-
Notifications
You must be signed in to change notification settings - Fork 2
/
analysis--berkeley-202a-hw01.tex
398 lines (323 loc) · 17.8 KB
/
analysis--berkeley-202a-hw01.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
\section{Math 202a - HW1 - Dan Davison - \texttt{[email protected]}}
% 2: You actually managed to show that C((0, 1), R) is not a metric space, and the metric is only defined on bounded functions. Of course, this is a failure of compactness, so this works -- but it's worth noting that you can even find a discrete uncountable set of *bounded* continuous functions on (0, 1).
% Aidan Backus, Sep 10 at 6:37am
% 6: That I(\phi^-) < I(\phi^+) for every \phi^\pm does not imply that \sup_{\phi^-} I(\phi^-) < \inf_{\phi^+} I(\phi^+). Just look at the sets (-\infty, 0) and (0, \infty) as an example. In fact, in this case, f is Riemann integrable, since there are \phi^+ such that I(\phi^+) is arbitrarily small.
% Aidan Backus, Sep 14 at 2:27pm
% 2. 10 4. 10 6. 7 7. 10
% Ian Francis, Sep 16 at 10:37am
\begin{mdframed}
\includegraphics[width=400pt]{img/analysis--berkeley-202a--homework-1-a75a.png}
\end{mdframed}
\begin{proof}
$d$ is a metric if it satisfies (I) $d(f,f) = 0$, (II) $d(f,g) = d(g, f)$, and (III) $d(f,g) + d(g, h) \le d(f, h)$.
(I) is satisfied: $d(f, f) = \int_{[0,1]}|f(x) - f(x)| \dx = 0$.
(II) is satisfied:
\begin{align*}
d(f, g)
&= \int_{[0,1]}|f(x) - g(x)| \dx \\
&= \int_{[0,1]}|g(x) - f(x)| \dx \\
&= d(g, f),
\end{align*}
(III) is satisfied:
\begin{align*}
d(f, g) + d(g, h)
&= \int_{[0,1]} |f(x) - g(x)| \dx + \int_{[0,1]} |g(x) - h(x)| \dx \\
&= \int_{[0,1]} |f(x) - g(x)| + |g(x) - h(x)| \dx \\
&\le \int_{[0,1]} |f(x) - g(x) + g(x) - h(x)| \dx \\
&= \int_{[0,1]} |f(x) - h(x)| \dx \\
&= d(f, h).
\end{align*}
\end{proof}
\newpage
\begin{mdframed}
\includegraphics[width=400pt]{img/analysis--berkeley-202a--homework-1-d1d3.png}
\end{mdframed}
\begin{enumerate}
\item
\begin{proof}
$d$ is a metric on the function space $\mathcal C\([0, 1], \R\)$ if it satisfies (I) $d(f,f) = 0$,
(II) $d(f,g) = d(g, f)$, and (III) $d(f,g) + d(g, h) \le d(f, h)$.
(I) is satisfied: $\sup_{x\in [0,1]} |f(x) - f(x)| = \sup_{x\in [0,1]} 0 = 0$.
(II) is satisfied:
\begin{align*}
d(f, g)
&= \sup_{x \in [0,1]}|f(x) - g(x)| \\
&= \sup_{x \in [0,1]}|g(x) - f(x)| \\
&= d(g, f),
\end{align*}
(III) is satisfied:
\begin{align*}
d(f, g) + d(g, h)
&= \sup_{x \in [0,1]} |f(x) - g(x)| + \sup_{x \in [0,1]} |g(x) - h(x)| \\
&= \sup_{x \in [0,1]} \Big(|f(x) - g(x)| + |g(x) - h(x)|\Big) \\
&\le \sup_{x \in [0,1]} |f(x) - h(x)| \\
&= d(f, h).
\end{align*}
\end{proof}
\item
\begin{claim*}
$\mc C\([0, 1], \R\)$ is separable.
\end{claim*}
\begin{proof}
Let $\mc C$ be the set of continuous functions $[0, 1] \to \R$.
Fix an arbitrary function $f \in \mc C$ and fix some $\epsilon > 0$.
Define $g^*_n: [0, 1] \to \R$ as follows:
\begin{enumerate}
\item For all $i \in 0, 1, 2, \ldots, n$ set $x_i = i/n$.
\item For all $i \in \{0, 1, 2, \ldots, n\}$, set $y^*_i = f(x_i)$. (Note that $y^*_i$ is in general not a rational
number; we will account for this later.)
\item For all $i \in \{1, 2, \ldots, n\}$ draw a straight line segment connecting $(x_{i-1}, y^*_{i-1})$
and $(x_i, y^*_i)$.
\item Define $g^*_n: [0, 1] \to \R$ to be the function whose graph was just drawn. (It is possible to give an
explicit procedure for computing $g^*_n(x)$ by finding the interval in which $x$ lies and then using linear
interpolation.)
\end{enumerate}
We now modify the definition of the family of approximating functions so that the $y$-coordinates of the
endpoints are rational. Define $g_n: [0, 1] \to \R$ as follows:
\begin{enumerate}
\item Construct the sets of points $\{(x_i, y^*_i) ~|~ i \in \{0, 1, 2, \ldots, n\}\}$ as above.
\item For $i \in \{0, 1, 2, \ldots, n\}$ set $y_i$ equal to a rational number in the
interval $(y^*_i - \epsilon/4, y^*_i)$. (Such a rational number exists: for example, set $k$ equal to the
smallest natural number such that $1/k < \epsilon/4$, and then set $j$ equal to the smallest natural number
such that $j/k > y^*_i - \epsilon/4$. Then $j/k$ is a rational number in $(y^*_i - \epsilon/4, y^*_i)$.)
\item For all $i \in \{1, 2, \ldots, n\}$ draw a straight line segment connecting $(x_{i-1}, y_{i-1})$
and $(x_i, y_i)$.
\item Define $g_n: [0, 1] \to \R$ to be the function whose graph was just drawn.
\end{enumerate}
Note that, since $f$ is continuous on a compact domain, $f$ is uniformly continuous. Fix some $\delta > 0$ such
that $|x - x'| < \delta \implies |f(x) - f(x')| < \epsilon$ for all $(x, x') \in [0, 1]^2$.
Set $m$ equal to the smallest natural number such that $1/m < \delta/2$ and note
that $|f(x) - g_n(x)| < \epsilon$ for all $x \in [0, 1]$ due to the uniform continuity of $f$. (Informally,
this is true because we can view uniform continuity as stating that a rectangle of base $\delta$ and
height $\epsilon$ can be positioned over the graph at any point such that the graph intersects the left and
right edges of the rectange but does not otherwise leave the rectangle. Our piecewise affine function
consists of straight line segments that fit within such rectangles.) Therefore $d(g_n, f) < \epsilon$ for
all $n \geq m$ and so $\{g_n | n \in \N\}$ is dense in $\mc C$.
Finally we must show that $\{g_n ~|~ n \in \N\}$ is countable. Note that $g_n$ is piecewise affine for a
given $n$, and that the $x$-coordinates of the endpoints are fixed. Thus for a given $n$, the cardinality
of $\{g_n\}$ is equal to the cardinality of the set of possible $y$-coordinates. The latter set is $\Q^n$.
Thus the cardinality of $\{g_n ~|~ n \in \N\}$ is equal to the cardinality of the
set $\bigcup_{n\in \N} \Q^n$. This is a countable union of countable sets and is therefore countable.
\end{proof}
\item
\begin{proof}
Let $f_s: (0, 1) \to \R$ be given by $f_s(x) = \frac{1}{r(s)x}$, where $s \in 2^\N$ and $r(s) \in [0, 1]$ is
the real number corresponding to $s$, i.e. the number $r(s) = 0.d_1d_2d_3\ldots$ where
\begin{align*}
d_i =
\begin{cases}
1, ~~~~ \text{if} ~~~~ i \in s,\\
0, ~~~~ \text{otherwise}.
\end{cases}
\end{align*}
Note that for real $a, b$ we have
\begin{align*}
\frac{1}{ax} - \frac{1}{bx} = \frac{b - a}{abx}
\end{align*}
and therefore the supremum distance between any two elements $f_{s_1}$ and $f_{s_2}$ is unbounded
as $x \to 0$.
\end{proof}
\item
\begin{proof}
Let $\mc C$ be the set of continuous functions $f: (0, 1) \to \R$.
Assume for a contradiction that $\mc C$ is separable. Let $\mc G \subset \mc C$ be a countable dense set of
functions.
Recall that in part (3) we found an uncountable set $\mc H \subset \mc C$ with the property that every pair
of elements in $\mc H$ is at supremum distance at least one.
But this is a contradiction, since we can establish a bijection between $\mc H$ and $\mc G$ as follows:
Pick an element $h_1 \in \mc H$. Since $\mc G$ is dense in $\mc C$, there exists $g_1 \in \mc G$ such
that $d(h_1, g_1) < 1/2$. Now pick $h_2 \in \mc H$ such that $h_1 \neq h_2$. Again, there
exists $g_2 \in \mc G$ such that $d(h_2, g_2) < 1/2$. Furthermore, by the triangle
inequality, $g_2 \neq g_1$. Continuing in this fashion, on the $i$-th iteration we pick $h_i \in \mc H$ and
find a nearby $g_i \in \mc G$ such that $d(h_i, g_i) < 1/2$, and by the triangle inequality conclude
that $g_i \neq g_j$ for all $j < i$.
Thus we can associate each element of $\mc H$ with a unique element of $\mc G$ and conclude that the
cardinality of $\mc G$ equals that of $\mc H$, which is that of the power set of the natural numbers.
But $\mc G$ is countable; a contradiction. Therefore no such countable dense set $\mc G$ exists and $\mc C$
is not separable.
\end{proof}
But can find bounded fns with image [0, 1]
\begin{mdframed}
\includegraphics[width=400pt]{img/analysis--berkeley-202a-hw-7459.png}
\end{mdframed}
\end{enumerate}
\newpage
\begin{mdframed}
\includegraphics[width=400pt]{img/analysis--berkeley-202a--homework-1-8349.png}
\end{mdframed}
\begin{proof}
Let $E_m^a = \{x \in [0, 1] : |f_m(x) < a|\}$ and let $T = \bigcap_{k=1}^\infty \bigcup_{l=1}^\infty \bigcap_{m > l} E_m^{1/k}$.
Informally, $E_m^a$ is the set of points for which $f_m$ is within $a$ of zero.
Let $f_n: [0, 1] \to \R$ be a sequence of functions and let $S \subseteq [0, 1]$ be the set of points $x$
such that $f_n(x) \to 0$ as $n \to \infty$.
First we prove that $x \in S \implies x \in T$.
So let $x \in S$. Then from the definition of limit we have
\begin{align*}
&\forall \epsilon>0 ~~~~ \exists l \in \N ~~~~ \forall m \geq l ~~~~ x \in E_m^\epsilon \\
\iff &\forall \epsilon>0 ~~~~ \exists l \in \N ~~~~ x \in \bigcap_{m \geq l} E_m^\epsilon \\
\iff &\forall \epsilon>0 ~~~~ x \in \bigcup_{l=1}^\infty \bigcap_{m \geq l} E_m^\epsilon \\
\iff & x \in \bigcap_{k=1}^\infty \bigcup_{l=1}^\infty \bigcap_{m \geq l} E_m^{1/k} = T,
\end{align*}
as required.
Secondly we prove that $x \in T \implies x \in S$.
So let $x \in T$. We have
\begin{align*}
x \in \bigcap_{k=1}^\infty \bigcup_{l=1}^\infty \bigcap_{m \geq l} E_m^{1/k},
\end{align*}
which is equivalent to the statement
\begin{align*}
\forall k>0 ~~~~ \exists l \in \N ~~~~ \forall m \geq l ~~~~ |f_m(x)| < \frac{1}{k}.
\end{align*}
Let $\epsilon > 0$ be a real number. Then there exists $k \in \N$ such that $\frac{1}{k} < \epsilon$. Therefore we have
\begin{align*}
\forall \epsilon>0 ~~~~ \exists l \in \N ~~~~ \forall m \geq l ~~~~ |f_m(x)| < \epsilon
\end{align*}
which is equivalent to $x \in S$, as required.
\end{proof}
\begin{proof}
Let $S \subseteq [0, 1]$ be the set of points $x$ for which $f_n(x)$ converges as $n \to \infty$. Since every
convergent sequence in the reals is Cauchy, we have that $x \in S$ is equivalent to
\begin{align*}
\forall \epsilon > 0 ~~~~ \exists l \in \N ~~~~ \forall m \geq l ~~~~ \forall n \geq l ~~~~ |f_m(x) - f_n(x)| < \epsilon,
\end{align*}
which is equivalent to
\begin{align*}
\forall \epsilon > 0 ~~~~ x \in \bigcup_{l=1}^\infty \bigcap_{m \geq l} \bigcap_{n \geq l} E_{m,n}^\epsilon.
\end{align*}
Therefore, we have that $x \in S$ implies
\begin{align*}
x \in \bigcap_{k=1}^\infty \bigcup_{l=1}^\infty \bigcap_{m \geq l} \bigcap_{n \geq l} E_{m,n}^{1/k}.
\end{align*}
As before, the reverse implication also holds since, for any given $\epsilon > 0$, we can find a $k \in \N$
such that $\frac{1}{k} < \epsilon$.
\end{proof}
\newpage
\begin{mdframed}
\includegraphics[width=400pt]{img/analysis--berkeley-202a--homework-1-f175.png}
\end{mdframed}
\begin{proof}
Let $X \subset [0, 1]$ be the subset of real numbers without any 6 in their decimal expansion. Let $x \in X$ and let
\begin{align*}
d_n(x) =
\begin{cases}
0, ~~~~ n\text{-th decimal place of }x \text{ is }0\\
1, ~~~~ \text{otherwise}.
\end{cases}
\end{align*}
Define $f: X \to [0, 1]$ by setting $f(x)$ equal to the real number whose binary expansion
is $0.d_1(x)d_2(x)\cdots$.
Note that for any real number $\om \in [0, 1]$, there exists $x \in X$ such that $f(x) = \om$. To find such
an $x$, we could for example choose the number whose decimal expansion is equal to the binary expansion
of $\om$.
Therefore $f$ is a non-injective surjection from $X$ to the reals in $[0, 1]$, and so the cardinality of $X$
is at least that of the reals. Since $X \subset \R$ we conclude that the cardinality of $X$ is equal to that
of the reals.
\end{proof}
\newpage
\begin{mdframed}
\includegraphics[width=400pt]{img/analysis--berkeley-202a--homework-1-5192.png}
\end{mdframed}
\begin{proof}
Let $X \subseteq [0, 1]$ be the set of points at which $f$ is continuous. We want to show that $f$ is
continuous at $x$ iff $x$ is not rational.
Suppose for a contradiction that $f$ is continuous at a rational point $x = p/q$, where $p, q$ are
non-negative integers. Then $f(x) = 1/q$. But there will always be irrational points within a given
distance $\delta$ of $x$, now matter how small $\delta$ is, and at such an irrational point $x'$ we
have $|f(x) - f(x')| = |1/q - 0| = 1/q$. Therefore $f$ is not continuous at $x$ since the definition of
continuity does not hold for $\epsilon < 1/q$.
Now let $x$ be irrational, so that $f(x) = 0$. Fix an arbitrary $\epsilon > 0$. We want to show that there
exists a $\delta$ such that $1/q < \epsilon$ for any rational point $p/q$ lying within $\delta$ of $x$,
where $p/q$ is in reduced terms. If $\epsilon > 1/2$ then any $\delta$ will work, so
assume $\epsilon \leq 1/2$. Let $k$ be the largest natural number such that $1/k \geq \epsilon$, let $i$ be
the largest natural number such that $i/k < x$ and let $j$ be the smallest natural number such
that $j/k > x$. Then a choice of $\delta = \frac{1}{2}\min\{x - i/k, j/k - x\}$ will work to prove continuity
of $f$ at irrational $x$.
\end{proof}
\begin{mdframed}
\includegraphics[width=400pt]{img/analysis--berkeley-202a-hw-af4c.png}
\end{mdframed}
\begin{mdframed}
\includegraphics[width=400pt]{img/analysis--berkeley-202a-hw-5fee.png}
\end{mdframed}
\newpage
\begin{mdframed}
\includegraphics[width=400pt]{img/analysis--berkeley-202a--homework-1-f5e8.png}
\end{mdframed}
\begin{definition*}
$g: [0, 1] \to \R$ is Riemann integrable if
\begin{align*}
\sup_{\phi^-} I(\phi^-) = \inf_{\phi^+} I(\phi^+).
\end{align*}
Here $\phi^-$ and $\phi^+$ are step functions adapted to some
partition $0 \leq x_1 \leq x_2 \leq \ldots \leq x_{n-1} \leq 1$, such that $\phi(x) = c_i$
for $x \in (x_{i-1}, x_i)$. $I(\phi)$ is (informally) the area under the step function $\phi$:
\begin{align*}
I(\phi) = \sum_{i=1}^n c_i(x_i - x_{i-1}).
\end{align*}
And the supremum is over all minorants $\phi^- \leq g$ and the infimum is over all
majorants $\phi^+ \geq g$, where the length $n$ of the partition is allowed to vary as well as the constant
values $\{c_1, c_2, \ldots, c_n\}$ of the step function within each segment.
\end{definition*}
\begin{enumerate}
\item
\begin{claim*}
The specified function $f$ is not Riemann integrable.
\end{claim*}
\begin{proof}
Consider the first segment of any partition: $(0, x_1)$. No matter how small $x_1$ is, there
exists $n \in \N$ such that $1/n < x_1$. Therefore for all majorants we have $c_1 \geq 1$ and yet for all
minorants we have $c_1 \leq 0$. So, when restricted to this first segment, we have $I(\phi^-) > I(\phi^+)$
for all $\phi^-, \phi^+$ and, since every majorant is elsewhere less than every minorant, it is not
possible that $\sup_{\phi^-} I(\phi^-) = \inf_{\phi^+} I(\phi^+)$ and hence the Riemann integral is
undefined.
\end{proof}
\item
\begin{claim*}
$\int_0^1 f > 0$
\end{claim*}
\begin{proof}
Suppose for a contradiction that $\int_0^1 f = 0$. Fix an arbitrary minorant $\phi^-$, adapted to a
partition of length $n$. Then we have that $\sum_{i=1}^n c_i(x_i - x_{i-1}) \leq 0$.
Since $x_i \geq x_{i-1}$ for all $i$, and since $x_0 = 0 < x_n = 1$, it must be the case
that $x_i - x_{i-1} > 0$ for some $i$, and therefore that $c_i > 0$ for some $i$. Therefore $f$ vanishes at
at least one point. This contradiction proves that $\int_0^1 f \neq 0$.
To see that it's not negative, note that for every majorant $\phi^+$ we have $x_i - x_{i-1} \geq 0$
and $c_i > 0$ for all $i$ and therefore $I(\phi^+) = \sum_{i=1}^n c_i(x_i - x_{i-1}) \geq 0$. Therefore $\int_0^1 f > 0$.
\end{proof}
\end{enumerate}
\newpage
\begin{mdframed}
\includegraphics[width=400pt]{img/analysis--berkeley-202a--homework-1-a577.png}
\end{mdframed}
\begin{proof}
Suppose for a contradiction that $[0, 1] \subset \R$ is countable. Fix an enumeration $\{u_n ~|~ n \in \N\}$
of the elements of $[0, 1]$.
If $u_1 > u_2$ then relabel them so that $u_1 < u_2$. Set $U_1 = (u_1, u_2)$.
Continue examining the numbers in the enumeration (starting at $u_3$) until two numbers have been encountered
that are both in $(u_1, u_2)$. Form an interval from this pair and label it $U_2$. Continue examining the
numbers in the enumeration until two numbers are encountered that are both in $U_2$; label this
interval $U_3$. Continue in this fashion indefinitely.
We will write $(U_{i1}, U_{i2})$ to refer to the endpoints of interval $i$.
There are two cases:
\begin{enumerate}
\item {\bf The process terminates.}\\
Then there is a last interval $U_L = (U_{L1}, U_{L2})$. It is possible that there is one (but not more than
one) element $u^*$ of the original enumeration that is present in the interval $U_L$. If that is so, then
every element of $U_L \setminus \{u^*\}$ is a real number not in the original enumeration; otherwise every
element of $U_L$ is a real number not in the original enumeration.
\item {\bf The process does not terminate.}\\
Note that the sequence of interval lower bounds $(U_{i1})_{i\in\N}$ forms a strictly increasing sequence
bounded above by $u_2$ and that the sequence of interval upper bounds $(U_{i2})_{i\in\N}$ forms a strictly
decreasing sequence bounded below by $u_1$. By the Monotone Convergence theorem, both sequences converge:
let these limits be $\alpha$ and $\beta$ respectively. There are two cases:
\begin{enumerate}
\item {\bf $\alpha < \beta$}\\
Then every element of $(\alpha, \beta)$ is a real number not in the original enumeration.
\item {\bf $\alpha = \beta$}\\
Then $\alpha$ is a real number not in the original enumeration.
\end{enumerate}
\end{enumerate}
In all cases, we found a real number that was not present in the original enumeration. But this is a
contradiction, since the original enumeration contains all real numbers. Therefore no such enumeration exists
and the real numbers are not countable.
\end{proof}