Table of Contents

Typical Examples

We present here three examples that are fairly typical of junior-level coursework for physics majors, followed in each case by a brief discussion, emphasizing the challenges students have when applying their prior mathematical knowledge. Bear in mind that these examples arise shortly after students have learned the underlying mathematical concepts.

— TD & CAM, 3/15/14

The electric field above a line segment

Find the electric field a distance $z$ above the midpoint of a straight line segment of length $2L$ that carries a uniform line charge $\lambda$.

(David J. Griffiths, Introduction to Electrodynamics, 4th edition, Pearson, 2013, pages 64–65) 1)

Background

The electric field $\EE$ due to a point charge $q$ at the origin is \begin{equation} \EE = \frac{1}{4\pi\epsilon_0} \frac{q\,\rhat}{r^2} \end{equation} This leads through the superposition principle to the electric field of a linear charge distribution $\lambda$ along a curve $C$ being \begin{equation} \EE = \frac{1}{4\pi\epsilon_0} \int_C \frac{\lambda\,\widehat{\Delta r}\,d\ell'}{|\vec{\Delta r}|^2} \end{equation} where $\EE$ is regarded as a function of position $\rr$, and the integral is over the position $\rr'$ along the curve. Thus, $\lambda$ depends on $\rr'$, and $d\ell'=|d\rr'|$.

Solution

We have \begin{align} \rr=z\,\zhat \qquad \rrp&=x\,\xhat \qquad d\ell'=dx \nonumber\\ \vec{\Delta r}=\rr-\rrp=z\,\zhat-x\,\xhat &\qquad |\vec{\Delta r}|=\sqrt{z^2+x^2} \nonumber\\ \widehat{\Delta r}=\frac{\vec{\Delta r}}{|\vec{\Delta r}|} &= \frac{z\,\zhat-x\,\xhat}{\sqrt{z^2+x^2}} \end{align} Thus, \begin{align} \EE &= \frac{1}{4\pi\epsilon_0} \int_{-L}^L \frac{\lambda}{z^2+x^2} \frac{z\,\zhat-x\,\xhat}{\sqrt{z^2+x^2}}dx \nonumber\\ \EE &= \frac{\lambda}{4\pi\epsilon_0} \left[ z\,\zhat \int_{-L}^L \frac{1}{(z^2+x^2)^{3/2}}\,dx -\xhat \int_{-L}^L \frac{x}{(z^2+x^2)^{3/2}}\,dx \right] \nonumber\\ &= \frac{\lambda}{4\pi\epsilon_0} \left[ z\,\zhat \left( \frac{x}{z^2\sqrt{z^2+x^2}} \right) \Bigg|_{-L}^L -\xhat \left( -\frac{1}{\sqrt{z^2+x^2}} \right) \Bigg|_{-L}^L \right]\ \nonumber\\ &= \frac{1}{4\pi\epsilon_0} \frac{2\lambda L}{z\sqrt{z^2+L^2}}\zhat \end{align}

For points far from the line ($z\gg L$), \begin{equation} E \cong \frac{1}{4\pi\epsilon_0} \frac{2\lambda L}{z^2} \end{equation} This makes sense: From far away the line looks like a point charge $q=2\lambda L$. In the limit $L\mapsto\infty$, on the other hand, we obtain the field of an infinite straight wire: \begin{equation} E = \frac{1}{4\pi\epsilon_0} \frac{2\lambda}{z} \end{equation}

Discussion

Solutions of Schrödinger's Equation

Find the general solution of the Schrödinger equation for a time-independent Hamiltonion.

(David H. McIntyre, Quantum Mechanics: A Paradigms Approach, Pearson Addison-Wesley, 2012, pages 68–70)

Background

The time evolution of a quantum system is determined by the Hamiltonion or total energy operator $H$ through the Schrödinger equation \begin{equation} i\hbar\frac{d}{dt}|\psi(t)\rangle = H|\psi(t)\rangle \end{equation} The energy eigenvalue equation is \begin{equation} H|E_n\rangle = E_n|E_n\rangle \end{equation} The eigenvectors of the Hamiltonion form a complete basis, so that \begin{equation} |\psi(t)\rangle = \sum_n c_n(t) |E_n\rangle \end{equation} We assume that $H$, and hence $|E_n\rangle$ and $E_n$, are independent of $t$.

Solution

Substitute the general state $\psi(t)$ above into the Schrödinger equation: \begin{equation} i\hbar\frac{d}{dt}\sum_n c_n(t)|E_n\rangle = H \sum_n c_n(t)|E_n\rangle \end{equation} and use the energy eigenvalue equation to obtain \begin{equation} i\hbar\sum_n\frac{dc_n(t)}{dt}|E_n\rangle = \sum_n c_n(t)E_n|E_n\rangle \end{equation} Each side of this equation is a sum over all the energy states of the system. To simplify this equation, we isolate single terms in these two sums by taking the inner product of the ket on each side with one particular ket $|E_k\rangle$. The orthonormality condition $\langle E_k|E_n\rangle=\delta_{kn}$ then collapses the sums: \begin{align} \langle E_k|i\hbar\sum_n\frac{dc_n(t)}{dt}|E_n\rangle &= \langle E_k|\sum_n c_n(t)E_n|E_n\rangle \\ i\hbar\sum_n\frac{dc_n(t)}{dt} \langle E_k|E_n\rangle &= \sum_n c_n(t)E_n\langle E_k|E_n\rangle \\ i\hbar\sum_n\frac{dc_n(t)}{dt} \delta_{kn} &= \sum_n c_n(t)E_n\delta_{kn} \\ i\hbar\frac{dc_k(t)}{dt} &= c_k(t)E_k \end{align} We are left with a single differential equation for each of the possible energy states of the system $k=1,2,3,…$. This first-order differential equation can be rewritten as \begin{equation} \frac{dc_k(t)}{dt} = -i\frac{E_k}{\hbar} c_k(t) \end{equation} whose solution is a complex exponential \begin{equation} c_k(t) = c_k(0) e^{-iE_k/\hbar} \end{equation} We have denoted the initial condition as $c_k(0)$, but we denote it simply as $c_k$ hereafter. Each coefficient in the energy basis expansion of the state obeys the same form of the time dependence, but with a different exponent due to the different energies. The time-dependent solution for the full state vector is summarized by saying that if the initial state of the system at time $t=0$ is \begin{equation} |\psi(0)\rangle = \sum_n c_n|E_n\rangle \end{equation} then the time evolution of this state under the time-independent Hamiltonian $H$ is \begin{equation} |\psi(t)\rangle = \sum_n c_ne^{-iE_n/\hbar}|E_n\rangle \end{equation}

Discussion

Maxwell relations

Derive the Maxwell relation associated with internal energy.

(Ashley H. Carter, Classical and Statistical Thermodynamics, Prentice-Hall, 2001, page 134)

Background

For homogeneous systems, with a well-defined temperature $T$ and pressure $P$, the first law of thermodynamics becomes the fundamental thermodynamic relation \begin{equation} dU = T\,dS - P\,dV \end{equation} where $V$ is the volume of the system, $S$ its entropy, and $U$ denotes the internal energy.

Solution

The internal energy $U$ is a state variable whose differential is exact. Since the value of a mixed second partial derivative is independent of the order in which the differentiation is applied, we have \begin{equation} dU = T\,dS + (-P)\,dV = \left(\frac{\partial U}{\partial S}\right)_V dS + \left(\frac{\partial U}{\partial V}\right)_S dV \end{equation} The exactness of $dU$ immediately gives \begin{equation} \frac{\partial^2U}{\partial V\,\partial S} = \left(\frac{\partial T}{\partial V}\right)_S = \frac{\partial^2U}{\partial S\,\partial V} = -\left(\frac{\partial P}{\partial S}\right)_V \end{equation} The equality of the first derivatives, \begin{equation} \left(\frac{\partial T}{\partial V}\right)_S = -\left(\frac{\partial P}{\partial S}\right)_V \end{equation} is known as a Maxwell relation. Its utility lies in the fact that each of the partials is a state variable that can be integrated along any convenient reversible path to obtain differences in values of the fundamental state variables between given equilibrium states.

Discussion

1) With apologies to David Griffiths, we have used $\vec{\Delta r}$ in place of “script-r vector”, which we have been unable to typeset.