-
Notifications
You must be signed in to change notification settings - Fork 1
/
Copy pathch5.tex
849 lines (774 loc) · 51.5 KB
/
ch5.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
684
685
686
687
688
689
690
691
692
693
694
695
696
697
698
699
700
701
702
703
704
705
706
707
708
709
710
711
712
713
714
715
716
717
718
719
720
721
722
723
724
725
726
727
728
729
730
731
732
733
734
735
736
737
738
739
740
741
742
743
744
745
746
747
748
749
750
751
752
753
754
755
756
757
758
759
760
761
762
763
764
765
766
767
768
769
770
771
772
773
774
775
776
777
778
779
780
781
782
783
784
785
786
787
788
789
790
791
792
793
794
795
796
797
798
799
800
801
802
803
804
805
806
807
808
809
810
811
812
813
814
815
816
817
818
819
820
821
822
823
824
825
826
827
828
829
830
831
832
833
834
835
836
837
838
839
840
841
842
843
844
845
846
847
848
849
% -----------------------------------------------------------
%
% CHAPTER 5 of Quantum information with atoms and photons
%
% written by many people, Jan 2023
% -----------------------------------------------------------
\section{Two-level atoms and coherent light}
\label{sec:atom_coherent_light}
Consider an system with an atom with only two energy levels, $\ket{g}$ and $\ket{e}$, separated by the quantity $\hbar\omega_{eg}$. A generic state of the system is given by the tensor product of the states in the two subspaces; in particular:
\begin{align*}
\ket{\psi_\text{atom}} &\equiv \ket{\psi_A} \in \{\ket{g}, \ket{e}\} \\
\ket{\psi_\text{light}} &\equiv \ket{\psi_L} = \ket{\psi_{k_1}} \otimes \ket{\psi_{k_2}} \otimes ... \otimes \ket{\psi_{k_L}} \otimes ... \\
\ket{\psi} &= \ket{\psi_A} \otimes \ket{\psi_L}
\end{align*}
For the light state, one can then assume to have all the modes in the vacuum state except for one, leading to the state $\ket{\psi_L} = \ket{0_{k_1}} \otimes \ket{0_{k_2}} \otimes ... \otimes \ket{\psi_{k_L}} \otimes ... $
Then it is possible to assume that the state $\ket{\psi_{k_L}}$ is a coherent state $\ket{\alpha_{k_L}}$, with the following properties:
\begin{itemize}
\item there are many photons, which means mathematically that $\left<\hat{n}_{k_L}\right> \equiv \Bar{n} \gg 1$;
\item the electric field is classical, i.e. with negligible fluctuations, which means that its average value is simply an oscillating function $\langle \hat{\Vec{E}}\rangle= \Vec{\mathcal{E}} \text{cos}(\omega_{k_L} t)$, where $\omega_{k_L} \equiv \omega$ is the only frequency of the problem (this can be derived also from the first propriety).
\end{itemize}
From these assumptions, the contributions to the Hamiltonian can be rewritten as:
\begin{align}
H_\text{atom} &= \hbar \omega_{eg} \ket{e}\bra{e} \\
H_\text{light} &= \hbar \omega_c \hat{a}^\dagger \hat{a} \\
H_\text{atom-light} &= - \Vec{\mathcal{E}} \text{cos}(\omega t) \cdot \left[ \Vec{d}_{eg} \ket{e}\bra{g} + \Vec{d}^*_{eg} \ket{g}\bra{e} \right]
\end{align}
where $\Vec{d}_{eg} = \bra{e} \Vec{d} \ket{g}$. One can also introduce the quantity
\begin{align}
\Omega \equiv \frac{1}{\hbar} \, \Vec{d}_{eg} \cdot \Vec{\mathcal{E}},
\end{align}
called \textit{Rabi frequency}. After this considerations, the complete Hamiltonian becomes
\begin{equation}
H(t) = \hbar \omega_{eg} \ket{e}\bra{e} + \hbar \Omega \cos(\omega t) \left[ \ket{e}\bra{g} + \text{h.c.} \right]
\label{eq:Hatom_light}
\end{equation}
In section (\ref{subsec:1.5.2}) the problem of a driven qubit with $\omega \gg \omega_{eg}$ (for both the weak drive case $\omega \gg \Omega$ and the strong drive case $\omega \simeq \Omega$) is solved, but a resolution is possible also for the case $\omega \simeq \omega_{eg}$. This regime is generally indicated with \textit{resonant light condition}. \\
To deal with this situation, it is useful to introduce the \textit{detuning}
\begin{align}
\delta = \omega_{eg} - \omega.
\end{align}
Hence three energy scales are present and they are associated to three frequencies: $\omega$, $\delta$ and $\Omega$.
%The resonant light condition brings to $\omega \gg \delta$, but we make the further assumption, later justified, that $\omega \gg \Omega$.
\subsection{Rotating wave approximation in the high-frequency regime}
Consider a system in which $\omega \gg \delta, \Omega$. To solve the dynamics, one has to go into a new convenient frame, in which the separation of energy scales is visible, which means to find mathematically a new Hamiltonian $$H'(t)= H_0 + V(t).$$
This Hamiltonian can be found going into the \textit{rotating frame}, described by the operator $$R = e^{\text{i} \omega t \ket{e} \bra{e}},$$ from which the new Hamiltonian is given by equation (\ref{eq:transfH}). \\
One can expand the $R$ operator from its definition to find a more convenient representation:
\begin{align}
R &= {\mathbb{1} + {i} \omega t \ket{e} \bra{e} + \frac{({i} \omega t)^2}{2 !} \underbrace{(\ket{e} \bra{e})^2}_{\ket{e} \bra{e}} + \sum_{n>2} \frac{(\text{i} \omega t)^n}{n !} \underbrace{(\ket{e} \bra{e})^n}_{\ket{e} \bra{e}} } = \\
&= \mathbb{1} + \left( {i} \omega t + \frac{({i} \omega t)^2}{2 !} + \sum_{n>2} \frac{({i} \omega t)^n}{n !} \right) \ket{e} \bra{e} = \\
= & \mathbb{1} + \left( - 1 + \underbrace{ 1 + {i} \omega t + \frac{({i} \omega t)^2}{2 !} + \sum_{n>2} \frac{({i} \omega t)^n}{n !} }_{e^{\text{i} \omega t}} \right) \ket{e} \bra{e} = \\
= & \mathbb{1} + \left( e^{{i} \omega t} -1 \right) \ket{e} \bra{e}.
\end{align}
We write the contributions to $H'(t)$, i.e.
\begin{align*}
R \ket{e} \bra{g} R^{\dagger} &= \left( \mathbb{1} + e^{\text{i} \omega t} -1\right) \ket{e} \bra{g} \left[ \mathbb{1} + \left( e^{\text{i} \omega t} -1 \right) \ket{e} \bra{e}\right]^{\dagger} = \\
&= e^{\text{i} \omega t} \ket{e} \bra{g} \\
R \ket{g} \bra{e} R^{\dagger} &= \left[ \mathbb{1} + \left( e^{\text{i} \omega t} -1 \right) \ket{e} \bra{e}\right] \ket{g} \bra{e} \left( \mathbb{1} + e^{- \text{i} \omega t} -1\right) = \\
&= e^{- \text{i} \omega t} \ket{g} \bra{e} \\
R\Dot{R}^{\dagger} = & \left[ \mathbb{1} + \left( e^{{i} \omega t} -1 \right) \ket{e} \bra{e}\right] \left[ {i} \omega e^{{i} \omega t} \ket{e} \bra{e}\right]^{\dagger} = \\
&= \left[ -{i} \omega e^{ - i \omega t} - \text{i} \omega + {i} \omega e^{ - {i} \omega t} \right] \ket{e} \bra{e} = \\
&= - {i} \omega \ket{e} \bra{e}
\end{align*}
and eventually
\begin{align*}
H'(t) &\equalexpl{\ref{eq:transfH}} R(t) H(t) R^\dagger(t) - i \hbar R(t) \frac{dR^\dagger(t)}{dt} =\\
&= \hbar\omega_{eg} \ket{e}\bra{e} + \hbar\Omega\cos(\omega t)e^{i\omega t} \ket{e}\bra{g} + \text{h.c.} -i\hbar(-i\omega\ket{e}\bra{e})=\\
&= \hbar \delta \ket{e}\bra{e} + \left[\hbar \Omega \cos(\omega t) e^{{i} \omega t} \ket{e}\bra{g} + \text{h.c.} \right] = \\
&= \hbar \delta \ket{e}\bra{e} + \left[\hbar \Omega \frac{e^{ {i} \omega t} + e^{ - {i} \omega t}}{2} e^{ \text{i} \omega t} \ket{e}\bra{g} + \text{h.c.} \right] = \\
&= \hbar \delta \ket{e}\bra{e} + \left[ \frac{\hbar \Omega}{2} \left(1 + e^{ 2 {i} \omega t} \right) \ket{e}\bra{g} + \text{h.c.} \right]
\end{align*}
This Hamiltonian can be rewritten in terms of the time-dependent part $V(t)$ and the time-independent one $H_0$, with
\begin{align}
H_0 &= {\hbar \delta \ket{e}\bra{e} + \frac{\hbar \Omega}{2} \left[\ket{e}\bra{g} + \ket{g}\bra{e} \right]} \\
V(t) & = {\frac{\hbar \Omega}{2} \left[e^{2 {i} \omega t} \ket{e}\bra{g} + e^{- 2 {i} \omega t} \ket{g}\bra{e} + \text{h.c.} \right]}.
\end{align}
To solve the problem, one can use the results obtained in section \ref{sec:time_dep_sys} in the case of a system with a weak drive in the high-frequency regime; this corresponds to the initial assumption $\omega \gg \delta, \Omega$. In particular, it is possible to deduce the final Hamiltonian using (\ref{eq:final_ham}):
\begin{equation}
\label{eq:Hreslight}
H'= H_0 + \frac{1}{\hbar \omega} \left[V, V^{\dagger}\right] = \hbar \delta \ket{e}\bra{e} + \frac{\hbar \Omega}{2} \left[\ket{e}\bra{g} + \ket{g}\bra{e} \right]
\end{equation}
The term $\left[V, V^{\dagger}\right]$ is responsible for the micro-motion of the system, due to the fact that it goes like $\Omega^2/\omega$. Equation (\ref{eq:Hreslight}) is the final Hamiltonian in the new frame for the resonant light regime ($\omega \gg \delta$, or $\omega \simeq \omega_{eg}$) and in the high-frequency regime ($\omega \gg \Omega$). It is important to notice that we have taken into account only two frequencies ($\Omega$ and $\delta$) in order to study the dynamics of the system. This is the so-called \textit{Rotating-wave approximation}. \\
A final comment is about the fact that, in order to have a simpler notation, one can introduce the following \emph{jump operators}, responsible of the jump from a level to another:
\begin{align}
\sigma_+ \equiv \ket{e} \bra{g} \qquad \text{and} \qquad \sigma_- \equiv \ket{g} \bra{e}.
\end{align}
Therefore, the Hamiltonian in the rotating frame with $\omega \gg \delta, \Omega$ is
\begin{align}
\label{eq:Hrw}
H = \hbar \delta \ket{e}\bra{e} + \frac{\hbar \Omega}{2} \sigma_+ + \frac{\hbar \Omega}{2} \sigma_-.
\end{align}
\section{Jaynes-Cummings model for a two-level atom in a cavity}
\begin{figure}[h]
\centering
\includegraphics[width=0.65\linewidth]{images/AtomCavity.png}
\caption{Atom inside the cavity in a quantum regime.}
\label{fig:atomcavity}
\end{figure}
\subsection{Single-mode cavity}
Consider a cavity with only one mode of the electromagnetic field (\textit{single-mode approximation}) with frequency $\omega_c$ and an atom with only two energy levels; the motion of its center of mass motion can be ignored. The system is presented in figure (\ref{fig:atomcavity}) and the Hamiltonian of this system is made of
\begin{align*}
H_\text{light} & = H_{EM} = \hbar \omega_c \crt{} \dsr{} \\
H_\text{atom} & = H_A = \hbar \omega_{eg} \ket{e}\bra{e} \\
H_\text{atom-light} & = H_{AL}= -\hat{\Vec{d}} \cdot \hat{\Vec{E}}\\
\end{align*}
Using the expression (\ref{eq:efFP}) for the electric field obtained for a Fabry-Perot cavity for a single mode with frequency $\omega_c = c k_c $
\begin{align*}
\hat{\Vec{E}} = -i \Vec{\epsilon} \sqrt{\frac{\hbar \omega_c}{\varepsilon_0 V}} \sin{(k_c z)} \left( \hat{a}^\dagger - \hat{a} \right),
\end{align*}
the contribution of the atom-light interaction becomes
\begin{align*}
H_{AL} & = - \left( \Vec{d}_{eg} \ket{e}\bra{g} + \Vec{d}^*_{eg} \ket{g}\bra{e} \right) \cdot \left[-i \Vec{\epsilon} \sqrt{\frac{\hbar \omega_c}{\varepsilon_0 V}} \sin(k_c z) \left(\crt{} - \dsr{}\right)\right] = \\
&= i \dfrac{\hbar}{2} (\tilde{g} \sigma_{+} + \tilde{g}^{\ast} \sigma_{-}) \left(\crt{} - \dsr{}\right)
\end{align*}
where the notation previously introduced is used and
\begin{equation*}
\tilde{g} = \frac{2\Vec{d}_{eg}^* \cdot \Vec{\epsilon}}{\hbar} \sqrt{\frac{\hbar \omega_c}{\varepsilon_0 V}} \sin(k_c z)
\end{equation*}
is called \textit{vacuum Rabi frequency}. \\
The total Hamiltonian becomes
\begin{equation}
H = \hbar \omega_c \crt{} \dsr{} + \hbar \omega_{eg} \ket{e}\bra{e} + \frac{i \hbar}{2} (\tilde{g} \sigma_{+} +\tilde{g}^{\ast} \sigma_{-}) \left(\crt{} - \dsr{}\right).
\end{equation}
Consider a situation near resonance regime ($\omega_c \simeq \omega_{eg}$); we know that it is possible to simplify the problem using the rotating-wave approximation, as seen in the previous section. This approximation leads to the elimination of the terms $\sigma_{+} \crt{}$ and $\sigma_{-} \dsr{}$ in $H_{AL}$. More in detail, one can move to a new frame where the energy is conserved through the transformation
\begin{equation}
R = \underbrace{\exp{{i \omega_{eg} \ket{e}\bra{e} t}}}_{\text{\footnotesize atom}} \, \underbrace{\exp{i \omega_c t \hat{a}^{\dagger} \hat{a}}}_{\text{\footnotesize cavity}} \equiv R_A \otimes R_{EM}
\end{equation}
Using equation (\ref{eq:transfH}), the new Hamiltonian for the atom-light interaction is
\begin{align}
H'_{AL} = \frac{i \hbar}{2} \Tilde{g} \, R \sigma_{+} \hat{a}^{\dagger} R^{\dagger} - \frac{i \hbar}{2} \Tilde{g} \, R \sigma_{+} \hat{a} R^{\dagger} + \text{h.c.}
\label{eq:atom-light-new}
\end{align}
The first term gives
\begin{align*}
R \sigma_{+} \hat{a}^{\dagger} R^{\dagger} &= {R_A \sigma_{+} R_A^{\dagger}} \otimes R_{EM} \hat{a}^{\dagger} R_{EM}^{\dagger} = \\
&= e^{i \omega_{eg} t } \sigma_{+} \left[ e^{i \omega_c t \hat{a}^{\dagger} \hat{a}} \hat{a}^{\dagger} e^{-i \omega_c t \hat{a}^{\dagger} \hat{a}} \right]
\end{align*}
and we recognize
\begin{align*}
e^{i \omega_c t \hat{a}^{\dagger} \hat{a}} \hat{a}^{\dagger} e^{-i \omega_c t \hat{a}^{\dagger} \hat{a}} \equiv \hat{a}_H^{\dagger}(t)
\end{align*}
which is the creation operator in the Heisenberg picture for a free harmonic oscillator. Since one can write
\begin{align*}
\hat{a}_H^{\dagger}(t) = \hat{a}_H^{\dagger}(0) \, e^{i \omega_c t},
\end{align*}
then
\begin{align}
R \sigma_{+} \hat{a}^{\dagger} R^{\dagger} = e^{i (\omega_{eg} + \omega_c) t} \sigma_{+} \hat{a}_H^{\dagger}(0)
\end{align}
In the same way, one obtains the second term of Eq. \ref{eq:atom-light-new}
\begin{equation}
R \sigma_{+} \hat{a} R^{\dagger} = e^{i (\omega_{eg} - \omega_c ) t} \sigma_{+} \hat{a}_H(0)
\end{equation}
Finally, in the in the rotating frame the new Hamiltonian for the atom-light interaction becomes
\begin{equation*}
H'_{AL}(t) = \underbrace{\frac{i \hbar}{2} \Tilde{g} {e^{i (\omega_{eg} + \omega_c) t}} \sigma_{+} \hat{a}^{\dagger}}_{V^{AL}(t)} + \underbrace{\frac{i \hbar}{2} \Tilde{g}
e^{i (\omega_{eg} - \omega_c )t} \sigma_{+} \hat{a} }_{H^{AL}_0} ~+~ \text{h.c.}
\end{equation*}
One can notice that the term $H_0$ is independent of time when $\omega_c \simeq \omega_{eg}$, while the term $V(t)$ goes like $e^{2 i \omega_c t}$. Thus,
\begin{equation*}
H'_{AL}(t) = \frac{i \hbar}{2} \left[ \tilde{g} e^{2 i \omega_c t} \sigma_+ \hat{a}^\dagger + \tilde{g} \sigma_{+} \hat{a} + \text{h.c.} \right].
\end{equation*}
The last step consists of going back to the original frame; this leads to
\begin{align*}
H_{AL} = -\frac{i \hbar}{2} \tilde{g} \sigma_+ \hat{a} + \frac{i \hbar}{2} \tilde{g}^* \sigma_- \hat{a}^\dagger,
\end{align*}
from which one can notice that the terms $\sigma_+ \hat{a}^\dagger$ and $\sigma_- \hat{a}$ do not contribute to the dynamics.
Finally, the rotating-wave approximation is equivalent to the condition
\begin{align*}
\abs{\omega_{eg} + \omega_c} \gg \abs{\omega_{eg} - \omega_c}, \tilde{g}
\end{align*}
and the total Hamiltonian, which describes the so-called \textbf{Jaynes-Cummings model}, becomes
\begin{align}
H = \hbar \omega_c \hat{a}^\dagger \hat{a} + \hbar \omega_{eg} \ket{e}\bra{e} -\frac{i \hbar}{2} \tilde{g} \sigma_+ \hat{a} + \frac{i \hbar}{2} \tilde{g}^* \sigma_- \hat{a}^\dagger.
\end{align}
\\
An alternative form to the model can be obtained by redefining $\Tilde{g} = g e^{i \phi}$, hence
\begin{equation*}
-i \Tilde{g} \sigma_{+} \hat{a} = g \sigma_{+} \left(-i e^{i\phi} \hat{a} \right) \equiv {g} \sigma_+ \hat{a}',
\end{equation*}
where a unitary transformation allows to introduce $\hat{a}' = -i e^{i\phi} \hat{a}$ and $\hat{a}'^\dagger = i\hat{a}^\dagger e^{-i \phi}$. This implies that $\hat{a}^\dagger \hat{a} = \hat{a}'^\dagger \hat{a}'$ and hence
\begin{equation} \label{eq:JCModel}
H = \hbar \omega_c \hat{a}'^\dagger \hat{a}' + \hbar \omega_{eg} \ket{e}\bra{e} + \frac{g \hbar}{2} ( \sigma_+ \hat{a}' + \sigma_- \hat{a}'^\dagger) \qquad \text{with} \qquad g \in \mathbb{R}.
\end{equation}
This is the model for an atom interacting with a single-mode, nearly resonant ($\omega_{eg} \simeq \omega_{c}$) cavity mode within the rotating-wave approximation, ignoring any dissipation process such as spontaneous emission or any input or output from the cavity.
\subsection{Multi-mode cavity}
To treat a multi-mode, nearly resonant cavity, it is sufficient to consider:
\begin{align*}
\sigma_{+} a_n^{\dagger} \quad & \longrightarrow \quad e^{i(\omega_{eg} + n \omega_c) t} \sigma_{+} a^{\dagger} \\
\sigma_{+} a_n \quad & \longrightarrow \quad e^{i(\omega_{eg} - n \omega_c) t} \sigma_{+} a
\end{align*}
In this case only $n=1$ would give some relevant terms, the others would be negligible.
\subsection{Dynamics of the Jaynes-Cummings model}
The next step consists of solving the Jaynes-Cummings model; the physics field studying this kind of systems is called \textit{cavity quantum electro-dynamics}.
In order to study the dynamics of the system, one can start from a basis formed with the eigenstates of
\begin{equation*}
H_0 = \hbar \omega_c \hat{a}'^\dagger \hat{a}' + \hbar \omega_{eg} \ket{e}\bra{e}.
\end{equation*}
They are
\begin{align*}
\ket{g}_A \otimes \ket{n}_{EM} \equiv \ket{g, n} \qquad \text{and} \qquad \ket{e}_A \otimes \ket{n}_{EM} \equiv \ket{e, n},
\end{align*}
where $\ket{n}$ is the $n$-photon state in the cavity.
Taking the atom ground energy as reference value for the zero, the eigenvalues of the eigenstates are given by:
\begin{align*}
\ket{g, n} \quad & \longrightarrow \quad E_n^{(g)} = E_A^g + E_{EM} = 0 + n \hbar \omega_c = \hbar \omega_c n \\
\ket{e, n} \quad & \longrightarrow \quad E_n^{(e)} = E_A^e + E_{EM} = \hbar \omega_{eg} + n \hbar \omega_c = \hbar (\omega_{eg} + \omega_c n )
\end{align*}
These energy levels are reported in figure \ref{fig:JCeigen}, where the decoupling $\delta \equiv \omega_{eg} - \omega_c$ is introduced.
\begin{figure}[h]
\centering
\includegraphics[width=0.57\linewidth]{images/eigenstatesAL.png}
\caption{Eigenstates of the James-Cummings model (neglecting the atom-light interaction).}
\label{fig:JCeigen}
\end{figure}
Introducing the interaction between the atom and the cavity, which is described by the last term in Eq. \ref{eq:JCModel}
\begin{equation*}
H_{AL} = \frac{\hbar g}{2} ({\sigma_{+} \hat{a}} + {\sigma_{-} \hat{a}^{\dagger}}).
\end{equation*}
the action of this (full) Hamiltonian on the states of the basis is:
\begin{align*}
\sigma_+ \hat{a} \ket{g,n} & = \sqrt{n} \ket{e}\bra{g} \ket{g,n-1} = \sqrt{n} \ket{e,n-1}, \\
\sigma_+ \hat{a} \ket{e,n} & \propto \ket{e}\bra{g} \ket{e,n-1} = 0, \\
\sigma_- \hat{a}^\dagger \ket{g,n} & \propto \ket{g}\bra{e} \ket{g,n+1} = 0, \\
\sigma_- \hat{a}^\dagger \ket{e,n} & = \sqrt{n} \ket{g}\bra{e} \ket{e,n+1} = \sqrt{n} \ket{g,n+1},
\end{align*}
from which one deduces that $H_{AL}$ only couples
\begin{align*}
\ket{g, n} \quad \longleftrightarrow \quad \ket{e,n-1} \qquad \text{and} \qquad
\ket{e, n} \quad \longleftrightarrow \quad \ket{g,n+1}.
\end{align*}
The first relation corresponds to the process of \textit{light absorption}, in which one photon is absorbed by the atom which jumps from the ground state to the excited state, while the second one described the process of \textit{light emission}, in which one photon is emitted from the atom that goes from the excited state to the ground state.
\begin{figure}[t]
\centering
\includegraphics[width=0.67\linewidth]{images/EmsissionAbsorption.png}
\caption{Processes of light emission and light absorption for an atom with two energy levels: $\ket{g}$ and $\ket{e}$. }
\label{fig:AbEm}
\end{figure}
The problem ends up being a set of two-levels systems (dotted rectangles in figure \ref{fig:JCeigen}) plus the vacuum state $\ket{g,0}$. Each one of these systems is associated to an Hamiltonian $H_n$ which, in the basis formed by $\ket{g,n}$ and $\ket{e,n-1}$, is given by
\begin{align*}
H_n & =
\begin{pmatrix}
n \hbar \omega_c & \dfrac{g \hbar \sqrt{n}}{2} \\
\dfrac{\hbar g \sqrt{n}}{2} & \hbar \omega_{eg} + (n-1) \hbar \omega_c
\end{pmatrix} = \\
&= n \hbar \omega_c \mathbb{1} +
\begin{pmatrix}
0 & \dfrac{g \hbar \sqrt{n}}{2} \\
\dfrac{g \hbar \sqrt{n}}{2} & \hbar (\omega_{eg} - \omega_c)
\end{pmatrix} = \\
&= n \hbar \omega_c \mathbb{1} +
\begin{pmatrix}
0 & \dfrac{g \hbar \sqrt{n}}{2} \\
\dfrac{g \hbar \sqrt{n}}{2} & \hbar \delta
\end{pmatrix}
\end{align*}
Therefore, the total hamiltonian is a $2 \times 2$ block diagonal matrix with the first element of the diagonal being zero (corresponding to $\ket{g, 0}$):
\begin{equation*}
H = \left(
\begin{array}{ccccc}
\mathbf{0} & & & \\
& H_1 & & \\
& & \ddots & \\
& & & H_n & \\
& & & & \ddots
\end{array}
\right)
\end{equation*}
As for each $H_n$, one can compute the eigenvalues:
\begin{equation}
E_\pm^{(n)} = n \hbar \omega_c + \frac{1}{2} \hbar \delta \pm \frac{\hbar}{2} \sqrt{\delta^2 + g^2 n} \;\;.
\label{eq:energies}
\end{equation}
This result shows that there is a further separation between $\ket{g, n}$ and $\ket{e, n-1}$, indeed the eigenvectors corresponding to each eigenvalue are
\begin{align*}
\ket{+}_n &= \cos \left(\frac{\theta}{2} \right) \ket{g, n} + \sin\left( \frac{\theta}{2} \right) \ket{e, n-1} \\
\ket{-}_n &= - \sin\left( \frac{\theta}{2} \right) \ket{g, n} + \cos\left( \frac{\theta}{2} \right) \ket{e, n-1}
\end{align*}
with
\begin{equation*}
\theta = \arctan \left( \frac{g \sqrt{n}}{\delta}\right).
\end{equation*}
A schematic view of the separation between the eigenstates of the Jaynes-Cumming model is reported in figure \ref{fig:newJCeigen}.
\begin{figure}[t]
\centering
\includegraphics[width=0.6\linewidth]{images/newEigenValue.png}
\caption{Separation between the eigenstates of the James-Cumming model due to the atom-light interaction.}
\label{fig:newJCeigen}
\end{figure}
It is important to notice that $\ket{+}_n$ and $\ket{-}_n$ are superpositions of atom and photons modes and they are called \textbf{dressed states}. This superposition is maximum when $\theta = \pi/2$, which is equivalent to the condition $\delta \to 0$ in which $\sin(\theta/2) = \cos{\theta/2}$. When the decoupling is very small, the system is resonant and the separation between eigenvalues, evaluated using (\ref{eq:energies}), becomes $\Delta E_n = \hbar g \sqrt{n}$. This shows that the Rabi oscillations are now completely quantized since the energy separation between states depends on the number of photons $n$. The separation trend $\omega_n = \Delta E_n$ is reported in figure \ref{fig:omegan}, from which one can notice that the best condition to solve the problem (bigger separation of energies, or lower density) is for low $n$ (far from the classical regime).
\begin{figure}[h!]
\centering
\includegraphics[width=0.8\linewidth]{images/omega_n.png}
\caption{Frequency associated to the separation of state as a function of the number of photons in the cavity.}
\label{fig:omegan}
\end{figure}
\subsection{Quantum Rabi oscillations}
Given an initial state $\ket{\psi_0} = \ket{e, n-1} $, it is possible to evaluate the probability that the atom remains in the excited state and the number of photons in conserved (always in the case of a cavity with resonant light, i.e. $\delta \approx 0$):
\begin{align*}
P_e^{(n-1)}(t) &\propto \abs{\exp{-iE_+ t/\hbar} + \exp{-iE_- t/\hbar}}^2 \\
&\propto \abs{\exp{-\frac{i}{2} g \sqrt{n}t } + \exp{\frac{i}{2} g \sqrt{n}t }}^2 \\
& = \cos^2{\left(\sqrt{n} \, g t\right)}.
\end{align*}
This trend (shown in Figure \ref{fig:Pe}) represents perfect oscillations, also called \textit{quantum Rabi oscillations}.
% EXPLAIN: only if they are fast compared to the decay time of the system (this happens when the photons exit the cavity).
\begin{figure}[t]
\centering
\includegraphics[width=0.7\linewidth]{images/PeJC.png}
\caption{Probability of finding the atom in the excited state starting from a situation in which the atom is already in the excited state and the number of photons in the cavity are $n$ or $n'$, with $n'<n$. }
\label{fig:Pe}
\end{figure}
\subsection{Collapses and revivals of the atomic population}
Consider the case in which the initial configuration is $\ket{\psi_0} = \ket{e} \otimes \ket{\alpha}$, with $\ket{\alpha}$ given by equation (\ref{eq:cohstate}), and let $\Bar{n} = \abs{\alpha}^2$. The probability of having an excited state at a time starting from $\psi_0$ is
\begin{align}
P_e(t) = |\bra{e}\ket{\psi(t)}|^2 = \sum_{n=1} \abs{c_{n-1}(0)}^2 P_e^{(n-1)}(t),
\label{eq:Pe}
\end{align}
where
\begin{align*}
\abs{c_{n-1}(0)}^2 = \abs{e^{-\abs{\alpha}^2/2 } \frac{\alpha^{n-1}}{\sqrt{(n-1)!}}}^2 = e^{-\abs{\alpha}^2} \frac{\alpha^{2(n-1)}}{(n-1)!}
\end{align*}
indicate the initial occupation probability and $P_e^{(n-1)}$ is similar to the result obtained in the previous section
\begin{align*}
P_e^{(n-1)} = \cos^2{(\sqrt{\bar{n}+ \delta n} \, gt )}.
\end{align*}
In the last expression, $\delta n$ is introduced to indicate the spread of the distribution which describes the number of photons. Moreover, one can notice that $\delta n/\bar{n} \sim 1/\sqrt{\bar{n}} \to 0$ when $\bar{n} \to \infty$. In this limit, the expression for $P_e^{(n-1)}$ can be rewritten as
\begin{align*}
P_e^{(n-1)} & = \cos^2{\left( \sqrt{\bar{n}} gt \sqrt{1 + \frac{\delta n}{\bar{n}}} \right)} \simeq \\
& \simeq \cos^2{\left( \sqrt{\bar{n}} gt \left( 1 + \frac{\delta n}{2\bar{n}} \right) \right)} = \\
& = \cos^2{\left( \sqrt{\bar{n}} gt + \frac{\delta n}{2\sqrt{\bar{n}}}gt\right)} = \\
& = \frac{1}{2} + \frac{1}{2} \cos{\left(2 \sqrt{\bar{n}} g t + \frac{\delta n}{\sqrt{\bar{n}}}gt \right)} \;\;,
\end{align*}
therefore equation (\ref{eq:Pe}) leads to
\begin{equation*}
P_e = \frac{1}{2} + \frac{1}{2} \sum_{n=1} \abs{c_{n-1}(0)}^2 \left[ \cos{(2 \sqrt{\bar{n}} gt)} \cos{\left( \frac{\delta n}{\sqrt{\bar{n}}} gt \right)} - \sin{(2 \sqrt{\bar{n}} gt)} \sin{\left( \frac{\delta n}{\sqrt{\bar{n}}} gt \right)}\right].
\end{equation*}
Two frequencies are identified from it:
\begin{align}
\omega_R \equiv 2 \sqrt{\bar{n}} g \qquad \text{and} \qquad \omega_C \equiv g \frac{\delta n}{\sqrt{\bar{n}}}.
\end{align}
Having a large average number of photons, one can easily see that $\omega_R \gg \omega_C$ and thus, in the interval $0 < t < (2 \pi)/\omega_C$, only the oscillations with frequency $\omega_R$ are relevant. To better understand what happens, the expression for $P_e(t)$ has to be rewritten in a more convenient form:
\begin{align}
P_e(t) & = \frac{1}{2} + \frac{1}{2} \sum_{n=0}^\infty e^{-\bar{n}} \frac{\bar{n}^n}{n!} \left[ \cos{\left( 2 \sqrt{\bar{n}} gt + gt \frac{\delta n}{\sqrt{\bar{n}}} \right)}\right] = \label{eq:st1}\\
&= \frac{1}{2} + \frac{1}{2} \sum_n e^{-\bar{n}} \frac{\bar{n}^n}{n!} \frac{1}{2} \left( \exp{i 2 \sqrt{\bar{n}} gt} \exp{igt\frac{n-\bar{n}}{\sqrt{\bar{n}}}} + \text{c.c.} \right) = \nonumber \\
& = \frac{1}{2} + \frac{1}{4}\left[ e^{-\bar{n}} \sum_n \left( \frac{\bar{n}^n}{n!} \exp{igt\frac{n}{\sqrt{\bar{n}}}}\right) \exp{i2\sqrt{\bar{n}}gt - igt\frac{\bar{n}}{\sqrt{\bar{n}}}} + \text{c.c.} \right] = \nonumber\\
& = \frac{1}{2} + \frac{1}{4} \left[ e^{-\bar{n}} \exp{\bar{n}e^{igt/\sqrt{\bar{n}}}} e^{i \sqrt{\bar{n}}gt} + \text{c.c.} \right] \equalexpl{we develop the exponent of exponent} \nonumber\\
& \simeq \frac{1}{2} + \frac{1}{4} \left[ e^{-\bar{n}} \cdot e^{\bar{n}} \, e^{i\sqrt{\bar{n}}gt} \, e^{-g^2 t^2/2} \cdot e^{i\sqrt{\bar{n}}gt} \, + \text{c.c.} \right] = \nonumber\\
&= \frac{1}{2} + \frac{1}{4} \left[ e^{2i\sqrt{\bar{n}}gt} e^{-\frac{1}{2} g^2 t^2} + \text{c.c.} \right] = \nonumber \\
& = \frac{1}{2} + \frac{1}{2} \cos{\left(2 \sqrt{\bar{n}}gt\right)} e^{-\frac{1}{2}g^2 t^2} \label{eq:stfin}
\end{align}
The cosine term represents the oscillations with frequency $\omega_R$, while the exponential coefficient corresponds to a damping with \textit{collapse time} $\tau_C \sim \sqrt{2}/g$. It is important to underline that the last term does not correspond properly to a damping, but it is better to consider it as a dephasing, since it has been obtained summing over all the possible frequencies.
\begin{center}
\includegraphics[scale=0.6]{img/Rabi_scales_2.pdf}
\end{center}
From the previous analysis, two time scales have been identified: the one associated to the Rabi frequency ($t_\text{Rabi}$) and the one of the exponential damping ($\tau_C$). In addition to that, there is another time scale that must be considered, which is obtained taking times such that
\begin{align}
\frac{gt}{\sqrt{\bar{n}}} = 2 \pi m \qquad \text{with} \qquad m \in \mathbb{N}.
\label{eq:revival}
\end{align}
Inserting this value in (\ref{eq:st1}), one obtains $P_e \sim \cos{(4 \pi m\bar{n} + 2 \pi m \delta n)} \sim 1$, and hence the probability reaches again its maximum value every $t$ which satisfies (\ref{eq:revival}), i.e after the so-called \textit{revival time}. An example of the full dynamics is reported in the following figure.
\begin{center}
\includegraphics[scale=0.7]{img/Rabi_scales_1.pdf}
\end{center}
\section{Two-level atom coupled with a bath}
Consider a two-level atom coupled with a bath in which many modes are available. This means that, if the atom is in the excited state, there are several decay channels in the bath. The Hamiltonian for this system is made of $H_A$, $H_{EM}$ and $H_{AL}$ which can be written in the rotating-wave approximation
\begin{align}
H_{AL} = -\vec{d}_{eg} \sigma_+ \hat{\vec{E}}^{(+)} - \vec{d}_{eg}^* \sigma_- \hat{\vec{E}}^{(-)}
\end{align}
with
\begin{align}
\hat{\vec{E}}^{(+)} &= -i \sum_{\vec{k}, \lambda} \sqrt{\frac{\hbar \omega_k}{2 \varepsilon_0 V}} \vec{\epsilon}_\lambda \,\hat{a}_{\vec{k},\lambda}, \\
\hat{\vec{E}}^{(-)} &= i \sum_{\vec{k}, \lambda} \sqrt{\frac{\hbar \omega_k}{2 \varepsilon_0 V}} \vec{\epsilon}_\lambda \, \hat{a}_{\vec{k},\lambda}^\dagger.
\end{align}
These terms correspond to the excitation and de-excitation of the atom after the interaction with light.
This problem can be solved using the Lindblad Master Equation (\ref{eq:cos}) (also called \textit{Nakajima-Zwanzig Equation}) presented in section \ref{sec:ME} and considering
\begin{align*}
{S'}_\alpha(t') \equiv \sigma_+(t') \qquad \text{and} \qquad {S'}_\beta^\dagger(t) \equiv \sigma_-(t)
\end{align*}
\begin{align*}
\hat{B}_\alpha \equiv i \vec{d}_{eg}^* \cdot \hat{\vec{E}}^{(-)} \qquad \text{and} \qquad \hat{B}_\beta^\dagger \equiv - i \vec{d}_{eg} \cdot \hat{\vec{E}}^{(+)}.
\end{align*}
Moreover,
\begin{align*}
\sigma_-(t') = \sigma_-(0) \, e^{-i \omega_{eg} t'} \qquad \text{and} \qquad \sigma_+(t) = \sigma_+(0) \, e^{i \omega_{eg} t}.
\end{align*}
Therefore, the first term of equation (\ref{eq:cos}) becomes
\begin{align*}
& \int_0^t dt'\, G_{\alpha\beta}(t-t') \sigma_-(0) e^{-i \omega_{eg} t'} {\rho'}_S(t') \sigma_+(0) e^{i \omega_{eg} t} = \\
& = \int_0^t dt'\, G_{\alpha\beta}(t-t') e^{i \omega_{eg}(t-t')} \sigma_-(0) \rho'_S(t') \sigma_+(0);
\end{align*}
introducing the variable $\tau \equiv t-t'$ and assuming that $\rho(\tau)$ is smoothly evolving over the time scale at which $G_{\alpha \beta}(\tau)$ decays ($\rho'_S(t-\tau) \simeq \rho'_S(t)$), it can be written as
\begin{align*}
- \left[\int_t^0 d\tau \, G_{\alpha \beta}(\tau) e^{i \omega_{eg}\tau} \right]\sigma_- \rho'_S(t) \sigma_+.
\end{align*}
For the same reason, the lower limit of this integral can be taken $t\to\infty$. Therefore, all the terms in the Master Equation in this assumption has an integral coefficient with the generic form
\begin{align*}
\mathcal{G}(\omega_{eg}) \equiv \int_0^\infty d \tau \, G_{--}(\tau) e^{i \omega_{eg} \tau} \qquad \text{with} \qquad G_{--}(\tau) = \langle \left( \vec{d}_{eg} \cdot \hat{\vec{E}}^{(+)}(\tau) \right) \left( \vec{d}_{eg}^* \cdot \hat{\vec{E}}^{(-)}(0)\right) \rangle_B
\end{align*}
An element of this expectation value can be written as
\begin{align*}
\langle \hat{a}_{\vec{k},\lambda}(\tau) \hat{a}^\dagger_{\vec{k}',\lambda'}(0)\rangle_B &= \langle e^{-i \omega_k t} \hat{a}_{\vec{k},\lambda}(0) \hat{a}^\dagger_{\vec{k}',\lambda'}(0)\rangle_B = \\ &= \langle \hat{a}_{\vec{k},\lambda}(0) \hat{a}^\dagger_{\vec{k}',\lambda'}(0)\rangle_B = \\
& = \delta_{\vec{k},\vec{k}'} \delta_{\lambda,\lambda'} \left(\mathbb{1}-\langle \hat{a}^\dagger_{\vec{k}',\lambda'} \hat{a}_{\vec{k},\lambda}(0)\rangle_B\right).
\end{align*}
The term $\langle \hat{a}^\dagger_{\vec{k}',\lambda'}\hat{a}_{\vec{k},\lambda}(0)\rangle_B$ can be ignored since it corresponds to the average number of background photons in the bath (like the CMB = cosmic microwave background). Therefore,
\begin{align}
\langle \hat{a}_{\vec{k},\lambda}(\tau) \hat{a}^\dagger_{\vec{k}',\lambda'}(0)\rangle_B \simeq \delta_{\vec{k},\vec{k}'} \delta_{\lambda,\lambda'}
\end{align}
and
\begin{align*}
\qquad G_{--}(\tau) = \frac{\hbar \omega_k}{2 \varepsilon_0 V} \sum_{\vec{k},\lambda} e^{-i\omega_k t} \left( \vec{d}_{eg} \cdot \vec{\epsilon}_\lambda \right)\left( \vec{d}_{eg}^* \cdot \vec{\epsilon}_\lambda \right).
\end{align*}
Using the continuum limit to rewrite the summation over $\vec{k}$ and considering the projections of $\vec{d}_{eg}$ and $\vec{d}_{eg}^*$ on $\vec{\epsilon}_1$ and $\vec{\epsilon}_2$, one as
\begin{align*}
\qquad G_{--}(\tau) &= \frac{\hbar \omega_k}{2 \varepsilon_0 V} \frac{1}{(2\pi)^2} \int d^3 \vec{k} \, e^{-i\omega_k t} \sum_\lambda \abs{({d}_{eg})_\lambda}^2 = \\
& = \frac{\hbar \omega_k}{2 \varepsilon_0 V} \frac{1}{(2\pi)^2} \int d^3 \vec{k} \, e^{-i\omega_k t} \abs{\vec{d}_{eg}}^2 \sin^2{\theta},
\end{align*}
where $\theta$ is the angle between $\vec{d}_{eg}$ and the plane of $\vec{\epsilon}_1$ and $\vec{\epsilon}_2$. From this result, the spherical coordinates (with the solid angle $d \Omega$) can be used to rewrite the integral over $d^3 \vec{k}$ and the complete expression is
\begin{align*}
\mathcal{G}(\omega_{eg})
&= \int_0^\infty d \tau \, \int \frac{d \Omega}{(2 \pi)^3} \sin^2 \theta \int_0^\infty dk \, k^2 \frac{\hbar \omega}{2 \varepsilon_0} e^{-i(\omega-\omega_{eg})\tau} |{\vec{d}_{eg}}|^2 = \\
&= \frac{\hbar c |{\vec{d}_{eg}}|^2 }{2 \varepsilon_0 (2\pi)^3} \int d\Omega \, \sin^2 \theta \int_0^\infty dk \, k^3 \int_0^\infty d \tau \, e^{-i(\omega-\omega_{eg})\tau} = \\
&= \frac{2 \hbar |{\vec{d}_{eg}}|^2 }{3 \varepsilon_0 (2 \pi)^3 c^3} \int_0^\infty d\omega \, \omega^3 \int_0^\infty d\tau \, e^{-i(\omega-\omega_{eg})\tau},
\end{align*}
where the relation $\omega_k \equiv \omega = c k$ is used. The last integral is quite tricky. In the following, we will consider its real and imaginary parts.\\
To write the complete Master Equation, the other correlators (similar to $G_{\alpha \beta}$) should be taken into account; it is possible to prove that they are all null. Hence the final expression is
\begin{align*}
\dot{\rho}'_S(t) &= \frac{1}{\hbar^2} \bigg\{ \real( \mathcal{G}(\omega_{eg})) \left( \sigma_-\rho_S \sigma_+ - \sigma_+ \sigma_- \rho_S \right) ~+ \\
&~~+ \real( \mathcal{G}(\omega_{eg})) \left( \sigma_- \rho_S \sigma_+ - \rho_S \sigma_+ \sigma_- \right) ~+ \\
&~~+i \mathfrak{I}(\mathcal{G}(\omega_{eg})) (\rho_S \sigma_+ \sigma_- - \sigma_+ \sigma_- \rho)\bigg\}.
\end{align*}
The final step consists of going back to the Schr\"odinger picture to write
\begin{equation}
\begin{split}
\dot{\rho_S} &= \frac{i}{\hbar} \left[ \rho_S,H_S \right] + \frac{i}{\hbar} \mathfrak{I} (\mathcal{G}(\omega_{eg})) \left[ \rho_S, \sigma_+ \sigma_-\right]~+ \\
&~~+ \frac{i}{\hbar^2} \real(\mathcal{G}(\omega_{eg})) \left( \sigma_- \rho_S \sigma_+ - \bigg\{ \sigma_+ \sigma_-, \rho_S \bigg\}\right)
\end{split}
\end{equation}
\noindent Two final observations can be done:
\begin{itemize}
\item Since $\sigma_+ \sigma_- = \ket{e}\bra{e}$, one can introduce a re-normalized Hamiltonian
\begin{align*}
\bar{H}_S = \ket{e}\bra{e} \left( \hbar \omega_{eg} + \frac{1}{\hbar} \mathfrak{I}(\mathcal{G}(\omega_{eg})) \right).
\end{align*}
The last term is divergent and this problem can be fixed considering the relativistic contributions to the Hamiltonian (in QED different techniques are used to solve it). If this is done, tha imaginary part is just a constant energy shift, called \textit{Lamb shift}.
\item Introducing
\begin{equation}
\Gamma_{eg} \equiv \frac{2}{\hbar^2} \real (\mathcal{G}(\omega_{eg}) = \frac{|\vec{d}_{eg}|^2 \omega^3_{eg}}{3 \pi \varepsilon_0 \hbar c^3}
\end{equation}
\end{itemize}
one can define the \textit{inverse lifetime} of the excited state
\begin{align}
\Gamma_e \equiv{\sum_g} \Gamma_{eg}
\end{align}
\section{Two-level atom in a dissipative system}
Consider a two-level atom, a source of coherent light and a dissipation described by the quantity $\Gamma$. As seen, in the rotating frame with $\omega \gg \delta, \Omega$, the Hamiltonian is given by equation (\ref{eq:Hrw}), while the Lindblad Master Equation is
\begin{equation}
\frac{\partial}{\partial t} {\rho} = \frac{i}{\hbar} [\rho,H] + \frac{\Gamma}{2} [2 \sigma_- \rho \sigma_+ - \sigma_+ \sigma_- \rho - \rho \sigma_+ \sigma_-].
\end{equation}
The latter can be used to study the evolution of $\langle \sigma_-\rangle$ and $\langle \sigma_+\rangle$, starting from
\begin{align*}
\frac{\partial}{\partial t} \langle \sigma_-\rangle = \frac{\partial}{\partial t} \text{Tr}[\rho \sigma_-] = \text{Tr}\left[\frac{\partial}{\partial t} {\rho} \sigma_-\right]
\end{align*}
and inserting it in the Master Equation
\begin{align*}
\frac{\partial}{\partial t} \langle \sigma_-\rangle &= \text{Tr} \left\{ \left[ \frac{i}{\hbar} (\rho H - H \rho) + \frac{\Gamma}{2} \left( 2 \sigma_- \rho \sigma_+ - \sigma_+ \sigma_- \rho - \rho \sigma_+ \sigma_-\right)\right] \sigma_- \right\} = \\
& = \frac{i}{\hbar} \text{Tr} \left[ \rho \left(\hbar \delta \ket{e}\bra{e} + \frac{\hbar \Omega}{2} \sigma_+ + \frac{\hbar \Omega}{2} \sigma_- \right) \sigma_- \right] + \\
&~~~ -\frac{i}{\hbar} \text{Tr} \left[ \left(\hbar \delta \ket{e}\bra{e} + \frac{\hbar \Omega}{2} \sigma_+ + \frac{\hbar \Omega}{2} \sigma_- \right) \rho \sigma_- \right] + \\
& ~~~ +\frac{\Gamma}{2} \text{Tr} \left[ 2 \sigma_- \rho \sigma_+ \sigma_- - \sigma_+ \sigma_- \rho \sigma_- - \rho \sigma_+ \sigma_- \sigma_- \right] = \\
&= \frac{i}{\hbar} \text{Tr}\left[ \frac{\hbar \Omega}{2} \rho \sigma_+ \sigma_-\right] - \frac{i}{\hbar} \text{Tr}\left[ \hbar \delta \underbrace{\ket{e}\bra{e} \rho \sigma_-}_{\rho \sigma_-} \right] - \frac{i}{\hbar} \text{Tr}\left[ \frac{\hbar \Omega}{2} \sigma_+ \rho\sigma_-\right] + \\
& ~~~ +\frac{\Gamma}{2} \text{Tr} \left[ \underbrace{\ket{e}\bra{e} \rho \sigma_-}_{\rho \sigma_-}\right] = \\
& = \frac{i}{\hbar} \frac{\hbar \Omega}{2} \text{Tr}[\rho \ket{e}\bra{e}] - \frac{i}{\hbar} \hbar \delta \, \text{Tr}[\rho \sigma_-] - \frac{i}{\hbar} \frac{\hbar \Omega}{2} \text{Tr} \left[ \ket{g}\bra{g} \rho \right] - \frac{\Gamma}{2} \text{Tr} \left[ \rho \sigma_- \right] = \\
&= i \frac{\Omega}{2} \text{Tr} [\rho \sigma^z] - i \delta \, \text{Tr} [\rho \sigma_-] - \frac{\Gamma}{2} \text{Tr} [\rho \sigma_-] = \\
&= i\frac{\Omega}{2} \langle \sigma^z \rangle - \left(i \delta +\frac{\Gamma}{2} \right) \langle \sigma_- \rangle .
\end{align*}
The evolution equation for $\langle \sigma_+ \rangle$ is obtained from the complex conjugate of this result, while a similar procedure leads to the evolution equation for $\langle \sigma^z \rangle$. These equations can be put in matrix form:
\begin{equation}
\frac{\partial}{\partial t}
\begin{pmatrix}
\langle \sigma_- \rangle \\ \langle \sigma_+ \rangle \\ \langle \sigma^z \rangle
\end{pmatrix} =
\begin{pmatrix}
-i\delta -{\Gamma}{2} & 0 & {i\Omega}/{2} \\
0 & i\delta -{\Gamma}/{2} & -{i\Omega^*}/{2} \\
i \Omega^* & -i \Omega & -\Gamma
\end{pmatrix}
\begin{pmatrix}
\langle \sigma_- \rangle \\ \langle \sigma_+ \rangle \\ \langle \sigma^z \rangle
\end{pmatrix} - \begin{pmatrix}
0 \\ 0 \\ \Gamma
\end{pmatrix}
\end{equation}
or equivalently
\begin{align}
\frac{\partial}{\partial t} \vec{S} = M \vec{S} - \vec{R}.
\end{align}
These equations are called \textbf{Optical Bloch Equations}.
\subsection{Stationary states of the Optical Bloch Equations}
Non trivial stationary states are given by
\begin{align*}
\vec{S}_\infty = M^{-1} \vec{R},
\end{align*}
from which
\begin{align*}
\langle \sigma_- \rangle_\infty & = -\frac{\Omega (2 \delta + i \Gamma)}{\Gamma^2 + 4 \delta^2 + 2 \abs{\Omega}^2} \\
\langle \sigma_+ \rangle_\infty & = -\frac{\Omega (2 \delta - i \Gamma)}{\Gamma^2 + 4 \delta^2 + 2 \abs{\Omega}^2} \\
\langle \sigma^z \rangle_\infty & = -\frac{\Gamma^2 + 4 \delta^2}{\Gamma^2 + 4 \delta^2 + 2 \abs{\Omega}^2}.
\end{align*}
One can also introduce the \textit{saturation parameter}
\begin{align}
\mathcal{S} \equiv \frac{2\dfrac{\abs{\Omega}^2}{\Gamma^2}}{1+\dfrac{4\delta^2}{\Gamma^2}},
\end{align}
which allows to rewrite the previous results as
\begin{align}
\abs{\langle \sigma_- \rangle_\infty }^2 = \frac{\mathcal{S}}{2(1+\mathcal{S})^2} \qquad \text{and} \qquad \langle \sigma^z \rangle_\infty = \frac{1}{1+\mathcal{S}}
\end{align}
\\
\noindent In addition to that, one can evaluate the probability of being in an excited state
\begin{align*}
P_e = \rho_{ee} = \text{Tr}[\rho \ket{e}\bra{e}] = \text{Tr}\left[ \rho \left( \frac{\mathbb{1}}{2} + \frac{\sigma^z}{2} \right)\right] = \frac{1}{2} \text{Tr}[\rho] + \frac{1}{2} \langle \sigma^z \rangle = \frac{1}{2} \langle \sigma^z \rangle
\end{align*}
and, using the results for the stationary states,
\begin{align*}
P_e^\infty = \lim_{t\to \infty} P_e = \frac{1}{2}\langle \sigma^z \rangle_\infty = \frac{1}{2(1+\mathcal{S})} =\frac{\Omega^2}{2\Omega^2 + 4 \delta^2 + \Gamma^2}.
\end{align*}
\begin{center}
\scalebox{1.4}{ \input{img/Rabi_PE.tex} }
\end{center}
\noindent From this expression, we recognize that two regimes are possible:
\begin{itemize}
\item weak drive:
\begin{align*}
\Omega \ll \Gamma \qquad \implies \qquad P_e \simeq \frac{\Omega^2}{4 \delta^2 + \Gamma^2},
\end{align*}
which is Lorentzian with width $\sim \Gamma$;
\item strong drive:
\begin{align*}
\Omega \gg \Gamma \qquad \implies \qquad P_e \simeq \frac{\Omega^2}{4 \delta^2 + 2\Omega^2},
\end{align*}
which is Lorentzian with width $\sim \Omega$;
\end{itemize}
\section{Three-level atoms and Raman coupling}
Consider a three-level atom, with two quasi-ground states (meaning that they are very stable) $\ket{g_i}, i = 0,1\;$, and a short-lived excited state $\ket{e}$; assume that transitions between the quasi-ground states are forbidden. A schematic view is reported in figure \ref{fig:lambda}.
\begin{figure}[h]
\centering
\includegraphics[width=0.3\linewidth]{images/lambda_system.png}
\caption{$\Lambda$ system with suppressed transition between the ground states. }
\label{fig:lambda}
\end{figure}
The Hamiltonian for the atom and for the atom-light interaction are
\begin{align*}
H_{A} &= -\hbar \omega_{g_1} \ket{g_1}\bra{g_1} -\hbar \omega_{g_2} \ket{g_2}\bra{g_2} + 0\\
H_{AL} &= - \hat{\vec{d}} \cdot \hat{\vec{E}},
\end{align*}
where the dipole transition term evaluates as
\begin{equation*}
\begin{aligned}
{\vec{d}} = -e{\vec{r}} ~\big\rvert_{g_1,g_2,e} & = -e\ket{g_1}\bra{g_1}\vec{r}\ket{e}\bra{e} \;\;- e\ket{g_2}\bra{g_2}\vec{r}\ket{e}\bra{e} \;\; + \text{h.c.} \\
&\equiv d_{eg_1}^*\ket{g_1}\bra{e} \;\;+\;\; d_{eg_2}^*\ket{g_2}\bra{e} \;\; + \text{h.c}.
\end{aligned}
\end{equation*}
In this derivation, we will write the electric field is written in a slightly different way with respect to the previous chapters:
\begin{equation*}
\vec{E} =
\sum_{\vec{k},\lambda} \sqrt{\frac{\hbar\omega_k}{2\varepsilon_0}}\vec\epsilon_{\lambda} \, 2 \mathfrak{Im} \left[u_{\vec{k},\lambda}(\vec{r}_\text{atom})\hat{a}_{\vec{k},\lambda}\right],
\end{equation*}
since for a generic operator $A$ one can write $\mathfrak{Im} A = (A-A^\dagger)/2i$. Notice that $\vec{r}_\text{atom}$ is evaluated in center-of-mass coordinate and not in the coordinates of the electron.
Consider two independent lasers to control the system
\begin{equation*}
\ket{\psi_\text{lasers}} = \ket{0} \otimes \dots \otimes \ket{0} \otimes \underbrace{\ket{\alpha_1}}_{\substack{\textrm{first laser}\\k_1,\lambda_1}} \otimes \ket{0} \otimes \dots \otimes \ket{0} \otimes \underbrace{\ket{\alpha_2}}_{\substack{\textrm{second laser}\\k_2,\lambda_2}} \otimes \ket{0} \otimes \dots,
\end{equation*}
and assume that:
\begin{itemize}
\item $\vec{E}$ is perceived by the system as an averaged value
\begin{align*}
\bra{\psi_\text{lasers}}\vec{E}\ket{\psi_\text{lasers}} & = \sqrt{\frac{\hbar\omega_1}{2 \varepsilon_0}}
\vec\epsilon_{\lambda_1} 2 \mathfrak{Im} \left(u_{\vec{k}_1,\lambda_1}\right) \bra{\alpha_1}\hat{a}_{\vec{k}_1,\lambda_1} \ket{\alpha_1} + \left(\substack{\text{same for}\\\vec{k}_2, \lambda_2 }\right)\\
& = \sqrt{\frac{2\hbar\omega_1}{\varepsilon_0}}
\vec\epsilon_{\lambda_1} |u_{\vec{k}_1,\lambda_1}||\alpha_1| \cos\left[(c k_1)t + \phi_1\right] + \left(\substack{\text{same for}\\\vec{k}_2, \lambda_2 }\right),
\end{align*}
\item the lasers are powerful and hence the state $\ket{\psi}_\text{laser}$ is not much affected by the absorption of one photon.
\end{itemize}
The average value of the electric field can be rewritten as
\begin{align}
\langle \vec{E}\rangle =
\vec{\mathcal{E}}_1 \cos\left(\omega_1t + \phi_1\right) +
\vec{\mathcal{E}}_2 \cos\left(\omega_2t + \phi_2\right).
\end{align}
From the previous considerations
\begin{align*}
H_{AL} & = - \hat{\vec{d}} \cdot \langle \vec{E}\rangle = \\
&= -\left(
\vec{d}_{eg_1}\ket{e}\bra{g_1} \;+\;
\vec{d}_{eg_2}\ket{e}\bra{g_2} \;+\;
\vec{d}_{eg_1}^*\ket{g_1}\bra{e} \;+\;
\vec{d}_{eg_2}^*\ket{g_2}\bra{e}
\right)
\cdot\\
&\qquad \cdot\left(
\vec{\mathcal{E}}_1 \cos\left(\omega_1t + \phi_1\right)
\;+\;
\vec{\mathcal{E}}_2 \cos\left(\omega_2t + \phi_2\right)
\right).
\end{align*}
A formal simplification can be done by introducing the the coefficients
\begin{equation*}
\Omega_1 = \frac{\vec{d}_{eg_1} \cdot \vec{\mathcal{E}}_1}{\hbar}
\qquad\qquad
\Omega_2 = \frac{\vec{d}_{eg_2} \cdot \vec{\mathcal{E}}_1}{\hbar}
\end{equation*}
\begin{equation*}
\cancel{\Omega}_1 = \frac{\vec{d}_{eg_1} \cdot \vec{\mathcal{E}}_2}{\hbar}
\qquad\qquad
\cancel{\Omega}_2 = \frac{\vec{d}_{eg_2} \cdot \vec{\mathcal{E}}_2}{\hbar}
\end{equation*}
in order to write
\begin{align*}
H_{AL} &= -\hbar \bigg[ \ket{e}\bra{g_1} \left( \Omega_1 \cos{(\omega_1 t + \phi_1)} + \cancel{\Omega}_1 \cos{(\omega_2 t + \phi_2)}\right) + \\
&~~~+ \ket{e}\bra{g_2} \left( \Omega_2 \cos{(\omega_2 t + \phi_2)} + \cancel{\Omega}_2 \cos{(\omega_1 t + \phi_1)}\right) + \text{h.c.} \bigg]
\end{align*}
Eventually, the full Hamiltonian in the laboratory frame, sorting the basis in the order $\{\ket{g_1}, \ket{e}, \ket{g_2}\}$, is
\begin{equation*}
H^{lab} = -\hbar \left(
\begin{array}{c;{5pt/5pt}c;{5pt/5pt}c}
\omega_{eg_1} & \text{c.c.} & 0\\ \hdashline[5pt/5pt]
\Omega_1\cos\left(\omega_1t + \phi_1\right) +
& \multirow{2}{*}{0} & \multirow{2}{*}{\text{c.c.}}\\
+ \cancel{\Omega}_1\cos\left(\omega_2t + \phi_2\right) &&\\ \hdashline[5pt/5pt]
\multirow{2}{*}{0}& \Omega_2^*\cos\left(\omega_2t + \phi_2\right) + & \multirow{2}{*}{$\omega_{eg_2}$}\\
&+\cancel{\Omega}_2^*\cos\left(\omega_1t + \phi_1\right)&\hspace*{3cm}
\end{array}
\right)
\end{equation*}
Now a change of reference frame is performed, for convenience,
$$H^{lab} \quad \longrightarrow \quad H^{rot} = U H^{lab}U^\dagger + i \hbar \dot{U} U^\dagger,$$
applying the time-dependent rotation
\begin{equation*}
U(t) = \begin{pmatrix}
e^{-i(\omega_1t + \phi_1)} & 0 & 0\\
0 & 1 & 0\\
0 & 0 & e^{-i(\omega_2t + \phi_2)}
\end{pmatrix}
\;\; .
\end{equation*}
Hence, the two terms composing the rotated hamiltonian are
\begin{align*}
UH^{lab}U^\dag & = - \hbar U
\begin{pmatrix}
\omega_{eg_1}e^{i(\omega_1...)} & (\Omega_1 ... + \cancel{\Omega}_1...) & 0\\
(\Omega_1 ... + \cancel{\Omega}_1...)e^{i(\omega_1...)} & 0 & (\Omega_2 ... + \cancel{\Omega}_2...)e^{i(\omega_2...)}\\
0 & (\Omega_2 ... + \cancel{\Omega}_2...) & \omega_{eg_2}e^{i(\omega_2...)}\\
\end{pmatrix}
\end{align*}
and
\begin{equation*}
i\hbar\dot{U}U^\dag = \hbar
\begin{pmatrix}
\omega_1 & 0 & 0\\
0 & 0 & 0\\
0 & 0 & \omega_2
\end{pmatrix}.
\end{equation*}
The final result is
\begin{equation*}
H_{rot} = -\hbar \left(
\begin{array}{ccc}
\omega_{eg_1} -\omega_1 & \text{c.c.} & 0\\ %\hdashline[6pt/6pt]
(\Omega_1 ... + \cancel{\Omega}_1...)e^{i(\omega_1t+\phi_1)}
& 0 & \text{c.c.}\\ %\hdashline[6pt/6pt]
0 & (\Omega_2 ... + \cancel{\Omega}_2...)e^{-i(\omega_2t+\phi_2)} & \omega_{eg_2}-\omega_2\\
\end{array}
\right)
\end{equation*}
The following step consists of defining $\delta_1 \equiv \omega_{eg_1} -\omega_1$ and $\delta_2 \equiv \omega_{eg_2} -\omega_2$ and in breaking the full Hamiltonian in two pieces such that
$$H(t) = H_0 + V(t)$$
It is trivial that
\begin{align*}
H_0 & = - \hbar
\begin{pmatrix}
\delta_1 & \Omega_1^*/2 & 0\\
\Omega_1/2 & 0 & \Omega_2/2\\
0 & \Omega_2^*/2 & \delta_2\\
\end{pmatrix}
\end{align*}
and
\begin{align}
V(t) = -\hbar \left(
\begin{array}{c;{6pt/6pt}c;{6pt/6pt}c}
0 & \text{c.c.} & 0 \\ \hdashline[6pt/6pt]
\dfrac{\Omega_1}{2} e^{2i(\omega_1t+\phi_1)} +
& \multirow{4}{*}{0} & \multirow{4}{*}{\text{c.c.}}\\
+ \dfrac{\cancel{\Omega}_1}{2} e^{i(\omega_1+\omega_2)t} e^{i(\phi_1+\phi_2)} + & & \\
+ \dfrac{\cancel{\Omega}_1}{2} e^{i(\omega_1-\omega_2)t} e^{i(\phi_1-\phi_2)} & & \\\hdashline[6pt/6pt]
\multirow{2}{*}{0} & \text{something}& \multirow{2}{*}{0} \\
& \text{similar} &
\end{array}
\right)
\end{align}
As we already know, this formulation allows to describe the problem in terms of an effective Hamiltonian
\begin{equation}
H_{\text{eff}} = H_0 +
\frac{1}{\hbar \tilde\omega}\left[V,V^\dag\right] + \mathcal{O}(\omega^{-2})
\label{eq:ch6:lambdaform}
\end{equation}
where $\tilde{\omega}$ is the oscillation frequency of $V(t)$. The term ${1}/(\hbar {\tilde\omega}) \left[V,V^\dag\right]$ could be neglected if the timescales of $H_0$ are much larger than ${1}/{\tilde\omega}$ or equivalently, if the energy scales of $H_0$ ($\Omega_1$, $\Omega_2$, $\delta_1$, $\delta_2$) and $V$ ($\cancel{\Omega}_1$, $\cancel{\Omega}_2$) are much smaller than $\hbar\tilde\omega$. \\
Basically, this implies that
$$\Omega_1,\; \Omega_2,\; \cancel{\Omega_1},\; \cancel{\Omega_2}, |\delta_1|,\; |\delta_2| ~
\ll ~\omega_1,\; \omega_2,\; \omega_1+\omega_2,\; |\omega_1-\omega_2|.$$
The most strict bound is on $|\omega_1-\omega_2|$, which physically means that the frequencies of the two lasers can not be too close.
% By a physical point of view, this condition is met if $g_1$ and $g_2$ are not too close!
Adding a constant $\hbar\delta_1$, and neglecting the second term of equation (\ref{eq:ch6:lambdaform}), the effective Hamiltonian of the system is
\begin{equation*}
H_{\text{eff}} = H_0 = \hbar \begin{pmatrix}
0 & \bigg\rvert\dfrac{\Omega_1}{2}\bigg\rvert & 0\\
\bigg\rvert\dfrac{\Omega_1}{2}\bigg\rvert & \delta_1 & \bigg\rvert\dfrac{\Omega_2}{2}\bigg\rvert\\
0 & \bigg\rvert\dfrac{\Omega_2}{2}\bigg\rvert & \Delta
\end{pmatrix}
\end{equation*}
where $\Delta \equiv \delta_1 - \delta_2$ and $\ket{e}$. Adding a phase to $\ket{e}$, it is possible to obtain real values for $\Omega_1$ and $\Omega_2$.
Finally, in order for this problem to be equivalent to the previously solved three-level problem (section \ref{sec:timedep}), we must make additional hypothesis:
$$ \Delta, \Omega_j \ll \delta_j \qquad \qquad j = 1, 2.$$
\begin{tcolorbox} [breakable, enhanced]
The hypothesis constraints are usually met in the experiments with these timescales:
\begin{equation*}
\underbrace{\Delta, \Omega_j}_{\sim MHz} \;
\ll
\underbrace{|\delta_1|,\;|\delta_2|}_{\sim GHz}
\ll
\underbrace{\omega_1,\;\omega_2,\;|\omega_1-\omega_2|}_{ \sim THz}
\end{equation*}
\end{tcolorbox}
\noindent Now the problem is formally equivalent to the $\Lambda$-system. The time-independent second-order perturbation theory can be applied considering
\begin{equation*}
H_\text{unpert} = \hbar \begin{pmatrix}
0 & 0 & 0\\
0 &\delta_1 \simeq \delta_2 & 0\\
0 & 0 & 0
\end{pmatrix}
\end{equation*}
and
\begin{equation*}
H_\text{pert}^{(A)} =
\frac{\hbar}{2}\begin{pmatrix}
0&\Omega_1&0\\
\Omega_1&0&\Omega_2\\
0&\Omega_2&0
\end{pmatrix}
\qquad\qquad
H_\text{pert}^{(B)} =
\hbar\begin{pmatrix}
0 & 0 & 0\\
0 & 0 & 0\\
0 & 0 & \Delta
\end{pmatrix}.
\end{equation*}
The final Hamiltonian projected on the ground state space is
\begin{equation*}
\begin{aligned}
H_\text{final}^{\ket{g_1}\ket{g_2}} &= -\frac{\hbar}{4\delta}
\begin{pmatrix}
\Omega_1^2 & \Omega_1\Omega_2\\
\Omega_1\Omega_2 & \Omega_2^2
\end{pmatrix} + \hbar
\begin{pmatrix}
0&0\\
0&\Delta
\end{pmatrix} = \\
&=
-\hbar \frac{\Omega_1\Omega_2}{4\delta}\sigma_x
+\hbar\left(
\frac{\Omega_2^2 - \Omega_1^2}{8\delta}-\Delta
\right) \sigma_z + \text{const}.
\end{aligned}
\end{equation*}
From the practical point of view, the systems transits between the ground states by absorbing and emitting a photon at the same time (one by each laser, at different frequencies). \\
The following sections will explain how to move to an environment with many atoms.