-
Notifications
You must be signed in to change notification settings - Fork 1
/
ch4.tex
688 lines (620 loc) · 41.5 KB
/
ch4.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
246
247
248
249
250
251
252
253
254
255
256
257
258
259
260
261
262
263
264
265
266
267
268
269
270
271
272
273
274
275
276
277
278
279
280
281
282
283
284
285
286
287
288
289
290
291
292
293
294
295
296
297
298
299
300
301
302
303
304
305
306
307
308
309
310
311
312
313
314
315
316
317
318
319
320
321
322
323
324
325
326
327
328
329
330
331
332
333
334
335
336
337
338
339
340
341
342
343
344
345
346
347
348
349
350
351
352
353
354
355
356
357
358
359
360
361
362
363
364
365
366
367
368
369
370
371
372
373
374
375
376
377
378
379
380
381
382
383
384
385
386
387
388
389
390
391
392
393
394
395
396
397
398
399
400
401
402
403
404
405
406
407
408
409
410
411
412
413
414
415
416
417
418
419
420
421
422
423
424
425
426
427
428
429
430
431
432
433
434
435
436
437
438
439
440
441
442
443
444
445
446
447
448
449
450
451
452
453
454
455
456
457
458
459
460
461
462
463
464
465
466
467
468
469
470
471
472
473
474
475
476
477
478
479
480
481
482
483
484
485
486
487
488
489
490
491
492
493
494
495
496
497
498
499
500
501
502
503
504
505
506
507
508
509
510
511
512
513
514
515
516
517
518
519
520
521
522
523
524
525
526
527
528
529
530
531
532
533
534
535
536
537
538
539
540
541
542
543
544
545
546
547
548
549
550
551
552
553
554
555
556
557
558
559
560
561
562
563
564
565
566
567
568
569
570
571
572
573
574
575
576
577
578
579
580
581
582
583
584
585
586
587
588
589
590
591
592
593
594
595
596
597
598
599
600
601
602
603
604
605
606
607
608
609
610
611
612
613
614
615
616
617
618
619
620
621
622
623
624
625
626
627
628
629
630
631
632
633
634
635
636
637
638
639
640
641
642
643
644
645
646
647
648
649
650
651
652
653
654
655
656
657
658
659
660
661
662
663
664
665
666
667
668
669
670
671
672
673
674
675
676
677
678
679
680
681
682
683
% -----------------------------------------------------------
%
% CHAPTER 4 of Quantum information with atoms and photons
%
% written by many authors Jan 2023
% -----------------------------------------------------------
The structure of the energy levels of a given atomic system can be different according to the resolution adopted to resolve it and which type of interactions have been considered. In particular, three structures are present:
\begin{table}[h!]
\centering
\begin{tabular}{c|c|c|c}
\toprule
$~$ \textbf{Structure} $~$ & $~$ \textbf{Energy} $~$ & $~$ \textbf{Frequency} $~$ & $~$ \textbf{Source} $~$ \\
\hline
\textit{Gross} & 1-10 eV & \makecell{$>$ 200 THz \\ (visible light)} &\makecell{nucleus-electron interaction \\ electron-electron interaction \\ and kinetic energy} \\ %\hline
\textit{Fine} & 10$^{-3}$-$10^{-2}$ eV & \makecell{$\sim$ 1 THz \\ (far infra-red)} & \makecell{spin-orbit interaction \\ (relativistic effect)}\\ %\hline
\textit{Hyperfine} & 10$^{-6}$-$10^{-5}$ eV & \makecell{$\sim$ 1 GHz \\ (micro-wave)} & \makecell{magnetic dipole coupling \\ (nucleus-electron magnetic \\ interaction)} \\
\bottomrule
\end{tabular}
\caption{Atomic energy structures whit typical energies, frequencies and source of the spectra. }
\label{tab:en_lev}
\end{table}
In the following, only the first two structures are considered.
\section{From the Hydrogen atom to the Alkali}
An \textit{hydrogenoid atomic system} (which includes hydrogen atoms and Alkali atoms) is made of two elements:
\begin{itemize}
\item a heavy attractor (the nucleus and the electrons of the inner shell) characterized by a mass $m_H$ and a momentum $\vec{p}_H$;
\item a light attractor (the outer electron) characterized by a mass $m_e$ and a momentum $\vec{p}_e$.
\end{itemize}
The Hamiltonian of the system is
\begin{align}
H = \frac{\abs{\vec{p}_H}^2}{2 m_H} + \frac{\abs{\vec{p}_e}^2}{2 m_e} + V(\abs{\vec{r}_H - \Vec{r_e}}),
\label{eq:Hatoms1}
\end{align}
Where the first two terms are the kinetic energies of the heavy and light attractors, while the potential depends only on the modulus of the distance between them (radial potential).
Moreover, for $\vec{r}_H$, $\vec{r}_e$, $\vec{p}_H$ and $\vec{p}_e$
\begin{align}
[({r}_H)_j,({p}_H)_k] &= i \hbar \delta_{jk}, \label{eq:comm1} \\
[({r}_e)_j,({p}_e)_k] &= i \hbar \delta_{jk}, \label{eq:comm2} \\
[({r}_e)_j,({p}_H)_k] &= [(\vec{r}_H)_j,(\vec{p}_e)_k] = 0. \label{eq:comm3}
\end{align}
The Hamiltonian $H$ in (\ref{eq:Hatoms1}) can be rewritten introducing the following change of coordinates:
\begin{align*}
\text{Centre-of-mass coordinate} \qquad \vec{R} &\equiv \dfrac{\vec{r}_H m_H + \vec{r}_e m_e }{m_H + m_e} \\
\text{Relative coordinate} \qquad \vec{r} &\equiv \vec{r}_e - \vec{r}_H \\
\text{Centre-of-mass momentum} \qquad \vec{P} &\equiv \vec{p}_H + \vec{p}_e \\
\text{Relative momentum} \qquad \vec{p} &\equiv \dfrac{\dfrac{\vec{p}_e}{m_e} - \dfrac{\vec{p}_H}{m_H}}{\dfrac{1}{m_e} + \dfrac{1}{m_H}}
\end{align*}
and from relations (\ref{eq:comm1}), (\ref{eq:comm2}) and (\ref{eq:comm3}), one can obtain
\begin{align}
[R_j,P_k] = i \hbar \delta_{jk}.
\end{align}
Therefore, the Hamiltonian in (\ref{eq:Hatoms1}) becomes
\begin{align}
H = \frac{|{\vec{P}}|^2}{2 m_+} + \frac{|{\vec{p}}|^2}{2 m} + V(\abs{\underbrace{\vec{r}_H - \Vec{r_e}}_{\vec{r}}}),
\label{eq:Hatoms2}
\end{align}
where
\begin{align*}
m_+ \equiv m_H + m_e \simeq m_H \qquad \text{and} \qquad \frac{1}{m} \equiv \left( \frac{1}{m_H} + \frac{1}{m_e} \right) \simeq \frac{1}{m_e}
\end{align*}
are the total and reduced mass, respectively. \\
In order to rearrange again (\ref{eq:Hatoms2}), it is necessary to introduce the \textit{orbital angular momentum operator}.
\subsection{Orbital angular momentum operator}
The angular momentum is a pseudo-vector defined as
\begin{equation}
\Vec{L} = \Vec{r} \, \cross \, \Vec{p} = \begin{pmatrix}
r_yp_z - r_zp_y \\
r_zp_x - r_xp_z \\
r_xp_y - r_yp_x \\
\end{pmatrix},
\label{eq:Ldef}
\end{equation}
where the general $i$-th component is given by
\begin{align}
L_i = -i\hbar \hat{r}\cross \nabla = -i \hbar \left(x_j \frac{\partial}{\partial x_k} - x_k \frac{\partial}{\partial x_j}\right) \qquad i,j,k = x,y,z.
\label{eq:Li}
\end{align}
The operator $L$ satisfies the commutation rules
\begin{equation}
[L_i,L_j]=i \hbar \epsilon_{ijk} L_k.
\label{eq:commu1}
\end{equation}
\begin{tcolorbox} [breakable, enhanced]
\textbf{Explicit evaluation of $[L_x,L_y]$}
\begin{align*}
[L_x,L_y] &= -\hbar^2\bigl((y\partial_z - z \partial_y)(z \partial_x - x \partial_z) - (z \partial_x - x \partial_z)(y \partial_z - z \partial_y)\bigl) = \\
&= -\hbar^2\bigl[\bigl(y\partial_z(z \partial_x) - y \partial_z (x \partial_z) -
z \partial_y (z \partial_x) + z \partial_y (x \partial_z) \bigl) + \\
& \qquad -\bigl(z \partial_x (y \partial_z) - z \partial_x (z \partial_y) - x \partial_z (y \partial_z ) + x \partial_z (z \partial_y)\bigl) \bigl] = \\
&= -\hbar^2\bigl[\bigl(y \partial_x (\partial_z z) + yz \partial_z \partial_x - yx \partial^2_z - z^2 \partial_y \partial_x + zx \partial_y \partial_z \bigl) = \\
& \qquad - \bigl(zy \partial_x \partial_z - z^2 \partial_x \partial_y - xy \partial_z^2 + x \partial_y (\partial_z z) + xz \partial_x \partial_y \bigl) \bigl] = \\
&= \hbar^2 ( x \partial_y - y \partial_x ) = \hbar^2 \frac{L_z}{-i\hbar} = i\hbar L_z
\end{align*}
\end{tcolorbox}
Relation (\ref{eq:commu1}) shows that $\vec{L}$:
\begin{itemize}
\item has the mathematical structure of a Lie algebra and the $\epsilon_{ijk}$ are its structure constants;
\item plays a central role in quantum problems involving rotational symmetry and its most general and fundamental definition is the \textit{generator of rotations}
\begin{align*}
R(\theta,\varphi,\chi) = \exp{\frac{i}{\hbar}L_z \chi} \exp{\frac{i}{\hbar}L_x \theta} \exp{\frac{i}{\hbar}L_z \varphi},
\end{align*}
where $\theta$, $\varphi$ and $\chi$ are the conventional Euler angles.
\end{itemize}
The other essential commutator rule is
\begin{equation}
[L_i,L^2] = 0, \qquad \forall i = x,y,z.
\label{eq:commu2}
\end{equation}
\begin{tcolorbox}[breakable, enhanced]
\textbf{Explicit evaluation of $[L_z,L^2]$}
\begin{align*}
[L_z,L^2] &= [L_z,L_x^2] + [L_z,L_y^2] = \\
&= [L_z,L_x]L_x + L_x[L_z,L_x] + [L_z,L_y]L_y + L_y[L_z,L_y] = \\
&= i\hbar (L_y L_x + L_x L_y - L_xL_y - L_y L_x) = 0
\end{align*}
\end{tcolorbox}
If one fixes the $z$-axis as the reference direction (defining suitable spherical coordinates), relation (\ref{eq:commu2}) ensures that there is a common basis of eigenstates for $L_z$ and $L^2$. These eigenstates $\psi$ are functions of two labels $l$ and $m$ such that
\begin{align}
L_z \psi(l,m) = \hbar m \, \psi(l,m) \qquad \text{and} \qquad L^2 \psi(l,m) = f(l) \, \psi(l,m),
\label{eq:eigenvaleq}
\end{align}
where $f(l)$ is a function of $l$.
Note that $\psi(l,m)$ could have some degeneracy if only $l$ and $m$ are specified. However, this will not occur for the orbital angular momentum as we will show that its eigenfunctions are spherical harmonics that form a full base of $L^2$.
\subsubsection{Raising and lowering operators}
At this point, it is convenient to define the \textit{raising} and \textit{lowering operators}
\begin{equation}
L_\pm = L_x \pm i L_y
\label{eq:lpm}
\end{equation}
with commutation rules
\begin{equation}
[L_+,L_-] = 2 \hbar \, L_z \qquad \text{and} \qquad [L_z,L_\pm] = \pm \hbar \, L_\pm.
\end{equation}
\begin{tcolorbox}[breakable, enhanced]
\textbf{Explicit evaluation of $[L_+,L_-]$}
\begin{align*}
[L_+,L_-] & = L_+ L_- - L_- L_+ = (L_x + iL_y)(L_x - iL_y) - (L_x - iL_y)(L_x + iL_y) = \\
& = (L_x^2 -i L_xL_y +i L_yL_x + L_y^2) - (L_x^2 + iL_xL_y - iL_yL_x + L_y^2) = \\
& = -2i [L_x,L_y] = 2\hbar L_z
\end{align*}
\end{tcolorbox}
\begin{tcolorbox}[breakable, enhanced]
\textbf{Explicit evaluation of $[L_z,L_\pm]$}
\begin{align*}
[L_z,L_\pm] &= L_z(L_x \pm iL_y)-(L_x \pm iL_y)L_z = \\
& = L_zL_x\pm iL_zL_y-L_xL_z\mp iL_yL_z = \\
& = i\hbar L_y \pm \hbar L_x = \pm \hbar L_\pm
\end{align*}
\end{tcolorbox}
From (\ref{eq:lpm}), one can notice that $L_\pm$ are not Hermitian since $(L_\pm)^\dagger = L_\mp$. In addition, it is possible to write
\begin{align}
L^2 = L_x^2 + L_y^2 + L_z^2 = L_- L_+ + L_z^2 + \hbar L_z = L_+ L_- + L_z^2 - \hbar L_z
\label{eq:L2}
\end{align}
and to conclude that
\begin{align}
[L^2,L_\pm] = 0.
\end{align}
\begin{tcolorbox} [breakable, enhanced]
\textbf{Proof of (\ref{eq:L2})} \\
The terms $L_- L_+$ and $L_+ L_-$ are evaluated explicitly:
\begin{align*}
L_- L_+ &= (L_x - i L_-)(L_x + i L_-) = L_x^2 + i L_x L_y - i L_y L_x + L_y^2 = \\
&= L_x^2 + L_y^2 +i [L_x,L_y] = L_x^2 + L_y^2 - \hbar L_z, \\
L_+ L_- &= (L_x + i L_-)(L_x - i L_-) = L_x^2 - i L_x L_y + i L_y L_x + L_y^2 = \\
& = L_x^2 + L_y^2 -i [L_x,L_y] = L_x^2 + L_y^2 + \hbar L_z.
\end{align*}
From these result, relation (\ref{eq:L2}) follows immediately.
\end{tcolorbox}
\begin{tcolorbox} [breakable, enhanced]
\textbf{Explicit evaluation of $[L^2,L_\pm]$}
\begin{align*}
[L^2,L_\pm] &= [L_+L_-+L_z^2-\hbar L_z,L_\pm] = \\
& = [L_+L_-,L_\pm] + [L_z^2,L_\pm] - \hbar[L_z,L_\pm] = \\
&= [L_+,L_\pm]L_- + L_+[L_-,L_\pm] + [L_z,L_\pm]L_z + L_z [L_z,L_\pm] - \hbar [L_z,L_\pm].
\end{align*}
For clarity, it is now convenient to compute separately the cases $L_+$ and $L_-$:
\begin{align*}
[L^2,L_+] &= [L_+,L_+]L_- + L_+[L_-,L_+] + [L_z,L_+]L_z + L_z [L_z,L_+] - \hbar [L_z,L_+] = \\
&= -2\hbar L_+L_z + \hbar L_+L_z + \hbar L_zL_+ - \hbar^2 L_+ = \\
&= \hbar(L_zL_+-L_+L_z)-\hbar^2L_+ = \\
&= \hbar[L_z,L_+] - \hbar^2 L_+ = \hbar^2 (L_+ - L_+) = 0, \\
[L^2,L_-] &= [L_+,L_-]L_- + L_+[L_-,L_-] + [L_z,L_-]L_z + L_z [L_z,L_-] - \hbar [L_z,L_-] = \\
& = 2\hbar L_zL_- - \hbar L_-L_z - \hbar L_zL_- + \hbar^2 L_- = \\
& = -\hbar(L_zL_--L_-L_z)+\hbar^2L_+ = \\
& = -\hbar[L_z,L_-] + \hbar^2 L_- = -\hbar^2 (L_- - L_-) = 0.
\end{align*}
\end{tcolorbox}
It is now possible to study the effect of $L_z$ and $L^2$ on a state $\psi(l,m)$ on which acts $L_\pm$:
\begin{align}
L^2 \left( L_\pm \psi (l,m) \right) &= L_\pm \left( L^2 \psi (l,m) \right) = f(l) \left( L_\pm \psi (l,m) \right) \label{eq:prop1} \\
L_z \left( L_\pm \psi (l,m) \right) &= L_\pm L_z \psi (l,m) + [L_z, L_\pm] \, \psi (l,m) = \nonumber \\
&= \hbar m \left( L_\pm \psi (l,m) \right) \pm \hbar \left( L_\pm \psi (l,m) \right) = \nonumber \\
&= \hbar (m \pm 1) \left( L_\pm \psi (l,m) \right) \label{eq:prop2}
\end{align}
From the result in (\ref{eq:prop1}), one can conclude that $L_\pm \psi (l,m)$ is an eigenstate of $L^2$ and also that it does not impact on his eigenvalues. While from (\ref{eq:prop2}) it is possible to identify two cases:
\begin{itemize}
\item \textbf{Case 1} \\
Equation (\ref{eq:prop2}) is trivially satisfied if $L_\pm \psi (l,m) = 0$. This case suggests that there must exist two values of $m$, $m_\text{max}$ and $m_\text{min}$, such that
\begin{align}
L_+\psi(l,m_\text{max}) = 0 \qquad \text{and} \qquad L_- \psi(l,m_\text{min}) = 0.
\end{align}
From this
\begin{align*}
L^2 \psi(l,m_\text{max}) &= (L_- L_+ + L_z^2 + \hbar L_z) \, \psi(l,m_\text{max}) = \\
&= L_-(L_+\psi(l,m_\text{max})) + L_z^2\psi(l,m_\text{max}) +\hbar L_z \psi(l,m_\text{max}) = \\
&= \hbar^2 m_\text{max} (m_\text{max} + 1) \, \psi(l,m_\text{max}) = f(l) \, \psi(l,m_\text{max}), \\
L^2 \psi(l,m_\text{min}) &= (L_+L_- + L_z^2 - \hbar L_z) \, \psi(l,m_\text{min}) = \\
&= L_+(L_-\psi(l,m_\text{min})) + L_z^2\psi(l,m_\text{min}) -\hbar L_z \psi(l,m_\text{min}) = \\
&= \hbar^2 m_\text{min} (m_\text{min} - 1) \, \psi(l,m_\text{min}) = f(l) \, \psi(l,m_\text{min}).
\end{align*}
Notice that the results presents the same function $f(l)$ because the function $\psi$ has the same $l$ value in both the calculations. Therefore,
\begin{equation}
f(l) = \hbar^2 m_\text{max} (m_\text{max} + 1) = \hbar^2 m_\text{min} (m_\text{min} - 1), \qquad m_\text{max} - m_\text{min} \in \mathbb{N}
\label{eq:sistmaxmin}
\end{equation}
\begin{tcolorbox}
\textbf{Solution of (\ref{eq:sistmaxmin})} \\
The general solution can be found by introducing a number $a \in \mathbb{N}$ in such a way that
\begin{align*}
\begin{cases}
m_\text{min} = m_\text{max} - a \\
m_\text{max} (m_\text{max} + 1) = m_\text{min} (m_\text{min} - 1)
\end{cases}
\end{align*}
By substituting the first equation in the second one
\begin{align*}
& m_\text{max}^2 + m_\text{max} - (m_\text{max} - a)^2 + (m_\text{max} - a) = \\
& = m_\text{max}^2 + m_\text{max} - m_\text{max}^2 -a^2 +2 m_\text{max}a + m_\text{max} - a = 0
\end{align*}
and hence
\begin{align*}
2m_\text {max}(a+1) = a(a+1) \qquad & \implies \qquad m_\text{max} = \frac{a}{2}, \\
& \implies \qquad m_\text{min} = -\frac{a}{2}.
\end{align*}
\end{tcolorbox}
This implies that $m_\text{max}$ and $m_\text{min}$ can only be integer or half integer: $m_\text{max}$, $m_\text{min} \in \dfrac{\mathbb{N}}{2}$.
From now on, consider
\begin{align*}
l \equiv m_\text{max} \qquad \implies \qquad & f(l) = \hbar^2 l(l+1) \qquad \text{and} \qquad l \in \frac{\mathbb{N}}{2}
\end{align*}
\item \textbf{Case 2} \\
Since the eigenvalue of equation (\ref{eq:prop2}) is $\hbar(m \pm 1)$, $L_\pm \psi (l,m)$ must be proportional to a an eigenfunction $\psi'(m,l+1)$ associated to that eigenvalue. Therefore, using the bra-ket notation, one can write
\begin{align}
& L_+ \ket{l,m} = \alpha \, \ket{l,m + 1} \qquad \text{with} \qquad \alpha \in \mathbb{C} \label{eq:L+a} \\
& L_- \ket{l,m+1} = \beta \, \ket{l,m} \qquad \text{with} \qquad \beta \in \mathbb{C} \label{eq:L-a}
\end{align}
and notice that
\begin{align*}
\beta = \bra{l,m} L_- \ket{l,m+1} = \left( \bra{l,m+1} L_+ \ket{l,m} \right)^* = \alpha^*.
\end{align*}
Therefore, using this result and the expression (\ref{eq:L2}) for $L^2$,
\begin{align*}
L_- L_+ \ket{l,m} &= \abs{\alpha}^2 \ket{l,m} = (L^2 - L_z^2 - \hbar L_z) \ket{l,m} = \\
& = \hbar^2 [l(l+1) - m^2 - m)] \ket{l,m} \\
& = \hbar^2 [l(l+1) - m(m+1)] \ket{l,m} \\
L_+ L_- \ket{l,m+1} &= \abs{\beta}^2 \ket{l,m+1} = (L^2 - L_z^2 + \hbar L_z) \ket{l,m+1} = \\
& = \hbar^2 [l(l+1) - m^2 - m)] \ket{l,m+1} \\
& = \hbar^2 [l(l+1) -m(m+1)] \ket{l,m+1}
\end{align*}
If $\alpha$ and $\beta$ are real, one can conclude that
\begin{align}
\alpha = \beta = \hbar \sqrt{l(l+1) - m(m+1)}
\label{eq:alphabeta}
\end{align}
Hence, unifying (\ref{eq:L+a}), (\ref{eq:L-a}) and (\ref{eq:alphabeta})
\begin{align}
L_+ \ket{l,m} &= \hbar \sqrt{l(l+1) - m(m+1)} \ket{l,m + 1} \label{eq:L+action}\\
L_- \ket{l,m} &= \hbar \sqrt{l(l+1) - m(m-1)} \ket{l,m -1} \label{eq:L-action}
\end{align}
The action of $L_+$ and $L_-$ on a state $\ket{l,m}$ is shown in figure \ref{fig:lowrai}.
\begin{figure}[h!]
\centering
\includegraphics[width=0.35\linewidth]{images/Lowering_Raising_operators.png}
\caption{Action of the lowering and raising operators $L_-$ and $L_+$; they lower and raise by 1 the quantum number $m$ of the state on which act.}
\label{fig:lowrai}
\end{figure}
\end{itemize}
Summarizing the results
\begin{align}
& L^2 \psi(l,m) = \hbar^2 l(l+1) \psi(l,m) \qquad \text{with} \qquad l \in \frac{\mathbb{N}}{2} \\
& L_z \psi(l,m) = \hbar m \psi(l,m), \qquad \text{with} \qquad m \in \, \{-l, \, ..., \, +l \}
\end{align}
It is important to notice that these rules follow only from the Lie algebra of the angular momentum operators; by concerning the orbital angular momentum defined in (\ref{eq:Ldef}), it is possible to deduce further restrictions on $l$. Indeed, introducing
\begin{align*}
r_1 = \frac{r_x+p_y}{\sqrt{2}}, \qquad p_1 = \frac{r_x-p_y}{\sqrt{2}}, \qquad r_2 = \frac{p_x-r_y}{\sqrt{2}} \qquad \text{and} \qquad p_2 = \frac{p_x+r_y}{\sqrt{2}}
\end{align*}
the operator $L_z$ (from (\ref{eq:Li})) becomes
\begin{align*}
L_z = r_x p_y - r_y p_x = \frac{1}{2}\left(r_1^2 + p_1^2\right) - \frac{1}{2}\left(r_2^2+p_2^2\right).
\end{align*}
It resembles the Hamiltonian of two decoupled harmonic oscillators ($r_1$, $p_1$, $r_1$ and $r_2$ satisfy the same commutation rules) and therefore one can deduce the form of the energy spectrum and write
\begin{align*}
L_z \, \psi(l,m) = \hbar m \, \psi(l,m) = \hbar \left[ \left(n_1 +\frac{1}{2}\right) - \left(n_2 + \frac{1}{2}\right)\right] \psi(l,m) = \hbar (n_1-n_2) \, \psi(l,m).
\end{align*}
Since $n_1, n_2 \in \mathbb{N}$, therefore $m \equiv n_1 - n_2 \in \mathbb{N}$ and also $l \in \mathbb{N}$.
\subsubsection{$L$ in spherical coordinates}
The orbital angular momentum operator can be written in spherical coordinates; indeed, if one introduces three variables
\begin{align*}
r \in [0,\infty), \qquad \theta \in [0,\pi) \qquad \text{and} \qquad \varphi \in [0, 2\pi)
\end{align*}
and the versors
\begin{align*}
\hat{r} = \begin{pmatrix} \sin{\theta} \cos{\varphi} \\ \sin{\theta} \sin{\varphi} \\ \cos{\theta} \end{pmatrix}, \qquad \hat{\theta} = \begin{pmatrix} \cos{\theta} \cos{\varphi} \\ \cos{\theta} \sin{\varphi} \\ -\sin{\theta} \end{pmatrix} \qquad \text{and} \qquad \hat{\varphi} = \begin{pmatrix} -\sin{\varphi} \\ \cos{\varphi} \\ 0 \end{pmatrix},
\end{align*}
the gradient becomes
\begin{align}
\vec{\nabla} = \hat{r} \frac{\partial}{\partial r} + \hat{\theta} \frac{1}{r} \frac{\partial}{\partial \theta} + \hat{\varphi} \frac{1}{r \sin{\theta}} \frac{\partial}{\partial\varphi}
%\label{eq:laplacian_spher}
\end{align}
and $\vec{L}$ can be written starting form (\ref{eq:Ldef})
\begin{align*}
\vec{L} &= r \, \hat{r} \cross (-i \hbar) \vec{\nabla} = \\
&= -i\hbar \left( \hat{r} \cross \hat{r} r \frac{\partial}{\partial r} + \hat{r} \cross \hat{\theta} r \frac{1}{r} \frac{\partial}{\partial \theta} + \hat{r} \cross \hat{\varphi} r \frac{1}{r \sin{\theta}} \frac{\partial}{\partial\varphi} \right) = \\
&= -i \hbar \left( \hat{\varphi} \frac{\partial}{\partial \theta} - \hat{\theta} \frac{1}{r \sin{\theta}} \frac{\partial}{\partial\varphi} \right).
\end{align*}
From this expression, one can conclude that the orbital angular momentum operator acts only on a sphere surface because there is no dependence on $r$.
Also the component of $\vec{L}$ can be expressed in spherical coordinates
\begin{align*}
L_x = \vec{L} \cdot \vec{x} &= -i \hbar \left( \hat{x} \cdot \hat{\varphi} \frac{\partial}{\partial \theta} - \hat{x} \cdot \hat{\theta} \frac{1}{\sin{\theta}} \frac{\partial}{\partial \varphi}\right) = i\hbar \left( \sin{\varphi} \frac{\partial}{\partial \theta} + \cos{\theta} \cos{\varphi} \frac{\partial}{\partial \varphi} \right), \\
L_y = \vec{L} \cdot \vec{y} &= -i \hbar \left( \hat{y} \cdot \hat{\varphi} \frac{\partial}{\partial \theta} - \hat{y} \cdot \hat{\theta} \frac{1}{\sin{\theta}} \frac{\partial}{\partial \varphi}\right) = i\hbar \left( -\cos{\varphi} \frac{\partial}{\partial \theta} + \cos{\theta} \cos{\varphi} \frac{\partial}{\partial \varphi} \right), \\
L_z = \vec{L} \cdot \vec{z} &= -i \hbar \left( \hat{z} \cdot \hat{\varphi} \frac{\partial}{\partial \theta} - \hat{z} \cdot \hat{\theta} \frac{1}{\sin{\theta}} \frac{\partial}{\partial \varphi}\right) = i\hbar \frac{\partial}{\partial \varphi},
\end{align*}
and similarly for the operators $L_\pm$
\begin{align*}
L_+ &= L_x + i L_y = i \hbar (\cos{\varphi} + i \sin{\varphi}) \left( \cot{\theta} \frac{\partial}{\partial \varphi} -i \frac{\partial}{\partial \theta} \right) = i\hbar e^{i \varphi} \left( \cot{\theta} \frac{\partial}{\partial \varphi} -i \frac{\partial}{\partial \theta} \right), \\
L_- &= L_x - i L_y = i \hbar (\cos{\varphi} - i \sin{\varphi}) \left( \cot{\theta} \frac{\partial}{\partial \varphi}+ i \frac{\partial}{\partial \theta} \right) = i\hbar e^{-i \varphi} \left( \cot{\theta} \frac{\partial}{\partial \varphi}+i \frac{\partial}{\partial \theta} \right).
\end{align*}
With these results and by using the relation in (\ref{eq:L2}) it is possible to obtain the expression for $L^2$ in spherical coordinates:
\begin{align*}
L^2 = -\hbar^2 \left[\frac{1}{\sin{\theta}} \frac{\partial}{\partial \theta} \left( \sin{\theta} \frac{\partial}{\partial \theta} \right) + \frac{1}{\sin^2{\theta}} \frac{\partial^2}{\partial \varphi^2}. \right]
\end{align*}
If one considers the Laplacian operator
\begin{align}
\nabla^2 = \frac{1}{r^2} \frac{\partial}{\partial r} \left( r^2 \frac{\partial}{\partial r}\right) + \frac{1}{r^2 \sin{\theta}} \frac{\partial}{\partial \theta} \left( \sin{\theta} \frac{\partial}{\partial \theta} \right) + \frac{1}{r^2 \sin^2{\theta}} \frac{\partial^2}{\partial\varphi^2},
\label{eq:laplacian_spher}
\end{align}
it is possible to notice that it can be written in terms of $L^2$ as
\begin{align*}
\nabla^2 = \frac{1}{r^2} \frac{\partial}{\partial r} r^2 \frac{\partial}{\partial r}+ \frac{L^2}{r^2}.
\end{align*}
These considerations are useful to obtain an important expression for a generic Hamiltonian, indeed
\begin{align}
H = \frac{p^2}{2m}+V(\Vec{r}) = -\frac{\hbar^2}{2m} \nabla^2 + V(\Vec{r}) = -\frac{\hbar^2}{2m}\frac{1}{r^2}\frac{\partial}{\partial r} r^2 \frac{\partial}{\partial r} + \frac{L^2}{2m r^2} + V(\vec{r}).
\label{eq:Hatoms3}
\end{align}
When the potential is a function of just $\abs{\vec{r}}$ ($V(\vec{r}) = V(r)$), $H$ can be decomposed in its radial and angular parts and hence the eigenfunctions $\psi(r,\theta,\varphi)$ can be written as
\begin{align}
\psi_{n,l,m}(r,\theta,\varphi) = R_{n,l}(r) \, Y_{l,m}(\theta,\varphi),
\label{eq:psi_compl}
\end{align}
where $n$ is an additional quantum number which is discussed in the following section.
\subsubsection{Angular part of $\psi(r,\theta,\varphi)$}
Firstly, consider only the angular part of $H$ in (\ref{eq:Hatoms3}) and search for the functions $Y_{l,m}(\theta,\varphi)$, called \textit{spherical harmonics}, that give the angular distribution of the states $\ket{l,m}$; since $H$ commutes with $L^2$ and $L_z$, $Y_{l,m}(\theta,\varphi)$ are also eigenstates of $L_z$:
\begin{equation*}
L_z \, Y_{l,m}(\theta,\varphi) = \hbar m Y_{l,m}(\theta,\varphi).
\end{equation*}
Given the expression for $L_z$ in spherical coordinates, the solution can be written in the form
\begin{equation}
Y_{l,m}(\theta, \varphi) = P_l^m(\theta) e^{im\varphi},
\end{equation}
where $P_l^m(\theta)$ is the Legendre polynomial associated to $Y_{l,m}(\theta, \varphi)$. To find the explicit expression of $P_l^m$, one can start from the case $m = l$ and remember that:
\begin{equation*}
L_+ \, Y_{l,l} = 0 \quad \implies \quad \left( \cot{\theta} \frac{\partial}{\partial \varphi} - i \frac{\partial}{\partial \theta}\right) e^{il\varphi} P_l^l (\theta) = \left( l \cot{\theta} - \frac{\partial}{\partial \theta} \right) P_l^l(\theta)= 0.
\end{equation*}
This is a differential equation which has an unique solution for $l \in \mathbb{N}$:
\begin{equation}
P_l^l(\theta)=\sin^l{(\theta)}.
\end{equation}
From this expression, one can evaluate all the other spherical harmonics action of the lowering operator reported in (\ref{eq:L-action}):
\begin{equation*}
Y_{l,m-1}(\theta, \varphi) = \frac{1}{\sqrt{l(l+1)-m(m-1)}} L_- \, Y_{l,m}(\theta, \varphi).
\end{equation*}
\begin{tcolorbox}[breakable, enhanced]
\textbf{Some spherical harmonics} \\
The first few spherical harmonics are:
\begin{itemize}
\item \textbf{$l = 0$}
\begin{align*}
Y_{0,0} &= \sqrt{\frac{1}{4\pi}}
\end{align*}
\item \textbf{$l = 1$}
\begin{align*}
Y_{1,-1} &= \sqrt{\frac{3}{8\pi}} \sin{\theta} e^{-i\varphi} \\
Y_{1,0} &= \sqrt{\frac{3}{4\pi}} \cos{\theta} \\
Y_{1,1} &= -\sqrt{\frac{3}{8\pi}} \sin{\theta)} e^{i\varphi}
\end{align*}
\item \textbf{$l = 2$}
\begin{align*}
Y_{2,-2} &= \frac{1}{4} \sqrt{\frac{15}{2\pi}} \sin^2{\theta} e^{-2i\varphi} \\
Y_{2,-1} &= \frac{1}{2} \sqrt{\frac{15}{2\pi}}\sin{\theta} \cos{\theta} e^{-i\varphi} \\
Y_{20} &= \frac{1}{4} \sqrt{\frac{5}{\pi}} (3 \cos^2{\theta}-1) \\
Y_{2,1} &= -\frac{1}{2} \sqrt{\frac{15}{2\pi}}\sin{\theta} \cos{\theta} e^{i\varphi} \\
Y_{2,2} &= \frac{1}{4} \sqrt{\frac{15}{2\pi}} \sin^2{\theta} e^{2i\varphi}
\end{align*}
\end{itemize}
The normalization condition is:
\begin{equation*}
\int Y^*_{l,m} \, Y_{l',m'} \, \sin{\theta} \, d\theta d\varphi = \delta_{l,l'} \delta_{m,m'}.
\end{equation*}
The solutions with $l=0$ are called \textit{s orbitals} (which stands for ``sharp", as they have no degeneracy), those with $l=1$ are called \textit{p orbitals} (which stands for ``principal", as they correspond to the brightest lines in the spectra) and those with $l=2$ are called \textit{d orbitals} (which stands for ``diffuse", due to their high degeneracy).
\end{tcolorbox}
\subsubsection{Radial part of $\psi(r,\theta,\varphi)$}
The function $\psi_{n,l,m}(r,\theta,\varphi)$, as written in (\ref{eq:psi_compl}), comprehends also a radial part, which depends on the quantum numbers $n$ and $l$. In order to study it, consider the Hamiltonian $H$ in (\ref{eq:Hatoms3}) and take into account the fact that $Y_{l,m}$ is an eigenstate of $L^2$ with eigenvalue $\hbar^2 l(l+1)$; with these consideration, the Schr\"odinger equation can be written in the form:
\begin{align*}
H \psi_{n,l,m}(r,\theta,\varphi) &= \left(-\frac{\hbar^2}{2m}\frac{1}{r^2}\frac{\partial}{\partial r}\left(r^2 \frac{\partial}{\partial r}\right) + V(r)+ \frac{\hbar^2 l(l+1)}{2m r^2}\right)Y_{l,m}(\theta,\varphi) R_{n,l}(r) = \\
&= E \, \psi_{n,l,m}(r,\theta,\varphi).
\end{align*}
It is customary to introduce a function $P_{n,l}(r) = r \, R_{n,l}(r)$ and to obtain a modified radial Schr\"odinger equation
\begin{equation}
\left(-\frac{\hbar^2}{2m} \frac{\partial^2}{\partial r ^2}+ V_\text{eff}(r)\right) P_{n,l}(r)= E \, P_{n,l}(r),
\label{eq:eqP}
\end{equation}
where the $V_\text{eff}$ contains both the radial potential and the centrifugal effect given by orbital angular momentum.
\subsection{Spectrum of the Hydrogen atom}
Equation (\ref{eq:eqP}) is a one-dimensional differential equation and it is similar to the one obtained for the Hydrogen atom if
\begin{align}
V_\text{eff}(r) = \frac{\hbar^2 l(l+1)}{2 m r^2} - \frac{Z_\text{eff} e^2}{4 \pi \epsilon_0 r},
\end{align}
in which $Z_\text{eff}$ represents the charge (in units of $e$) of the attractor part and this case is $Z_\text{eff} = 1$. The behaviour of this potential is reported in figure \ref{fig:eff_potential}, while the solutions of the discussed Schr\"odinger equation constitute the energy levels reported in the spectrum in figure (\ref{fig:hydro_spectrum}).
\begin{figure}[h!]
\centering
\includegraphics[width=0.45\linewidth]{images/Effective_potential.png}
\caption{Behaviour or the effective potential; the repulsive part for low values of $r$ is associated to the angular component, the attractive part for high values of $r$ is associated to the electric part.}
\label{fig:eff_potential}
\end{figure}
\begin{figure}[h!]
\centering
\includegraphics[width=0.5\linewidth]{images/Hydrogen_spectrum.png}
\caption{First levels of the Hydrogen electron spectrum with the relative energies and orbital names.}
\label{fig:hydro_spectrum}
\end{figure}
\subsection{Alkali atoms}
\begin{figure}[t]
\centering
\includegraphics[width=0.58\linewidth]{images/Alkali_potential.png}
\caption{Potential felt by the outer electron of a Sodium atom; for small distances an additional attractive term with respect to the Hydrogen potential is present and the total potential is shifted down, while for large distances the electron sees only the potential of a Hydrogen core.}
\label{fig:pot_Alkali}
\end{figure}
\begin{figure}[b]
\centering
\includegraphics[width=0.61\linewidth]{images/Alkali_spectrum.png}
\caption{First levels of the Sodium electron spectrum with the relative energies and orbital names; also the reference energies from the Hydrogen spectrum are reported.}
\label{fig:alkali_spectrum}
\end{figure}
Hydrogen atoms are not used in everyday applications because they are light and volatile at temperatures little higher than the absolute zero. To avoid this, heavier atoms, like Alkali and in particular $^{11} \text{Na}_{\sim 22}$, are used.
\noindent The Hamiltonian associated to these atoms is
\begin{equation}
H_\text{Alkali} = \sum_{j} \left( - \frac{\hbar^2}{2m} \nabla_j^2 - \frac{Ze^2}{4 \pi \epsilon_0 \left| \Vec{r}_j\right|}\right) + \sum_{j\neq j'} \frac{e^2}{4 \pi \epsilon_0 \left| \Vec{r}_j - \Vec{r}_{j'}\right|},
\label{eq:alkali_hamiltonian}
\end{equation}
where the summations are performed over the electrons. It can be be rewritten in many better ways, through some approximations. One could express $H_\text{Alkali}$ using the \textit{Hartree-Fock approximation}, but the entanglement information would be lost, due to the fact that this approximation provides the use of the \textit{mean field}.
Another possibility is the \textit{Shell model approximation}, which is a harsher approximation with respect to the previous one: the idea is that every electron sees only the core screened by the previous electrons. The behaviour of this potential for the outer electron of Sodium atom is reported in figure \ref{fig:pot_Alkali}. This model also gives a good explanation of the energy shift for atomic orbitals observed in alkali atoms with respect to the Hydrogen levels (figure \ref{fig:alkali_spectrum}); indeed, shifting momentarily towards a semi-classical picture with the electron as a proper particle orbiting around the nucleus, one could imagine that two processes take place:
\begin{itemize}
\item as the electron gets closer to the nucleus, the electrostatic potential is stronger and the energy shift more pronounced;
\item electrons with higher $l$ number experience a stronger centrifugal potential, bringing them farther away from the nucleus, resulting in a weaker electrostatic attraction and smaller energy shift.
\end{itemize}
\section{Interaction of Alkali atoms with light}
The interaction between charged particles and the electromagnetic field is governed by the Lorentz force $\Vec{F_L}=q(\Vec{E}+\Vec{v}\cross\Vec{B})$. The Lagrangian of a system where this is the only force in play is
\begin{equation*}
L(\Vec{r}, \Dot{\Vec{r}}, t)=\frac{1}{2} m \Dot{r}^2 + q \Dot{\Vec{r}}\cdot\Vec{A}- q \phi,
\end{equation*}
where the last part is equivalent to $J_\mu A^\mu$ in relativistic notation. Moving to the Hamiltonian view, since the canonical momentum is defined as
\begin{equation*}
p_i=\frac{\partial L}{\partial \Dot{\Vec{r}}} \qquad \implies \qquad \Vec{p}=m\Dot{\Vec{r}}+q\Vec{A}.
\end{equation*}
The Hamiltonian in this \textit{minimal coupling} case can thus be written as
\begin{equation}
H= \frac{(\Vec{p}-q\Vec{A})^2}{2m}+q\phi,
\label{eq:Hatomlight}
\end{equation}
where both the part relative to the motion of the atom and the one associated to the interaction with light are present.
\subsection{Power-Zienau transformation}
When using light at wavelengths such that $k^{-1}\gg a_0$ (where $a_0$ is the Bohr radius, i.e. the average of the electron-nucleus distance), as is the case with visible light,
\begin{align*}
\Vec{A}(\Vec{r}, t) \qquad \longrightarrow \qquad \Vec{A}({0}, t).
\end{align*}
This means that if the light wavelength is much larger than the interaction region (approximated by $a_0$), the vector potential is expected to vary little enough to be approximated with its central value. Within this approximation, which is called \textit{long-wavelength approximation}, $\Vec{A}$ commutes with momentum $\vec{p}$, since it no longer depends on position, i.e.
\begin{align*}
\left[\vec{A}(\vec{r}),\, \vec{p} \right] \neq 0 \qquad \qquad \left[\vec{A}(0),\, \vec{p} \right] = 0.
\end{align*}
% Power-Zienam transformation
The atom-light Hamiltonian (\ref{eq:Hatomlight}) can be rewritten by defining an unitary operator, the \textit{Power-Zienau transformation}
\begin{equation*}
T = \exp{-\frac{i}{\hbar} q \, \Vec{r}\cdot\Vec{A}(0)}
\end{equation*}
and by using the using Hadamard's lemma
\begin{align*}
e^{-S} \, B \,e^S = B+[S,B]+\frac{1}{2} [S,[S,B]]+...~
\end{align*}
to evaluate $H_{AL}'= T \, H_{AL} \, T^\dagger$. The first step consists of evaluating
\begin{align*}
\Vec{p}\,'&=T\,\Vec{p}\,\,T^\dagger=\Vec{p}-\frac{i}{\hbar} q \left[\Vec{r}\cdot\Vec{A}(0), \, \Vec{p}\right] - \frac{q^2}{2\hbar^2} \left[\Vec{r}\cdot\Vec{A}(0),\left[\Vec{r}\cdot\Vec{A}(0),\Vec{p}\right]\right]+... = \\
&=\Vec{p}-\frac{i}{\hbar} q \Vec{A}(0) \, \left[\Vec{r},\,\Vec{p}\right]=\Vec{p}+q\Vec{A},
\end{align*}
given that $[\Vec{r}, \, \Vec{p}]=i\hbar \mathbb{1}$ and thus all higher-order commutators are null. Therefore,
\begin{equation*}
T (\Vec{p}-q\Vec{A})^2 T^\dagger = T (\Vec{p}-q\Vec{A}) T^\dagger \, T (\Vec{p}-q\Vec{A}) T^\dagger = (\vec{p} + q \vec{A} - q \vec{A}) \cdot \vec{p} = p^2,
\end{equation*}
because $\vec{A}'(0) = T \vec{A}(0) T^\dagger = \vec{A}(0)$. Since $\phi'(\vec{r}) = T \phi(\vec{r}) T^\dagger = \phi(\vec{r})$, the final expression for $H_{AL}$ is
\begin{align}
H'_{AL} = T H_{AL} T^\dagger = \frac{p^2}{2m} + q \phi.
\end{align}
\subsection{Dipole approximation}
\label{dipole}
The same transformation can be applied to the Hamiltonian of the electromagnetic field $H_\text{light} \equiv H_{EM}$, treated in the previous chapter. In particular, using the Hadamard's relation
\begin{align*}
H'_{EM} &= T H_{EM} T^\dagger = \\
&= H_{EM} - \frac{i}{\hbar} q \left[\vec{r} \cdot \vec{A}(0), \, H_{EM} \right] - \frac{q^2}{2\hbar^2} \left[ \vec{r} \cdot \vec{A}(0), \, \left[ \vec{r} \cdot \vec{A}(0), H_{EM} \right]\right] + ... = \\
&= H_{EM} - \frac{i}{\hbar} {r_i} \left[ {A_i}(0), H_{EM}\right] - \frac{q^2}{2\hbar^2} r_i \, r_j \left[ A_i(0), \left[ A_j(0), H_{EM} \right] \right] + ...,
\end{align*}
where the repeated indexes indicate a summation over the three components. Introducing the \textit{dipole operator} $\Vec{d}\equiv q\Vec{r},$ $H'_{EM}$ becomes
\begin{equation}
H'_{EM} = H_{EM} - \frac{i}{\hbar} d_i [A_i(0), H_{EM}]- \frac{1}{2\hbar^2} d_i d_j [A_i(0),[A_j(0), H_{EM}]]+ ...
\label{eq:Hem1}
\end{equation}
The commutator in the previous expression is evaluated inserting the relation (\ref{eq:potvec}) for $\Vec{A}$ with $u(\Vec{r} = 0) \approx 1$:
\begin{align*}
\left[\Vec{A}(0), H_{EM} \right] &= \sum_{\vec{k},\lambda} \sqrt{\frac{\hbar}{2 \varepsilon_0 \omega_k}} \frac{\vec{\epsilon}_\lambda}{\sqrt{V}} \left[ \left(a_{\vec{k},\lambda} + a_{\vec{k},\lambda}^\dagger\right), H_{EM}\right] = \\
&= \sum_{\vec{k},\lambda} \sqrt{\frac{\hbar}{2 \varepsilon_0 \omega_k}} \frac{\vec{\epsilon}_\lambda}{\sqrt{V}} \left[ \left(a_{\vec{k},\lambda} + a_{\vec{k},\lambda}^\dagger\right), \sum_{\vec{k}',\lambda'} \hbar \omega_{k'} a_{\vec{k}',\lambda}'^\dagger a_{\vec{k}',\lambda'} \right] = \\
&= \sum\limits_{\substack{\Vec{k},\lambda, \\ \Vec{k}', \lambda'}}
\sqrt{\frac{\hbar}{2\epsilon_0\omega_k}} \frac{\Vec{\epsilon}_\lambda}{\sqrt{V}} \hbar\omega_{k'} \left(\left[a_{\Vec{k},\lambda}^\dagger, n_{\Vec{k'},\lambda'}\right]+\left[a_{\Vec{k},\lambda}, n_{\Vec{k'},\lambda'}\right]\right)=\\
&=\hbar \sum_{\Vec{k}, \lambda} \sqrt{\frac{\hbar\omega_k}{2\epsilon_0}}~ \frac{\Vec{\epsilon}_{\Vec{k}}}{\sqrt{V}} \left(a_{\Vec{k},\lambda}-a_{\Vec{k},\lambda}^\dagger\right)= -i\hbar \Vec{E}(0),
\end{align*}
where the commutators are evaluated as
\begin{align*}
\left[a_{\Vec{k},\lambda}^\dagger, n_{\Vec{k'},\lambda'}\right] &= -a_{\vec{k},\lambda}^\dagger \delta_{\vec{k},\vec{k}'} \delta_{\lambda,\lambda'} \\
\left[a_{\Vec{k},\lambda}, n_{\Vec{k'},\lambda'}\right] &= a_{\vec{k},\lambda} \delta_{\vec{k},\vec{k}'} \delta_{\lambda,\lambda'}
\end{align*}
Then (\ref{eq:Hem1}) becomes
\begin{align}
H'_{EM} = H_{EM} - d_i \, E_i(0) - \frac{i}{2 \hbar} d_i d_j \left[ A_i(0), E_j(0)\right] \equiv H_{EM} - \vec{d} \cdot \vec{E}(0) + H_{dd}.
\end{align}
\\
%where other terms vanish since $[A_i,E_j] = 0$ for $i \neq j$ because
%\begin{align*}
% [\Vec{A},\Vec{E}] \propto \left[a_{\Vec{k},\lambda} + a_{\Vec{k},\lambda}^\dagger, a_{\Vec{k'},\lambda'}-a_{\Vec{k'},\lambda'}^\dagger\right] \sim -2 \left[a_{\Vec{k},\lambda}, a_{\Vec{k},\lambda}^\dagger\right] \delta_{\Vec{k}, \Vec{k'}}\delta_{\lambda, \lambda'} \sim -2 \delta_{\Vec{k}, \Vec{k'}}\delta_{\lambda, \lambda'}
%\end{align*}
%is no longer dependent on the field operators, the higher commutators are all zero; moreover, the second term only operates on the atom degrees of freedom via the dipole operators, resulting in a shift of the atom levels.
The complete Hamiltonian which takes into account the light and the light-matter interaction contributions after the considered transformation is
\begin{equation}
H = \frac{p^2}{2m}+ q\phi +H_{dd} + H_{EM} - \Vec{d}\cdot\Vec{E}(0) \equiv H_\text{atom} + H_\text{light} + H_\text{atom-light}
\label{eq:Hamfin}
\end{equation}
The first three terms define the total atom Hamiltonian, whose eigenstates are written as $\ket{n,l,m}$, the fourth is the light contribution, with eigenstates given by the Fock states, and the last is the first-order approximation for the atom-light interaction Hamiltonian. Higher orders can be obtained using further terms in the multipole expansion for $\Vec{A}$, i.e.
\begin{align*}
\Vec{A}(\Vec{r}) = \Vec{A}(\Vec{0})+ r_i \frac{\partial}{\partial x_i} \Vec{A}+\frac{1}{2} r_i r_j \frac{\partial^2}{\partial x_i \partial x_j} \Vec{A}+ ...
\end{align*}
\subsection{Selection rules}
Some considerations can be done about the dipole operator $\vec{d} = q \vec{r}$ which enters in the Hamiltonian for light-atom interaction.
\subsubsection{Diagonal elements}
Under the \textit{parity transformation} $\Tilde{R}$ defined as
\begin{align*}
\Tilde{R}\,\psi(x,y,z)&=\psi(-x, -y, -z) \\
\Tilde{R}\psi(r, \theta, \varphi)&=\psi(r, \pi-\theta, \pi+\varphi),
\end{align*}
with $\Tilde{R}^\dagger=\Tilde{R}^{-1}=\Tilde{R}$, one has
\begin{align*}
\Tilde{R}\, \Vec{r} \, \Tilde{R}^\dagger =-\Vec{r} \qquad &\Longleftrightarrow \qquad \{ \Tilde{R},\, \vec{r} \} = 0, \\
\Tilde{R}\, \Vec{p} \, \Tilde{R}^\dagger =-\Vec{p} \qquad &\Longleftrightarrow \qquad \{ \Tilde{R}, \, \vec{p} \} = 0, \\
\Tilde{R}\, \Vec{L} \, \Tilde{R}^\dagger =\vec{L} \qquad &\Longleftrightarrow \qquad [ \Tilde{R}, \, \vec{p} ] = 0. \\
\end{align*}
Therefore, $\vec{L}$, $L^2$ and $L_z$ share the eigenbasis $\ket{n,l,m}$; in particular, the eigenvalues for $\Tilde{R}$ are
\begin{align}
\Tilde{R}\ket{n,l,m} = (-1)^l \ket{n,l,m} \equiv e^{i \phi_{l,m}} \ket{n,l,m}
\end{align}
Therefore,
\begin{align*}
\bra{n,l,m}\Vec{r}\ket{n,l,m} &= \bra{n,l,m} \Tilde{R}^\dagger \Tilde{R} \Vec{r} \ket{n,l,m} = \\
&= - \left( \bra{n,l,m} \tilde{R}^\dagger \right) \vec{r} \left( \tilde{R} \ket{n,l,m}\right) =\\
&= -\bra{n,l,m} \vec{r} \ket{n,l,m},
\end{align*}
from which one deduces that the diagonal elements must all be zero.
\begin{equation*}
\bra{n,l,m}{\Vec{r}}\ket{n,l,m}=0.
\end{equation*}
\subsubsection{Selection rules on $l$}
From the canonical commutators one can derive the identity
\begin{equation*}
\left[L^2,\left[L^2, \vec{r}\right]\right]=2\hbar^2 \bigl\{\vec{r}, L^2\bigr\}.
\end{equation*}
Multiplying this by $\bra{n_1,l_1,m_1}$ on the left and by $\ket{n_2,l_2,m_2}$ on the right of both sides and bringing everything to the left one gets:
\begin{align*}
\hbar^4\Bigl((l_1(l_1+1))^2-2l_1(l_1+1)l_2(l_2+1)+(l_2(l_2+1))^2~+\\
-~2l_1(l_1+1)-2l_2(l_2+1)\Bigr)\bra{n_1,l_1,m_1}\Vec{r}\ket{n_2,l_2,m_2} = 0,
\end{align*}
which can be factorized in
\begin{equation*}
(l_1-l_2+1)(l_1-l_2-1)(l_1+l_2)(l_1+l_2+2)\bra{n_1,l_1,m_1}\Vec{r}\ket{n_2,l_2,m_2} = 0.
\end{equation*}
The last two parenthesis are irrelevant since they can not be null, hence from the first two one has that the only case in which the matrix element can be nonzero is when $\Delta l= l_1 - l_2 =\pm1$. This is a necessary, not sufficient, condition.
\subsubsection{Selection rules on $m$}
In order to obtain a similar relation for the quantum number $m$, one can assume that the electric field is along the $z-$axis and two cases are considered:
\begin{itemize}
\item \textbf{Case $\vec{\epsilon} \parallelslant \hat{z}$} \\
Since $[L_z, r_z]=0$, one immediately gets
\begin{equation*}
\bra{n_1,l_1,m_1}[L_z,\Vec{r}]\ket{n_2,l_2,m_2} =
0 = \hbar(m_1-m_2)\bra{n_1,l_1,m_1}\Vec{r}\ket{n_2,l_2,m_2} \;\;.
\end{equation*}
Thus the matrix element can be nonzero when $\Delta m = m_1 - m_2 =0$.
\item \textbf{Case $\vec{\epsilon} \perp \hat{z}$} \\
Since $[L_z, r_x ]= i\hbar y$ and $[L_z, r_y]=-i\hbar x$, which implies $[L_z, x\pm iy]=\pm\hbar(x\pm iy)$, one gets
\begin{align*}
\bra{n_1,l_1,m_1} \left( [L_z, x\pm iy] \mp \hbar(x\pm iy) \right) \ket{n_2,l_2,m_2} = 0
\end{align*}
from which
\begin{align*}
\hbar(m_1-m_2-1)\bra{n_1,l_1,m_1}(x+iy)\ket{n_2,l_2,m_2} &= 0, \\
\hbar(m_1-m_2+1)\bra{n_1,l_1,m_1}(x-iy)\ket{n_2,l_2,m_2} &= 0.
\end{align*}
Thus the matrix element can be null when the coefficients are null, hence $\Delta m = m_1 - m_2 = \pm 1$. It can be shown that the $+1$ case corresponds to left-handed circularly polarized light, while the $-1$ case corresponds to right-handed light.
\end{itemize}
An example of the application of this selection rules is given in figure (\ref{fig:sodium_spectrum}), in which the allowed transitions for the lowest level of Sodium spectrum are reported.
\begin{figure}[h]
\centering
\includegraphics[width=0.7\linewidth]{images/Sodium_transition.png}
\caption{First levels of the Sodium electron spectrum with the allowed transitions and the relative wavelengths.}
\label{fig:sodium_spectrum}
\end{figure}