summaryrefslogtreecommitdiff
path: root/source/know/concept/multi-photon-absorption/index.md
blob: af433ac89e7240738a9e6ab5b6bb363caf21e114 (plain)
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
---
title: "Multi-photon absorption"
sort_title: "Multi-photon absorption"
date: 2022-01-30
categories:
- Physics
- Optics
- Quantum mechanics
- Nonlinear optics
- Perturbation
layout: "concept"
---

Consider a quantum system where there are many eigenstates $\Ket{n}$,
e.g. atomic orbitals, for an electron to occupy.
Suppose an [electromagnetic wave](/know/concept/electromagnetic-wave-equation/)
passes by, such that its Hamiltonian gets perturbed by $\hat{H}_1$, given in the
[electric dipole approximation](/know/concept/electric-dipole-approximation/) by:

$$\begin{aligned}
    \hat{H}_1(t)
    = -\vu{p} \cdot \vb{E} \cos(\omega t)
    \approx -\vu{p} \cdot \vb{E} e^{-i \omega t}
\end{aligned}$$

Where $\vb{E}$ is the [electric field](/know/concept/electric-field/) amplitude,
and $\vu{p} \equiv q \vu{x}$ is the transition dipole moment operator.
Here, we have made the
[rotating wave approximation](/know/concept/rotating-wave-approximation/)
to neglect the $e^{i \omega t}$ term,
because it turns out to be irrelevant in this discussion.


We call the ground state $\Ket{0}$,
but other than that, the other states need *not* be sorted by energy.
However, we demand that the following holds
for all even-numbered states $\Ket{e}$ and $\Ket{e'}$,
and for all odd-numbered ($u$neven) states $\Ket{u}$ and $\Ket{u'}$:

$$\begin{aligned}
    \matrixel{e}{\hat{H}_1}{e'} = \matrixel{u}{\hat{H}_1}{u'} = 0
    \qquad \quad
    \matrixel{e}{\hat{H}_1}{u} \neq 0
\end{aligned}$$

This is justified for atomic orbitals thanks to
[Laporte's selection rule](/know/concept/selection-rules/).
Therefore, [time-dependent perturbation theory](/know/concept/time-dependent-perturbation-theory/)
says that the $N$th-order coefficient corrections are:

$$\begin{aligned}
    c_e^{(N)}(t)
    &= -\frac{i}{\hbar} \sum_{u}^{\mathrm{odd}} \int_0^t \matrixel{e}{\hat{H}_1(\tau)}{u} \: c_u^{(N-1)}(\tau) \: e^{i \omega_{eu} \tau} \dd{\tau}
    \\
    c_u^{(N)}(t)
    &= -\frac{i}{\hbar} \sum_{e}^{\mathrm{even}} \int_0^t \matrixel{u}{\hat{H}_1(\tau)}{e} \: c_e^{(N-1)}(\tau) \: e^{i \omega_{ue} \tau} \dd{\tau}
\end{aligned}$$

Where $\omega_{eu} = (E_e \!-\! E_u) / \hbar$.
For simplicity, the electron starts in the lowest-energy state $\Ket{0}$:

$$\begin{aligned}
    c_0^{(0)} = 1
    \qquad \qquad
    c_u^{(0)} = c_{e \neq 0}^{(0)} = 0
\end{aligned}$$

Finally, we prove the following useful relation for large $t$,
involving a [Dirac delta function](/know/concept/dirac-delta-function/) $\delta$:

$$\begin{aligned}
    \lim_{t \to \infty} \bigg| \frac{e^{i x t} - 1}{x} \bigg|^2
    = 2 \pi \: \delta(x) \: t
\end{aligned}$$

<div class="accordion">
<input type="checkbox" id="proof-sinc"/>
<label for="proof-sinc">Proof</label>
<div class="hidden" markdown="1">
<label for="proof-sinc">Proof.</label>
First, observe that we can rewrite the fraction using an integral:

$$\begin{aligned}
    \frac{e^{i x t} - 1}{x}
    = e^{i x t / 2} \frac{e^{i x t / 2} - e^{-i x t / 2}}{x}
    = i e^{i x t / 2} \int_{-t/2}^{t/2} e^{i x \tau} \dd{\tau}
\end{aligned}$$

By taking the limit $t \to \infty$,
it can be turned into a nascent Dirac delta function:

$$\begin{aligned}
    \lim_{t \to \infty} \frac{e^{i x t} - 1}{x}
    = \lim_{t \to \infty} i e^{i x t / 2} \frac{2 \pi}{2 \pi} \int_{-\infty}^{\infty} e^{i x \tau} \dd{\tau}
    = \lim_{t \to \infty} i 2 \pi e^{i x t / 2} \: \delta(x)
\end{aligned}$$

Consequently, the absolute value squared is as follows:

$$\begin{aligned}
    \lim_{t \to \infty} \bigg| \frac{e^{i x t} - 1}{x} \bigg|^2
    = 4 \pi^2 \delta^2(x)
\end{aligned}$$

However, a squared delta function $\delta^2$ is not ideal,
so we take a step back:

$$\begin{aligned}
    \delta^2(x)
    = \delta(x) \lim_{t \to \infty} \frac{1}{2 \pi} \int_{-t/2}^{t/2} e^{i x \tau} \dd{\tau}
    = \delta(x) \lim_{t \to \infty} \frac{t}{2 \pi}
\end{aligned}$$

Where we have set $x = 0$ according to the first delta function.
This gives the target:

$$\begin{aligned}
    \lim_{t \to \infty} \bigg| \frac{e^{i x t} - 1}{x} \bigg|^2
    = 4 \pi^2 \delta^2(x)
    = 2 \pi \: \delta(x) \: t
\end{aligned}$$
</div>
</div>


## One-photon absorption

To warm up, we start at first-order perturbation theory.
Thanks to our choice of initial condition,
nothing at all happens to any of the even-numbered states $\Ket{e}$:

$$\begin{aligned}
    c_e^{(1)}(t)
    &= -\frac{i}{\hbar} \sum_{u}^{\mathrm{odd}} \int_0^t \matrixel{e}{\hat{H}_1(\tau)}{u} \: c_u^{(0)} \: e^{i \omega_{eu} \tau} \dd{\tau}
    = 0
\end{aligned}$$

While the odd-numbered states $\Ket{u}$ have a nonzero correction $c_u^{(1)}$,
where $\vb{p}_{u0} = \matrixel{u}{\vu{p}}{0}$:

$$\begin{aligned}
    c_u^{(1)}(t)
    &= -\frac{i}{\hbar} \int_0^t \matrixel{u}{\hat{H}_1(\tau)}{0} \: c_0^{(0)} \: e^{i \omega_{u0} \tau} \dd{\tau}
    \\
    &= i \frac{\vb{p}_{u0} \cdot \vb{E}}{\hbar} \int_0^t e^{i (\omega_{u0} - \omega) \tau} \dd{\tau}
    \\
    &= i \frac{\vb{p}_{u0} \cdot \vb{E}}{\hbar} \bigg[ \frac{e^{i (\omega_{u0} - \omega) \tau}}{i (\omega_{u0} - \omega)} \bigg]_0^t
\end{aligned}$$

Consequently, the first-order correction
(in the rotating wave approximation) is given by:

$$\begin{aligned}
    \boxed{
        c_u^{(1)}(t)
        \approx \frac{\vb{p}_{u0} \cdot \vb{E}}{\hbar} \frac{e^{i (\omega_{u0} - \omega) t} - 1}{\omega_{u0} - \omega}
    }
\end{aligned}$$

Since $\big| c_u^{(1)}(t) \big|^2$ is the probability
of finding the electron in $\Ket{u}$,
its transition rate $R_u^{(1)}(t)$ is as follows,
averaged since the beginning $t = 0$:

$$\begin{aligned}
    R_u^{(1)}(t)
    = \frac{\big| c_u^{(1)}(t) \big|^2}{t}
    = \frac{1}{t} \bigg| \frac{\vb{p}_{u0} \cdot \vb{E}}{\hbar} \bigg|^2
    \cdot \bigg| \frac{e^{i (\omega_{u0} - \omega) t} - 1}{\omega_{u0} - \omega} \bigg|^2
\end{aligned}$$

For large $t \to \infty$, we can use the formula we proved earlier
to get [Fermi's golden rule](/know/concept/fermis-golden-rule/):

$$\begin{aligned}
    \boxed{
        R_u^{(1)}
        = 2 \pi \bigg| \frac{\vb{p}_{u0} \cdot \vb{E}}{\hbar} \bigg|^2 \delta(\omega_{u0} - \omega)
    }
\end{aligned}$$

This well-known formula represents **one-photon absorption**:
it peaks at $\omega_{u0} = \omega$, i.e. when one photon $\hbar \omega$
has the exact energy of the transition $\hbar \omega_{u0}$.
Note that this transition is only possible when $\matrixel{u}{\vu{p}}{0} \neq 0$,
i.e. for any odd-numbered final state $\Ket{u}$.


## Two-photon absorption

Next, we go to second-order perturbation theory.
Based on the previous result, this time
all odd-numbered states $\Ket{u}$ are unaffected:

$$\begin{aligned}
    c_u^{(2)}(t)
    &= -\frac{i}{\hbar} \sum_{e}^{\mathrm{even}} \int_0^t \matrixel{u}{\hat{H}_1(\tau)}{e} \: c_e^{(1)}(\tau) \: e^{i \omega_{ue} \tau} \dd{\tau}
    = 0
\end{aligned}$$

While the even-numbered states $\Ket{e}$ have the following correction,
using $\omega_{eu} \!+\! \omega_{u0} = \omega_{e0}$:

$$\begin{aligned}
    c_e^{(2)}(t)
    &= -\frac{i}{\hbar} \sum_{u}^{\mathrm{odd}} \int_0^t \matrixel{e}{\hat{H}_1(\tau)}{u} \: c_u^{(1)}(\tau) \: e^{i \omega_{eu} \tau} \dd{\tau}
    \\
    &= i \sum_{u}^{\mathrm{odd}} \frac{(\vb{p}_{eu} \cdot \vb{E}) (\vb{p}_{u0} \cdot \vb{E})}{\hbar^2 (\omega_{u0} - \omega)}
    \int_0^t e^{i (\omega_{eu} + \omega_{u0} - 2 \omega) \tau} - e^{i (\omega_{eu} - \omega) \tau} \dd{\tau}
    \\
    &= i \sum_{u}^{\mathrm{odd}} \frac{(\vb{p}_{eu} \cdot \vb{E}) (\vb{p}_{u0} \cdot \vb{E})}{\hbar^2 (\omega_{u0} - \omega)}
    \bigg[ \frac{e^{i (\omega_{e0} - 2 \omega) \tau}}{i (\omega_{e0} - 2 \omega)}
    - \frac{e^{i (\omega_{eu} - \omega) \tau}}{i (\omega_{eu} - \omega)} \bigg]_0^t
\end{aligned}$$

The second term represents one-photon absorption between $\Ket{u}$ and $\Ket{e}$.
We do not care about that, so we drop it, leaving only the first term:

$$\begin{aligned}
    \boxed{
        c_e^{(2)}(t)
        \approx \sum_{u}^{\mathrm{odd}} \frac{(\vb{p}_{eu} \cdot \vb{E}) (\vb{p}_{u0} \cdot \vb{E})}{\hbar^2 (\omega_{u0} - \omega)}
        \frac{e^{i (\omega_{e0} - 2 \omega) t} - 1}{\omega_{e0} - 2 \omega}
    }
\end{aligned}$$

As before, we can define a rate $R_e^{(2)}(t)$
for all transitions represented by this term:

$$\begin{aligned}
    R_e^{(2)}(t)
    = \frac{\big| c_e^{(2)}(t) \big|^2}{t}
    = \frac{1}{t} \bigg| \sum_{u}^{\mathrm{odd}} \frac{(\vb{p}_{eu} \cdot \vb{E}) (\vb{p}_{u0} \cdot \vb{E})}{\hbar^2 (\omega_{u0} - \omega)} \bigg|^2
    \cdot \bigg| \frac{e^{i (\omega_{e0} - 2 \omega) t} - 1}{\omega_{e0} - 2 \omega} \bigg|^2
\end{aligned}$$

Which for $t \to \infty$ takes a similar form to Fermi's golden rule,
using the formula we proved:

$$\begin{aligned}
    \boxed{
        R_e^{(2)}
        = 2 \pi \bigg| \sum_{u}^{\mathrm{odd}} \frac{(\vb{p}_{eu} \cdot \vb{E}) (\vb{p}_{u0} \cdot \vb{E})}{\hbar^2 (\omega_{u0} - \omega)} \bigg|^2
        \delta(\omega_{e0} - 2 \omega)
    }
\end{aligned}$$

This represents **two-photon absorption**, since it peaks at $\omega_{e0} = 2 \omega$:
two identical photons $\hbar \omega$ are absorbed simultaneously
to bridge the energy gap $\hbar \omega_{e0}$.
Surprisingly, such a transition can only occur when $\matrixel{e}{\vu{p}}{0} = 0$,
i.e. for any even-numbered final state $\Ket{e}$.
Notice that the rate is proportional to $|\vb{E}|^4$,
so this effect is only noticeable at high light intensities.


## Three-photon absorption

For third-order perturbation theory,
all even-numbered states $\Ket{e}$ are unchanged:

$$\begin{aligned}
    c_e^{(3)}(t)
    &= -\frac{i}{\hbar} \sum_{u}^{\mathrm{odd}} \int_0^t \matrixel{e}{\hat{H}_1(\tau)}{u} \: c_u^{(2)}(\tau) \: e^{i \omega_{eu} \tau} \dd{\tau}
    = 0
\end{aligned}$$

And the odd-numbered states $\Ket{u}$ get the following third-order corrections:

$$\begin{aligned}
    c_u^{(3)}(t)
    &= -\frac{i}{\hbar} \sum_{e}^{\mathrm{even}} \int_0^t \matrixel{u}{\hat{H}_1(\tau)}{e} \: c_e^{(2)}(\tau) \: e^{i \omega_{ue} \tau} \dd{\tau}
    \\
    &= i \sum_{e}^{\mathrm{even}} \sum_{u'}^{\mathrm{odd}}
    \frac{(\vb{p}_{ue} \cdot \vb{E}) (\vb{p}_{eu'} \cdot \vb{E}) (\vb{p}_{u'0} \cdot \vb{E})}{\hbar^3 (\omega_{u'0} - \omega) (\omega_{e0} - 2 \omega)}
    \int_0^t e^{i (\omega_{ue} + \omega_{e0} - 3 \omega) \tau} - e^{i (\omega_{ue} - \omega) \tau} \dd{\tau}
    \\
    &= i \sum_{e}^{\mathrm{even}} \sum_{u'}^{\mathrm{odd}}
    \frac{(\vb{p}_{ue} \cdot \vb{E}) (\vb{p}_{eu'} \cdot \vb{E}) (\vb{p}_{u'0} \cdot \vb{E})}{\hbar^3 (\omega_{u'0} - \omega) (\omega_{e0} - 2 \omega)}
    \bigg[ \frac{e^{i (\omega_{u0} - 3 \omega) \tau}}{i (\omega_{u0} - 3 \omega)}
    - \frac{e^{i (\omega_{ue} - \omega) \tau}}{i (\omega_{ue} - \omega)} \bigg]_0^t
\end{aligned}$$

Once again, the second term is uninteresting,
so we drop it and look at the first term only:

$$\begin{aligned}
    \boxed{
        c_u^{(3)}(t)
        \approx \sum_{e}^{\mathrm{even}} \sum_{u'}^{\mathrm{odd}}
        \frac{(\vb{p}_{ue} \cdot \vb{E}) (\vb{p}_{eu'} \cdot \vb{E}) (\vb{p}_{u'0} \cdot \vb{E})}
        {\hbar^3 (\omega_{u'0} - \omega) (\omega_{e0} - 2 \omega)}
        \frac{e^{i (\omega_{u0} - 3 \omega) t} - 1}{\omega_{u0} - 3 \omega}
    }
\end{aligned}$$

The resulting transition rate $R_u^{(3)}(t)$
is found to have the following familiar form:

$$\begin{aligned}
    R_u^{(3)}(t)
    = \frac{\big| c_u^{(3)}(t) \big|^2}{t}
    = \frac{1}{t} \bigg| \sum_{e}^{\mathrm{even}} \sum_{u'}^{\mathrm{odd}}
    \frac{(\vb{p}_{ue} \cdot \vb{E}) (\vb{p}_{eu'} \cdot \vb{E}) (\vb{p}_{u'0} \cdot \vb{E})}
    {\hbar^3 (\omega_{u'0} - \omega) (\omega_{e0} - 2 \omega)} \bigg|^2
    \cdot \bigg| \frac{e^{i (\omega_{u0} - 3 \omega) t} - 1}{\omega_{u0} - 3 \omega} \bigg|^2
\end{aligned}$$

Applying our formula to this yields the following analogue of Fermi's golden rule:

$$\begin{aligned}
    \boxed{
        R_u^{(3)}
        = 2 \pi \bigg| \sum_{e}^{\mathrm{even}} \sum_{u'}^{\mathrm{odd}}
        \frac{(\vb{p}_{ue} \cdot \vb{E}) (\vb{p}_{eu'} \cdot \vb{E}) (\vb{p}_{u'0} \cdot \vb{E})}
        {\hbar^3 (\omega_{u'0} - \omega) (\omega_{e0} - 2 \omega)} \bigg|^2 \delta(\omega_{u0} - 3 \omega)
    }
\end{aligned}$$

This represents **three-photon absorption**, since it peaks at $\omega_{u0} = 3 \omega$:
three identical photons $\hbar \omega$ are absorbed simultaneously
to bridge the energy gap $\hbar \omega_{u0}$.
This process is similar to one-photon absorption,
in the sense that it can only occur if $\matrixel{u}{\vu{p}}{0} \neq 0$.
The rate is proportional to $|\vb{E}|^6$,
so this effect only appears at extremely high light intensities.


## N-photon absorption

A pattern has appeared in these calculations:
in $N$th-order perturbation theory,
we get a term representing $N$-photon absorption,
with a transition rate proportional to $|\vb{E}|^{2N}$.
Indeed, we can derive infinitely many formulas in this way,
although the results become increasingly unrealistic
due to the dependence on $\vb{E}$.

If $N$ is odd, only odd-numbered destinations $\Ket{u}$ are allowed
(assuming the electron starts in the ground state $\Ket{0}$),
and if $N$ is even, only even-numbered destinations $\Ket{e}$.
Note that nothing has been said about the energies of these states
(other than $\Ket{0}$ being the minimum);
everything is determined by the matrix elements $\matrixel{f}{\vu{p}}{i}$.



## References
1.  R.W. Boyd,
    *Nonlinear optics*, 4th edition,
    Academic Press.
2.  R. Shankar,
    *Principles of quantum mechanics*, 2nd edition,
    Springer.