next up previous
Next: About this document ... Up: Wave Mechanics Previous: Simple Harmonic Oscillator


Angular Momentum

In classical mechanics, the vector angular momentum, L, of a particle of position vector $ {\bf r}$ and linear momentum $ {\bf p}$ is defined as

$\displaystyle {\bf L} = {\bf r}\times {\bf p}.$ (C.122)

In other words,

$\displaystyle L_x$ $\displaystyle = y p_z - z p_y,$ (C.123)
$\displaystyle L_y$ $\displaystyle = z p_x - x p_z,$ (C.124)
$\displaystyle L_z$ $\displaystyle = x p_y - y p_x.$ (C.125)

In quantum mechanics, the operators, $ p_i$ , that represent the Cartesian components of linear momentum, are represented as the spatial differential operators $ -{\rm i} \hbar \partial/\partial x_i$ . [See Equation (C.81)]. It follows that:

$\displaystyle L_x$ $\displaystyle = -{\rm i} \hbar\left(y \frac{\partial}{\partial z} - z \frac{\partial}{\partial y}\right),$ (C.126)
$\displaystyle L_y$ $\displaystyle = -{\rm i} \hbar\left(z \frac{\partial}{\partial x} - x \frac{\partial}{\partial z}\right),$ (C.127)
$\displaystyle L_z$ $\displaystyle = -{\rm i} \hbar\left(x \frac{\partial}{\partial y} - y \frac{\partial}{\partial x}\right).$ (C.128)

In addition, let

$\displaystyle L^{ 2} = L_x^{ 2}+L_y^{ 2}+L_z^{ 2}$ (C.129)

be the magnitude-squared of the angular momentum vector.

It is most convenient to work in terms of the standard spherical coordinates, $ r$ , $ \theta$ , and $ \phi$ . These are defined with respect to our usual Cartesian coordinates as follows:

$\displaystyle x$ $\displaystyle = r \sin\theta \cos\phi,$ (C.130)
$\displaystyle y$ $\displaystyle = r \sin\theta \sin\phi,$ (C.131)
$\displaystyle z$ $\displaystyle = r \cos\theta.$ (C.132)

We deduce, after some tedious analysis, that

$\displaystyle \frac{\partial}{\partial x}$ $\displaystyle = \sin\theta \cos\phi \frac{\partial}{\partial r} + \frac{\cos\...
...partial\theta} - \frac{\sin\phi}{r \sin\theta} \frac{\partial}{\partial\phi},$ (C.133)
$\displaystyle \frac{\partial}{\partial y}$ $\displaystyle = \sin\theta \sin\phi \frac{\partial}{\partial r} + \frac{\cos\...
...partial\theta} + \frac{\cos\phi}{r \sin\theta} \frac{\partial}{\partial\phi},$ (C.134)
$\displaystyle \frac{\partial}{\partial z}$ $\displaystyle = \cos\theta \frac{\partial}{\partial r} -\frac{\sin\theta}{r} \frac{\partial}{\partial \theta}.$ (C.135)

Making use of the definitions (C.126)-(C.129), after more tedious algebra, we obtain

$\displaystyle L_x$ $\displaystyle = - {\rm i} \hbar\left(-\sin\phi \frac{\partial}{\partial\theta} -\cos\phi \cot\theta \frac{\partial}{\partial\phi}\right),$ (C.136)
$\displaystyle L_y$ $\displaystyle = - {\rm i} \hbar\left(\cos\phi \frac{\partial}{\partial\theta} -\sin\phi \cot\theta \frac{\partial}{\partial\phi}\right),$ (C.137)
$\displaystyle L_z$ $\displaystyle = -{\rm i} \hbar \frac{\partial}{\partial\phi},$ (C.138)

as well as

$\displaystyle L^2 = -\hbar^{ 2}\left[\frac{1}{\sin\theta}\frac{\partial}{\part...
...ight) + \frac{1}{\sin^2\theta}\frac{\partial^{ 2}}{\partial\phi^{ 2}}\right].$ (C.139)

We, thus, conclude that all of our angular momentum operators can be represented as differential operators involving the angular spherical coordinates, $ \theta$ and $ \phi$ , but not involving the radial coordinate, $ r$ .

Let us search for an angular wavefunction, $ Y_{l,m}(\theta,\phi)$ , that is a simultaneous eigenstate of $ L^{ 2}$ and $ L_z$ . In other words,

$\displaystyle L^{ 2} Y_{l,m}$ $\displaystyle = l (l+1) \hbar^{ 2} Y_{l,m},$ (C.140)
$\displaystyle L_z Y_{l,m}$ $\displaystyle =m \hbar Y_{l,m},$ (C.141)

where $ l$ and $ m$ are dimensionless constants. We also want the wavefunction to satisfy the normalization constraint

$\displaystyle \oint \vert Y_{l,m}\vert^{ 2} d{\mit\Omega} = 1,$ (C.142)

where $ d{\mit\Omega}= \sin\theta d\theta d\phi$ is an element of solid angle, and the integral is over all solid angle. Thus, we are searching for well-behaved angular functions that simultaneously satisfy,

$\displaystyle \left[\frac{1}{\sin\theta}\frac{\partial}{\partial\theta}\left( \...
...+ \frac{1}{\sin^2\theta}\frac{\partial^{ 2}}{\partial\phi^{ 2}}\right]Y_{l,m}$ $\displaystyle =-l (l+1) Y_{l,m},$ (C.143)
$\displaystyle \frac{\partial Y_{l,m}}{\partial\phi}$ $\displaystyle ={\rm i} m Y_{l,m},$ (C.144)
$\displaystyle \int_0^\pi\int_0^{2\pi}\vert Y_{l,m}\vert^{ 2} \sin\theta d\theta d\phi$ $\displaystyle = 1.$ (C.145)

As is well known, the requisite functions are the so-called spherical harmonics,

$\displaystyle Y_{l,m}(\theta,\phi) =(-1)^{ m}  \left[\frac{2 l+1}{4\pi} \frac{(l-m)!}{(l+m)!}\right]^{1/2} P_{l,m}(\cos\theta) {\rm e}^{ {\rm i} m \phi},$ (C.146)

for $ m\geq 0$ . Here, the $ P_{l,m}$ are known as associated Legendre polynomials, and are written

$\displaystyle P_{l,m}(u) = (-1)^{ l+m} \frac{(l+m)!}{(l-m)!} \frac{\left(1-u...
...m/2}}{2^{ l} l!}\left(\frac{d}{du}\right)^{l-m} \left(1-u^{ 2}\right)^{ l}.$ (C.147)

for $ m\geq 0$ . Note that

$\displaystyle Y_{l,-m} = (-1)^m Y^{ \ast}_{l,m}.$ (C.148)

Finally, the constant $ l$ is constrained to take non-negative integer values, whereas the constant $ m$ is constrained to take integer values in the range $ -l\leq m\leq +l$ . Thus, $ l$ is a quantum number that determines the value of $ L^{ 2}$ . In fact, $ L^{ 2}= l (l+1) \hbar^{ 2}$ . Likewise, $ m$ is a quantum number that determines the value of $ L_z$ . In fact, $ L_z=m \hbar$ .

The classical Hamiltonian of an extended object spinning with constant angular momentum $ {\bf L}$ about one of its principal axes of rotation is

$\displaystyle H = \frac{L^{ 2}}{2 I},$ (C.149)

where $ I$ is the associated principal moment of inertia. Let us assume that the quantum-mechanical Hamiltonian of an object spinning about a principal axis of rotation has the same form. We can solve the energy eigenvalue problem,

$\displaystyle H \psi = E \psi,$ (C.150)

by writing $ \psi(r,\theta,\phi) = R(r) Y_{l,m}(\theta,\phi)$ , where $ R(r)$ is arbitrary. It immediately follows from Equation (C.140) that

$\displaystyle E = \frac{l (l+1) \hbar^{ 2}}{2 I}.$ (C.151)

In other words, the energy levels are quantized in terms of the quantum number, $ l$ , that specifies the value of $ L^{ 2}$ .

The classical Hamiltonian of an extended object spinning about a general axis is

$\displaystyle H = \frac{L_x^{ 2}}{2 I_x} + \frac{L_y^{ 2}}{2 L_y} + \frac{L_z^{ 2}}{2 L_z},$ (C.152)

where the Cartesian axes are aligned along the body's principal axes of rotation, and $ I_x$ , $ I_y$ , $ I_z$ are the corresponding principal moments of inertia. Suppose that the body is axially symmetric about the $ z$ -axis. It follows that $ I_x=I_y$ . The previous Hamiltonian can be written

$\displaystyle H = \frac{L^{ 2}}{2 I_x}+\left(\frac{1}{2 I_z}-\frac{1}{2 I_x}\right)L_z^{ 2}.$ (C.153)

Let us assume that the quantum-mechanical Hamiltonian of an axisymmetric object spinning about an arbitrary axis has the same form as the previous Hamiltonian. We can solve the energy eigenvalue problem, $ H \psi=E \psi$ , by writing $ \psi(r,\theta,\phi) = R(r) Y_{l,m}(\theta,\phi)$ , where $ R(r)$ is arbitrary. It immediately follows from Equations (C.140) and (C.141) that

$\displaystyle E = \frac{[l (l+1)-m^{ 2}] \hbar^{ 2}}{2 I_x} + \frac{m^{ 2} \hbar^{ 2}}{2 I_z}.$ (C.154)

In other words, the energy levels are quantized in terms of the quantum number, $ l$ , that specifies the value of $ L^{ 2}$ , as well as the quantum number, $ m$ , that determines the value of $ L_z$ . If $ I_z\ll I_x$ (in other words, if the object is highly elongated along its symmetry axis) then the spacing between energy levels corresponding to different values of $ m$ becomes much greater than the spacing between energy levels corresponding to different values of $ l$ . In this case, it is plausible that the system remains in the $ m=0$ state (because it cannot acquire sufficient energy to reach the $ m=1$ state), so that

$\displaystyle E = \frac{l (l+1) \hbar^{ 2}}{2 I_x}.$ (C.155)


next up previous
Next: About this document ... Up: Wave Mechanics Previous: Simple Harmonic Oscillator
Richard Fitzpatrick 2016-01-25