking about the area under the curve for their product.)

  2. Show that ∫ 0 Ο€ sin ⁑ n x sin ⁑ l x d x \displaystyle \int_0^\pi \sin nx \sin lx dx = 0 for integer n n and l l if l β‰  n l\ne n by explicitly evaluating the integral.

  3. Compute ∫ 0 Ο€ sin ⁑ 2 n x d x \displaystyle\int_0^\pi \sin^2 nx dx for

    1. any n β‰  0 n\ne 0 and

    2. for integer n n .

Comment:

The motivation for giving you these problems is to help you remember that

∫ 0 Ο€ sin ⁑ n x sin ⁑ l x d x = 0 \int_0^\pi \sin nx \sin lx dx = 0 if l β‰  n l\ne n which is important for Fourier sum problems.

A way to reason this out without doing a calculation is by considering a plot of the integrand sin ⁑ n x sin ⁑ l x \sin nx\sin lx over the interval 0 ≀ x ≀ Ο€ 0\le x\le \pi by plotting two sin ⁑ \sin curves for l β‰  m l\ne m and thinking about what the product of the curves looks like; over half interval the product will be positive and over the other half of the interval it will be negative, so the integral will be zero.

When l = m l=m , the integrand is sin ⁑ 2 m x \sin^2 mx , which is always positive, so its integral is not zero over any interval of x x .

Show that

V ( x , y ) = C 1 e βˆ’ Ο€ x / a sin ⁑ Ο€ y / a + C 2 e βˆ’ 2 Ο€ x / a sin ⁑ 2 Ο€ y / a + C 3 e βˆ’ 3 Ο€ x / a sin ⁑ 3 Ο€ y / a + … V(x,y) = C_1e^{-\pi x/a}\sin\pi y/a + C_2e^{-2\pi x/a}\sin 2\pi y/a + C_3e^{-3\pi x/a}\sin 3\pi y/a + …

satisfies Laplace’s equation.

Answer

From a previous problem, we know that equations of this form of each term satisfy Laplace’s equation. Here will show this again. Define these terms as

V n = C n e βˆ’ n Ο€ x / a sin ⁑ n Ο€ y / a V_n=C_ne^{-n\pi x/a}\sin n\pi y/a

Then for any n n ,

βˆ‡ 2 V n = =   βˆ‡ 2 ( C n e βˆ’ n Ο€ x / a sin ⁑ [ n Ο€ y / a ] ) =   C n sin ⁑ [ n Ο€ y / a ] βˆ‚ 2 e βˆ’ n Ο€ x / a βˆ‚ x 2 + C n e βˆ’ n Ο€ x / a βˆ‚ 2 sin ⁑ [ n Ο€ y / a ] βˆ‚ y 2 =   C n sin ⁑ [ n Ο€ y / a ] ( βˆ’ n Ο€ x / a ) ( βˆ’ n Ο€ x / a ) e βˆ’ n Ο€ x / a + C n e βˆ’ n Ο€ x / a ( n Ο€ x / a ) ( βˆ’ n Ο€ x / a ) sin ⁑ [ n Ο€ y / a ] =   0 \begin{aligned} \nabla^2 V_n = & \\ = & \text{ } \nabla^2\big(C_ne^{-n\pi x/a}\sin [n\pi y/a]\big) \\ = & \text{ } C_n\sin [n\pi y/a]\frac{\partial^2 e^{-n\pi x/a}}{\partial x^2} + C_ne^{-n\pi x/a}\frac{\partial^2 \sin [n\pi y/a]}{\partial y^2} \\ = & \text{ } C_n \sin [n\pi y/a] (-n\pi x/a)(-n\pi x/a)e^{-n\pi x/a} + C_n e^{-n\pi x/a}(n\pi x/a)(-n\pi x/a)\sin [n\pi y/a] \\ = & \text{ } 0 \end{aligned}

The Laplacian of a sum can be written as the sum of Laplacians

βˆ‡ 2 ( V 1 + V 2 + … ) = βˆ‡ 2 V 1 + βˆ‡ 2 V 2 + … \displaystyle\nabla^2 (V_1+V_2+…) = \nabla^2 V_1 + \nabla^2 V_2 + …

Because βˆ‡ 2 V n = 0 \nabla^2 V_n=0 , we conclude

βˆ‡ 2 ( V 1 + V 2 + … ) = 0 + 0 + … = 0 \displaystyle\nabla^2 (V_1+V_2+…) = 0 + 0 + … = 0

In this step, we start with an equation that satisfies two or three of the boundary conditions, but cannot satisfy the remaining one(s). Then,

  1. Re–write one of the remaining boundary conditions as a sum of equations that satisfy the other boundary conditions and

  2. Apply Fourier’ trick: Multiply the equation from 1. by a sin ⁑ \sin or cos ⁑ \cos function and integrate.

The equation

V ( x , y ) = C e βˆ’ n Ο€ x / a sin ⁑ n Ο€ y / a V(x,y) = Ce^{-n\pi x/a}\sin n\pi y/a with n = 0 , Β± 1 , … n=0, \pm 1, …

satisfied three of the four boundary conditions in Example 3.3 but there is no C C or n n that can satisfy the last boundary condition V ( 0 , y ) = V o V(0,y)=V_o :

V o = C e βˆ’ n Ο€ 0 / a sin ⁑ n Ο€ y / a = C sin ⁑ n Ο€ y / a V_o = Ce^{-n\pi 0/a}\sin n\pi y/a=C\sin n\pi y/a

As a result, we try a sum of terms with n = 1 , 2 , … n=1, 2, …

V o = C 1 sin ⁑ Ο€ y / a + C 2 sin ⁑ 2 Ο€ y / a + C 3 sin ⁑ 3 Ο€ y / a + … V_o = C_1\sin \pi y/a + C_2\sin 2\pi y/a + C_3\sin 3\pi y/a + …

Multiplying this by d y sin ⁑ l Ο€ y / a dy\sin l\pi y/a and integrating from y = 0 y=0 to y = a y=a gives

Equation A:

∫ 0 a V o d y sin ⁑ ( l Ο€ y / a ) = \displaystyle \int_0^a V_ody \sin (l\pi y/a) =

∫ 0 a V o d y sin ⁑ ( l Ο€ y / a ) = + C 1 ∫ 0 a d y sin ⁑ ( l Ο€ y / a ) sin ⁑ ( Ο€ y / a ) \displaystyle \phantom{\int_0^a V_ody \sin (l\pi y/a) =}\phantom{+}C_1\int_0^a dy \sin (l\pi y/a) \sin (\pi y/a)

∫ 0 a V o d y sin ⁑ ( l Ο€ y / a ) = + C 2 ∫ 0 a d y sin ⁑ ( l Ο€ y / a ) sin ⁑ ( 2 Ο€ y / a ) \displaystyle \phantom{\int_0^a V_ody \sin (l\pi y/a) =}+C_2\int_0^a dy\sin( l\pi y/a) \sin (2\pi y/a)

∫ 0 a V o d y sin ⁑ ( l Ο€ y / a ) = + C 3 ∫ 0 a d y sin ⁑ ( l Ο€ y / a ) sin ⁑ ( 3 Ο€ y / a ) \displaystyle \phantom{\int_0^a V_ody \sin (l\pi y/a) =}+C_3\int_0^a dy \sin (l\pi y/a) \sin (3\pi y/a)

∫ 0 a V o d y sin ⁑ ( l Ο€ y / a ) = + … \displaystyle\phantom{\int_0^a V_ody \sin (l\pi y/a) =}+ …

To show that C 1 , C 2 , C_1, C_2, and C 3 C_3 are

  • C 1 = 4 V o 1 Ο€ \displaystyle C_1=\frac{4V_o}{1\pi} ;

  • C 2 = 0 \displaystyle C_2=0 ; and

  • C 3 = 4 V o 3 Ο€ \displaystyle C_3=\frac{4V_o}{3\pi} ,

set l = 1 l=1 in Equation A and evaluate all of the integrals. This will give you C 1 C_1 . Next, set l = 2 l=2 in Equation A and evaluate all of the integrals. This will give you C 2 C_2 . Finally, set l = 3 l=3 in Equation A and evaluate all of the integrals to get C 3 C_3 .

In class, I started the problem of finding the potential V ( x , y ) V(x,y) inside of a long square duct with boundary conditions

  1. V ( 0 , y ) = 0 V(0,y) = 0 (left)

  2. V ( a , y ) = 0 V(a,y) = 0 (right)

  3. V ( x , 0 ) = 0 V(x,0) = 0 (bottom)

  4. V ( x , a ) = V o V(x,a) = V_o (top)

Finish this problem. That is, find the coefficients C n C_n in the equation

V ( x , y ) = βˆ‘ n = 0 ∞ C n sin ⁑ ( n Ο€ a x ) sinh ⁑ ( n Ο€ a y ) \displaystyle V(x,y)=\sum_{n=0}^\infty C_n \sin \left(\frac{n\pi}{a} x \right)\sinh \left(\frac{n\pi}{a} y\right)

I recommend that you do the full problem, but you may start the problem by beginning with the equation that I ended with after addressing the third boundary condition.

Answer

C n = 4 V o / ( n Ο€ sinh ⁑ n Ο€ ) C_n=4V_o/(n\pi\sinh n\pi) for n = 1 , 3 , … n=1,3,…

C n = 0 C_n=0 for n = 2 , 4 , … n=2,4,…

V ( x , y ) = 4 V o Ο€ βˆ‘ n = 1 , 3 , … ∞ sin ⁑ ( n Ο€ x / a ) sinh ⁑ ( n Ο€ y / a ) n sinh ⁑ ( n Ο€ ) \displaystyle V(x,y)=\frac{4V_o}{\pi}\sum_{n=1,3,…}^\infty \sin(n\pi x/a)\frac{\sinh(n\pi y/a)}{n\sinh(n\pi)}

In a previous problem, you found the C n C_n s in

V ( x , y ) = βˆ‘ n = 0 ∞ C n sin ⁑ ( n Ο€ a x ) sinh ⁑ ( n Ο€ a y ) \displaystyle V(x,y)=\sum_{n=0}^\infty C_n \sin \left(\frac{n\pi}{a} x \right)\sinh \left(\frac{n\pi}{a} y\right)

Use this equation with the values of C n C_n found in your solution and plot V ( x , a ) V(x,a) vs. x x in the range x = 0 x=0 to x = a x=a . Plot only the terms in the sum for n ≀ 10 n\le 10 .

Find the potential V ( x , y ) V(x,y) inside of a long rectangular duct, 0 ≀ x ≀ b 0 \le x\le b ; 0 ≀ y ≀ a 0 \le y\le a , with boundary conditions

  1. V ( 0 , y ) = V o V(0,y) = V_o (left)

  2. V ( b , y ) = 0 V(b,y) = 0 (right)

  3. V ( x , 0 ) = 0 V(x,0) = 0 (bottom)

  4. V ( x , a ) = 0 V(x,a) = 0 (top)

Note that this is a rectangular duct, so your solution should have an a a and b b .

Do this by following the steps that I used in class, which are also in the notes.

Answer

V ( x , y ) = 4 V o Ο€ βˆ‘ n = 1 , 3 , … ∞ 1 n ( cosh ⁑ ( n Ο€ x / a ) βˆ’ sinh ⁑ ( n Ο€ x / a ) tanh ⁑ ( n Ο€ b / a ) ) sin ⁑ ( n Ο€ y / a ) \displaystyle V(x,y)=\frac{4V_o}{\pi}\sum_{n=1,3,…}^\infty\frac{1}{n}\left(\cosh(n\pi x/a)-\frac{\sinh(n\pi x/a)}{\tanh(n\pi b/a)}\right)\sin(n\pi y/a)

This can be simplified to

V ( x , y ) = 4 V o Ο€ βˆ‘ n = 1 , 3 , … ∞ ( sinh ⁑ [ n Ο€ ( b βˆ’ x ) / a ] n sinh ⁑ ( n Ο€ b / a ) ) sin ⁑ ( n Ο€ y / a ) \displaystyle V(x,y)=\frac{4V_o}{\pi}\sum_{n=1,3,…}^\infty\left(\frac{\sinh[n\pi (b-x)/a]}{n\sinh(n\pi b/a)}\right)\sin(n\pi y/a)

Short–cut solution

One could have started assuming an equation of the form

V ( x , y ) = A sinh ⁑ ( m x βˆ’ B ) [ C sin ⁑ ( m x ) + D cos ⁑ ( m x ) ] V(x,y) = A\sinh(mx-B)[C\sin(mx)+D\cos(mx)]

which can be shown to be related to any of the four forms that were given in the noted. BCs 3. and 4. give D = 0 D=0 and m = n Ο€ / a m=n\pi/a , leaving

V ( x , y ) = A C sinh ⁑ ( n Ο€ x / a βˆ’ B ) [ sin ⁑ ( n Ο€ y / a ) ] V(x,y) = AC\sinh(n\pi x/a-B)[\sin(n\pi y/a)]

To satisfy BC 2., one needs sinh ⁑ ( n Ο€ x / a βˆ’ B ) = 0 \sinh(n\pi x/a-B)=0 , so B = n Ο€ b / a B=n\pi b/a and replacing A C AC with C C

V ( x , y ) = C sinh ⁑ ( n Ο€ ( x βˆ’ b ) / a ) [ sin ⁑ ( n Ο€ y / a ) ] V(x,y) = C\sinh(n\pi(x-b)/a)[\sin(n\pi y/a)]

To satisfiy BC 2., we use

V ( 0 , y ) = C 1 sinh ⁑ ( βˆ’ Ο€ b / a ) sin ⁑ ( Ο€ y / a ) + C 2 sinh ⁑ ( βˆ’ 2 Ο€ b / a ) sin ⁑ ( 2 Ο€ y / a ) + … V(0,y) = C_1\sinh(-\pi b/a)\sin(\pi y/a)+C_2\sinh(-2\pi b/a)\sin(2\pi y/a)+…

and Fourier’s trick as usual. In this case, we arrive at the simplified form given above.

Show that your answer to the previous problem is consistent with the solution for Griffiths Example 3.3 in the limit that a / b β†’ 0 a/b\rightarrow 0 .