-
Notifications
You must be signed in to change notification settings - Fork 2
/
physics--classical-mechanics.tex
917 lines (732 loc) · 35.6 KB
/
physics--classical-mechanics.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
850
851
852
853
854
855
856
857
858
859
860
861
862
863
864
865
866
867
868
869
870
871
872
873
874
875
876
877
878
879
880
881
882
883
884
885
886
887
888
889
890
891
892
893
894
895
896
897
898
899
900
901
902
903
904
905
906
907
908
909
910
911
912
913
914
915
916
917
\renewcommand{\r}{\vec{r}}
\section{Gravity}
\begin{tabular*}{1.0\linewidth}{l|l}
$M, m$ & masses of two bodies \\
$r$ & distance between bodies
\end{tabular*}
Newton's law of gravity: the force between two bodies is $F = G\frac{Mm}{r^2}$, where $G$ is the
gravitational constant.
Units: since $[F] = \frac{LM}{T^2}$, we have $[G] = \frac{LM}{T^2}\frac{L^2}{M^2} = \frac{L^3}{MT^2}$.
\begin{tabular*}{1.0\linewidth}{l|l}
$M_E = 5.972 \times 10^{24} kg$ & mass of the Earth \\
$R = 6.371 \times 10^6 m$ & radius of the Earth \\
$G = 6.67408 \times 10^{-11} m^3 kg^{-1} s^{-2}$ & gravitational constant\\
$m$ & mass of an object
\end{tabular*}
The acceleration of the object due to gravity at the Earth's surface is
\begin{align*}
g
= \frac{F}{m}
= G\frac{M_E}{R^2}
= \frac{6.67408 \times 10^{-11} \times 5.972 \times 10^{24}}{(6.371 \times 10^6)^2}
= 9.82 ms^{-2}.
\end{align*}
\blue{Note that, for gravity, the strength of the force itself depends on the mass $m$. So Newton's
Second Law for a body of mass $m$ near the Earth's surface is
\begin{align*}
ma &= F \\
ma &= mg \\
a &= g.
\end{align*}
}
\section{Force, energy, work}
The argument is\footnote{Taylor ch. 4}:
\begin{enumerate}
\item Suppose a particle is moving under the influence of a force, according to Newton's Second Law.
\item Consider the integral of the force over some section of the particle's
trajectory: $\int_{\vec{r_0 \to \vec{r_1}}} F(\vec{r}) \dif \vec{r}$.
\item The integral is equal to the difference in value of a certain quantity that depends on speed, evaluated at
the start and end point.
\item This quantity is $\frac{1}{2}mv^2$. It depends only on the particle's mass, and speed at a moment in
time. As long as the particle's motion obeys Newton's Second Law $m\ddot{\vec{r}} = F(\vec{r})$, then the
relationship between this quantity and the integral of force over space is just a fact of calculus.
\item The quantity $\frac{1}{2}mv^2$ is given the name \blue{kinetic energy} and the integral of force over a path
in space is given the name \blue{work}.
\item Thus the \blue{Work - KE Theorem}: when a particle is moved along a path according to Newton's Second Law,
then the work done by the force is equal to the difference in kinetic energy.
\item This can be used, for example, to calculate the new velocity, given knowledge of the work done and the
initial velocity.
\item So the work done when moving from $\vec{r_0}$ to $\vec{r}$ is equal to the difference in kinetic energy at
those locations\footnote{This implies that every position is associated with a unique speed?}:
\begin{align*}
W(\vec{r_0} \to \vec{r})
&= \int_{\vec{r_0}}^{\vec{r}} F(\vec{r}) \dif \vec{r} \\
&= \frac{1}{2}mv^2 - \frac{1}{2}mv_0^2.
\end{align*}
\item Therefore the following quantity is constant at all points along the path:
\begin{align*}
E
&= \frac{1}{2}mv^2 - \int_{\vec{r_0}}^{\vec{r}} F(\vec{r}) \dif \vec{r} \\
&= \frac{1}{2}mv_0^2.
\end{align*}
\end{enumerate}
Suppose we start at position $x_0$ with speed $v_0$. Some time later we find ourselves at position
$x_1$ with speed $v_1$. Then the change in our KE is equal to the work done by the force along the
path we followed, $T(v_1) - T(v_0) = W(x_0 \to x_1)$. So, for this path, we have that the quantity
\begin{align*}
T(v_1) - W(x_0 \to x_1)
\end{align*}
is constant and equal to $T(v_0)$.
Now suppose that in this system the work done along any path depends only on the endpoints of the
path, and not otherwise on the actual path followed. Then the quantity
\begin{align*}
T(v) - W(x_0 \to x)
\end{align*}
is constant, where $v$ is the speed at $x$.
{\bf Intuition:} the quantity that is conserved is
\begin{align*}
\text{(new KE at some position)} - \text{(change in KE that must have occurred to get there)}
\end{align*}
\subsection{3 dimensions}
The extension to 3-dimensions is as expected. We have
\begin{align*}
W(\r_1 \to \r_2) &= \int_{\r_1}^{\r_2} \F(\r) \cdot \d\r,
\end{align*}
and
\begin{align*}
\F = -\grad \vec{V}.
\end{align*}
\subsection{Non-mathematical explanation of potential energy and kinetic energy}
Suppose we're in a 1-dimensional universe (i.e our universe is just one straight line). There's a
force that's pushing to the left or to the right, with some strength, at every point along the
line. The force doesn't vary over time; it depends only on where we are along the line. And there's
no friction; the only thing accelerating us (in one direction or the other) is the force.
% We're going to be moved around under the influence of the force; apart from the force, there's
% nothing else that will make us move, and we can't prevent the force moving us.
Let $x$ be our location on the line, and let $F(x)$ be the value of the force at that location
(negative means it's pushing to the left).
We're going to keep track of how much force we are subject to while it moves us around. So, we keep
a running total of all the $F$ values we are subject to at all the locations along the line that we
are taken to. If the force is pushing to the right with some strength $F$ then our running total
goes up by $F$; if it's pushing to the left, then it goes down by $F$. Then, at the new location we
were taken to, we do the same thing according to the strength of the force there. Except, of course,
space is continuous and this doesn't really happen in discrete steps: instead we compute the
integral of the $F$ function over some region of the line.
Let $V(x)$ be the value of this integral, i.e. the net amount of force we encounter while being
moved from our starting position to some position $x$. Recall that, basically, force is something
that causes acceleration. In other words, it increases or decreases our velocity. So $E$ is in some
sense measuring our net gain in velocity as we move around the line.
Now, suppose we're at some point $x_0$ on the line. The force is pushing us to the right. Looking to
the right, the value of the force function over that region dictates that a certain change in our
velocity will occur while we traverse it. So there's some function $V(x)$ that says, if the force
moves us to $x$, how much change in our velocity this will result in.
Say we start of at $a$ with some amount $T$ of velocity.
So that's like a promised amount of
change.
The function that maps $x$ to the
change in velocity that will occur between here and $x$ is called the potential energy function
$V(x)$.
\subsection{Solving a gravity problem using dimensional analysis, Newton's Second Law, and Conservation of Energy}
\begin{mdframed}
\includegraphics[width=400pt]{img/physics--classical-mechanics--oxford--dynamics--example--gravity.png}
\end{mdframed}
Below we find the maximum height using two methods:
\begin{enumerate}
\item by using Newton's Second Law to write a second-order differential equation for the position
function, and solving it to give velocity and position functions,
\item by equating the kinetic energy at the beginning with the potential energy at the maximum
height (when velocity is zero).
\end{enumerate}
Conservation of energy is useful here because total energy ($T + V$) is constant along any
trajectory that satisfies N2L. The logic here is:
\begin{enumerate}
\item We know the trajectory satisfies N2L (because that's how the universe works)
\item Therefore, we know that $T + V$ is constant at any $x$ on the trajectory (one can prove that
$T+V$ is constant over any trajectory that satisfies N2L.)
\item At one point in the trajectory, we know the values of $T$ and $V$, and the expression for $V$
involves the quantity of interest $z_{max}$.
\item We also know the values of $T$ and $V$ at another point in the trajectory, not involving
$z_{max}$.
\item This gives us one equation for the one unknown.
\end{enumerate}
Momentum \textit{is} also conserved, but to see conservation of momentum we have to expand the
system to include the Earth and the particle\footnote{Conservation of momentum is related to N3L,
and the relevant N3L pair of forces are the Earth's gravitational attraction on the particle and
the particle's on the Earth, both of which have strength $G\frac{mM_E}{d^2}$, where $M_E$ is the
mass of the Earth and $d$ is the separation of the two particles, which is approximately the
radius of the Earth.}. I don't think conservation of momentum can be used to find the height at
which velocity is zero.
We'll use $v_0$ instead of $u$ for the initial velocity.
\begin{enumerate}
\item {\bf Solution via dimensional analysis}\\
We have
\begin{align*}
[g] &= LT^{-2} \\
[v_0] &= LT^{-1},
\end{align*}
and we want $z_{max}$ with units $L$. This suggests that the answer is $z_{max} = C\frac{v_0^2}{g}$,
where $C$
is dimensionless.\\
\item {\bf Solution via Newton's Second Law}\\
Applying N2L here gives $m\ddot{z}(t) = -mg$, i.e.
\begin{align*}
\ddot{z}(t) = -g.
\end{align*}
Applying FTC once gives the velocity function,
\begin{align}
\dot{z}(t) - \dot{z}(0) = \dot{z}(t) - v_0 &= \int_0^t \ddot{z}(t) \dt = -gt \nonumber \\
\dot{z}(t) &= v_0 - gt, \label{oxford-dynamics-example-gravity-velocity}
\end{align}
and applying FTC again gives the position function (trajectory),
\begin{align}
z(t) - z(0) = z(t) - 0 = \int_0^t \dot{z}(t) \dt = v_0t - \frac{1}{2}gt^2. \label{oxford-dynamics-example-gravity-position}
\end{align}
The maximum height occurs when the velocity is zero. From
\eqref{oxford-dynamics-example-gravity-velocity} we see that this occurs when $t = \frac{v_0}{g}$,
and from \eqref{oxford-dynamics-example-gravity-position} we see that at that time the height is
\begin{align*}
z_{max} = z\(\frac{v_0}{g}\) = \frac{v_0^2}{g} - \frac{1}{2}g\frac{v_0^2}{g^2} = \frac{1}{2}\frac{v_0^2}{g}.
\end{align*}
\item {\bf Solution via conservation of energy}\\
Initially, the particle has kinetic energy $T = \frac{1}{2}mv_0^2$.
Recall that the definition of potential energy relative to $z_0$ is $V(z) = -\int_{z_0}^z F(s) \ds$.
At its maximum height, gravity has done work
\begin{align*}
W = \int_0^{z_{max}} F(z) \dz = \int_0^{z_{max}} m(-g) \dz = -mgz_{max}
\end{align*}
on the particle (slowing it down). The potential energy at $z = z_{max}$ relative to $z = 0$ is
$mgz_{max}$.
So, we have\\
\begin{tabular}{l|l|l|}
Time & Kinetic energy & Potential energy \\
\hline
$0$ & $\frac{1}{2}mv_0^2$ & 0\\
$t_{z_{max}}$ & $0$ & $mgz_{max}$.
\end{tabular}
So, from conservation of energy, we have that
\begin{align*}
mgz_{max} &= \frac{1}{2}mv_0^2 \\
z_{max} &= \frac{1}{2}\frac{v_0^2}{g}.
\end{align*}
\end{enumerate}
\section{Force, energy, work, momentum}
Basically, energy measures an accumulation of force over some path in space. Equivalently, the force
at some location in space is a spatial gradient of energy at that location.
However, we also have the Second Law notion that force is the rate of change of momentum,
$F = \dot{p} = m\dot{v}$.
The Second Law is relating force to change in $x$ over time, whereas the notion of energy is
concerned with accumulating force along a spatial trajectory.
Suppose space $x$ is one-dimensional and there is a force $F(x)$ that depends only on position
$x$. We want to compute the integral of force over some interval in space (i.e. ``work''). Let
$v = \dxdt$. Since $F(x) = m\dvdt(x)$, we basically want to compute the following integral:
\begin{align*}
W = \int_{x_1}^{x_2} \dvdt(x) \dx.
\end{align*}
One approach is to write down the equation $\dvdt = \dxdt \dvdx = v\dvdx$ and thus argue that
\begin{align*}
W
= \int_{x_1}^{x_2} v\dvdx \dx
&= \int_{x_1}^{x_2} v\d v \\
&= \Big[\frac{1}{2}mv^2\Big]_{v(x_1)}^{v(x_2)},
\end{align*}
where we have ``canceled $\dx$'', and for some reason stopped writing the function argument syntax
``$(x)$''. And why are velocity and acceleration functions of position rather than of time here?
\red{What if the particle visits $x_1$ multiple times with different velocities so that $v(x)$ is
not well-defined?}
What do these formal manipulations of Leibniz notation actually mean?
Let's go back to the integral we want to compute:
\begin{align*}
W = \int_{x_1}^{x_2} \dvdt(x) \dx.
\end{align*}
So, we want to calculate the signed area under the graph of $\dvdt(x)$, over the interval
$[x_1, x_2]$.
~\\\hrule~\\
Suppose that we know the trajectory function $x(t)$ of a particle of mass $m$.
Therefore we also know $\dot{x}(t)$ and $\ddot{x}(t)$.
Further, suppose that we know that the net force acting on the particle depends only on its position
$x$. Let this force be $F(x)$.
We want to evaluate
\begin{align*}
\int_{x_1}^{x_2} F(x) \dx
\end{align*}
in terms of the trajectory function $x(t)$ and its derivatives (velocity and acceleration).
By definition of derivative, at time $t$, the increments $\dx$ and $\dt$ are related according to
$\dx = \dot{x}(t)\dt$, so we can rewrite the integral as
\begin{align*}
\int_{x_1}^{x_2} F(x(t)) \dot{x}(t)\dt.
\end{align*}
From Newton's Second Law, this is
\begin{align*}
\int_{x_1}^{x_2} m\ddot{x}(t) \dot{x}(t)\dt,
\end{align*}
or equivalently
\begin{align*}
\int_{x_1}^{x_2} m\dot{v}(t) v(t)\dt.
\end{align*}
Note that an antiderivative is available here, since $\ddt \frac{1}{2}v^2 = v\dot{v}$. Therefore
\begin{align*}
\int_{x_1}^{x_2} F(x) \dx = \Big[\frac{1}{2}mv(t)^2\Big]_{t_1}^{t_2},
\end{align*}
where $x(t_1) = x_1$ and $x(t_2) = x_2$. \red{But what if the particle visits $x_1$ multiple times,
i.e. $x(t_1) = x_1$ has multiple solutions?}
\subsection{Potential energy, conservative force, and work}
The \defn{work} done when a particle moves from $x_0$ to $x_1$ is defined to be
\begin{align*}
W(x_0 \to x_1) := \int_{x_0}^{x_1} F(x) \dx.
\end{align*}
If this is independent of the path taken between $x_0$ and $x_1$, then we say the force is
\defn{conservative} and define the \defn{potential energy} at $x$, relative to $x_0$ to be
\begin{align*}
V(x) &:= -\int_{x_0}^x F(x') \dx'.
\end{align*}
If the integral depends on the path taken, the potential energy is undefined.
\blue{A force moves something in space (causes an acceleration). Equivalently, something moves in
space because a small displacement in some direction is associated with a lower potential
energy. In fact, the force at a location \emph{is} the gradient in potential energy at that
location.}
\begin{align*}
F(x) &= -\frac{\d V(x)}{\dx} \\
V(x) &= -\int_{x_0}^x F(x') \dx'
\end{align*}
\todo{Understand this}:
A force that depends only on position in one dimension is always conservative, because the integral
depends only on the endpoints. But a force that depends on e.g. time or velocity is
non-conservative. Friction is such a force because although it looks like a constant force
($\mu mg$), its direction (sign) is the opposite of the direction of velocity, so it is in fact
velocity-dependent.
Also, how is this so:
\begin{quote}
\emph{Since friction always opposes the motion, the contributions to the $W = \int F \dx$ integral are
always negative, so there is never any cancellation. The result is therefore a large negative
number.}
\end{quote}
\subsection{Slowing down a moving object}
Facts:
\begin{enumerate}
\item The rate of change of momentum is determined by the net force acting on an object.
\item To slow something down, a force must be applied for some period of time.
\item To slow something which has momentum $p$, a strong force could be applied briefly, or a weaker
force for a longer time.
\item ``Slow something down'' in those sentences could be replaced with ``change the velocity''. The
change in velocity could be a change in direction, without a change in speed.
\end{enumerate}
Consider two bodies traveling in one dimension with masses and velocities $m_1, v_1$ and $m_2,
v_2$. Their momenta are identical: $m_1v_1 = m_2v_2$.
Suppose that mass $m_1 > m_2$, and therefore $v_1 < v_2$.
A force $F$ acts to slow them down. Over what length of time must the force be in effect?
The force causes accelerations (decelerations) $a_1 = F/m_1$ and $a_2 = F/m_2$. So we have velocity
functions
\begin{align*}
\dv{x_1}{t} (t) &= v_1 - \frac{F}{m_1}t \\
\dv{x_2}{t} (t) &= v_2 - \frac{F}{m_1}t,
\end{align*}
and we see that the velocities become equal to zero at the same time, i.e. when
$t = \frac{m_1v_1}{F} = \frac{m_2v_2}{F}$.
So, \blue{the length of time for which a force must be applied depends only on the momentum}; not on the
mass or velocity separately.
Now, how far does a mass travel while it is slowing down (i.e. over time $mv/F$)? This is
\begin{align*}
x &= \int_{t=0}^{t=mv/F} \(v - \frac{F}{m}t\) \dt \\
&= vt - \frac{F}{2m}t^2\Big|_{t=0}^{t=mv/F} \\
&= \frac{mv^2}{F} - \frac{F}{2m}\frac{m^2v^2}{F^2} \\
&= \frac{1}{2}mv^2/F.
\end{align*}
So, in order for a constant force $F$ to halt a mass $m$ with velocity $v$, the product of force and
the distance over which the force is applied (i.e. the work done) must equal the kinetic energy of
the mass $\frac{1}{2}mv^2$.
I.e. \blue{the distance over which the force must be applied depends on the kinetic energy}
$\frac{1}{2}mv^2$.
The lighter mass covers more distance while it is slowing down, because it has higher kinetic energy.
The following connections between force and momentum, and force and work/energy, make sense under
the FTC:
\begin{enumerate}
\item Force is the time derivative of momentum. The integral of force over time corresponds to change in momentum.
\item Force is the spatial derivative of potential energy. The integral of force over space
(i.e. work) corresponds to change in potential energy.
\end{enumerate}
\subsection{Example: gravitational potential energy}
Consider two point masses $M$ and $m$, separated by a distance $r$. Newton's law of gravitation
states that there is a force between them of magnitude $-GMm/r^2$ (the force is attractive, hence
the negative sign).
The potential energy of the system at separation $r$, relative to separation $r_0$, is
\begin{align*}
V(r) &= -\int_{r_0}^r F(r) \dr \\
&= -\int_{r_0}^r \frac{-GMm}{r^2} \dr \\
&= -\(\frac{GMm}{r} - \frac{GMm}{r_0}\).
\end{align*}
Typically in this situation we would choose $r_0 = \infty$ as the reference separation, so that
\begin{align*}
V(r) = -\frac{GMm}{r}.
\end{align*}
\todo{Do this in 3D.}
\blue{Potential energy, relative to $r = \infty$, decreases as the two masses get closer: so the
masses will approach each other. It's always negative because we have measured it relative to
$r = \infty$, and any separation is more favorable than infinitely large separation.}
\begin{question*}
What is the gravitational potential energy of a mass $m$ at a height $y$, relative to the Earth's
surface?
\end{question*}
\begin{proof}
Let the mass and radius of Earth be $M$ and $R$. Then
\begin{align*}
V(y) &= -\(\frac{GMm}{R + y} - \frac{GMm}{R}\) \\
&= GMm(\frac{R + y - R}{R^2 + Ry}) \\
&\approx \frac{GMmy}{R^2},
\end{align*}
for $y << R$. Recall that the gravitational acceleration of a particle of mass $m$ is $g = \frac{GM}{R^2}$. So
we can write this in terms of $g$ as
\begin{align*}
V(y) &\approx mgy.
\end{align*}
\end{proof}
\blue{This expression for potential energy decreases as the height $y$ decreases. It does still
depend on the inverse of the spatial separation, but this factor is approximately constant since
$y << R$. In any case, the mass falls to Earth, since smaller $y$ has lower potential energy. The
derivative of the potential energy is the familiar gravitational force
\begin{align*}
mg = \frac{GMm}{R^2}.
\end{align*}
}
\subsection*{Conservation of momentum}
The basic argument for conservation of momentum derives from Newton's 3rd law:
Suppose particle $a$ is exerting a force $\vec{F}_{ab}$ on particle $b$. Since
$\vec{F} = \dv{\vec{p}}{t}$, we have that
\begin{align*}
\int_{t=0}^t \vec{F}_{ab} \dt = \vec{p_b(t)} - \vec{p_b(0)}.
\end{align*}
From Newton's Third Law we have $\vec{F}_{ab} = -\vec{F}_{ba}$, which implies
$\vec{p_b(t)} - \vec{p_b(0)} = -\(\vec{p_a(t)} - \vec{p_a(0)}\)$ and therefore
\begin{align*}
\vec{p_a(t)} + \vec{p_b(t)} &= \vec{p_a(0)} + \vec{p_b(0)},
\end{align*}
i.e. total momentum is conserved.
Consider an 2-dimensional phase space defined by the velocities of 2 particles. Conservation of
momentum means that the pre- and post-collision system state in this phase space is constrained to
lie on a line: $m_1v_1 + m_2v_2 = \text{constant}$. Another equation (e.g. conservation of energy)
is needed to determine the post-collision velocities.
\subsection*{Collisions in 1-D}
A collision between two particles may be
\begin{itemize}
\item \defn{inelastic}: kinetic energy is lost, e.g. lumps of putty.\footnote{Taylor 3.1 Ex 3.1 p.84, }
\item \defn{elastic}: kinetic energy is conserved, e.g. billiard balls\footnote{Taylor ch4 p. 142, Morin 5.6 p.162 (using CM frame), Morin 5.7 p.164 (using conservation of KE)}
\end{itemize}
\subsubsection*{Perfectly inelastic collision\footnote{Taylor 3.1 p.84}}
The simplest case seems to be ``perfectly inelastic'' collision: two lumps of putty that stick
together (without generating any heat or sound). In this case we have one unknown (the
post-collision velocity), and one equation (conservation of momentum) suffices.
\red{Is kinetic energy conserved here? If not why not?}
\begin{mdframed}
\includegraphics[width=400pt]{img/physics--classical-mechanics--taylor--sec-3-1.png}
\end{mdframed}
We have
\begin{align*}
m_1\vec{v_1} + m_2\vec{v_2} &= (m_1 + m_2)\vec{v} \\
\vec{v} &= \frac{m_1\vec{v_1} + m_2\vec{v_2}}{m_1 + m_2} \\
&= \alpha\vec{v_1} + (1 - \alpha)\vec{v_2},
\end{align*}
where $\alpha = \frac{m_1}{m_1 + m_2}$.
\red{What happens if we do this with conservation of energy? Suppose we are in 1-D}
\begin{align*}
\frac{1}{2}m_1v_1^2 + \frac{1}{2}m_2v_2^2 &= \frac{1}{2}(m_1 + m_2)v^2 \\
v &= \sqrt{\frac{m_1v_1^2 + m_2v_2^2}{m_1 + m_2}}.
\end{align*}
So, letting $v_M$ and $v_E$ be the momentum- and energy- based post-collision velocity answers, we
have
\begin{align*}
v_E^2 &= \alpha v_1^2 + (1 - \alpha)v_2^2 \\
v_M^2 &= \(\alpha v_1 + (1 - \alpha)v_2\)^2.
\end{align*}
Since $x \mapsto x^2$ is a convex function (\ref{convexity}), we have that $v_E > v_M$.
\red{Why is this -- why does assuming conservation of kinetic energy lead to a higher post-collision
velocity than assuming conservation of momentum?} \blue{Is this because squashing the lumps of
putty together necessarily loses KE, so if we assume KE is conserved then we are overestimating
the post-collision KE, which is why the post-collision velocity $v_E$ is greater than the estimate
$v_M$ based on conservation of momentum?}
\subsubsection*{Example: Elastic collision (Morin p.162)}
\begin{mdframed}
\includegraphics[width=400pt]{img/physics--classical-mechanics--morin--sec-5-6-ex.png}
\end{mdframed}
\textit{For consistency with the 3blue1brown ``colliding blocks'' section below we will use
$(m_1, m_2)$ instead of $(m, M)$, and in contrast to the diagram we will have $m_1$ initially
stationary and $m_2$ moving right to left at constant velocity $v_2$ (which is negative, since it
is moving right to left).}
This can be solved in two ways:
\begin{enumerate}
\item Using the center of mass (CM) frame (Morin p.162)
\item Using conservation of kinetic energy (Morin p. 164, )
\end{enumerate}
Let $m_1 = m$ the post-collision masses and velocities be $(m_1, v_1')$ and $(m_2, v_2')$.
{\bf Solution using conservation of kinetic energy}\\
From conservation of momentum and KE we have the system of equations
\begin{align}
m_2v_2 &= m_1v_1' + m_2v_2' \\
\frac{1}{2}m_2v_2^2 &= \frac{1}{2}m_1v_1'^2 + \frac{1}{2}m_2v_2'^2.
\end{align}
Now, we solve for $v_1'$ and $v_2'$...
\begin{verbatim}
#+begin_src mathematica :results raw pp
Solve[m2 v2 == m1 v1p + m2 v2p && m2 v2^2 == m1 v1p^2 + m2 v2p^2, {v1p, v2p}]
#+end_src
#+RESULTS:
: {{v1p -> 0, v2p -> v2}, {v1p -> (2*m2*v2)/(m1 + m2), v2p -> (-(m1*v2) + m2*v2)/(m1 + m2)}}
\end{verbatim}
...and the solution is
\begin{align*}
v_1' &= \frac{2m_2}{m_1 + m_2}v_2 \\
v_2' &= \frac{m_2 - m_1}{m + M}v_2.
\end{align*}
So the stationary mass moves in the direction of impact, and for the moving mass:
\begin{itemize}
\item if it's smaller, it reverses its direction,
\item if it's the same mass, it stops dead,
\item if it's larger, it continues in the same direction.
\end{itemize}
Furthermore, the stationary mass always moves away faster than the moving mass. For values of the
stationary mass approaching zero, the ratio of their speeds approaches 2.
Now\footnote{\url{https://www.youtube.com/watch?v=HEfHFsfGXjs}}, what about if they are both in
motion when they collide? In that case we have
\begin{align}
m_1v_1 + m_2v_2 &= m_1v_1' + m_2v_2' \\
\frac{1}{2}m_1v_1^2 + \frac{1}{2}m_2v_2^2 &= \frac{1}{2}m_1v_1'^2 + \frac{1}{2}m_2v_2'^2.
\end{align}
\begin{verbatim}
#+begin_src mathematica :results raw pp
Solve[m1 v1 + m2 v2 == m1 v1p + m2 v2p &&
m1 v1^2 + m2 v2^2 == m1 v1p^2 + m2 v2p^2,
{v1p, v2p}]
#+end_src
#+RESULTS:
: {{v1p -> v1, v2p -> v2}, {v1p -> (m1*v1 - m2*v1 + 2*m2*v2)/(m1 + m2), v2p -> (2*m1*v1 - m1*v2 + m2*v2)/(m1 + m2)}}
\end{verbatim}
\begin{align*}
v_1' &= \frac{2m_2v_2 - v_1(m_2 - m_1)}{m_1 + m_2} \\
v_2' &= \frac{2m_1v_1 + v_2(m_2 - m_1)}{m_1 + m_2}.
\end{align*}
\newpage
\subsubsection{3blue1brown $\pi$ in colliding blocks}
This 3blue1brown video\footnote{\url{https://www.youtube.com/watch?v=HEfHFsfGXjs}} involves
perfectly elastic collisions of two blocks, and a wall. The block closest to the wall (mass $m_2$)
starts off stationary, and the outer block (mass $m_1$) moves towards it with constant velocity. We
count the number of collisions between blocks, and between the inner block and the wall, for varying
values of the mass $m_1$ of the outer block.
\begin{minted}{python3}
import sys
from dataclasses import dataclass
@dataclass
class Blocks:
m1 = None # mass of right block (supplied as input)
m2 = 1.0 # mass of left block
v1 = -1.0 # initial velocity of right block
v2 = 0.0 # left block starts off stationary
def will_collide_again(self):
return abs(self.v2) > self.v1
def collide(self):
v1, v2 = self.v1, self.v2
m1, m2 = self.m1, self.m2
self.v1 = (2 * m2 * v2 + v1 * (m1 - m2)) / (m2 + m1)
self.v2 = (2 * m1 * v1 - v2 * (m1 - m2)) / (m2 + m1)
def simulate(self):
n = 0
while True:
self.collide()
n += 1
if self.will_collide_again():
# Bounce off wall
self.v2 = -self.v2
n += 1
else:
break
if self.v2 < 0:
# Bounce off wall
n += 1
print(n)
if __name__ == "__main__":
assert not sys.argv[2:]
blocks = Blocks()
blocks.m1 = float(sys.argv[1])
blocks.simulate()
\end{minted}
Here are the counts for $m_1$ taking values that are integer powers of 100:
\begin{verbatim}
$ for i in 0 1 2 3 4 5 6; do python 3blue1brown-blocks.py $((100**i)); done
3
31
314
3141
31415
314159
3141592
\end{verbatim}
OK, so. The problem statement is
\begin{mdframed}
A vertical wall stands at the left edge of a frictionless surface. There are two blocks on the
surface: the left block is stationary and has mass $m_2$, and the right block, with mass $m_1$, is
moving towards the left block at constant (negative) velocity.
Show that for $n \in \N$, if $m_1/m_2 = b^{2n}$ then the total number of collisions (including the
left block bouncing off the wall) is equal to the integer formed from the first $n+1$ digits
of the base-$b$ expansion of $\pi$. So:\\~\\
\begin{tabular}{l|l}
$m_1$ & \#collisions \\
\hline
$m_2$ & 3 \\
$10^2m_2$ & 31 \\
$10^4m_2$ & 314 \\
$10^6m_2$ & 3141 \\
\ldots
\end{tabular}
All collisions are ``elastic'', i.e. momentum and kinetic energy are conserved. You may use the
following result from classical mechanics for elastic collisions:
\textit{Suppose that an object with mass $m_1$ and velocity $v_1$ collides elastically with an
object with mass $m_2$ and velocity $v_2$. The post-collision velocities are
\begin{align*}
v_1' &= \frac{2m_2v_2 - v_1(m_2 - m_1)}{m_1 + m_2} \\
v_2' &= \frac{2m_1v_1 + v_2(m_2 - m_1)}{m_1 + m_2}.
\end{align*}
}
When the left block bounces off the wall, its new velocity has the same magnitude and opposite
sign.
\end{mdframed}
Note that $E := \frac{1}{2}m_1v_1^2 + \frac{1}{2}m_2v_2^2$ is constant throughout the evolution of
the system (i.e. over all collisions).
Suppose we define a phase space $x := v_1$, $y := v_2$. Then the system is constrained to points
satisfying $Ax^2 + By^2 = E$, where $A = m_1/2$ and $B = m_2/2$.
This is an ellipse. Instead, we define the phase space:
\begin{align*}
u_1 &= \alpha_1v_1 \\
u_2 &= \alpha_2v_2,
\end{align*}
where $\alpha_i = \sqrt{m_i/2}$.
Now we have $u_1^2 + u_2^2 = E$ and the system is constrained to lie on a circle of radius $E$ centered
at the origin.
The $(u_1, u_2)$ space is just a change of basis in which the basis vectors point in the same
direction but are scaled differently: to transform a point in $(v_1, v_2)$-space to $(u_1, u_2)$ one
multiplies by
\begin{align*}
\matMMxNN{\alpha_1}{0}
{0}{\alpha_2}.
\end{align*}
At time $t=0$ we have $v_1 < 0$ and $v_2 = 0$, so the state of the system corresponds to the point
on the circle at $-\pi$.
Now consider conservation of momentum. Whenever there is a collision, we have
$m_1v_1 + m_2v_2 = m_1v_1' + m_2v_2'$. In other words, suppose the system is at $(v_1, v_2)$ and
define $p = m_1v_1 + m_2v_2$. Then the possible new velocities lie on a line
\begin{align*}
m_1v_1' + m_2v_2' &= p \\
v_2' &= -\frac{m_1}{m_2}v_1' + \frac{p}{m_2}.
\end{align*}
This will still be a straight line in $(u_1, u_2)$ space: from
\ref{equation-of-line-under-change-of-basis} we have that
\begin{align*}
u_2'
&= -\frac{\alpha_1m_1}{\alpha_2m_2}u_1' + \frac{p}{\alpha_2m_2} \\
&= -\frac{\beta_1}{\beta_2}u_1' + \frac{p}{\beta_2},
\end{align*}
where $\beta_i = \sqrt{m_i^3/2}$.
Initially, $v_1 = 0$, and $v_2 < 0$, and $p = m_2v_2 < 0$, and $E = \frac{1}{2}m_2v_2^2$.
If $m_1 = m_2 = m$ then $u_2' = 0$ and
$u_1' = \frac{p}{\beta_1} = \frac{mv_2}{\sqrt{m^3/2}} = \frac{v_2}{\sqrt{m/2}}$. \red{This doesn't
seem right, we should have $u_1'^2 = E$}.
\begin{mdframed}
\includegraphics[width=400pt]{img/3blue1brown--pi-colliding-blocks-1.png}
\end{mdframed}
\newpage
If at any time $|v_1'| - v_2' \leq 0$ then the right block is escaping to the right and the left
block will never catch up with it. Otherwise $|v_1'| - v_2' > 0$ and there will be another
collision.
Initially, $v_1 = 0$, and $v_2 < 0$, and we have post-collision velocities
\begin{align*}
v_1' &= \frac{2m_2}{m_1 + m_2}v_2 \\
v_2' &= \frac{m_2 - m_1}{m_1 + m_2}v_2.
\end{align*}
Thus after the first collision the further-collisions criterion is
\begin{align*}
(m_1 + m_2)(|v_1'| - v_2')
&= 2m_2(-v_2) - (m_2 - m_1)v_2 \\
&= v_2(m_1 - 3m_2).
\end{align*}
Therefore there will be another collision if
\begin{align*}
|v_1'| - v_2' &> 0 \\
v_2(m_1 - 3m_2) &> 0 \\
m_1 - 3m_2 &< 0 \\
m_1 &< 3m_2.
\end{align*}
Note that we are only considering $m_2 \geq m_1$, so this is always true and there are 3 collisions.
\section{Projectile motion}
A projectile of mass $m$ is released with velocity $v_0$ at an angle $\theta$ to the ground. There
is no air resistance.
Note that vertical and horizontal motion are independent of each other. The equations of motion are:
\begin{itemize}
\item {\bf Vertical}\\
$\ddot{y}(t) = -g$, with initial condition $\dot{y}(0) = v_0\sin\theta$.
\end{itemize}
\subsection{Using integration / FTC to solve the equation of motion}
Informally, since the acceleration is constant, the solution for the velocity function must be
$\dot{y}(t) = v_0\sin\theta - gt$, and therefore by integration the solution for vertical
position is $y(t) = y_0 + v_0\sin\theta t -\frac{1}{2}gt^2$.
More formally, we want to identify the set of functions $y$ that are consistent with the facts:
\begin{align*}
\ddot{y}(t) &= -g \\
\dot{y}(0) &= v_0 \\
y(0) &= y_0.
\end{align*}
The first step is to identify the set of first-derivative functions $\dot{y}$ that fit the
facts. Using only the fact that the second derivative is a constant $-g$, we conclude that the
first derivative can be any linear function: $\dot{y}(t) = C - gt$. Then using the initial
velocity, we narrow this further to $\dot{y}(t) = v_0 - gt$.
More formally... the ``antiderivative'' operation maps a single function to a (infinite) set of
functions.
\begin{align*}
\int \ddot{y}(t) \dt = \int -g \dt = C - gt.
\end{align*}
But the operation of integration maps an $\R \to \R$ function to $\R$.
We have an $\R \to \R$ function $\ddot{y}(t) = -g$. Antidifferentiating tells us that a family of
linear velocity functions are consistent with the acceleration function. What does integration
tell us? It tells us the ``net amount'' of acceleration that has accumulated between time $0$ and
time $t$. (And the FTC tells us that antidifferentiating gives us a trick to find this net amount
easily.)
It's easier to think about velocity and distance. ``Net amount of accumulated velocity'',
i.e. area under the velocity graph, corresponds to net displacement. E.g. $70$ mph $\times$ $2$
hrs equals $140$ miles displacement. What does that familiar calculation correspond to formally?
$140$ miles displacement is saying $x(2) - x(0) = 140$. So the statement is that
\begin{align}
x(t) - x(0) &= \int_{t'=0}^{t'=t} \dxdt(t') \dt' \label{pcm-ftc} \\
&= \int_{t'=0}^{t'=t} 70 \dt'. \nonumber
\end{align}
In this case, \eqref{pcm-ftc} is common sense, because the velocity is constant. But for an
arbitrary velocity function, it would be invoking the FTC.
To compute that integral we can either
\begin{enumerate}
\item Use common sense again: visualize it as the area of a rectangle with height $70$ and width
$t$.
\item Invoke the FTC a second time: notice that area is increasing linearly with $t$ with a slope
of $70$, so the answer to the integral is given by the difference in value at $0$ and at $t$ of
some function which has a slope equal to the integrand. Obvious for a constant integrand, but
the FTC says that that line of thought still holds when the integrand is any well-behaved
function. In other words, we need to antidifferentiate: find a function that has derivative
$70$. In other words, we need to solve a differential equation: $f' = 70$.
\end{enumerate}
So in the velocity/distance problem, the facts were
\begin{align*}
\dot{x} &= 70 \\
x(0) &= 0,
\end{align*}
we wanted to know the function $x$, and we found the solution $x(t) = 70t$ by solving the equation
\begin{align*}
x(t) - x(0) = \int_0^t 70 \dt.
\end{align*}
Returning to the vertical acceleration of the projectile, the area under the acceleration graph
corresponds to net change in velocity. Recall that the facts are
\begin{align*}
\ddot{y}(t) &= -g \\
\dot{y}(0) &= v_0 \\
y(0) &= y_0.
\end{align*}
We want to know the function $y$. So, invoking the FTC, we write down the equation
\begin{align*}
\int_{t'=0}^{t'=t}\ddot{y}(t') \dt' &= \dot{y}(t) - \dot{y}(0) \\
\int_{t'=0}^{t'=t}-g \dt' &= \dot{y}(t) - v_0.
\end{align*}
Then, to solve this equation, we invoke FTC again, identifying $-gt$ as an antiderivative, and
conclude that
\begin{align*}
-gt'\Big|_{t'=0}^{t'=t} &= \dot{y}(t) - v_0 \\
\dot{y}(t) &= v_0 - gt.
\end{align*}
Then, we repeat the procedure, writing down
\begin{align*}
\int_{t'=0}^{t'=t}\dot{y}(t') \dt' &= y(t) - y(0) \\
\int_{t'=0}^{t'=t}v_0 - gt \dt' &= y(t) - y_0 \\
v_0t' - \frac{1}{2}gt'^2\Big|_{t'=0}^{t'=t} &= y(t) - y_0 \\
y(t) &= y_0 + v_0t - \frac{1}{2}gt^2.
\end{align*}