-
Notifications
You must be signed in to change notification settings - Fork 2
/
analysis--berkeley-202a-hw09.tex
479 lines (341 loc) · 18.4 KB
/
analysis--berkeley-202a-hw09.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
\section{Math 202A - HW9 - Dan Davison - \texttt{[email protected]}}
% In Bass 8.5, most of you guessed the counterexample f(x)=−(xlogx)−1 or some similar variant. To motivate this
% counterexample, note that x↦x−1 is not integrable near 0 but x↦x−1+ε is. You want something which is not
% integrable, but which is "weakly integrable" (in the sense that it satisfies the conclusion of Bass 8.4). So
% the natural thing to do is search for functions whose growth rate is between x−1 and x−1+ε near the origin,
% which is what this counterexample gives.
% However, when it came time to actually integrate f, a lot of you struggled, and came up with complicated
% solutions. The good news, that some of you picked up on, is Theorem 9.1 in Bass: the Lebesgue and Riemann
% integrals agree whenever both exist. A short proof of this for f: since we're taking an improper Riemann
% integral, we can restrict to the domain I=[δ,1] and take δ→0. Since the Riemann integral exists (since f is
% continuous on the compact set I) it is equal to the limit of lower Riemann sums ∫fη as the mesh |η| of a
% partition η tends to 0. But fη→f pointwise, since fη was sampled from a continuous function (so if |η| is small
% then |f−fη| is also small; this is a typical epsilon-delta argument). But f was Lebesgue integrable on I and
% dominated the fη, so that implies that ∫fη→∫f.
% Another issue here is that some folks recognized that integrating f amounts to computing the Riemann integral,
% but then fumbled on computing the antiderivative. It is not F(x)=loglogx as you might naively expect, since F
% blows up at 1 and is ill-defined (complex-valued and depends on a choice of branch) on [0,1). This can be fixed
% with some calculus, however.
% As for Bass 10.2, let me emphasize that convergence in the metric d that we introduced is not equivalent to
% convergence almost everywhere. It is true that convergence almost everywhere implies convergence in measure
% (hence in the metric d), and convergence in measure (hence in the metric d) implies convergence almost
% everywhere along a subsequence, but the typewriter sequence is a counterexample that prevents us from promoting
% this convergence on a subsequence to convergence almost everywhere.
% 8.5. I don't think this works (I don't even see why that would make the integral diverge). The motivation here is the condition here *almost* holds for 1/x, if we were to just perturb it slightly -- so the correct answer is something like -1/x log x. (-8)
% Aidan Backus, Nov 14 at 9:57am
% 10.2. Small nitpick on the triangle inequality -- you definitely want < rather than \leq. Much more egregiously, convergence in measure is not equivalent to convergence a.e. Since \mu(X) < \infty, convergence a.e. implies convergence in measure, but you can only get the converse along a subsequence (the counterexample is a typewriter sequence). (-4)
% Aidan Backus, Nov 15 at 10:15am
% 8.4) 10 8.5) 2 10.2) 6 10.7) 6
% Ian Francis, Nov 24 at 1:44am
% 10.7) (edited) For the ending concern about 2eps, this is not an issue: If 0 \leq x < 2\eps for all \eps, then x=0. It seems like you're heading in the right direction, but, one comment: - When you define your set F, you want to make it a countable union of countable intersections of unions, but \cup_{\eps>0} is an uncountable union, so you wouldn't be able to conclude that \mu(F) = 0 just because each E_{n, \eps} has measure 0. Instead of \eps, use 1/k. This would make it countable. A countable union of measure zero sets necessarily has measure 0 by sub-additivity. (-4) Otherwise you have a lot of the key ingredients here. But, you could simplify it quite a bit, and that would probably lead to less self-confusion.
% Ian Francis, Nov 24 at 2:19am
\begin{mdframed}
\includegraphics[width=400pt]{img/analysis--berkeley-202a-hw09-210c.png}
\end{mdframed}
\begin{proof}
Let $t_1, t_2, \ldots \in \R$ with $\limn t_n = \infty$ and define $A_n = \{x ~:~ f(x) > t_n\}$.
Thus we have $\lim_{t\to\infty} t ~ \mu\big(\{x ~:~ f(x) \geq t\}\big) = \limn t_n \mu(A_n)$, since the limit
is the same along any sequence.
Note that $f\ind_{A_n} \to 0$, since $A_n \downarrow \emptyset$, and note also that $f\ind_{A_n} \leq f$ for
all $n$.
Therefore
\begin{align*}
\limn t_n \mu(A_n)
&\leq \limn \int f\ind_{A_n} \\
&= \int \limn f\ind_{A_n} &\text{by the dominated convergence theorem}\\
&= \int \limn 0 \\
&= 0.
\end{align*}
\end{proof}
\newpage
\begin{mdframed}
\includegraphics[width=400pt]{img/analysis--berkeley-202a-hw09-b0e8.png}
\end{mdframed}
% 8.5. I don't think this works (I don't even see why that would make the integral diverge). The motivation
% here is the condition here *almost* holds for 1/x, if we were to just perturb it slightly -- so the correct
% answer is something like -1/x log x. (-8)
\begin{proof}
My original thought was to take the graph of $y = x^{-1/2}$ and place a copy of it over every rational,
scaled to fit in $\eps/2^k$ where $k$ indexes an enumeration of the rationals, somehow stitching them
together. However, while something like that might make the integral diverge, I actually don't think it
solves the limit requirement, and may well not make sense and indeed I haven't accomplished it.
\end{proof}
\newpage
\begin{mdframed}
\includegraphics[width=400pt]{img/analysis--berkeley-202a-hw09-113a.png}
\end{mdframed}
\begin{proof}
[Incomplete]
Let $g: [0, 1] \to \R$ be a function that is not constant a.e.
We will construct an $f$ such that $\int_0^1 f = 0$ and $\int_0^1 fg \neq 0$, thus proving the result by
contradiction.
Let $f(x) = x - 1/2$. Note that $\int_0^1 f = 0$.
If $\int_0^1 fg \neq 0$ then we are done.
Alternatively, we have $\int_0^1 fg = 0$, or equivalently
\begin{align*}
\int_0^{1/2} fg + \int_{1/2}^1 fg = 0.
\end{align*}
I feel that it should be possible to solve the problem by contradiction like this: i.e. by specifying a
procedure that modifies $f$ to produce an $f^*$ such that $\int_0^1 f^*g \neq 0$, while preserving the property that $\int_0^1 f^* = 0$. In particular I note that we
are free to cut out two vertical strips of the graph of $f$ and exchange them: this will preserve the value
of the integral.
Unfortunately I failed to complete the question again despite thinking about it for ages :) I'm including a
few more of the incomplete approaches to this problem that I thought about below.
\end{proof}
\begin{proof}
[another incomplete thought]
We have that for every continuous function $f$ with $\int_0^1 f = 0$ then $\int_0^1 fg = 0$.
It follows that for every continuous function $f$ with $\int_0^1 f = 0$ and for all $a, b \in \R$
\begin{align*}
\int_0^1 af + bfg = 0.
\end{align*}
It is given that $g$ is bounded, and we see that $f$ is bounded also since it is continuous on a compact set.
\end{proof}
\begin{proof}
[another incomplete thought]
Let $\ms F = \{f: [0, 1] \to \R ~:~ f \text{~is continuous}, \int_0^1 f(x) \dx = 0 \}$.
Is $\ms F$ (in an appropriate sense) full rank, such that if a function $g$ is orthogonal to every element
of $f$ then $g$ must be the zero function? (And then some argument allowing $g$ to be any constant a.e.
function).
\end{proof}
% \begin{proof}
% Let $f_1 \neq f_2$ be continuous with $\int_0^1 f_1 = \int_0^1 f_2 = 0$.
% We have that
% \begin{align*}
% \int_0^1 f_1g = \int_0^1 f_2g = 0.
% \end{align*}
% For example, let $f_n(x) = n(x - 1/2)$. Then
% \begin{align*}
% \int_0^1 f_n(x) \dx = n\int_0^1 x \dx - \frac{n}{2} = \frac{n}{2} - \frac{n}{2} = 0.
% \end{align*}
% Let
% \begin{align*}
% g(x) =
% \begin{cases}
% -c & x \leq 1/2 \\
% +c & x > 1/2.
% \end{cases}
% \end{align*}
% Then
% \begin{align*}
% \int_0^1 f_n(x) g(x)
% = -cn\int_0^{1/2}(x - 1/2) \dx + cn\int_{1/2}^1(x - 1/2) \dx
% = 0.
% \end{align*}
% \end{proof}
% \begin{remark}
% As a sanity check note that the required behavior with continuous $f$ does obtain
% \begin{enumerate}
% \item if $g$ is equal to a constant a.e.: if $c \in \R$ and $g = c$ a.e. then $\int_0^1 f g = c \int_0^1 f = 0$,
% \item if $f = 0$.
% \end{enumerate}
% \end{remark}
\begin{proof}
[another incomplete thought]
Suppose $g: [0, 1] \to [0, \infty]$ is bounded and measurable.
Let $\lambda(E) = \int_E g(x) \dx$ for a Borel set $E \subseteq [0, 1]$.
Then by HW7 Ex. 1 we have that $\lambda$ is a measure on $[0, 1]$ and
\begin{align*}
\int_0^1 f(x)g(x) \dx = \int_0^1 f \d\lambda.
\end{align*}
Thus the question is equivalent to positing the existence of a measure $\lambda$ such that for every
continuous function $f$ with $\int_0^1 f(x) \dx = 0$ we have
\begin{align*}
\int_0^1 f \d\lambda = 0.
\end{align*}
\end{proof}
% Suppose condition A(f, g) is true whenever f satisfies condition B. Prove that g satisfies condition C.
% Claim: If (f satisfies condition B) => A(f, g), then g satisfies condition C.
% Proof:
% We must show
% (1) if not B(f) then C(g)
% (2) if B(f) and
% \begin{claim*}
% Suppose $g: [0, 1] \to \R$ is bounded and measurable.
% Suppose further that, if $f$ is continuous and $\int_0^1 f = 0$, then
% \begin{align*}
% \int_0^1 f g = 0.
% \end{align*}
% Prove that $g$ is equal to a constant a.e.
% \end{claim*}
% \begin{intuition}
% We have an unknown function $g: [0, 1] \to \R$.
% We are told that whenever $g$ is integrated against a continuous function $f$ for which $\int_0^1 f = 0$, the
% result is zero.
% We must prove that $g$ is equal to a constant a.e.
% \end{intuition}
% \begin{proof}
% Suppose $g$ is not equal to a constant a.e.
% \end{proof}
\newpage
\begin{mdframed}
\includegraphics[width=400pt]{img/analysis--berkeley-202a-hw09-955c.png}
\end{mdframed}
\begin{claim*}
Henstock-Kurzweil $\int \ind_\Q \neq 0$
\end{claim*}
\begin{proof}
[There is something wrong with my argument as it basically proves that $\int \ind_\Q \neq 0$]
We claim that $\int \ind_\Q = 0$.
Let $\eps > 0$.
We must prove that there exists a gauge $\delta$ such that for all tagged partitions $(d, p)$ if $(d, p)$
is $\delta$-fine then
\begin{align*}
\sum_{i=1}^n (d_i - d_{i-1})\ind_\Q(p_i) < \eps.
\end{align*}
Equivalently we must prove that there exists a gauge $\delta$ such that
\begin{align*}
\max_{(d, p) \delta\text{-fine}} \Bigg( \sum_{\{i ~:~ i \in \{1, \ldots, n\}, p_i \in \Q\}} (d_i - d_{i-1}) \Bigg) < \eps,
\end{align*}
where we were able to remove the absolute value operator because the $d_i$ are increasing.
Note that the tagged partition that maximises this quantity will be one for which $p_i \in \Q$ for
all $i \in \{1, \ldots, n\}$.
Therefore we must prove that there exists a gauge $\delta$ such that
\begin{align*}
\max_{(d, p) \delta\text{-fine}, ~\forall i ~ p_i \in \Q} \sum_{i=1}^n (d_i - d_{i-1}) < \eps,
\end{align*}
Note that if $(d, p)$ is a tagged partition such that $p_i \notin \Q$ for all $i \in \{1, \ldots, n\}$, then
the condition is true. Therefore we need only consider tagged partitions for which some $p_i$ are rational.
In fact it seems to me that we need to show it's true for tagged partitions for which all $p_i$ are rational,
since this will maximises the sum. However, in that case we will simply have
\begin{align*}
\sum_{i=1}^n (d_i - d_{i-1}) = 1.
\end{align*}
\red{TODO} I'm confused about how I'm misunderstanding the definition. I can see solutions to this problem online,
such as \url{https://www.math.unm.edu/~crisp/courses/math402/spring15/HKintegralStevenJocelyn.pdf}, which I studied but my confusion remained (the
presentation there is very similar to here).
\end{proof}
% \begin{proof}
% \begin{align*}
% \hline
% &\forall \eps > 0 ~~ \exists \delta ~~ \forall (d, p) ~~ \delta\text{-fine}(d, p) ~~ \Big|\sum_{\{i ~:~ i \in \{1, \ldots, n\}, p_i \in \Q\}} (d_i - d_{i-1})\Big| < \eps
% \end{align*}
% \begin{align*}
% &\eps > 0 \\
% \hline
% &\exists \delta ~~ \forall (d, p) ~~ \delta\text{-fine}(d, p) ~~ \Big|\sum_{\{i ~:~ i \in \{1, \ldots, n\}, p_i \in \Q\}} (d_i - d_{i-1})\Big| < \eps
% \end{align*}
% \end{proof}
\newpage
\begin{mdframed}
\includegraphics[width=400pt]{img/analysis--berkeley-202a-hw09-b4e1.png}
\end{mdframed}
% 10.2. Small nitpick on the triangle inequality -- you definitely want < rather than \leq.
% Much more egregiously, convergence in measure is not equivalent to convergence a.e. Since \mu(X) < \infty,
% convergence a.e. implies convergence in measure, but you can only get the converse along a subsequence (the
% counterexample is a typewriter sequence). (-4)
\begin{claim*}
$d$ is a metric on the space of measurable functions except for the fact that $d(f, g) = 0$ implies $f = g$
a.e., not necessarily everywhere.
\end{claim*}
\begin{proof}~\\
\begin{enumerate}
\item {\bf identity of indiscernibles}\\
Note that $\frac{|f - g|}{1 + |f - g|} \geq 0$. Therefore
if $d(f, g) = \int\frac{|f - g|}{1 + |f - g|} = 0$ then $|f - g| = 0$ a.e., and therefore $f = g$ a.e.
Conversely, if $f = g$ a.e. then $|f - g| = 0$ a.e. and we have $d(f, g) = 0$.
\item {\bf symmetry}\\
$d(f, g) = \int \frac{|f - g|}{1 + |f - g|} = \int \frac{|g - f|}{1 + |g - f|} = d(g, f)$
\item {\bf triangle inequality}
\begin{align*}
d(f, h)
&= \int \frac{|f - h|}{1 + |f - h|} \\
&= \int \frac{|f - g + g - h|}{1 + |f - g + g - h|} \\
&\leq \int \frac{|f - g| + |g - h|}{1 + |f - g| + |g - h|} \\
&< \int \frac{|f - g|}{1 + |f - g|} + \int \frac{|g - h|}{1 + |g - h|} \\
&= d(f, g) + d(g, h)
\end{align*}
\end{enumerate}
\end{proof}
\begin{lemma}\label{10-2-lemma}
$d(f_n, f) \to 0$ if and only if $f_n \to f$ a.e.
\end{lemma}
\begin{proof}
Since $0 \leq \frac{|f_n - f|}{1 + |f_n - f|} \leq 1$ we may appply the dominated convergence theorem,
yielding
\begin{align}
\limn d(f_n, f)
&= \limn \int \frac{|f_n - f|}{1 + |f_n - f|} \nonumber\\
&= \int \limn \frac{|f_n - f|}{1 + |f_n - f|} \nonumber\\
&= \int \frac{\limn |f_n - f|}{1 + \limn |f_n - f|}. \label{10-2-eqn}
\end{align}
First suppose that $d(f_n, f) \to 0$. Then, since for all $n$ the integrand on the RHS of \eqref{10-2-eqn} is
non-negative, and the denominator of the integrand on the RHS of \eqref{10-2-eqn} is strictly positive, we
have $\limn |f_n - f| = 0$ a.e. or in other words $f_n \to f$ a.e.
Converself, suppose that $f_n \to f$ a.e. Then the integrand on the RHS of \eqref{10-2-eqn} is zero and we
have $d(f_n, f) \to 0$.
\end{proof}
\begin{claim*}
If $d(f_n, f) \to 0$ then $f_n \to f$ in measure.
\end{claim*}
\begin{proof}
From lemma \ref{10-2-lemma} we have $f_n \to f$ a.e. Therefore $f_n \to f$ in measure.
\end{proof}
\begin{claim*}
If $f_n \to f$ in measure then $d(f_n, f) \to 0$.
\end{claim*}
\begin{proof}
Suppose that $f_n \to f$ in measure.
Let $\eps > 0$ and define $A_{\eps, n} = \{x ~:~ |f_n(x) - f| > \eps\}$.
Since $0 \leq \ind_{A_{\eps, n}} \leq 1$ and $\int 1 = \mu(X) < \infty$ we may apply the dominated
convergence theorem, yielding
\begin{align*}
0 = \limn \mu(A_{\eps, n}) = \limn \int \ind_{A_{\eps, n}} = \int \limn \ind_{A_{\eps, n}}.
\end{align*}
Since the integrand $\limn \ind_{A_{\eps, n}}$ is non-negative we have
\begin{align*}
\limn \ind_{A_{\eps, n}}(x) = 0 \ae,
\end{align*}
or equivalently
\begin{align*}
\limn |f_n(x) - f(x)| \leq \eps \ae
\end{align*}
Thus since $\eps$ was arbitrary we have
\begin{align*}
\limn |f_n(x) - f(x)| = 0 \ae
\end{align*}
and therefore
$f_n \to f$ a.e.
The claim then follows from lemma \ref{10-2-lemma}.
\end{proof}
\newpage
\begin{mdframed}
\includegraphics[width=400pt]{img/analysis--berkeley-202a-hw09-ea17.png}
\end{mdframed}
% 10.7) (edited) For the ending concern about 2eps, this is not an issue: If 0 \leq x < 2\eps for all \eps, then
% x=0. It seems like you're heading in the right direction, but, one comment: - When you define your set F, you
% want to make it a countable union of countable intersections of unions, but \cup_{\eps>0} is an uncountable
% union, so you wouldn't be able to conclude that \mu(F) = 0 just because each E_{n, \eps} has measure 0. Instead
% of \eps, use 1/k. This would make it countable. A countable union of measure zero sets necessarily has measure
% 0 by sub-additivity. (-4) Otherwise you have a lot of the key ingredients here. But, you could simplify it
% quite a bit, and that would probably lead to less self-confusion.
\begin{proof}
It suffices to show that the sequence $f_n$ is Cauchy a.e.
The condition for $f_n(x)$ not Cauchy is that there exists $\eps$ such that for all $N$ there
exists $m, n \geq N$ such that $|f_m(x) - f_n(x)| \geq \eps$.
Equivalently, there exists $\eps$ such that for all $N$ there exists $n \geq N$ such
that $\origsup_{m \geq n} |f_m(x) - f_n(x))| \geq \eps$.
Let $E_{n, a} = \{x ~:~ \origsup_{m \geq n} |f_m(x) - f_n(x)| \geq a\}$.
Then the condition for $f_n(x)$ not Cauchy is that there exists $\eps$ such that for all $N$ there
exists $n \geq N$ such that $x \in E_{n, \eps}$.
Let $F = \bigcup_{\eps > 0} \bigcap_{N=1}^\infty \bigcup_{n \geq N} E_{n, \eps}$.
Then if $x \in F$ then $f_n(x)$ is not Cauchy.
We want to show that $\mu(F) = 0$.
Since $g_n$ converges in measure to $0$ we have by definition that for all $\eps > 0$
\begin{align*}
\limn \mu(E_{n, \eps}) = 0.
\end{align*}
Equivalently for all $\eps > 0$ and $\eta > 0$ there exists $N$ such that $\mu(E_{n, \eps}) < \eta$ for
all $n \geq N$.
[Argh! I thought I'd be able to finish this one properly but it looks like I'm getting tired and running out
of time. The remainder of this proof is just a vague sketch of where I was trying to go.]
Note that $E_{n, \eps}$ includes (a) points $x$ at which the sequence $\(|f_m(x) - f_n(x)|\)_{m=n}^\infty$
exceeds $\eps$ in supremum finitely many times only and (b) points $x$ at which it never stops
exceeding $\eps$ in supremum. Let's denote category (b) as $E^*_{n, \eps}$. This set contains the $x$ at
which $\(f_n(x)\)_{n=1}^\infty$ is not Cauchy.
Since $E^*_{n, \eps} \subseteq E_{n, \eps}$ we have $\mu(E^*_{n, \eps}) < \eta$, but $\eta$ was arbitrary hence $\mu(E^*_{n, \eps}) = 0$.
I was hoping to formally link $E^*_{n, \eps}$ with the set $F$ defined above.
I'm also kind of concerned that there's a factor-of-two argument that I haven't used yet:
if $\origsup_{m\geq n} |f_m(x) - f_n(x)| < \eps$, then for independent pairs $j, k \geq n$ we
have $|f_j(x) - f_k(x)| < 2\eps$.
\end{proof}