-
Notifications
You must be signed in to change notification settings - Fork 1
/
Copy pathch5_addendum.tex
246 lines (219 loc) · 12.3 KB
/
ch5_addendum.tex
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63
64
65
66
67
68
69
70
71
72
73
74
75
76
77
78
79
80
81
82
83
84
85
86
87
88
89
90
91
92
93
94
95
96
97
98
99
100
101
102
103
104
105
106
107
108
109
110
111
112
113
114
115
116
117
118
119
120
121
122
123
124
125
126
127
128
129
130
131
132
133
134
135
136
137
138
139
140
141
142
143
144
145
146
147
148
149
150
151
152
153
154
155
156
157
158
159
160
161
162
163
164
165
166
167
168
169
170
171
172
173
174
175
176
177
178
179
180
181
182
183
184
185
186
187
188
189
190
191
192
193
194
195
196
197
198
199
200
201
202
203
204
205
206
207
208
209
210
211
212
213
214
215
216
217
218
219
220
221
222
223
224
225
226
227
228
229
230
231
232
233
234
235
236
237
238
239
240
241
242
243
244
245
\section{Lifting energy-level degeneracy}
The Hamiltonian for the outer electron of an Hydrogen-like atom can be written as
\begin{align*}
H = \frac{p^2}{2m} + V_\text{eff},
\end{align*}
where:
\begin{align*}
V_\text{eff} = - \frac{e^2}{4 \pi \varepsilon_0 r} \qquad &\text{for the Hydrogen atoms;} \\
V_\text{eff} = - \frac{Z_\text{eff} e^2}{4 \pi \varepsilon_0 }\left( \frac{1}{r} - \mathcal{C}(r)\right)
\qquad &\text{with $Z_\text{eff} = 1$ for the neutral alkali atoms} \\
&\text{or $Z_\text{eff} = 2$ for single-charged alkali-earth ions.}
\end{align*}
The term $\mathcal{C}(r)$ is a corrective term which is relevant for lower $l$.
The energy levels are
\begin{align*}
E_{n,l} &= - \frac{m_e e^4 Z_\text{eff}^2}{2 (4 \pi \varepsilon_0)^2 \hbar^2} \left( \frac{1}{n^2} + \tilde{\mathcal{C}}_{n,l}(r)\right) = \\
& = -\frac{m_e c^2}{2} Z_\text{eff}^2 \underbrace{\left( \frac{e^2}{4 \pi \varepsilon_0 \hbar c}\right)^2}_{\alpha^2} \left( \frac{1}{n^2} + \tilde{\mathcal{C}}_{n,l}(r) \right),
\end{align*}
where $Z_\text{eff}$ is the residual charge at the outer shell and $\alpha$ is the fine structure constant. From the last expression, one can notice that the degeneracy in $n$ and $l$ is resolved. Nevertheless in many applications it is useful to have complete resolved energy levels and to obtain this situation the quantum numbers $m$ and $s_z$ are taken into account. In particular, the effects studied to remove this degeneracy are two:
\begin{itemize}
\item \textit{Spin-orbit coupling}: it is an effect already present in the studied atoms and it derives from the fine structure;
\item \textit{Zeeman effect}: it derives from the application of an external magnetic field.
\end{itemize}
\subsection{Spin-orbit coupling}
%\begin{center}
%\input{img/oscillator_electron.tex}
%\end{center}
Consider the an Hydrogen-like atom in a laboratory frame in absence of magnetic fields and imagine that the outer electron moves with velocity $\vec{v}$. If the problem is studied from the frame in which the electron is at rest, the latter experiences an effective magnetic field
\begin{align}
\vec{B}_{el} = -\frac{1}{c^2} \vec{v} \times \vec{E},
\label{eq:B}
\end{align}
where $\vec{E}$ is the core electric field measured in the laboratory frame. Since $\vec{E}$ is radial, one can write
\begin{align*}
\vec{E} = \frac{\vec{\nabla}V}{e} = \frac{\vec{r}}{e {r}} \frac{\partial V_\text{eff}}{\partial r} ,
\end{align*}
while $\vec{v} = \vec{p}/m_e$ (in non-relativistic approximation). Therefore, equation (\ref{eq:B}) becomes
\begin{align}
\vec{B}_{el} = -\frac{1}{c^2} \left( \frac{\vec{p}}{m_e} \times \vec{r} \right) \frac{1}{e {r}} \frac{\partial V_\text{eff} }{\partial r} = \frac{1}{e m_e c^2 r} \frac{\partial V_\text{eff}}{\partial r} \vec{L},
\end{align}
where $\vec{L}$ is the angular momentum in the laboratory frame. \\
This magnetic field can couple with the magnetic dipole moment associated to the electron spin, given by
\begin{align*}
\vec{\mu}_s = -\frac{e}{2 m_e} g_s \vec{S} = -\mu_B g_s \vec{S},
\end{align*}
where $g_s$ is the electron $g$-factor ($g_s = 2$ from Dirac free theory) and $\mu_B = e\hbar/(2 m_e)$ is the Bohr magneton. In particular, the additional term to the Hamiltonian of the electron evaluated in the electron reference frame is
\begin{align}
H_\text{spin-orbit}^{el} \equiv H_{SO}^{el} = - \vec{\mu}_s \cdot \vec{B}_{el} = \frac{\mu_B g_s}{e m c^2 \hbar} \left( \frac{1}{r} \frac{\partial V_\text{eff}}{\partial r} \right)\vec{S} \cdot \vec{L}.
\end{align}
The following step consists of returning to the laboratory frame. The explicit calculation leads to the following terms
\begin{align}
H_\text{spin-orbit}^{lab} \equiv H_{SO}^{lab} = \frac{\mu_B (g_S-1)}{e m c^2 \hbar} \left( \frac{1}{r} \frac{\partial V_\text{eff}}{\partial r} \right)\vec{S} \cdot \vec{L}.
\end{align}
Using the general expression for $V_\text{eff}$ presented above, the spin-orbit Hamiltonian is
\begin{align}
H_{SO}^{lab} = \frac{\mu_B (g_S-1)Z_\text{eff} e}{4 \pi \varepsilon_0 m c^2 \hbar} \left( \frac{1}{r^3} - \frac{\partial \mathcal{C}}{\partial r} \right) \vec{S} \cdot \vec{L},
\end{align}
or more in general (omitting the reference to the laboratory frame)
$$H_{SO} = \Bigl( \textrm{some function of } r\Bigr) \; \vec{L}\cdot\vec{S}.$$
The basis formed by elements like $\ket{l,m,s,s_z}$ does not diagonalize this Hamiltonian. In order to do this, one has to introduce the \textit{total angular momentum}
\begin{align}
\vec{J} = \vec{L} + \vec{S}
\end{align}
and the new useful basis becomes $\ket{l,s,j,j_z}$ (indeed $L^2$, $S^2$, $J^2$ and $J_z$ commute). Hence from
\begin{align*}
\begin{pmatrix}
L^2 \\ L_z \\ S^2 \\ S_z
\end{pmatrix}
\ket{l,m,s,s_z} =
\begin{pmatrix}
\hbar^2 l(l+1) \\ \hbar m \\ \hbar^2 s(s+1) \\ \hbar s_z
\end{pmatrix}
\ket{l,m,s,s_z}
\end{align*}
one passes to
\begin{align*}
\begin{pmatrix}
L^2 \\ S^2 \\ J^2 \\ J_z
\end{pmatrix}
\ket{l,s,j,j_z} =
\begin{pmatrix}
\hbar^2 l(l+1) \\ \hbar^2 s(s+1) \\ \hbar^2 j(j+1) \\ \ \hbar j_z
\end{pmatrix}
\ket{l,s,j,j_z}
\end{align*}
with
\begin{align*}
j = l +\dfrac{1}{2}, l - \dfrac{1}{2} \qquad \text{and} \qquad j_z = -j, ..., j.
\end{align*}
In principle, one should evaluate the Clebsch-Gordan coefficient to move from one basis to another, but in this case a simpler approach is used. \\
The term $\vec{L} \cdot \vec{S}$ in $H_{SO}$ can be rewritten by noticing that
\begin{align*}
{J}^2 & = (\vec{L} + \vec{S})^2 = {S}^2 + {L}^2 + \vec{S}\cdot\vec{L} + \vec{L}\cdot\vec{S} = {S}^2 + {L}^2 + 2\vec{L}\cdot\vec{S},
\end{align*}
where the last equation is obtained from the fact that $\vec{L}$ and $\vec{S}$ commute because the operators act on different spaces. From this
\begin{equation*}
\vec{L}\cdot\vec{S} = \frac{1}{2}\left( {J}^2 - {L}^2 - {S}^2\right)
\end{equation*}
and hence
\begin{align}
H_{SO} \simeq \Bigl( \textrm{some function of } r\Bigr) \left( {J}^2 - {L}^2 - {S}^2\right),
\end{align}
which is clearly diagonalized by $\ket{n,l,s,j,j_z}$.
Once the new basis is found, the first-order degeneracy removing Hamiltonian is given by (\ref{eq:H1})
\begin{align*}
H^{(1)} = \Pi_{n,l,s=1/2}\, H_{SO} \, \Pi_{n,l,s=1/2}
\end{align*}
while the associated energy shift is
\begin{align*}
\Delta E_{{SO}}^{(1)} & = \bra{n,l,j,j_z} H_{SO} \ket{n,l,j,j_z} = \\
& \propto \underbrace{ \bra{n,l}{\left( \frac{1}{r}\frac{\partial V_{\text{eff}}}{\partial r} \right)}\ket{n,l} }_{\text{just an integral $>$ 0}} \left( j(j+1) - l(l+1) - \underbrace{s(s+1)}_{3/4} \right)
\end{align*}
This expression shows that some degeneracies are removed, as levels with same $l$ are now split by means of $j = l \pm \frac{1}{2}$.
\begin{center}
\input{img/split_l_spin.tex}
\end{center}
A final note is about the fact that in this procedure the relativistic effects are neglected. Nevertheless, they do not determine, at this level, a splitting but a shift of all energy levels of the order of $2\pi/\alpha^2$.
\begin{tcolorbox} [breakable, enhanced]
\textbf{Note about notation}\\
A state with quantum numbers $n, l, s, j$ is labelled as $$n^{2s+1}l_j$$
where $l$ is replaced with a capital letter according to its value. The first four values are associated to:
\begin{align*}
l=0 \rightarrow S \qquad & l=1 \rightarrow P\\
l=2 \rightarrow D \qquad & l=3 \rightarrow F
\end{align*}
For instance, $4^2S_{1/2} \Longleftrightarrow \ket{n=4,l=0,s=1/2,j=1/2}$.
\end{tcolorbox}
\begin{tcolorbox} [breakable, enhanced]
\textbf{Fine structure of Sodium} ($Z=11$) \\
A partial lifting of the degeneracy in Sodium spectrum is reported below. The energy levels $3^2 P_{1/2}$ and $3^2 P_{3/2}$ are two-time and four-time degenerate, hence the degeneracy is only partially removed.
\begin{center}
\input{img/sodium_3_transitions.tex}
\end{center}
\end{tcolorbox}
From the Sodium example it is evident that, to have a complete resolution of the energy levels, something different must be introduced.
\subsection{Zeeman effect}
Consider the system described in the previous section with an external classical magnetic field $\vec{B}$. The interaction is described by the Hamiltonian
\begin{equation*}
H_B = -\vec{\mu}_s\cdot \vec{B} -\vec{\mu}_l\cdot \vec{B} = \frac{\mu_B}{\hbar} \left( g_s \vec{S} + g_l \vec{L}\right) \cdot \vec{B}
\label{eq:HB}
\end{equation*}
where
\begin{equation*}
g_s \simeq 2.0023 \qquad \text{and} \qquad g_l \simeq \frac{m_{\textrm{reduced}}}{m_{e^-}} = \frac{1}{1 + m_e/m_\text{reduced}}\lesssim 1
\end{equation*}
are the electron spin $g$-factor and the orbital $g$-factor respectively.
In (\ref{eq:HB}), $\vec{\mu}_s$ is the spin magnetic dipole moment while $\vec{\mu}_l$ is the orbital magnetic dipole moment.
To simplify even more, one can orient $\vec B$ along the $\vec{z}$ axis and hence
\begin{equation*}
H_B = \frac{\mu_B}{\hbar} \left( 2S_z + L_z \right) B_z
\end{equation*}
\subsubsection{Case 1: Strong $\vec B$ regime}
By ``strong $B$ regime" one refers to a situation in which the energies associated to the magnetic field are much bigger than those of spin-orbit, namely $\Delta E_B \gg \Delta E_{SO}$.
In this case, $n,l,m$ and $s_z$ are good quantum numbers and the computation of the energy shift is straightforward:
\begin{align*}
\Delta E_{n,l,m,s_z} &= \bra{n,l,m,s_z} H_B \ket{n,l,m,s_z} =\\
& = \frac{\mu_B B}{\hbar} \bra{n,l,m,s_z} 2S_z + L_z \ket{n,l,m,s_z} = \mu_B (2s_z + m)B
\end{align*}
In this regime, the splitting is called \textit{Paschen-Back effect}.
\begin{tcolorbox} [breakable, enhanced]
\textbf{Paschen-Back effect for a $P$ level}
\begin{center}
\input{img/paschen-back-P.tex}
\end{center}
The degeneracy of $\ket{m,s_z} = \ket{1, -1/2}, \ket{-1, 1/2}$ is eventually removed by relativistic effects.
\end{tcolorbox}
\subsubsection{Case 2: Weak $\vec B$ regime}
On the other hand, one might consider the range of energies in which $\Delta E_B \lesssim \Delta E_{SO}$; in this case, the quantum numbers $n,l,j$ and $j_z$ are used:
\begin{align*}
\Delta E_{n,l,j,j_z} & = \bra{n,l,j,j_z} H_B \ket{n,l,j,j_z} =\\
& = \frac{\mu_B B}{\hbar} \bra{n,l,j,j_z} g_sS_z + g_lL_z \ket{n,l,j,j_z} = \\
& = \frac{\mu_B B}{\hbar} \bra{n,l,j,j_z} (g_s-g_l)S_z + g_lJ_z \ket{n,l,j,j_z} = \\
&= \frac{\mu_B B}{\hbar} \bigg[ (g_s-g_l)\bra{n,l,j,j_z}S_z\ket{n,l,j,j_z} + g_l\hbar j_z \bigg]
\end{align*}
From the last line, one notices that there is not a straightforward way to evaluate $\bra{n,l,j,j_z}S_z\ket{n,l,j,j_z}$ no eigenvector for $S_z$ are present. To compute this term one can use the \textit{projection theorem} according to which the expectation value of a quantity $\vec{A}$ that transforms as a vector can be written as
\begin{equation*}
\langle \vec A \rangle = \Bigl\langle \left(J^2\right)^{-1} \left( \vec{J}\cdot \vec{A}\right) \vec{J} \Bigr\rangle
\end{equation*}
Therefore
\begin{equation*}
\begin{aligned}
\bra{n,l,j,j_z}S_z\ket{n,l,j,j_z} & =
\bra{n,l,j,j_z}
\left({J}^2\right)^{-1}
\left( {\vec{J}\cdot\vec{S}} \right)
{J_z}
\ket{n,l,j,j_z} = \\
& = \frac{j_z}{\hbar j(j+1)} \bra{n,l,j,j_z} \left( {\vec{J}\cdot\vec{S}} \right) \ket{n,l,j,j_z} = \\
& = \frac{j_z}{2 \hbar j(j+1)} \bra{n,l,j,j_z} \left( {J}^2 + {S}^2 - {L}^2\right) \ket{n,l,j,j_z} = \\
& = \frac{j(j+1)+s(s+1)-l(l+1)}{j(j+1)} \hbar j_z
\end{aligned}
\end{equation*}
Eventually, the energy gap can be written as
\begin{equation}
\Delta E_{n,l,j,j_z} = \mu_B j_z B_z g_j(j,l)
\label{eq:degrem}
\end{equation}
where
$$g_j(j,l) = g_l + (g_s-g_l) \frac{j(j+1)-l(l+1) + 3/4}{j(j+1)}$$
since $s=1/2$.
In this case, one completely resolves the degeneracy; this result is called \textit{anomalous Zeeman effect}.
\begin{tcolorbox} [breakable, enhanced]
\textbf{Degeneracy removal for Sodium} \\
Consider the following degenerate energy levels for Sodium:
\begin{align*}
S_{1/2} \qquad &\longrightarrow \qquad g_j = 2 \\
P_{1/2} \qquad &\longrightarrow \qquad g_j = 2/3 \\
P_{3/2} \qquad &\longrightarrow \qquad g_j = 4/3
\end{align*}
These values of $g_j$ removed the degeneracy according to equation (\ref{eq:degrem}).
\begin{center}
\input{img/ex-weakB-jump.tex}
\end{center}
\end{tcolorbox}
\subsubsection{Selection rules}
A final comment is about the selection rules that must be changed. The Wigner-Eckart theorem allows to find the new selection rules:
\begin{align}
\Delta j &= 0, \pm 1 \quad \text{except when}\quad j=j'=0 \qquad \text{and} \qquad \Delta j_z &= 0, \pm 1
\end{align}