-
Notifications
You must be signed in to change notification settings - Fork 0
/
Copy pathAMresolved2.tex
601 lines (540 loc) · 40.9 KB
/
AMresolved2.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
\documentclass[twocolumn,english,pra,aps,superscriptaddress,floatfix]{revtex4-1}
\usepackage{amsthm}
\usepackage{amsmath}
\usepackage{graphicx}
\usepackage{amssymb}
%\usepackage{esint}
\usepackage{bm}
\usepackage{latexsym}
\makeatletter
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% Textclass specific LaTeX commands.
\@ifundefined{textcolor}{}
{%
\definecolor{BLACK}{gray}{0}
\definecolor{WHITE}{gray}{1}
\definecolor{RED}{rgb}{1,0,0}
\definecolor{GREEN}{rgb}{0,1,0}
\definecolor{BLUE}{rgb}{0,0,1}
\definecolor{CYAN}{cmyk}{1,0,0,0}
\definecolor{MAGENTA}{cmyk}{0,1,0,0}
\definecolor{YELLOW}{cmyk}{0,0,1,0}
}
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% User specified LaTeX commands.
%\numberwithin{equation}{section}
\makeatother
\usepackage{babel}
\usepackage{braket}
\begin{document}
\author{N. Miladinovic}
\affiliation{Department of Physics and Astronomy, McMaster University, 1280 Main
St.\ W., Hamilton, ON, L8S 4M1, Canada}
\author{D.\ H.\ J.\ O'Dell}
\affiliation{Department of Physics and Astronomy, McMaster University, 1280 Main
St.\ W., Hamilton, ON, L8S 4M1, Canada}
\title{Abraham-Minkowski and the HMW phase}
\begin{abstract}
\label{sec:abstract}
We show that Abraham $P_A=D\times B$ and the Minkowski $P_M=\frac{1}{c^2}E\times H$ electromagnetic momentum are both correct. They are linked to two different representations of the Schr\"{o}dinger equation which differ from each other through a phase factor given by the HMW topological phase.
\end{abstract}
\pacs{42.50.Ex, 42.50.Pq, 42.60.Da, 42.82.Et, 03.67.Lx}
\maketitle
\section{The Lorentz Force}
\label{sec:lorentz}
We begin with the Lorentz force acting on a linear medium due to an electromagnetic plane wave of the form $\mathbf{E}(\mathbf{x},t)=\mathbf{\mathcal{E}}(\mathbf{x},t)\cos{(\omega t -\mathbf{k\cdot x})}$. In this exposition, we will be making use of Maxwell's general equations:
\begin{eqnarray}
&&\mathrm{\mathbf{\nabla}\cdot\mathbf{E}=\frac{1}{\epsilon_0}\rho} \\
&&\mathrm{\mathbf{\nabla}\times\mathbf{E}=-\frac{\partial \mathbf{B}}{\partial t}} \\
&&\mathrm{\mathbf{\nabla}\cdot\mathbf{B}=0} \\
&&\mathrm{\mathbf{\nabla}\times\mathbf{B}=\mu_0\mathbf{J}+\mu_0 \epsilon_0\frac{\partial \mathbf{E}}{\partial t}}
\end{eqnarray}
Along with Maxwell's equations in matter:
\begin{eqnarray}
&&\mathrm{\mathbf{\nabla}\cdot\mathbf{D}=\rho_f} \\
&&\mathrm{\mathbf{\nabla}\times\mathbf{E}=-\frac{\partial \mathbf{B}}{\partial t}} \\
&&\mathrm{\mathbf{\nabla}\cdot\mathbf{B}=0} \\
&&\mathrm{\mathbf{\nabla}\times\mathbf{H}=\mathbf{J}_f+\frac{\partial \mathbf{D}}{\partial t}}
\end{eqnarray}
The general Lorentz force is given by
\begin{eqnarray}
&&\mathrm{\mathbf{f}_i=\rho \mathbf{E}_i+\left(\mathbf{J} \times \mathbf{B}\right)_i}\nonumber \\
&&=\mathrm{\left(\epsilon_0\mathbf{\nabla} \cdot \mathbf{E}\right)\mathbf{E}_i+\left(\frac{1}{\mu_0}\left(\mathbf{\nabla}\times \mathbf{B}\right)\times\mathbf{B}-\epsilon_0\frac{\partial \mathbf{E}}{\partial t} \times \mathbf{B}\right)_i}
\end{eqnarray}
Here we have made use of Eq.\ (1) and Eq.\ (4) in the second line. Using the vector identity $\mathbf{A}\times\left(\mathbf{\nabla}\times \mathbf{A}\right)=\frac{1}{2}\mathbf{\nabla}A^2-\left(\mathbf{A}\cdot\mathbf{\nabla}\right)\mathbf{A}$, along with Eq.\ (3) we can rewrite this as
\begin{eqnarray}
&&\mathrm{\mathbf{f}_i=\epsilon_0\left(\mathbf{\nabla} \cdot \mathbf{E}\right)\mathbf{E}_i-\frac{1}{2\mu_0}\mathbf{\nabla}B^2+\frac{1}{\mu_0}\left(\mathbf{B}\cdot\mathbf{\nabla}\right)\mathbf{B}_i} \nonumber \\
&&\mathrm{-\epsilon_0\frac{\partial}{\partial t} \left(\mathbf{E}\times \mathbf{B}\right)_i-\frac{1}{2}\epsilon_0\mathbf{\nabla}E^2+\epsilon_0\left(\mathbf{E}\cdot\mathbf{\nabla}\right)\mathbf{E}_i}
\end{eqnarray}
For a plane wave $\mathbf{E}(\mathbf{x},t)=\mathbf{\mathcal{E}}(\mathbf{x},t)\cos{(\omega t -\mathbf{k\cdot x})}$ the first, third, and last term will drop out since the electromagnetic field doesn't have a longitudinal component. We are therefore left with
\begin{equation}
\mathrm{\mathbf{f}_i=-\frac{1}{2}\mathbf{\nabla}\left(\epsilon_0 E^2+\frac{1}{\mu_0}B^2\right)-\epsilon_0\frac{\partial}{\partial t} \left(\mathbf{E}\times \mathbf{B}\right)_i}
\end{equation}
Let's now do the same thing, but instead consider the Lorentz force $\tilde{\mathbf{f}}$ acting on only the free charges in the material. Following the same procedure used for the general Lorentz force
\begin{eqnarray}
&&\mathrm{\tilde{\mathbf{f}}_i=\rho_f \mathbf{E}_i+\left(\mathbf{J_f} \times \mathbf{B}\right)_i}\nonumber \\
&&=\mathrm{\left(\mathbf{\nabla} \cdot \mathbf{D}\right)\mathbf{E}_i+\left(\left(\mathbf{\nabla}\times \mathbf{H}\right)\times\mathbf{B}-\frac{\partial \mathbf{D}}{\partial t} \times \mathbf{B}\right)_i} \nonumber \\
&&\mathrm{=\left(\mathbf{\nabla} \cdot \mathbf{D}\right)\mathbf{E}_i-\frac{1}{2}\mathbf{\nabla}\left(\mathbf{H}\cdot\mathbf{B}\right)+\left(\mathbf{H}\cdot\mathbf{\nabla}\right)\mathbf{B}_i} \nonumber \\
&&\mathrm{-\frac{\partial}{\partial t} \left(\mathbf{D}\times \mathbf{B}\right)_i-\frac{1}{2}\mathbf{\nabla}\left(\mathbf{D}\cdot\mathbf{E}\right)+\left(\mathbf{D}\cdot\mathbf{\nabla}\right)\mathbf{E}_i}
\end{eqnarray}
Once again, we drop the first, third, and last term due to electromagnetic plane wave being longitudinal. We then arrive at
\begin{equation}
\mathrm{\tilde{\mathbf{f}}_i=-\frac{1}{2}\mathbf{\nabla}\left(\mathbf{D}\cdot\mathbf{E}+\mathbf{H}\cdot\mathbf{B}\right)-\frac{\partial}{\partial t} \left(\mathbf{D}\times \mathbf{B}\right)_i}
\end{equation}
What happens now if we wish to find the Lorentz force $\check{\mathbf{f}}$ acting on the bound charges?
\begin{eqnarray}
&&\mathrm{\check{\mathbf{f}}_i=\rho_b \mathbf{E}_i+\left(\mathbf{J_b} \times \mathbf{B}\right)_i}\nonumber \\
&&=-\mathrm{\left(\mathbf{\nabla} \cdot \mathbf{P}\right)\mathbf{E}_i+\left(\left(\mathbf{\nabla}\times \mathbf{M}\right)\times\mathbf{B}+\frac{\partial \mathbf{P}}{\partial t} \times \mathbf{B}\right)_i} \nonumber \\
&&\mathrm{=-\left(\mathbf{\nabla} \cdot \mathbf{P}\right)\mathbf{E}_i-\frac{1}{2}\mathbf{\nabla}\left(\mathbf{M}\cdot\mathbf{B}\right)+\left(\mathbf{M}\cdot\mathbf{\nabla}\right)\mathbf{B}_i} \nonumber \\
&&\mathrm{+\frac{\partial}{\partial t} \left(\mathbf{P}\times \mathbf{B}\right)_i+\frac{1}{2}\mathbf{\nabla}\left(\mathbf{P}\cdot\mathbf{E}\right)-\left(\mathbf{P}\cdot\mathbf{\nabla}\right)\mathbf{E}_i}
\end{eqnarray}
Dropping the first, third and last term again yields
\begin{equation}
\mathrm{\check{\mathbf{f}}_i=\frac{1}{2}\mathbf{\nabla}\left(\mathbf{P}\cdot\mathbf{E}-\mathbf{M}\cdot\mathbf{B}\right)+\frac{\partial}{\partial t} \left(\mathbf{P}\times \mathbf{B}\right)_i}
\end{equation}
We see that $\mathbf{f}=\tilde{\mathbf{f}}+\check{\mathbf{f}}$. What does tell us about the Abraham-Minkowski momenta?
Consider the case in which $M=0$. The Lorentz force equations Eq.\ (11), Eq.\ (13), and Eq.\ (15) tell us that the force due to the electromagnetic momentum carried by the plane waves is
\begin{equation}
\mathrm{-\epsilon_0\frac{\partial}{\partial t} \left(\mathbf{E}\times \mathbf{B}\right)_i=-\frac{\partial}{\partial t} \left(\mathbf{D}\times \mathbf{B}\right)_i+\frac{\partial}{\partial t} \left(\mathbf{P}\times \mathbf{B}\right)_i}
\end{equation}
or
\begin{equation}
\mathrm{\frac{\partial}{\partial t}\mathbf{S}_{Min}=\frac{\partial}{\partial t}\mathbf{S}_{Abr}+\frac{\partial}{\partial t}\left(\mathbf{P}\times \mathbf{B}\right)}
\end{equation}
We have arrived at the well known relationship between the Abraham and Minkowski momentum
\begin{equation}
\mathrm{\mathbf{S}_{Min}=\mathbf{S}_{Abr}+\left(\mathbf{P}\times \mathbf{B}\right)}
\end{equation}
The derivation gives some insight into the partitioning of electromagnetic momenta into the Abraham and the Minkowski representations. To better understand the significance of each term let's first consider the energy contained in a linear dielectric medium. The work required to assemble the dielectric system includes 3 different terms. The first is the work required to bring in all the free charges from infinity and assemble them in the proper position. The second term is the work required to assemble the bound charges. The third term is the work required to stretch the polarized atoms - also known as the Stark energy. This total energy is given by
\begin{equation}
\mathrm{\frac{1}{2}\left(\mathbf{D}\cdot\mathbf{E}+\mathbf{H}\cdot\mathbf{B}\right)}
\end{equation}
If instead we are interested in the energy required in assembling only the free and the bound charges in the system piece by piece, the energy is given by
\begin{equation}
\mathrm{\frac{1}{2}\left(\epsilon_0 E^2+\frac{1}{\mu_0}B^2\right)}
\end{equation}
Therefore we see that Eq.\ (11) describes the force on the material as a result of free and bound charges. Eq.\ (15) on the other hand includes the force due to the polarized fields.
\section{The Lagrangian}
\label{sec:Lagrange}
We begin with the Lagrangian for a polarizable, atom interacting with an external electromagnetic field in the lab frame \cite{wei95}
\begin{equation}
\mathrm{L=\frac{1}{2}\mathrm{M}\mathbf{v}^2 + \frac{1}{2}\alpha E^2+\mathbf{m} \cdot B +\int\left(\frac{\epsilon_0}{2}E^2+ \frac{1}{2\mu_0}B^2\right)\,dV} \nonumber \\
\label{lagrangian1}
\end{equation}
Where $\alpha$ is the atomic polarizability, and $m$ is the magnetic moment. In the atom's reference frame the Lorentz transformed fields are $\bar{\mathbf{E}}=\mathbf{E}+\mathbf{v}\times\mathbf{B}$ and $\bar{\mathbf{B}}=\mathbf{B}-\epsilon_0\mu_0\left(\mathbf{v}\times\mathbf{E}\right)$ (to first order in v/c). Switching into the atom's reference frame the Lagrangian becomes
\begin{eqnarray}
\mathrm{L}&=&\mathrm{\frac{1}{2}\mathrm{M}\mathbf{v}^2 + \frac{1}{2}\alpha\left(\mathbf{E}+\mathbf{v}\times\mathbf{B}\right)^2 +\mathbf{m}\cdot\left(\mathbf{B}-\epsilon_0\mu_0\left(\mathbf{v}\times\mathbf{E}\right)\right) } \nonumber \\
&+&\mathrm{\int\left(\frac{\epsilon_0}{2}\left(\mathbf{E}+\mathbf{v}\times\mathbf{B}\right)^2+ \frac{1}{2\mu_0}B^2\right)\,dV}
\label{lagrangian2}
\end{eqnarray}
Expanding this and droppings terms of order $\mathrm{v^2/c^2}$ gives
\begin{eqnarray}
\mathrm{L}&=&\mathrm{\frac{1}{2}\mathrm{M}\mathbf{v}^2 + \frac{1}{2}\alpha E^2 +\mathbf{m}\cdot\mathbf{B}}\nonumber \\
&+&{\int\left(\frac{\epsilon_0}{2}E^2+ \frac{1}{2\mu_0}B^2\right)\,dV+\mathbf{v}\cdot\left(\frac{\mathbf{m}\times\mathbf{E}}{c^2}\right)} \nonumber \\
&+&\mathrm{-\mathbf{v}\cdot\left(\alpha\mathbf{E}\times\mathbf{B}\right)-\epsilon_0\int\mathbf{v}\cdot\left(\mathbf{E}\times\mathbf{B}\right)\,dV}
\label{lagrangian3}
\end{eqnarray}
The last 2 terms may be written as
\begin{equation}
\mathrm{-\int\mathbf{v}\cdot\left(\mathbf{D}\times\mathbf{B}\right)\,dV}
\end{equation}
Here $\mathbf{D}$ is the displacement field. The corresponding Hamiltonian is
\begin{eqnarray}
\mathrm{H}&=&\mathrm{\frac{1}{2M}\left(\mathbf{P}+\int\left(\mathbf{D}\times\mathbf{B}\right)\,dV-\frac{\mathbf{m}\times\mathbf{E}}{c^2}\right)^2} \nonumber \\
&-&\mathrm{\frac{1}{2}\alpha E^2-\mathbf{m}\cdot\mathbf{B}+\int\left(\frac{\epsilon_0}{2}E^2+ \frac{1}{2\mu_0}B^2\right)\,dV}
\label{hamiltonian1}
\end{eqnarray}
Where $P$ is the canonical momentum of system. The corresponding Schr\"{o}dinger equation is
\begin{eqnarray}
\mathrm{i\hbar\dot{\psi}}&=&\mathrm{\frac{1}{2M}\left(\mathbf{P}+ \mathbf{S}_A-\frac{\mathbf{m}\times\mathbf{E}}{c^2}\right)^2\,\Psi-\frac{1}{2}\alpha E^2\,\psi}\nonumber \\
&-&\mathrm{\mathbf{m}\cdot\mathbf{B}\,\psi+\left(\int\left[\frac{\epsilon_0}{2}E^2+ \frac{1}{2\mu_0}B^2\right]\,dV\right)\,\psi}
\label{schrodinger1}
\end{eqnarray}
Where $\mathrm{\mathbf{S}_A=\int \left(\mathbf{D}\times\mathbf{B}\right) \, dV}$ is the Abraham momentum. We preform a unitary transformation by writing the wave function as
\begin{equation}
\mathrm{\psi=\Psi\exp{\left[-\frac{\mathrm{i}}{\mathrm{\hbar}}\int_{\gamma[0,\mathbf{r}]}\mathbf{S}_A\cdot d\mathbf{r}'\right]}}
\label{abrahamrep}
\end{equation}
Where $\gamma$ is an arbitrary path between $[0,\mathbf{r}]$. Substituting this into Eq.\ (\ref{schrodinger1}) and rearranging yields
\begin{eqnarray}
&&\mathrm{i\hbar\dot{\Psi}=\frac{1}{2M}\left(\mathbf{P}-\frac{\mathbf{m}\times\mathbf{E}}{c^2}\right)^2\Psi-\left(\frac{1}{2}\alpha E^2+\mathbf{m}\cdot\mathbf{B}\right)\Psi} \nonumber \\
&&\mathrm{+\left(\int\left[\frac{\epsilon_0}{2}E^2+\frac{1}{2\mu_0}B^2\right]\,dV\right)\Psi-\left(\frac{\partial}{\partial t}\int \mathbf{S}_A\,\mathrm{dV}\right)\Psi} \nonumber \\
\label{schrodinger2}
\end{eqnarray}
We will call this the Abraham representation, and will come back to it shortly, but first we explore another representation. Instead of the unitary transformation given in Eq.\ (\ref{abrahamrep}), let us choose
\begin{equation}
\mathrm{\psi=\Psi\exp{\left[-\frac{\mathrm{i}}{\mathrm{\hbar}}\int_{\gamma[0,\mathbf{r}]}\mathbf{S}_M\cdot d\mathbf{r}'\right]}}
\label{directrep}
\end{equation}
Where $\mathbf{S}_M=\int \left(\epsilon_0\mu_0\,\mathbf{E}\times\mathbf{H}\right) \, dV$ is the Minkowski momentum. Plugging this into Eq.\ (\ref{schrodinger1}) and rearranging yields
\begin{eqnarray}
&&\mathrm{i\hbar\dot{\Psi}=\frac{1}{2M}\left(\mathbf{P}-\mathbf{d}\times\mathbf{B}\right)^2\Psi-\left(\frac{1}{2}\alpha E^2+\mathbf{m}\cdot\mathbf{B}\right)\Psi} \nonumber \\
&&\mathrm{+\left(\int\left[\frac{\epsilon_0}{2}E^2+\frac{1}{2\mu_0}B^2\right]\,dV\right)\Psi-\left(\frac{\partial}{\partial t}\int \mathbf{S}_M\,\mathrm{dV}\right)\Psi} \nonumber \\
\label{schrodinger3}
\end{eqnarray}
This representation will be called the Minkowski representation.
Both the Abraham representation Eq.\ (\ref{schrodinger2}) and the Minkowski representation Eq.\ (\ref{schrodinger3}) are valid in describing the interaction between an electrically neutral, polarizable atom and an electromagnetic field. Although both are valid, the initial wave function to the system differs from one representation to another. Suppose we know the ground state wave function of the system beforehand $\psi_0$. Which formulation of the Hamilton should we use? If we use the Abraham Hamiltonian and naively plug in $\Psi=\psi_0$, we would get an incorrect result. Using the Abraham Hamiltonian forces us to use the initial wave function \cite{boyd}
\begin{equation}
\mathrm{\Psi=\psi_0\exp{\left[-\frac{\mathrm{i}}{\mathrm{\hbar}}\int_0^\mathbf{x} \left(\mathbf{d}\times\mathbf{B}\right) \, \mathrm{dx'}\right]}}
\end{equation}
On the other hand, we are free to use $\Psi=\psi_0$ if we decide to use the direct coupling representation.
\vspace{5mm}
What Eq.\ (\ref{schrodinger3}) and Eq.\ (\ref{schrodinger5}) tell us is that using the direct coupling representation would yield us the Minkowski momentum, while the semi-coupled representation would give the Abraham momentum. Of course we arrive at the well known relation between the A/B momentum and the kinetic/canonical momentum of the polarizable atom.
\begin{equation}
\mathrm{M\dot{\mathbf{r}}+\int \mathbf{\mathcal{S}}_{Abr}\,dV=\mathbf{P}+\int \mathbf{\mathcal{S}}_{Min}\,dV}
\end{equation}
Another interesting feature is that the HMW phase
\begin{equation}
\mathrm{\phi_{HMW}=\exp{\left[-\frac{\mathrm{i}}{\mathrm{\hbar}}\int_0^\mathbf{x}(\mathbf{d}\times\mathbf{B})\,\mathrm{dx'}\right]}}
\label{HMWphase}
\end{equation}
appears naturally in the semi-coupled representation. As was noted above, the HMW phase must be present whenever one uses the canonical momentum rather than the kinetic.
\vspace{5mm}
Perhaps most surprising of all is the manner in which the momentum - be it Abraham or Minkowski - came about in this calculation. We began with a Lagrangian Eq.\ (\ref{lagrangian1}) describing the energy present in the coupled atom-electromagnetic system. Here the Euler-Lagrange equations yield no electromagnetic momentum to speak of
\begin{equation}
\mathrm{M\ddot{\mathbf{x}}=\mathbf{\nabla}\int\frac{1}{2}\left(\mathbf{D\cdot E}+\mathbf{H\cdot B}\right)}
\end{equation}
It was only after we included the Lorentz correction $\mathbf{v}\times\mathbf{B}$ in Eq.\ (\ref{lagrangian2}) that the Euler-Lagrange equation coughed up the contribution due to the electromagnetic momentum.
\begin{equation}
\mathrm{M\ddot{\mathbf{x}}=\mathbf{\nabla}\int\frac{1}{2}\left(\mathbf{D\cdot E}+\mathbf{H\cdot B}\right)+\frac{\partial}{\partial t}\left(\mathbf{D}\times\mathbf{B}\right)}
\end{equation}
Suppose now that the fields $\mathbf{D}$ and $\mathbf{B}$ are static and act perpendicular to each other. We would say that clearly the fields carry no momentum, and yet a momentum term appears in the Euler-Lagrange equation. What is going on?
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
\section{Appendix: Abraham representation expansion}
In this section we go through the derivation of the Abraham Hamiltonian given by Eq.\ (\ref{schrodinger2}). Eq.\ (\ref{schrodinger3}) can be similarly derived. The Schrodinger equation Eq.\ (\ref{schrodinger1}) is given by
\begin{eqnarray}
\mathrm{i\hbar\dot{\psi}}&=&\mathrm{\frac{1}{2M}\left(\mathbf{P}+ \mathbf{S}_A-\frac{\mathbf{m}\times\mathbf{E}}{c^2}\right)^2\,\Psi-\frac{1}{2}\alpha E^2\,\psi}\nonumber \\
&-&\mathrm{\mathbf{m}\cdot\mathbf{B}\,\psi+\left(\int\left[\frac{\epsilon_0}{2}E^2+ \frac{1}{2\mu_0}B^2\right]\,dV\right)\,\psi}
\label{schrodingerappendix1}
\end{eqnarray}
Here we have written
\begin{equation}
\mathrm{\mathbf{S}_A=\int \left(\mathbf{D}\times\mathbf{B}\right) \, dV}
\end{equation}
We then apply the a unitary transformation to the wave function and write $\psi$ as
\begin{equation}
\mathrm{\psi=\Psi\exp{\left[-\frac{\mathrm{i}}{\mathrm{\hbar}}\int_{\gamma[0,\mathbf{r}]}\mathbf{S}_M\cdot d\mathbf{r}'\right]}}
\label{abrahamrepappendix}
\end{equation}
Where the integral is a path integral between $[0,\mathrm{\mathbf{r}}]$
This yields
\begin{eqnarray}
&&\mathrm{\mathrm{i}\mathrm{\hbar}\dot{\Psi}+\Psi\,\left(\frac{\partial}{\partial t}\mathrm{\int_0^\mathbf{x} \mathbf{S}_A \, \mathrm{dx'}}\right)}\nonumber \\
&=&\mathrm{-\hbar^2(\nabla^2\Psi)+2i\hbar(\mathbf{\nabla}\Psi)\mathbf{S}_A+\Psi {S_A}^2+i\hbar\Psi(\mathbf{\nabla}\mathbf{S}_A)} \nonumber \\
&&\mathrm{-2i\hbar(\mathbf{\nabla}\Psi)\mathbf{S}_A-i\hbar\Psi(\mathbf{\nabla}\mathbf{S}_A)-2\Psi{\mathbf{S}_A}^2} \nonumber \\
&&\mathrm{-2i\hbar(\mathbf{\nabla}\Psi)(\mathbf{d}\times\mathbf{B})-i\hbar\Psi(\mathbf{\nabla}(\mathbf{d}\times\mathbf{B}))-2\Psi\mathbf{S}_A(\mathbf{d}\times\mathbf{B}) }\nonumber \\
&&\mathrm{+\Psi\left({\mathbf{S}_A}^2+(\mathbf{d}\times\mathbf{B})^2+2\mathbf{S}_A(\mathbf{d}\times\mathbf{B})\right)}\nonumber \\
&&\mathrm{-\Psi\int\frac{1}{2}\left(\mathbf{D\cdot E +H\cdot B}\right)\,dV}
\label{schrodingera2}
\end{eqnarray}
Here we have omitted writing out the phase factor
\begin{equation}
\mathrm{\exp{\left[-\frac{\mathrm{i}}{\mathrm{\hbar}}\int_0^\mathbf{x}\mathbf{S}_A\,\mathrm{dx'}\right]}}
\label{phasefactor}
\end{equation}
as it appears multiplying every term and will be factored out. Canceling terms leaves us with
\begin{eqnarray}
&&\mathrm{\mathrm{i}\mathrm{\hbar}\dot{\Psi}+\Psi\,\left(\frac{\partial}{\partial t}\mathrm{\int_0^\mathbf{x} \mathbf{S}_A \, \mathrm{dx'}}\right)}\nonumber \\
&=&\mathrm{-\hbar^2(\nabla^2\Psi) -2i\hbar(\mathbf{\nabla}\Psi)(\mathbf{d}\times\mathbf{B})-i\hbar\Psi(\mathbf{\nabla}(\mathbf{d}\times\mathbf{B}))}\nonumber \\
&&\mathrm{+\Psi(\mathbf{d}\times\mathbf{B})^2-\Psi\int\frac{1}{2}\left(\mathbf{D\cdot E +H\cdot B}\right)\,dV}
\label{schrodingera3}
\end{eqnarray}
Factoring terms, this can be rearranged into
\begin{equation}
\mathrm{i\mathrm{\hbar}\dot{\Psi}=\mathrm{\left(\frac{\left(\mathbf{P}+\mathbf{d}\times\mathbf{B}\right)^2}{2M}-\int\frac{1}{2}\left(\mathbf{D\cdot E}+\mathbf{H\cdot B}\right)\,dV-\frac{\partial}{\partial t}\int_0^\mathbf{x} \mathbf{S}_A\,\mathrm{dx'}\right)\Psi}}
\label{schrodingera4}
\end{equation}
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
\section{Appendix: The G\"{o}ppert-Mayer Transformation }
We begin with the minimal coupling Hamiltonian for a system of charges interacting with an electromagnetic field
\begin{equation}
\mathrm{H(t)=\sum_{j}\frac{1}{2m_j}\left[\mathbf{p}_j-q_j \,\mathbf{A}(\mathbf{0},t)\right]^2+V_{c}(\mathbf{x})}
\label{appendixminimalhamiltonian}
\end{equation}
where $\mathbf{A}(\mathbf{x},t)$ is the vector potential, $e_j$ is the charge, and $V_{\mathrm{c}}(\mathbf{x})$ is the scalar potential energy of the system. In the long-wavelength approximation, we have assumed the spatial variation of $\mathbf{A}(\mathbf{x},t)$ is negligible. We therefore choose the location of the system of charges considered to be at $\mathbf{x}=0$ and set $\mathbf{A}(\mathbf{x},t)=\mathbf{A}(\mathbf{0},t)$. The corresponding Schr\"{o}dinger equation for the minimal coupling Hamiltonian is given by
\begin{equation}
\mathrm{i\hbar \dot{\psi}(\mathbf{x},t)=\left[\sum_{j}\frac{1}{2m_j}\left[\mathbf{p}_j-q_j\, \mathbf{A}(\mathbf{0},t)\right]^2+V_{c}(\mathbf{x})\right]\psi(\mathbf{x},t)}
\label{minimalschrodinger}
\end{equation}
The unitary transformation responsible for giving rise to the electric dipole interaction (direct coupling representation) is given by the G\"{o}ppert-Mayer transformation (GMT)
\begin{equation}
\mathrm{\Theta (t)=\mathrm{\exp{\left[\frac{\mathrm{i}}{\mathrm{\hbar}}\sum_jq_j\,\mathbf{r}_j\cdot \mathbf{A}(\mathbf{0},t) \right]}=\exp{\left[\frac{\mathrm{i}}{\mathrm{\hbar}}\mathbf{d}\cdot \mathbf{A}(\mathbf{0},t) \right]}}}
\label{gmtransformation}
\end{equation}
where
\begin{equation}
\mathbf{d}=\sum_{j} e_j \,\mathbf{r}_j
\end{equation}
We rewrite the wave function $\psi$ as
\begin{equation}
\mathrm{\psi(\mathbf{x},t)=\Theta(t)\Psi(\mathbf{x},t)=\exp{\left[\frac{\mathrm{i}}{\mathrm{\hbar}}\mathbf{d}\cdot \mathbf{A}(\mathbf{0},t) \right]}\Psi(\mathbf{x},t)}
\end{equation}
Substituting this into Eq.\ (\ref{minimalschrodinger}) yields
This yields
\begin{eqnarray}
&&\mathrm{\mathrm{i}\mathrm{\hbar}\dot{\Psi}(\mathbf{x},t)\Theta(t)+\Psi(\mathbf{x},t)\Theta(t)\,\left(\mathbf{d}\cdot\mathbf{E}(\mathbf{0},t)\right)=}\nonumber \\
&&\mathrm{-\Psi(\mathbf{x},t)\Theta(t)\left(i\sum_j\frac{e_j}{2m_jc}\mathbf{A}(\mathbf{0},t)\right)^2}\nonumber \\
&-&\mathrm{\nabla\Psi(\mathbf{x},t)\Theta(t)\left(2i\hbar\sum_j\frac{e_j}{2m_j}\mathbf{A}(\mathbf{0},t)\right)}\nonumber \\
&-&\mathrm{\hbar^2\nabla^2\Psi(\mathbf{x},t)\Theta(t)+2\Psi(\mathbf{x},t)\Theta(t)\left(i\sum_j\frac{e_j}{2m_j}\mathbf{A}(\mathbf{0},t)\right)^2}\nonumber \\
&+&\mathrm{\nabla\Psi(\mathbf{x},t)\Theta(t)\left(2i\hbar\sum_j\frac{e_j}{2m_j}\mathbf{A}(\mathbf{0},t)\right)}\nonumber \\
&+&\mathrm{\Psi(\mathbf{x},t)\Theta(t)\left(i\sum_j\frac{e_j}{2m_j}\mathbf{A}(\mathbf{0},t)\right)^2 +\Psi(\mathbf{x},t)\Theta(t)\mathbf{V}_c(\mathbf{x})}\nonumber \\
\end{eqnarray}
Where we have used
\begin{equation}
\mathrm{\mathbf{E}(\mathbf{0},t)=-\frac{\partial \mathbf{A}(\mathbf{0},t)}{\partial t}}
\label{constitutive1}
\end{equation}
Canceling terms and rearranging leaves us with
\begin{eqnarray}
\mathrm{i\hbar\dot{\Psi}(\mathbf{x},t)=\left[\frac{\mathbf{P}^2}{2M}-\mathbf{d}\cdot\mathbf{E}(\mathbf{0},t)+\mathbf{V}_c(\mathbf{x})\right]\Psi(\mathbf{x},t)}
\label{directschrodinger}
\end{eqnarray}
Where $\mathbf{P}=\sum_j\mathbf{p}_j$ and $M=\sum_j m_j$
Here we have arrived at the direct coupling representation of the Hamiltonian. This Hamiltonian however, does not include the radiation energy of the fields themselves. Our treatment of the minimal coupling Hamiltonian Eq.\ (\ref{minimalschrodinger}) may be extended further by including the radiation energy of the fields themselves
\begin{equation}
\mathrm{H_R=\frac{1}{2}\int\left( \epsilon_0\mathbf{E}^2(\mathbf{x},t)+\frac{\mathbf{B}^2(\mathbf{x},t)}{\mu_0}\right)\,dV=\sum_j\hbar\omega_j\left({a_j}^{\dagger}\, a_j+\frac{1}{2}\right)}
\end{equation}
The question then arises, how does the radiation Hamiltonian transform under the G\"{o}ppert-Mayer unitary transformation transformation? Clearly if the fields are treated classically, the GMT Eq.\ (\ref{gmtransformation}) will commute with the electric field. If however, we consider a quantized field, this is no longer true. In order to determine how the fields transform under a quantized field, we must promote the vector potential to an operator \cite{thirunamachandran}
\begin{equation}
\mathrm{\mathbf{A}(\mathbf{x},t)=\sum_j\mathcal{A}_{\omega_j}\left[a_j\,\mathbf{\varepsilon}_j e^{i(\mathbf{k}_j\cdot\mathbf{x}-\omega t)}+{a_j}^{\dagger}\,\mathbf{\varepsilon}_j e^{-i(\mathbf{k}_j\cdot\mathbf{x}-\omega t)}\right]}
\end{equation}
where $\varepsilon$ is the polarization, and
\begin{equation}
\mathrm{\mathcal{A}_{\omega_j}=\left[\frac{\hbar}{2\epsilon_0 L^3\omega_j}\right]^{\frac{1}{2}}}
\end{equation}
The transverse electric field in the Coulomb gauge is given by
\begin{equation}
\mathrm{\mathbf{E}_{\perp}(\mathbf{x},t)=i\sum_{j,\mu}\mathcal{E}_{\omega_j}\left[a_j\,\mathbf{\varepsilon}_j e^{i(\mathbf{k}_j\cdot\mathbf{x}-\omega t)}-{a_j}^{\dagger}\,\mathbf{\varepsilon}_j e^{-i(\mathbf{k}_j\cdot\mathbf{x}-\omega t)}\right]}
\end{equation}
where
\begin{equation}
\mathrm{\mathcal{E}_{\omega_j}=\left[\frac{\hbar\omega_j}{2\epsilon_0 L^3}\right]^{\frac{1}{2}}}
\end{equation}
From here it is necessary to determine the commutation relation between the vector potential and electric field.
\begin{eqnarray}
&&\mathrm{[\mathbf{A}(\mathbf{x},t),\mathbf{E}_{\perp}(\mathbf{x}',t)]=} \nonumber \\
&&\mathrm{\frac{i\hbar}{2\epsilon_0 L^3}\sum_{j_{\perp},j'_{\perp}}\left(\left[a^{\dagger}_{j},a_{j'}\right]\,e^{i\mathbf{k}_{j'}\cdot\mathbf{x'}}e^{-i\mathbf{k}_{j}\cdot\mathbf{x}}\right)+} \nonumber \\
&&\mathrm{\frac{i\hbar}{2\epsilon_0 L^3}\sum_{j_{\perp},j'_{\perp}}\left(\left[a^{\dagger}_{j'},a_{j}\right]\,e^{i\mathbf{k}_{j}\cdot\mathbf{x}}e^{-i\mathbf{k}_{j'}\cdot\mathbf{x'}}\right)=} \nonumber \\
&&\mathrm{\frac{i\hbar}{2\epsilon_0 L^3}\sum_{j_{\perp}}\left(e^{i\mathbf{k}_j\cdot(\mathbf{x}-\mathbf{x}')}+e^{-i\mathbf{k}_j\cdot(\mathbf{x}-\mathbf{x}')}\right)} \nonumber \\
\end{eqnarray}
Where we have made use of the relation $[a_j,a^{\dagger}_{j'}]=\delta_{j,j'}$. In the summation, the notation $j_{\perp}$ indicates that we are taking the sum over the transverse modes. We convert our sum into an integral through \cite{loudonbook}
\begin{equation}
\mathrm{\sum_{k}\rightarrow \frac{L^3}{(2\pi)^3}\int\,d^3k}
\end{equation}
and we make use of the transverse delta function
\begin{equation}
\mathrm{\delta_{\perp}(\mathbf{x}-\mathbf{x}')=\frac{1}{(2\pi)^3}\int\,d^3k_{\perp}\,e^{i\mathbf{k}_j\cdot(\mathbf{x}-\mathbf{x}')}}
\end{equation}
Therefore we find
\begin{equation}
\mathrm{[\mathbf{A}(\mathbf{x,t}),\mathbf{E}_{\perp}(\mathbf{x}',t)]=-\frac{i\hbar}{\epsilon_0}\delta_{\perp}(\mathbf{x}-\mathbf{x}')}
\label{comm1}
\end{equation}
In order to determine how the electric field transforms under the GMT $\Theta(t)$ we use the property that for any two operators $\mathbf{A}$ and $\mathbf{B}$
\begin{equation}
\exp{(i\mathbf{A})}\,\mathbf{B}\exp{(-i\mathbf{A})}=\mathbf{B}+i[\mathbf{A},\mathbf{B}]+...
\label{comm2}
\end{equation}
Using Eq.\ (\ref{comm1}) and Eq.\ (\ref{comm2}) we find
\begin{equation}
\mathrm{\Theta(t)\mathbf{E}_{\perp}(\mathbf{x},t)\Theta^{\dagger}(t)=\mathbf{E}_{\perp}(\mathbf{x},t)+\frac{1}{\epsilon_0}\mathbf{d}_{\perp}(\mathbf{x},t)=\frac{\mathbf{D}(\mathbf{x},t)}{\epsilon_0}}
\end{equation}
Where $\mathbf{D}$ is the displacement field. Note that for a neutral system $\nabla\cdot\mathbf{D}=0$ and therefore the displacement field is fully transverse which allows us to drop the perpendicular suffix.
It can easily be checked that
\begin{eqnarray}
&&\mathrm{\Theta(t)\mathbf{B}(\mathbf{x},t)\Theta^{\dagger}(t)=\mathbf{B}(\mathbf{x},t)} \\
&&\mathrm{\Theta(t)\mathbf{A}(\mathbf{x},t)\Theta^{\dagger}(t)=\mathbf{A}(\mathbf{x},t)} \\
&&\mathrm{\Theta(t)\mathbf{P}(\mathbf{x},t)\Theta^{\dagger}(t)=\mathbf{P}(\mathbf{x},t)}
\end{eqnarray}
This allows us to deduce that
\begin{equation}
\mathrm{\Theta(t)\mathbf{D}(\mathbf{x},t)\Theta^{\dagger}(t)=\epsilon_0\mathbf{E}(\mathbf{x},t)}
\end{equation}
We can now preform a generalized G\"{o}ppert-Mayer transformation for the total minimal coupling Hamiltonian
\begin{eqnarray}
&&\mathrm{H_{Min}=H_0+H_R}\nonumber \\
&&=\mathrm{\sum_{j}\frac{1}{2m_j}\left[\mathbf{p}_j-q_j \,\mathbf{A}(\mathbf{0},t)\right]^2+V_{c}(\mathbf{x})}\nonumber \\
&&\mathrm{+\frac{1}{2}\int\left( \epsilon_0\mathbf{E}^2(\mathbf{x},t)+\frac{\mathbf{B}^2(\mathbf{x},t)}{\mu_0}\right)\,dV}
\label{fullminimal}
\end{eqnarray}
We begin with the Schr\"{o}dinger equation
\begin{equation}
i\hbar\frac{\partial}{\partial t}\psi=H\,\psi
\end{equation}
and rewrite the wave function as
\begin{equation}
\mathrm{\psi=e^{\frac{i}{\hbar}\mathbf{d}\cdot\mathbf{A}}\Phi}
\end{equation}
This allows us to express the Schr\"{o}dinger equation in terms of the wave function $\Phi$
\begin{eqnarray}
&&\mathrm{i\hbar\frac{\partial}{\partial t}\Phi=i\hbar\frac{\partial}{\partial t}\left(e^{\frac{i}{\hbar}\mathbf{d}\cdot\mathbf{A}}\psi\right)}\nonumber \\
&=&\mathrm{i\hbar e^{\frac{i}{\hbar}\mathbf{d}\cdot\mathbf{A}}\frac{\partial}{\partial t}\psi-e^{\frac{i}{\hbar}\mathbf{d}\cdot\mathbf{A}}\left(\dot{\mathbf{d}}\cdot\mathbf{A}+\mathbf{d}\cdot\dot{\mathbf{A}}\right)\psi} \nonumber \\
&=&\mathrm{e^{\frac{i}{\hbar}\mathbf{d}\cdot\mathbf{A}}\,H\,\psi-e^{\frac{i}{\hbar}\mathbf{d}\cdot\mathbf{A}}\left(\dot{\mathbf{d}}\cdot\mathbf{A}+\mathbf{d}\cdot\mathbf{E}\right)\psi} \nonumber \\
&=&\mathrm{e^{\frac{i}{\hbar}\mathbf{d}\cdot\mathbf{A}}\,H\,e^{-\frac{i}{\hbar}\mathbf{d}\cdot\mathbf{A}}\Phi-e^{\frac{i}{\hbar}\mathbf{d}\cdot\mathbf{A}}\left(\dot{\mathbf{d}}\cdot\mathbf{A}+\mathbf{d}\cdot\mathbf{E}\right)e^{-\frac{i}{\hbar}\mathbf{d}\cdot\mathbf{A}}\Phi} \nonumber \\
\end{eqnarray}
The Hamiltonian $\mathrm{H_{Min}}$ transforms under the generalized GMT in the same way that that it did when the fields were considered classical, with the exception that $\mathbf{E}\rightarrow\mathbf{D}/\epsilon_0$. Therefore under the generalized GMT, the minimal coupling Hamiltonian Eq.\ (\ref{fullminimal}) is transformed into the direct coupling Hamiltonian
\begin{eqnarray}
&&\mathrm{H_{DC}=\frac{\mathbf{P}^2}{2M}-\mathbf{d}\cdot\mathbf{D}(\mathbf{0},t)}\nonumber \\
&&\mathrm{+\frac{1}{2}\int\left( \frac{\mathbf{D}^2}{\epsilon_0}(\mathbf{x},t)+\frac{\mathbf{B}^2(\mathbf{x},t)}{\mu_0}\right)\,dV} \nonumber \\
&&\mathrm{+\mathbf{V}_c(\mathbf{x})+\dot{\mathbf{d}}\cdot\mathbf{A}}
\end{eqnarray}
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
\section{Appendix: Dipole Direct Coupling Derivation}
In this section we go through the derivation of the direct coupling Hamiltonian. We begin with the Schr\"{o}dinger equation in the minimal coupling representation
\begin{equation}
\mathrm{i\hbar\dot{\psi}=\sum_j\left[\frac{1}{2m_j}\left(\mathbf{P}_j-q_j\mathbf{A}(\mathbf{x},t)\right)^2+\mathbf{V}_c(\mathbf{x})\right]\psi}
\label{minschrodinger1}
\end{equation}
Where $\mathbf{V}_c$ is the Coulomb potential. We write the wave function as
\begin{equation}
\mathrm{\psi=\Psi\exp{\left[-\frac{\mathrm{i}}{\mathrm{\hbar}}\sum_j\,\int q_j\mathbf{A}(\mathbf{x},t)\cdot\mathrm{dr}\right]}}
\label{directwavefunction}
\end{equation}
Plugging this into the Schr\"{o}dinger equation Eq.\ (\ref{minschrodinger1})
\begin{eqnarray}
&&\mathrm{\mathrm{i}\mathrm{\hbar}\dot{\Psi}+\Psi\,\frac{\partial}{\partial t}\left(\sum_j\int q_j\mathbf{A}(\mathbf{x},t) \cdot\mathbf{dr}\right)}\nonumber \\
&=&\mathrm{-\hbar^2(\nabla^2\Psi)+2i\hbar(\mathbf{\nabla}\Psi)\left(\sum_j q_j\mathbf{A}(\mathbf{x},t)\right)+\Psi \left(\sum_j q_j\mathbf{A}(\mathbf{x},t)\right)^2 }\nonumber \\
&&+\mathrm{i\hbar\Psi\mathbf{\nabla}\left(\sum_j q_j\mathbf{A}(\mathbf{x},t)\right)
-2i\hbar(\mathbf{\nabla}\Psi)\left(\sum_j q_j\mathbf{A}(\mathbf{x},t)\right)}\nonumber \\
&&\mathrm{-2\Psi\left(\sum_j q_j\mathbf{A}(\mathbf{x},t)\right)-i\hbar\Psi\mathbf{\nabla}\left(\sum_j q_j\mathbf{A}(\mathbf{x},t)\right)}\nonumber \\
&&+\mathrm{\Psi\left(\sum_j q_j\mathbf{A}(\mathbf{x},t)\right)^2+
\Psi\mathbf{V}_c(\mathbf{x})}
\label{minschrodinger2}
\end{eqnarray}
Here we have omitted writing out the phase factor
\begin{equation}
\mathrm{\exp{\left[-\frac{\mathrm{i}}{\mathrm{\hbar}}\sum_j\,\int q_j\mathbf{A}(\mathbf{x},t)\cdot\mathrm{dr}\right]}}
\label{phasefactor}
\end{equation}
as it appears multiplying every term and will be factored out. Canceling terms and rearranging leaves us with
\begin{equation}
\mathrm{i\mathrm{\hbar}\dot{\Psi}=\mathrm{\left[\frac{\mathbf{P}^2}{2M}-\frac{\partial}{\partial t}\left(\sum_j\int q_j\mathbf{A}(\mathbf{x},t) \cdot\mathbf{dr}\right]+\mathbf{V}_c(\mathbf{x})\right)\Psi}}
\end{equation}
Where $\mathbf{P}=\sum_j \mathbf{P}_j$, and $M=\sum_j m_j$. We now make use of the relationship $\partial_t \mathbf{A}(\mathbf{x},t)=-\mathbf{E}(\mathbf{x},t)$ to rewrite the direct coupling Hamiltonian as
\begin{equation}
\mathrm{i\mathrm{\hbar}\dot{\Psi}=\mathrm{\left(\frac{\mathbf{P}^2}{2M}-\left[\sum_j\int q_j\mathbf{E}(\mathbf{x},t) \cdot\mathbf{dr}\right]+\mathbf{V}_c(\mathbf{x})\right)\Psi}}
\end{equation}
Here we assume that we are dealing with a neutral dielectric material which allows us to write
\begin{equation}
\mathrm{\sum_j\int q_j\mathbf{E}(\mathbf{x},t) \cdot\mathbf{dr}=\sum_{\alpha}\int q_{\alpha}\left[\mathbf{E}_{+}-\mathbf{E}_{-}\right]_{\alpha}\cdot\mathbf{dr}}
\end{equation}
where we have now grouped the bound charges into the dipole pairings which are indexed by $\alpha$. This is nothing but the dipole energy $-\mathbf{d}\cdot\mathbf{E}$
where
\begin{equation}
\mathbf{d}=\sum_{j} e_j \,\mathbf{r}_j
\end{equation}
With this approximation we can write the direct coupling Hamiltonian in it's final form
\begin{equation}
\mathrm{i\mathrm{\hbar}\dot{\Psi}=\mathrm{\left(\frac{\mathbf{P}^2}{2M}-\mathbf{d}\cdot\mathbf{E}(\mathbf{x},t)+\mathbf{V}_c(\mathbf{x})\right)\Psi}}
\end{equation}
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
\section{Appendix: Gauge Transformation - Direct Coupling}
In this section, we outline the derivation of the direct coupling Hamiltonian through the use of the G\"{o}ppert-Mayer gauge. We begin with the minimal coupling Lagrangian
\begin{eqnarray}
\mathrm{L_{min}}&=&\mathrm{\frac{1}{2}\sum_n m_n\dot{\mathbf{r}}_n-V_c+\sum_n\left[q_n\dot{\mathbf{r}}_n\mathbf{A}(\mathbf{r}_n,t)-q_nU(\mathbf{r}_n,t)\right]} \nonumber \\
&+&\mathrm{\frac{\epsilon_0}{2}\int\left[\dot{\mathbf{A}}^2(\mathbf{r},t)-c^2\left(\nabla\times\mathbf{A}(\mathbf{r},t)\right)^2\right]d^3\mathbf{r}}
\label{minimallagrangian1}
\end{eqnarray}
Where $\mathbf{A}$ and $U$ are the vector and scalar potential for the external fields, and $V_{\mathrm{c}}$ is the Coulomb potential.
The corresponding Hamiltonian is given by
\begin{eqnarray}
\mathrm{H_{min}}&=&\mathrm{\sum_n \frac{1}{2m_n}\left[\mathbf{p}_n-q_n\mathbf{A}(\mathbf{r_n},t)\right]^2+V_{c}+\sum_nq_nU(\mathbf{r}_n,t)}\nonumber \\
&+&\mathrm{\frac{\epsilon_0}{2}\int\left[\dot{\mathbf{A}}^2(\mathbf{r},t)-c^2\left(\nabla\times\mathbf{A}(\mathbf{r},t)\right)^2\right]d^3\mathbf{r}}
\label{minimalhamiltonian1}
\end{eqnarray}
Here the conjugate momentum and conjugate field are given by
\begin{eqnarray}
&&\mathrm{\mathbf{p}_n=m_n\dot{\mathbf{r}}_n+q_n\mathbf{A}(\mathbf{r}_n,t)} \\
&&\mathrm{\mathbf{\Pi}(\mathbf{r},t)=\epsilon_0 \dot{\mathbf{A}}(\mathbf{r},t)=-\epsilon_0\mathbf{E}(\mathbf{r},t)}\\
\end{eqnarray}
Let us now preform the following gauge transformation
\begin{eqnarray}
&&\mathrm{\mathbf{A}'(\mathbf{r_n},t)=\mathbf{A}(\mathbf{r_n},t+\nabla\chi(\mathbf{r}_n,t)} \\
&&\mathrm{U'(\mathbf{r_n},t)=U(\mathbf{r_n},t)+\frac{\partial}{\partial t}\chi(\mathbf{r}_n,t)}
\end{eqnarray}
Substituting this into the minimal coupling Lagrangian Eq.\ (\ref{minimallagrangian1}), the transformed Lagrangian becomes
\begin{eqnarray}
\mathrm{L_{dir}}&=&\mathrm{\frac{1}{2}\sum_n m_n\dot{\mathbf{r}}_n-V_c+\sum_n\left[q_n\dot{\mathbf{r}}_n\mathbf{A}'(\mathbf{r}_n,t)-q_nU'(\mathbf{r}_n,t)\right]} \nonumber \\
&+&\mathrm{\frac{\epsilon_0}{2}\int\left[\dot{\mathbf{A}}^2(\mathbf{r},t)-c^2\left(\nabla\times\mathbf{A}(\mathbf{r},t)\right)^2\right]d^3\mathbf{r}} \nonumber \\
&=&\mathrm{L_{min}+\frac{d}{dt}\left[\sum_n q_n\,\chi(\mathbf{r}_n,t)\right]}
\label{minimallagrangian2}
\end{eqnarray}
This shows that the gauge transformation generated by $\chi$ is equivalent to adding
\begin{equation}
\mathrm{\frac{d}{dt}\left[\sum_n q_n\,\chi(\mathbf{r}_n,t)\right]}
\end{equation}
to the Lagrangian. The direct coupling Lagrangian is obtained through the G\"{o}ppert-Mayer generator
\begin{equation}
\mathrm{\chi(\mathbf{r},t)=-\int \mathbf{P}^{\perp}(\mathbf{r},t)\cdot\mathbf{A}(\mathbf{r},t)\,d^3\mathbf{r}}
\end{equation}
Where $\mathbf{P}(\mathbf{r})=\sum_n q_n\left(r-R_n\right)\delta(\mathbf{r}-\mathbf{R}_n)$ is the electric polarization vector field. The transformed conjugate momentum and conjugate field become
\begin{eqnarray}
&&\mathrm{\mathbf{p}_n=m_n\dot{\mathbf{r}}} \\
&&\mathrm{\mathbf{\Pi}(\mathbf{r},t)=\epsilon_0 \dot{\mathbf{A}}(\mathbf{r},t)-\mathbf{P}^{\perp}(\mathbf{r},t)=-\mathbf{D}(\mathbf{r},t)}
\end{eqnarray}
Where $\mathbf{D}$ is the displacement field. Note that for a neutral system $\nabla\cdot\mathbf{D}=0$ and therefore the displacement field is fully transverse which allows us to drop the perpendicular suffix. From here we can construct the transformed Hamiltonian
\begin{eqnarray}
\mathrm{H_{dir}}&=&\mathrm{\sum_n\mathbf{p}_n\cdot\dot{\mathbf{r}}_n+\int\Pi(\mathbf{r},t)\cdot\dot{\mathbf{A}}(\mathbf{x},t)\,d^3\mathbf{r}\,-L_{dir}} \nonumber \\
&=&\mathrm{\sum_n \frac{1}{2m_n}\mathbf{p}^2_n+V_{c}-\frac{1}{\epsilon_0}\int \mathbf{P}^{\perp}(\mathbf{r},t)\cdot\mathbf{D}(\mathbf{r},t)\,d^3\mathbf{r}}\nonumber \\
&+&\mathrm{\frac{\epsilon_0}{2}\int\left[\mathbf{D}^2(\mathbf{r},t)-c^2\left(\nabla\times\mathbf{A}(\mathbf{r},t)\right)^2\right]d^3\mathbf{r}}
\label{minimalhamiltonian1}
\end{eqnarray}
This is the direct coupling Hamiltonian.
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
\section{Appendix: Gauge Transformation}
The two state Hamiltonian is
\begin{equation}
\mathrm{H(\mathbf{r})}=\mathrm{\left(\frac{P^2}{2M}+V\right)I \,+\frac{\hbar\Omega}{2}}
\begin{bmatrix}
&\cos{\theta(\mathbf{r})}& &e^{-i\theta(\mathbf{r})}\sin{\theta(\mathbf{r})}& \\
&e^{i\theta(\mathbf{r})}\sin{\theta(\mathbf{r})}& &-\cos{\theta(\mathbf{r})}&
\end{bmatrix} \\
\end{equation}
where $I$ is the identity matrix, and $\theta(\mathbf{r})$ is defined by
\begin{eqnarray}
&&\mathrm{\cos{\left(2\theta(\mathbf{r})\right)}=\frac{-\delta}{\Omega_r(\mathbf{r})}} \\
&&\mathrm{\sin{\left(2\theta(\mathbf{r})\right)}=\frac{\Omega(\mathbf{r})}{\Omega_r(\mathbf{r})}}
\end{eqnarray}
Here $\Omega$ is the Rabi frequency and $\delta$ is the detuning between the laser and atomic frequency. The eigenenergies are given by
\begin{equation}
\mathrm{E_{\pm}=\pm\frac{\hbar\Omega(\mathbf{r})}{2}}
\end{equation}
and the corresponding eigenvectors are
\begin{eqnarray}
&&\mathrm{\ket{\chi_1(\mathbf{r})}=}
\begin{bmatrix}
\mathrm{\cos{\left(\frac{\theta(\mathbf{r})}{2}\right)} }\\
\mathrm{e^{i\phi(\mathbf{r})}\sin{\left(\frac{\theta(\mathbf{r})}{2}\right)}}
\end{bmatrix} \\
&&\mathrm{\ket{\chi_2(\mathbf{r})}=}
\begin{bmatrix}
\mathrm{-e^{-i\phi(\mathbf{r})}\sin{\left(\frac{\theta(\mathbf{r})}{2}\right)}}\\
\mathrm{\cos{\left(\frac{\theta(\mathbf{r})}{2}\right)} }
\end{bmatrix}
\end{eqnarray}
These states form an orthonormal basis, and thus we can expand the state vector $\ket{\Psi}$ in terms of the eigenvectors
\begin{equation}
\mathrm{\ket{\Psi(\mathbf{r},t)}=\sum_{l=1}^2\psi_j(\mathbf{r},t)\ket{\chi_1(\mathbf{r})}}
\end{equation}
\begin{thebibliography}{8}
\bibitem{wei95}{H. Wei, R. Han, and X. Wei, Phys. Rev. Lett. \textbf{75}, 2071 (1995).}
\bibitem{barnett}{S. M. Barnett, Phys. Rev. Lett. 104, 070401 (2010).}
\bibitem{chiao}{J. C. Garrison and R. Y. Chiao, Phys. Rev. A 70, 053826 (2004).}
\bibitem{mansuripur}{M. Mansuripur and A. R. Zakharian, Phys. Rev. E 79, 026608 (2009).}
\bibitem{ketterle}{G. K. Campbell, A. E. Leanhardt, J. Mun, M. Boyd, E. W. Streed, W. Ketterle, and D. E. Pritchard, Phys. Rev. Lett. 94, 170403 (2005).}
\bibitem{feng}{W. She, J. Yu, and R. Feng, Phys. Rev. Lett. 101, 243601 (2008).}
\bibitem{hinds}{E. Hinds and S. M. Barnett, Phys. Rev. Lett. 102, 050403 (2009).}
\bibitem{loudon}{S. M. Barnett and R. Loudon, Phil. Trans. R. Soc. A, 368 (2011).}
\bibitem{lang73}{R. Lang, M. O. Scully and W. E. Lamb, Phys. Rev. A \textbf{7}, 1788 (1973).}
\bibitem{domokos08}{J. K. Asboth, H. Ritsch, and P. Domokos, Pyhs. Rev. A \textbf{77}, 063424 (2008).}
\bibitem{wilkens94}{M. Wilkens, Phys. Rev. Lett. \textbf{72}, 5 (1994).}
\bibitem{griffiths}{David J. Griffiths, \textit{Introduction to Electrodynamics} Third Edition (Prentice Hall, New Jersey, 1999).}
\bibitem{cohentannoudji}{C. Cohen-Tannoudji, J. Dupont-Roc, G. Grynberg, \textit{Atom-Photon Interactions} (Wiley Professional, 1989).}
\bibitem{us}{N. Miladinovic, F. Hasan, N. Chisholm, E. A. Hinds, and D. H. J. O'Dell, Phys. Rev. A \textbf{84}, 043822 (2011).}
\bibitem{preble07}{S. F. Preble, Q. Xu, and M. Lipson, Nat. Photon. \textbf{1}, 293 (2007).}
\bibitem{pfeifer07}{Robert N. C. Pfeifer, Timo A. Nieminen, Norman R. Heckenberg, and Halina Rubinsztein-Dunlop, Rev. Mod. Phys. \textbf{79}(4) 1197-1216 (2007)}
\bibitem{loudon2}{M. Padgett, S. M. Barnett and R. Loudon, J. Mod. Opt.\textbf{50} ,1555 (2003).}
\bibitem{boyd}{K Rz\c{a}\`zewski and R. W. Boyd, J. Mod. Opt. 51, \textbf{8}, 1137 (2004).}
\bibitem{thirunamachandran}{D.P. Craig, and T. Thirunamachandran, \textit{Molecular Quantum Electrodynamics} (Dover Publications, 1984).}
\bibitem{loudonbook}{Rodney Loudon, \textit{The Quantum theory of Light} (Oxford, 1973).}
\bibitem{dalibard}{Jean Dalibard, and Fabrice Gerbier, Rev. Mod. Phys. \textbf{83}, 1523 (2011).}
\end{thebibliography}
\end{document}