-
Notifications
You must be signed in to change notification settings - Fork 2
/
Copy path2008fall_p2.tex
747 lines (696 loc) · 29.2 KB
/
2008fall_p2.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
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
%%%% Problem 1
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
\problem{1}
\subsubsection{Question}
% Keywords
\index{mechanics!Bead on a hoop}
A particle of mass $m$ is constrained to move without friction on a circular
wire of radius $R$ rotating with constant angular frequency $ω$ about a
vertical diameter. Gravity can not be neglected.
\begin{enumerate}[a)]
\item
Write down the Lagrangian for the system and the equations of motion.
\item
Find the equilibrium position(s) of the particle and determine
whether this position is stable.
\item
Calculate the frequency of small oscillations about any stable points.
\end{enumerate}
\begin{figure}[H]
\centering
\begin{tikzpicture}
% The axis
\draw [dashed,->] (0,-1) -- (0,4);
% The hoop
\draw (0,1.5) circle (1.5);
% The bead
\draw ($(0,1.5)!1.5cm!30:(0,0)$)
node[circle,fill=black,anchor=center,inner sep=2pt] {}
coordinate (bead);
% Label the angle
\draw [dashed] (0,1.5) -- (bead);
\draw ($(0,1.5)!0.5cm!(0,0)$)
coordinate (arc start)
arc (-90:-60:0.5cm)
coordinate (arc end);
\draw ($(arc start)!0.5!(arc end)$)
node [anchor=north] {$\,\,θ$};
\end{tikzpicture}
\end{figure}
\subsubsection{Answer}
To start constructing the Lagrangian and equations of motion, we first specify
the kinetic and potential energies. For the kinetic energy, there is an energy
associated with the rotation about the axis and one along the bead. These
combined to give
\begin{align*}
T &= \frac 12 m (Rω\sin θ)² + \frac 12 m (R\dot θ)² \\
{} &= \frac 12 m R² ω² \sin² θ + \frac 12 m R² {\dot θ}²
\end{align*}
The potential energy is all gravitational, so
\begin{align*}
V &= -mgR\cos θ
\end{align*}
where the zero point was taken to be at the center of the hoop to avoid adding
extra constant terms to the Lagrangian. Combining the two, we get
\begin{align}
\boxed{
\mathcal L = \frac 12 mR²ω²\sin² θ + \frac 12 mR²{\dot θ}² + mgR\cos θ
}
\end{align}
Taking the appropriate derivatives in $θ$, the equation of motion is
\begin{align}
\boxed{
\ddot θ = ω² \sin θ \cos θ - \frac{g}{R}\sin θ
}
\end{align}
In order to determine any possible stable points, we note that a stable point
is a place where the angle does not change in time. Since this also equates to
$\ddot θ = 0$, we set the equation above equal to zero and solve for the angles
which satisfy this condition. They end up being the trivial $θ = \{0, π\}$
where the sine function is zero as well as
\begin{align*}
\cos θ₀ &= \frac{g}{Rω²}
\end{align*}
The three stable points are then
\begin{align}
\boxed{ θ₀ = \left\{ 0, \arccos(\frac{g}{Rω²}), π \right\} }
\end{align}
To determine the stability of each, we must determine whether we get
oscillatory or exponential solutions to the differential equation of motion.
To do this, we suppose the angle $θ$ is composed of the equilibrium angle
$θ₀$ and a small perturbation $δ$. Expanding the equation in terms of this,
\begin{align*}
\ddot δ = ω²\sin(θ₀+δ)\cos(θ₀+δ) - \frac{g}{R}\sin(θ₀+δ)
\end{align*}
Using several trigonometric expansions, the equation can be expanded into the
form
\begin{align*}
\ddot δ = ω²\left[ \cos θ₀\sin θ₀ (\cos²δ - \sin²δ) + \cos δ\sin δ
(\cos²θ₀ - \sin²θ₀) \right] - \frac{g}{R}\left[ \sin θ₀\cos δ +
\cos θ₀\sin δ \right]
\end{align*}
For the case where $θ₀ = 0$,
\begin{align*}
\ddot δ &= ω² \cos δ \sin δ - \frac{g}{R} \sin δ
\intertext{Expanding to first order in $δ ≈ 0$,}
\ddot δ &= -(\frac{g}{R} - ω²)δ
\end{align*}
Therefore, the equilibrium point $θ₀ = 0$ is only stable if $ω <
\sqrt{\frac{g}{R}}$.
Likewise for for $θ₀ = π$,
\begin{align*}
\ddot δ &= (\frac{g}{R} + ω²) δ
\end{align*}
The coefficient on $δ$ will never be negative, so the angle $θ₀ = π$ will
be unstable under all conditions.
For the final angle where $θ₀ = \arccos(\frac{g}{Rω²})$, we must do several
substitutions and expansions. $\cos θ₀$ is trivial. $\sin θ₀$ ends up being
$\sqrt{Rω² - g²}/(Rω²)$ by triangle relations. If we substitute these in plus
do an expansion to first order for small $δ$, we get the equation
\begin{align*}
\ddot δ &= ω² \left[ \frac{g\sqrt{Rω²-g²}}{R²ω²} + δ \frac{2g²-Rω²}{R²ω²}
\right] - \frac{g}{R} \left[\frac{\sqrt{Rω²-g²}}{R²ω²} + δ
\frac{g}{Rω²} \right]
\end{align*}
If we consider only the homogeneous terms dependent on $δ$,
\begin{align*}
\ddot δ &= \frac{g² - Rω²}{R²ω²} δ
\end{align*}
This equation is stable if and only if the coefficient on $δ$ is negative, so
it must be that $ω > \frac{g}{\sqrt{R}}$.
In summary, the equilibrium points have the following conditions:
\begin{align}
\boxed{ θ₀ = 0 \quad\quad \text{Stable iff } ω < \sqrt{\frac gR} }
\end{align}
\begin{align}
\boxed{ θ₀ = \arccos(\frac{g}{Rω²}) \quad\quad \text{Stable iff }
ω > \frac{g}{\sqrt{R}} }
\end{align}
\begin{align}
\boxed{ θ₀ = π \quad\quad \text{Never stable} }
\end{align}
About the two stable points, we simply use the coefficient that has already
been isolated to determine the frequency of the oscillations about that point.
\begin{align}
\boxed{ ω₁ = \sqrt{\frac{g}{R} - ω²} \quad\quad\text{for $θ₀ = 0$} }
\end{align}
\begin{align}
\boxed{ ω₂ = \sqrt{\frac{Rω² - g²}{R²ω²}}\quad\quad\text{for $θ₀ =
\arccos(\frac{g}{Rω²})$} }
\end{align}
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
%%%% Problem 2
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
\problem{2}
\subsubsection{Question}
% Keywords
\index{electrostatics!Charged sphere in uniform electric field}
\index{Gauss' Law!Charged sphere in uniform electric field}
The general solution of the Laplace's equation for an electrostatic problem
having azimuthal symmetry can be written as
\begin{align*}
V(r,θ) &= \sum_{ℓ=0}^∞ (A_ℓ r^ℓ + \frac{B_ℓ}{r^{ℓ+1}}) P_ℓ(\cos θ)
\end{align*}
Now consider the follwoing problem. A solid spherical conductor of radius $R$
having charge $Q$ is placed in an otherwise uniform electric field $\vec E
= E₀\hat z$.
\begin{enumerate}[(a)]
\item
Qualtitatively describe the electric field inside and outside of the
sphere.
\item
Solve the problem and find the electric potential in th eregion outside
the sphere.
\end{enumerate}
\subsubsection{Answer}
To provide a qualitative description, we make use of several properties of
conductors. The electric field inside the conductor is guaranteed to be zero
in the limit of a perfect conductor which can move its electrons anywhere
they're needed to cancel any applied fields. For the region outside the
sphere, it is easiest to describe the region just outside the surface and
the region at infinitely large distances. Far away, the effects of the the
sphere are negligible and the electric field is the external uniform field.
Near the surface, though, all field lines are perpendicular to the surface;
therefore, the external field's lines are curved so that any intersections
occur perpendicular to the surface.
In summary
\par\fbox{\begin{minipage}{\textwidth}
\begin{enumerate}
\item
The field is uniform at large distances
\item
The field is perpendicular to the surface of the conductor at
the conductor's surface
\item
There is no field within the interior of the conductor
\end{enumerate}
\end{minipage}}
\vspace{\baselineskip}
In order to analyze the problem analytically, we make use of the
superposition principle to simplify the problem. Because the sphere carries
its own charge, we treat this case as a superposition of the two simpler
cases of a charged sphere in vacuum and that of a perfectly conducting,
grounded sphere in a uniform electric field.
Since we are only concerned with the potential outside the sphere, we can use
Gauss' Law to get the potential due to the charge $Q$. It is
\begin{align*}
V_Q(r,θ) &= \frac{Q}{4πε₀r} & r &> R
\end{align*}
The uniform field is the considered by satisfying the appropriate boundary
conditions to solve for the coefficients $A_ℓ$ and $B_ℓ$ in the general
solution given above. We start by converting the given electric field to a
potential. In Cartesian coordinates,
\begin{align*}
\vec E = E₀\hat z = -\vec ∇V_∞ \quad\quad\Rightarrow\quad\quad V_∞ &= -E₀z
\end{align*}
which when converted to spherical coordinates gives the potential as $r→∞$
\begin{align*}
V_∞ &= -E₀r\cos θ
\end{align*}
In the infinite distance limit, $B_ℓ/r^{ℓ+1} → 0$ so the boundary condition
equation becomes
\begin{align*}
-E₀r\cos θ &= \sum_{ℓ=0}^∞ A_ℓ r^ℓ P_ℓ(\cos θ) \\
\intertext{By the orthogonality of the Legendre polynomials, only $A₁$ is
non-zero:}
-E₀r\cos θ &= A₁r\cos θ \\
A₁ &= -E₀
\end{align*}
The general solution has thus been simplified to
\begin{align*}
V₀(r,θ) &= -E₀r\cos θ + \sum_{ℓ=0}^∞ \frac{B_ℓ}{r^{ℓ+1}}P_ℓ(\cos θ)
\end{align*}
By our choice of making use of the superposition principle, we have set the
potential to be zero at the surface, so at $r = R$, the boundary conditions
lets us solve for the values of the $B_ℓ$:
\begin{align*}
0 &= -E₀R\cos θ + \sum_{ℓ=0}^∞ \frac{B_ℓ}{R^{ℓ+1}}P_ℓ(\cos θ)
\intertext{The Legendre polynomial orthogonality again eliminates all
coefficients except $B₁$.}
E₀R\cos θ &= \frac{B₁}{R²}\cos θ \\
B₁ &= E₀R³
\end{align*}
This gives us the solution to the grounded sphere as
\begin{align*}
V₀(r,θ) &= -E₀r\cos θ \left[ 1 - (\frac{R}{r})³ \right]
\end{align*}
Therefore by superposition of both solutions, the potential in this situation
at all points outside the sphere is
\begin{align}
\boxed{ V(r,θ) = \frac{Q}{4πε₀r} - E₀r\cos θ \left[ 1 -
(\frac{R}{r})³ \right] }
\end{align}
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
%%%% Problem 3
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
\problem{3}
\subsubsection{Question}
% Keywords
\index{quantum!Reflection and transmission through a barrier}
Consider the transmission of a beam of particles of mass $m$ and momentum $p
= ℏk$, in one dimension, incident on a rectangular potential barrier of
height $V₀$ and extending from $x = 0$ to $x = L$, in the special case that
the energy $E$ of the incident particle \emph{is exactly equal} to the
barrier height $V₀$.
\begin{enumerate}[(a)]
\item
Calculate the transmission and reflection coefficients $T$ and $R$.
\item
Check some properties of your answers in (a): is probability
conserved? Do $T$ and $R$ have the expected limiting values for $L$
very large or very small?
\item
For what values of the de Broglie wavelength of th eparticles is th
etransmitted fraction equal to $1/2$?
\end{enumerate}
\subsubsection{Answer}
Consider a beam of particles incident from the left as shown in the figure
below:
\begin{figure}[H]
\centering
\begin{tikzpicture}
% Make the axes
\draw[dashed,<->] (-1,0) -- (8,0)
node [anchor=west] {$\hat x$};
\draw[dashed,<->] (2,-1) -- (2,2.5)
node [anchor=south] {$V$};
% Draw the potential
\draw (0,0) -- (2,0) -- (2,1.5) -- (5,1.5) -- (5,0) -- (7,0);
% Label the special values of note
\draw (2,0) node[anchor=north] {$x = 0$};
\draw (5,0) node[anchor=north] {$x = L$};
\draw (5,1.5) node[anchor=west] {$V = V₀$};
% Show the indicent particle beam
\draw[very thick,->] (0,1.5)
node[anchor=east] {$E = V₀$}
-- (1.9,1.5);
% Label the regions
\draw (1, 0.75) node {\Large Ⅰ};
\draw (3.5, 0.75) node {\Large Ⅱ};
\draw (6, 0.75) node {\Large Ⅲ};
\end{tikzpicture}
\end{figure}
Ignoring normalization for a minute, we know that in regions Ⅰ and Ⅲ that
the wavefunction is that of a free particle:
\begin{align*}
ψ(x) &= Ae^{ikx} + B^{-ikx}
& k² &= \frac{2mE}{ℏ²}
\end{align*}
In region Ⅱ, the energy $E$ cancels with the potential $V$ in the Schrödinger
equation, so the solution takes the form of a first order polynomial
\begin{align*}
ψ(x) &= Ax + B
\end{align*}
We will only be concerned with the reflection and transmission coefficients,
and knowing that they are defined in terms of a ratio of the wavefunction
amplitude for the reflected and transmitted components with respect to the
incident amplitude, we simplify our solution by directly setting the incident
particle amplitude to unity. Furthermore, we know that there is no leftward
traveling component in region Ⅲ. Assigning each component a unique and
appropriate unknown coefficient, the three wavefunctions are
\begin{align*}
ψ_{\text{Ⅰ}} &= e^{ikx} + re^{-ikx}\\
ψ_{\text{Ⅱ}} &= ax + b\\
ψ_{\text{Ⅲ}} &= te^{ikx}
\end{align*}
We find the values for $r$ and $t$ by applying continuity boundary conditions
at the interfaces between each solution. Starting at $x = 0$,
\begin{align*}
ψ_{\text{Ⅰ}}(0) &= ψ_{\text{Ⅱ}}(0)
& ψ_{\text{Ⅰ}}'(0) &= ψ_{\text{Ⅱ}}'(0) \\
1 + r &= b
& ik(1-r) &= a
\end{align*}
Then putting the values into $ψ_{\text{Ⅰ}}$ and solving the boundary conditions
at $x = L$,
\begin{align*}
ψ_{\text{Ⅱ}}(L) &= ψ_{\text{Ⅲ}}(L)
& ψ_{\text{Ⅱ}}'(L) &= ψ_{\text{Ⅲ}}'(L) \\
ik(1-r)L + 1+r &= te^{ikL}
& ik(1-r) &= ikte^{ikL}
\end{align*}
From the equation on the right, we solve for $t$ as a function of $r$ and
insert it into the condition on the left:
\begin{align*}
t &= (1-r)e^{-ikL} \\
ik(1-r)L + 1 + r &= ((1-r)e^{-ikL})e^{ikL} \\
r &= \frac{-ikL}{2-ikL}
\intertext{Plugged back into $t$ gives}
t &= \frac{2}{2-ikL}e^{-ikL}
\end{align*}
We then just take the complex square of both amplitudes to get the reflection
and transmission coefficients:
\begin{align}
\boxed{ R = |r|² = \frac{k²L²}{4 + k²L²} }
\end{align}
\begin{align}
\boxed{ T = |t|² = \frac{4}{4 + k²L²} }
\end{align}
These satisfy the requisite properties: the probabilities sum to unity so all
particles are accounted for, in the limit that the barrier vanishes no
particles are reflected and all are transmitted, and in the limit that the
barrier grows to infinite depth, all particles are reflected.
\begin{align}
\boxed{ R + T = 1}
\end{align}
\begin{align}
\boxed{
T \underset{L→0}{\longrightarrow} 1 \quad\quad
T \underset{L→∞}{\longrightarrow} 0
}
\end{align}
\begin{align}
\boxed{
R \underset{L→0}{\longrightarrow} 0 \quad\quad
R \underset{L→∞}{\longrightarrow} 1
}
\end{align}
This system can be tuned such that half of the particles are transmitted
through the barrier by changing the energy of the particles. To do so, we set
the transmission probability to $\frac 12$ and solve for the particles'
corresponding de Broglie wavelength.
\begin{align*}
\frac 12 &= \frac{4}{4+k²L²} \\
4 &= k²L² \\
k² &= \frac{4}{L²}
\intertext{Making use of the definition of $k²$ in terms of the energy,}
\frac{2mE}{ℏ²} &= \frac{4}{L²}
\intertext{Then writing the energy in terms of the de Broglie wavelength:}
\frac{2m}{ℏ²} \frac{4π²ℏ²}{2mλ²} &= \frac{4}{L²}
\end{align*}
Therefore, the particles' incident momentum can be tuned and half the particles
will be transmitted when
\begin{align}
\boxed{ λ = πL }
\end{align}
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
%%%% Problem 4
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
\problem{4}
\subsubsection{Question}
% Keywords
\index{quantum!Periodic array of potential wells}
\index{solid state!Periodic array of potential wells}
Consider a one-dimensional infinite array of points labeled by an index $n$
and separated by a fixed unit distance. At each point there is an identical
very deep and narrow potential well. Let $\ket{n}$ denote an eigenstate of a
\emph{single} well, with energy $E$.
\begin{enumerate}[(a)]
\item
Argue that if the wells are so narrow that the different sites can
be considered uncoupled, then $\ket{n}$ is an eigenstate of the
total Hamiltonian $H$ with eigenvalue $E$. What is its degeneracy?
Then show that the state $\ket{k}$ defined as
\begin{align*}
\ket{k} = \sum_{n=-∞}^∞ e^{ink} \ket{n}
\end{align*}
with $-π < k < π$ is an eigenstate of both $H$ and the translation
operator $T$ defined as $T\ket{n} = \ket{n+1}$. Find the respective
eigenvalues.
\item
Assume now that neighboring sites are weakly coupled so that the
total Hamiltonian can now be written as
\begin{align*}
H = \sum_{n=-∞}^∞ ( \ket{n}E\bra{n} - \ket{n+1}D\bra{n} -
\ket{n}D\bra{n+1} )
\end{align*}
where the coupling parameter $D$ is real and we assume that
$\braket{n}{n'} = δ_{n,n'}$. Show that $\ket{n}$ is no longer an
eigenstate of $H$ but that $\ket{k}$ still is. Find the eigenvalue.
\end{enumerate}
\subsubsection{Answer}
The only reasonable choice for the form of the total Hamiltonian $H$ is a
superposition of the Hamiltonian of individual sites.
\begin{align*}
H &= \sum_i H_i
\end{align*}
If we then operate on a state $\ket{n}$ with the total Hamiltonian,
\begin{align*}
H\ket{n} &= (\sum_i H_i)\ket{n} \\
{} &= \sum_i H_i\ket{n} \\
H\ket{n} &= H_i δ_{i,n} \ket{n}
\end{align*}
Only the $n$-th Hamiltonian will operate on $\ket{n}$, so the state is in fact
an eigenstate of the total Hamiltonian with an eigenvalue of $E$.
\begin{align}
\boxed{ H\ket{n} = E\ket{n}} \\
\boxed{ \text{$N$-fold degeneracy} }
\end{align}
Because each state $n$ has the same eigenvalue of $E$, the degeneracy is
equal to the number of sites. If there are $N$ sites in the array, then that
is also the degeneracy of the total system.
Similarly for the state $\ket{k}$ as defined will can be operated on by the
total Hamiltonian:
\begin{align*}
H\ket{k} &= H (\sum_n e^{ink}\ket{n}) \\
{} &= \sum_n e^{ink} H\ket{n}
\intertext{Then because we've already shown that $\ket{n}$ is an eigenstate
of $H$ with eigenvalue $E$}
{} &= \sum_n e^{ink} E\ket{n} \\
H\ket{k} &= E (\sum_n e^{ink}\ket{n})
\end{align*}
We find that $\ket{k}$ is also an eigenstate of the total Hamiltonian with
ane eigenvalue of $E$ as well.
\begin{align}
\boxed{ H\ket{k} = E\ket{k} }
\end{align}
Finally, we define a translation operator $T$ for $\ket{n}$ and determine its
effect on the state $\ket{k}$.
\begin{align*}
T\ket{k} &= T (\sum_n e^{ink}\ket{n}) \\
{} &= \sum_n e^{ink} T\ket{n} \\
{} &= \sum_n e^{ink} \ket{n+1} \\
\intertext{We can insert a factor of unity to extract a more useful form}
{} &= \sum_n e^{i(n+1)k}e^{-ik} \ket{n+1} \\
{} &= e^{-ik} \sum_n e^{i(n+1)k} \ket{n+1}
\end{align*}
and since $n ∈ (-∞,∞)$, the distinction between $n$ and $n+1$ is
inconsequential to the definition of $\ket{k}$. Therefore
\begin{align}
\boxed{ T\ket{k} = e^{-ik}\ket{k} }
\end{align}
The translation operator has a phase eigenvalue of $e^{-ik}$ when operating
on the Bloch wave function $\ket{k}$.
If the total Hamiltonian is then modified include nearest neighbor
interactions, the individual site wavefunctions $\ket{n}$ are no longer
eigenstates of the total Hamiltonian as shown by explicit calculation:
\begin{align*}
H\ket{n} &= \left[ \sum_{n'} \ket{n'}E\bra{n'} - \ket{n'+1}D\bra{n'}
- \ket{n'}D\bra{n'+1} \right] \ket{n} \\
{} &= \sum_{n'} \ket{n'}E\braket{n'}{n} -
\ket{n'+1}D\braket{n'}{n} - \ket{n'}D\braket{n'+1}{n} \\
{} &= \sum_{n'} δ_{n,n'}(E\ket{n'} - D\ket{n'+1}) - Dδ_{n,n'+1}\ket{n'} \\
{} &= E\ket{n} - D\ket{n+1} - D\ket{n-1}
\end{align*}
There are now three wavefunctions left with two different coeffients, so the
problem is not an eigenvalue problem.
\begin{align}
\boxed{ E\ket{n} - D\ket{n+1} - D\ket{n-1} = H\ket{n} ≠ λ\ket{n} }
\end{align}
Operating on the Bloch wavefunction, though
\begin{align*}
H\ket{k} &= \left[ \sum_{n'} \ket{n'}E\bra{n'} - \ket{n'+1}D\bra{n'}
- \ket{n'}D\bra{n'+1} \right] (\sum_n e^{ink}\ket{n}) \\
{} &= \sum_{n,n'} \ket{n'}Ee^{ink}\braket{n'}{n} -
\ket{n'+1}De^{ink}\braket{n'}{n} - \ket{n'}De^{ink}\braket{n'+1}{n} \\
{} &= \sum_{n,n'} Ee^{ink}δ_{n,n'}\ket{n'} - De^{ink}δ_{n,n'}\ket{n'+1} -
De^{ink}δ_{n,n'+1}\ket{n'} \\
\intertext{Consuming the summation over $n'$ to select a non-zero term in the
Kronecker delta leaves}
{} &= \sum_{n} Ee^{ink}\ket{n} - De^{ink}\ket{n+1} - De^{ink}\ket{n-1}
\intertext{We can then separate each term into a summation:}
{} &= E\sum_{n}(e^{ink}\ket{n}) - D\sum_{n}(e^{ink}\ket{n+1}) -
D\sum_{n}(e^{ink}\ket{n-1})
\intertext{Then using the same unity-factor trick as in demonstrating that
$\ket{k}$ is an eigenstate of $T$,}
{} &= E\sum_{n}(e^{ink}\ket{n}) - De^{-ik}\sum_{n}(e^{i(n+1)k}\ket{n+1}) -
De^{ik}\sum_{n}(e^{i(n-1)k}\ket{n-1})
\end{align*}
Each of these terms is the definition of $\ket{k}$, so we find that $\ket{k}$
is still an eigenstate of this new Hamiltonian with an eigenvalue of $E - 2D
\cos k$.
\begin{align}
\boxed{ H\ket{k} = (E - 2D\cos k)\ket{k} }
\end{align}
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
%%%% Problem 5
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
\problem{5}
\subsubsection{Question}
% Keywords
\index{thermodynamics!Mixing gases}
Two monatomic ideal gases, each occupying a volume $V = \SI{1}{\m³}$, are
separated by a removeable insulating partition. They have different
temperatures $T₁ = \SI{350}{\K}$ and $T₂ = \SI{450}{\K}$, and different
pressures $p₁ = \SI{e3}{\N\per\m²}$ and $p₂ = \SI{5e3}{\N\per\m²}$. The
partition is removed, and the gases are allowed to mix while remaining
thermally isolated from the outside.
\begin{enumerate}[(a)]
\item
What are the final temperature $T_f$ (in \si{\K}) and pressures $p_f$
(in \si{\N\per\m²})?
\item
What is the net change in entropy due to mixing (in \si{\J\per\K})?
\end{enumerate}
\subsubsection{Answer}
Starting with conservation of energy, the total internal energy after the
partition is removed must be the same as the sum of the internal energies of
both starting gases.
\begin{align*}
U &= U₁ + U₂ \\
\frac 32 Nk_B T_f &= \frac 32 N₁k_B T₁ + \frac 32 N₂k_B T₂ \\
T_f &= \frac{N₁T₁ + N₂T₂}{N₁ + N₂}
\end{align*}
where we made use of the fact that particle number must also be a conserved
quantity so that $N = N₁ + N₂$. The original particle numbers $N₁$ and $N₂$
can be determined from the ideal gas law in the initial state:
\begin{align*}
p₁V &= N₁k_B T₁ & p₂V &= N₂ k_B T₂ \\
N₁ &= \frac{p₁V}{k_B T₁} & N₂ &= \frac{p₂V}{k_B T₂}
\end{align*}
Plugging these into the final temperature $T_f$ above and simplifying gives
the value in terms of known quantities as
\begin{align*}
T_f &= \frac{p₁ + p₂}{p₁T₂ + p₂T₁} T₁T₂
\end{align*}
which when the numbers are plugged in gives a final temperature of
\begin{align}
\boxed{ T_f = \SI{429.55}{\K} }
\end{align}
We can then plug this temperature into a formulation of the ideal gas law
for the system after the partition has been removed to get the final pressure.
\begin{align*}
p_f (2V) &= (N₁+N₂) k_B T_f \\
p_f &= \frac{(N₁+N₂)k_B}{2V} \frac{N₁T₁ + N₂T₂}{N₁+N₂} \\
p_f &= \frac{k_B}{2V}(N₁T₁ + N₂T₂)
\end{align*}
Again substitution for $N₁$ and $N₂$ in terms of the original pressures and
temperatures gives
\begin{align*}
p_f &= \frac{p₁ + p₂}{2}
\end{align*}
which when the values are plugged in
\begin{align}
\boxed{ p_f = \SI{3e3}{\N\per\m²} }
\end{align}
From the Sackur-Tetrode euation, we can calculate the change in the entropy
from the beginning state to the final one. We start by simplifying the
equation to isolate constant factors:
\begin{align*}
S &= Nk_B \left\{\frac 52 + \ln\left[ \frac{V}{N}
(\frac{4πmU}{3Nh²})^{3/2} \right] \right\} \\
{} &= \frac 52 Nk_B + Nk_B \ln\left[ \frac{V}{N}
(\frac{4πmU}{3Nh²})^{3/2} \right] \\
{} &= Nk_B \left\{ \frac 52 + \frac 32 \ln(\frac{4πm}{3h²})
+ \ln\left[ \frac{V}{N}(\frac{U}{N})^{3/2} \right] \right\}
\end{align*}
Then with the change in entropy defined as
\begin{align*}
ΔS = S - S₁ - S₂
\end{align*}
we start calculating the sum term-by-term. For the first two constant terms,
the fact that $N = N₁ + N₂$ causes these terms to cancel with those in $S₁$
and $S₂$. That leaves us just with the last term.
\begin{align*}
ΔS &= (N₁+N₂)k_B
\ln\left[\frac{2V}{N}(\frac{\frac 32 Nk_B T_f}{N})^{3/2}\right]
- N₁k_B\ln\left[\frac{V}{N₁}(\frac{\frac 32 N₁k_B T₁}{N₁})^{3/2}\right]
- N₂k_B\ln\left[\frac{V}{N₂}(\frac{\frac 32 N₂k_B T₂}{N₂})^{3/2}\right]
\\
\intertext{By the ideal gas law, $V/N = k_BT/p$ which simplifies the expression
to}
{} &= N₁k_B\ln\left[
\frac{2k_B T_f}{p_f}(\frac 32 k_B T_f)^{3/2} ⋅
\frac{p₁}{k_B T₁}(\frac{1}{\frac 32 k_B T₁})^{3/2}
\right] + N₂k_B\ln\left[
\frac{2k_B T_f}{p_f}(\frac 32 k_B T_f)^{3/2} ⋅
\frac{p₂}{k_B T₂}(\frac{1}{\frac 32 k_B T₂})^{3/2}
\right]
\\
{} &= N₁k_B\ln\left[
\frac{p₁}{p_f}(\frac{T_f}{T₁})^{5/2}
\right] + N₂k_B\ln\left[
\frac{p₂}{p_f}(\frac{T_f}{T₂})^{5/2}
\right] + (N₁+N₂)k_B \ln 2
\end{align*}
Using the ideal gas law again to manipulate the coefficients $Nk_B = pV/T$,
\begin{align*}
ΔS &= V\left\{ \frac{p₁}{T₁}\ln\left[
\frac{p₁}{p_f}(\frac{T_f}{T₁})^{5/2}
\right] + \frac{p₂}{T₂}\ln\left[
\frac{p₂}{p_f}(\frac{T_f}{T₂})^{5/2}
\right] + \frac{p_f}{T_f} \ln 2 \right\}
\end{align*}
Plugging in the values for all these numbers as given or we determined, the
change in entropy is
\begin{align}
\boxed{ ΔS = \SI{7.55}{\J/\K} }
\end{align}
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
%%%% Problem 6
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
\problem{6}
\subsubsection{Question}
% Keywords
\index{special relativity!Rocket sending signal to Earth}
A rocket passes Earth at a speed $v = 0.6c$. When a clock on the rocket says
that one hour has elapsed since passing, the rocket sends a light signal
back to Earth.
\begin{enumerate}[(a)]
\item
Suppose that the Earth and rocket clocks were synchronized at zero
at the time passing. According to the \emph{Earth} clocks, when was
the signal sent?
\item
According to the \emph{Earth} clocks, when did the signa arrive back
on Earth?
\item
According to the \emph{rocket} clocks, how long after the rocket
passed did the signal arrive back on Earth?
\end{enumerate}
\subsubsection{Answer}
Let $β = v/c$.
\begin{enumerate}[(a)]
\item
Let $t'_{sent} = \SI{1}{\hour}$ be the time at which the rocket sent
the signal according to its own clock. Because from the Earth's
reference frame the rocket's clocks are running slow, the clocks on
Earth must show a time greater than the rocket's by a factor of $γ$.
\begin{align*}
t_{sent} &= γt'_{sent} \\
{} &= \frac{\SI{1}{\hour}}{\sqrt{1 - β²}}
\end{align*}
\begin{align}
\boxed{ t_{sent} = \frac 54 \,\si{\hour} = \SI{1.25}{\hour} }
\end{align}
\item
The returning light will traverse the intermediate distance at $c$, so
$t_{ret} = x/c$. The distance from the Earth is simply the velocity
times the time (in Earth's reference frame), so together,
\begin{align*}
t_{ret} &= \frac{vt_{sent}}{c} \\
{} &= βt_{sent} \\
t_{ret} &= \frac 34 \,\si{\hour} = \SI{0.75}{\hour}
\end{align*}
The total round trip time is then
\begin{align}
\boxed{ t_{tot} = t_{sent} + t_{ret} = \SI{2}{\hour} }
\end{align}
\item
Because the rocket is flying away from the Earth, the distance behind
it towards the earth appears to have been length expanded by a factor
of $γ$ compared to the distance that Earth would report. Therefore the
time for the signal to be sent from the rocket to Earth appears to the
rocket to be
\begin{align*}
t'_{ret} &= \frac{γx}{c} \\
{} &= γt_{ret} \\
t'_{ret} &= \frac{15}{16} \,\si{\hour} = \SI{0.9375}{\hour}
\end{align*}
Added to the 1 hour that the rocket observes as the time before it
sent the signal, the signal would arrive at earth at time
\begin{align*}
\boxed{ t'_{tot} = \frac{31}{16} \,\si{\hour} = \SI{1.9375}{\hour} }
\end{align*}
\end{enumerate}