-
Notifications
You must be signed in to change notification settings - Fork 0
/
main.tex
383 lines (283 loc) · 38.4 KB
/
main.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
\documentclass[12pt]{article}
\usepackage[margin=1in]{geometry}
\usepackage{mathtools}
\usepackage{amsthm}
\usepackage{amsfonts}
\usepackage{amssymb}
\usepackage{enumitem}
\include{MathMacrosJDH}
\theoremstyle{definition}
\newtheorem{innercustomthm}{Exercise}
\newenvironment{customthm}[1]
{\renewcommand\theinnercustomthm{#1}\innercustomthm}
{\endinnercustomthm}
\title{
\vspace{-2cm}
\author{Wojciech Aleksander Wołoszyn}
Solutions for
T. Jech's Set Theory \\
}
\begin{document}
\maketitle
\begin{customthm}{I.1.1}
Assume that $(a,b)=(c,d)$ or, in other words, $\set{\set{a},\set{a,b}} = \set{\set{c},\set{c,d}}$. By the axiom of extensionality, we have that $\set{a} \in \set{\set{c},\set{c,d}}$ and $\set{a,b} \in \set{\set{c},\set{c,d}}$. By the axiom of pairing, we have that $\set{a} = \set{c}$ or $\set{a} = \set{c,d}$ and $\set{a,b} = \set{c}$ or $\set{a,b} = \set{c,d}$. Suppose, $\set{a} = \set{c,d}$, then, by pairing and extensionality, $a=c=d$. So $\set{\set{a},\set{a,b}} = \set{\set{c},\set{c,d}}$ becomes $\set{\set{a},\set{a,b}} = \set{\set{a}}$, and, again by pairing and extensionality, $a=b$. And we have that $a=b=c=d$. Suppose $\set{a} = \set{c}$ and $\set{a} \neq \set{c,d}$. Then, by extensionality and pairing, $a=c \neq d$. If $\set{a,b} = \set{c}$, then $a=b=c$, which would imply $a=b=c=d$, contradiction. So $\set{a,b} = \set{c,d}$ and $a \neq b$. By the axioms of pairing and extensionality, we have that $a=c$ and $b=d$. Other direction is immediate.
\end{customthm}
\begin{customthm}{I.1.2}
Suppose we have that $P(X) \subset X$. Then, there must be a surjection $f \colon X \to P(X)$. But the set $\set{x \colon x \not\in f(x)}$ is not in $P(X)$.
\end{customthm}
\begin{customthm}{I.1.3}\label{ex:set-of-elements-of-inductive-set-is-inductive}
Let $Y=\set{x \in X \colon x \subset X}$. We show that $Y$ is inductive. First, it is obvious that $\emptyset \in Y$. Let $y \in Y$. Since $Y \subset X$, it follows that $y \in X$. But $X$ is inductive, so $y \cup \set{y} \in X$. Furthermore, $\set{y} \subset X$. By the definition of $Y$, it is also the case that $y \subset X$. So $y \cup \set{y} \subset X$. Thus, $y \cup \set{y} \in Y$.
Let $N$ be the smallest inductive set $N = \bigcap \set{X \colon X \text{ is inductive}}$. Set $Y = \set{x \in N \colon x \subset N}$. By the definition of $Y$, $Y \subset N$. But $N$ is minimal, so we have that $N \subset Y$. Hence, $Y=N$. And $N$ is transitive, because $y \in N$, implies $y \subset N$.
We show by induction that $n = \set{m \in N \colon m < n}$. Since $N$ is inductive, we have that $\emptyset \in N$. Assume that $n = \set{m \in N \colon m < n}$. Now, $n+1 = n \cup \set{n} = \set{m \in N \colon m < n} \cup \set{n} = \set{m \in N \colon m < n \text{ or } m=n} = \set{m \in N \colon m < n + 1}$.
\end{customthm}
\begin{customthm}{I.1.4}\label{ex:set-of-transitive-elements-of-inductive-set-is-inductive}
Let $Y = \set{x \in X \colon x \text{ is transitive}}$. First, since $X$ is inductive, $\emptyset \in X$. The empty set is transitive, so $\emptyset \in Y$. Let $y \in Y$. We know that $Y \subset X$, so $y \in X$. And we know that $y$ is transitive, so $y \subset X$. Hence $y \cup \set{y} \in X$. It remains to show that $y \cup \set{y}$ is transitive. Let $\gamma \in y \cup \set{y}$. Then $\gamma$ is either equal to $y$ or an element of $y$. If $\gamma = y$, then, in particular, $\gamma \subset y$. If $\gamma \in y$, then $\gamma \subset y$, because $y$ is transitive. And so, $\gamma \subset y \cup \set{y}$.
Set $Y = \set{x \in N \colon x \text{ is transitive}}$. By the definition of $Y$, $Y \subset N$. But since $Y$ is inductive and $N$ is a minimal inductive set, we also have that $N \subset Y$. Hence, $N=Y$ and every element $n$ of $N$ is transitive.
\end{customthm}
\begin{customthm}{I.1.5}\label{ex:set-of-transitive-elements-with-x-not-in-x-of-inductive-set-is-inductive}
Let $Y = \set{x \in X \colon x \text{ is transitive and } x \not\in x}$. First, since $X$ is inductive, $\emptyset \in X$. The empty set is transitive and has no element (particularly $\emptyset \not\in \emptyset)$), so $\emptyset \in Y$. Let $y \in Y$. From exercise~\ref{ex:set-of-transitive-elements-of-inductive-set-is-inductive}, the set $y \cup \set{y}$ is transitive. It is enough to show that $y \cup \set{y} \not\in y \cup \set{y}$. If the opposite were the case then either $y \cup \set{y}$ would be equal to $y$ or an element of $y$. If $y \cup \set{y} = y$, then $\set{y} \subset y$, hence $y \in y$, which is a contradiction. If $y \cup \set{y} \in y$, then, again $y \in y$.
Set $Y = \set{x \in N \colon x \text{ is transitive and } x \not\in x}$. By the definition of $Y$, $Y \subset N$. But since $Y$ is inductive and $N$ is a minimal inductive set, we also have that $N \subset Y$. Hence, $N=Y$ and every element $n$ of $N$ is such that $n \not\in n$. In other words, $n \neq n+1$ for each $n \in N$.
\end{customthm}
\begin{customthm}{I.1.6}
Let $Y = \set{x \in X \colon x \text{ is transitive and every nonempty } z \subset x \text{ has an $\in$-minimal element}}$. First, since $X$ is inductive, $\emptyset \in X$. The empty set is transitive and has no nonempty elements, so it is in $Y$. From exercise~\ref{ex:set-of-transitive-elements-of-inductive-set-is-inductive}, the set $y \cup \set{y} \in X$ is transitive. Let $z \subset y \cup \set{y}$ be nonempty. If $t \in z$, then $t \in y \cup \set{y}$. So either $t \in y$ or $t=y$, so $t \subset y$. So $t$ is either an $\in$-minimal or has $\in$-minimal element. But since $t \in z$, $z$ has $\in$-minimal element. And so, $y \cup \set{y} \in Y$.
\end{customthm}
\begin{customthm}{I.1.7}\label{ex:every-nonempty-subset-of-N-is-inductive}
Let $X$ be a nonempty subset of $N$ and let $n \in X$. If $n \cap X$ is nonempty, then there is a minimal element $m \in n \cap X \subset n$, because $n$ has a minimal element. If $n \cap X$ is empty, then $n$ is the minimal element. If it weren't then there would be an element $m \in X$ such that $m \in n$. But $n \cap X$ is empty, so this is impossible.
\end{customthm}
\begin{customthm}{I.1.8}
Let $Y = \set{x \in X \colon x = \emptyset \text{ or } x = y \cup \set{y} \text{ for some y}}$. First, the empty set is in $Y$. Let $\gamma \in Y$ be nonempty. There is some $y$ such that $\gamma = y \cup \set{y}$. Note that $\gamma \cup \set{\gamma}$ is also of this form and $\gamma \in Y$, so $\gamma \cup \set{\gamma} \in Y$.
\end{customthm}
\begin{customthm}{I.1.9}
$A$ is a subset of $N$, so $A \subset N$. Since $\emptyset \in A$ and $n \in A \implies n+1 = n \cup \set{n} \in A$, $A$ is inductive. By definition, $N$ is the minimal inductive set so $N \subset A$. Hence, $A=N$.
\end{customthm}
\begin{customthm}{I.1.10}
We proceed by induction. Element $0 = \emptyset \in N$ has no nonempty elements, so it's trivially $T$-finite. Suppose $n$ is $T$-finite. And let $X \subset P(n+1)$ be nonempty. Note that $X$ is either equal to $Y$ or to $Y \cup \set{n}$, for some subset $Y$ of $P(n)$. By the induction hypothesis, $Y$ has a $\subset$-maximal element $y$. If $X = Y$, $y$ is the maximal element of $X$. If $X = Y \cup \set{n}$, $y \cup \set{n}$ is the maximal element of $X$.
\end{customthm}
\begin{customthm}{I.1.11}
It is enough to observe that $n \in N$ is a proper subset of $n+1 \in N$.
\end{customthm}
\begin{customthm}{I.1.12}
Let $S$ be a finite set. And take any nonempty $X \subset P(S)$. We need to show that there is a $\subset$-maximal element in $X$. By the definition of finiteness, there is a bijection $f: S \to n$ for some $n \in N$. So there is an $x \in X$ such that $f(x)$ is a $\subset$-maximal element of $f(X)$. We claim that $x$ must be a $\subset$-maximal element in $X$. Otherwise there would be some $y \in X$ such that $x \subset y$, but $f(x) \not\subset f(y)$, which is absurd.
\end{customthm}
\begin{customthm}{I.1.13}
Let $S$ be infinite and set $X = \set{u \subset S \colon u \text{ is finite}}$. Note that $\emptyset \in X$, so $X$ is nonempty. Suppose $X$ has a $\subset$-maximal element $x$. Take $s \in S$ such that $s \not\in x$ (there must be such element as $S$ is infinite and $x$ is finite). But then $x \cup \set{s}$ is also finite, hence is in $X$. This is a contradiction with $x$ being the $\subset$-maximal element in $X$.
\end{customthm}
\begin{customthm}{I.1.14}
Let $\varphi(x)$ be a formula and let $X$ be arbitrary. Take $F = \set{(x,x) \colon \varphi(x)}$. By the Replacement Schema, $F(X) = \set{x \in X \colon \varphi(x)}$ for every $X$. Hence, the Separation Axioms follow from Replacement Schema.
\end{customthm}
\begin{customthm}{I.1.15}
\leavevmode
\begin{enumerate}
\item We have that $\forall X \exists Y (\forall x \in X) (\forall u \in X) u \in Y$. Set an $X$ and take $\varphi(u) = (\exists z \in X) u \in z$. By the Separation Schema and the above weaker version of the Union, there is a set $Y' = \set{u \in Y \colon \varphi(u)} = \set{u \in Y \colon (\exists z \in X) u \in z}$. But then, $u \in Y' \iff \exists z (z \in X \land u \in z)$. Hence $Y' = \bigcup X$.
\item We have that $\forall X \exists Y \forall u (u \subset X \implies u \in Y)$. Set an $X$ and take $\varphi(u) = u \subset X$. By the Separation Schema and the above weaker version of the Power Set, there is a set $Y' = \set{u \in Y \colon \varphi(u)} = \set{u \in Y \colon u \subset X}$. But then, $u \in Y' \iff u \subset X$. Hence $Y' = P(X)$.
\item Let $F$ be a class function. We have that $\forall X \exists Y F(X) \subset Y$. Set an $X$ and take $varphi(u) = \exists x \in X F(x,u)$. By the Separation Schema and the above weaker version of the Replacement, there is a set $Y' = \set{u \in Y \colon \varphi(u)} = \set{u \in Y \colon \exists x \in X F(x,u)} = F(X)$.
\end{enumerate}
\end{customthm}
\begin{customthm}{I.2.1}
Let $P,Q,R$ be partial orderings. Order $P$ is isomorphic to itself via the identity function. If $P$ is isomorphic to $Q$ via $f$, then $Q$ is isomorphic to $P$ via $f^{-1}$. Finally, if $P$ is isomorphic to $Q$ via $f$ and $Q$ is isomorphic to $R$ via $g$, then $P$ is isomorphic to $R$ via $g \circ f$.
\end{customthm}
\begin{customthm}{I.2.2}
Let $\alpha$ be a limit ordinal. Suppose there is a $\beta < \alpha$ such that $\beta+1 \geq \alpha$. Hence, $\beta + 1 = \alpha$. But that would mean that $\alpha$ is a successor cardinal, contradiction.
Let $\alpha$ be such that for every $\beta < \alpha$, $\beta + 1 < \alpha$. Then $\alpha$ must be a limit cardinal. If it weren't then there would be some $\beta_0 < \alpha$ such that $\alpha = \beta_0 + 1$, contradiction.
\end{customthm}
\begin{customthm}{I.2.3}
Let $X$ be inductive. First, $\emptyset \in X \cap \Ord$ -- for $\Ord$ by definition. for $X$ because it is inductive. Let $x \in X \cap \Ord$, then $x \cup \set{x} \in X \cap \Ord$ -- it is a basic fact that for $\Ord$ and it follows from inductivity for $X$. Hence, $X \cap \Ord$ is inductive.
By exercise~\ref{ex:set-of-elements-of-inductive-set-is-inductive}, $N$ -- the smallest inductive set -- is transitive. It is easy to see that $N$ is linearly ordered. And what is more, by exercise~\ref{ex:every-nonempty-subset-of-N-is-inductive}, every nonempty subset of $N$ has an $\in$-minimal element. Hence, $N$ is an ordinal number. Also, note that for any $n<N$, $n+1<N$, so it is a limit ordinal. Furthermore, it is the smallest limit ordinal other than the empty set: every $n \in N$ is such that $(n-1) + 1 \not\in n$.
\end{customthm}
\begin{customthm}{I.2.4}
The first implication (i$\to$ii) follows, because $N$ is inductive and infinite.
For the second (ii$\to$iii), assume that $X$ is infinite and consider the set $Y = \set{x \subset X \colon x \text{ is finite}}$. For each $x \in Y$, let $f(x)$ be a number of elements in $x$. Note that $f \colon Y \to \omega$ is a function. Let us prove this function is a surjection. Suppose $n \in Y$. There must be at least one $x \in f^{-1}(n)$. Otherwise, $X$ wouldn't be infinite. Hence, $f(Y) = \omega$ is a set.
The last implication (iii$\to$i), observe that $\omega$ is an inductive set. Indeed, $\emptyset \in \omega$ and whenever $\alpha \in \omega$, then $\alpha+1 \in \omega$.
\end{customthm}
\begin{customthm}{I.2.5}
A nonempty subset $\set{a_0,a_1,\ldots}$ of $W$, constructed from such a sequence, would not have a $\in$-minimal element.
\end{customthm}
\begin{customthm}{I.2.6}
Let $\alpha$ be an ordinal. Take $\alpha_0 = \alpha$ and $\alpha_{n+1} = \alpha_n + 1$. Define the ordinal $\beta = \bigcup \alpha_n = \sup \set{\alpha_n} = \lim \alpha_n$. We shall prove that $\beta$ is a limit ordinal and since $\alpha$ was chosen arbitrarily, there are arbitrarily large limit ordinals. It suffices to show that whenever $\gamma \in \beta$, then $\gamma + 1 \in \beta$. Let $\gamma \in \beta$. There is an $\alpha_n$ such that $\gamma < \alpha_n$ (otherwise $\gamma \geq \beta$). But then $\gamma + 1 < \alpha_n + 1 = \alpha_n < \beta$. This ends the proof.
\end{customthm}
\begin{customthm}{I.2.7}
\end{customthm}
\begin{customthm}{I.2.8}
By induction: \\ i) $\gamma = 0$, easy; successor: $\alpha \cdot (\beta + \gamma + 1) = \alpha \cdot (\beta + \gamma) + \alpha \cdot 1 = \alpha \cdot \beta + \alpha \cdot \gamma + \alpha \cdot 1 = \alpha \cdot \beta + \alpha (\gamma + 1)$; limit: $\alpha \cdot (\beta + \gamma) = \alpha \cdot (\beta + \lim \gamma_n) = \lim \alpha \cdot (\beta + \gamma_n) = \lim \alpha \cdot \beta + \alpha \cdot \gamma_n = \alpha \cdot \beta + \alpha \cdot (\lim \gamma_n) = \alpha \cdot \beta + \alpha \cdot \gamma$. \\
ii) $\gamma = 0$, easy; successor: $\alpha^{\beta + \gamma + 1} = \alpha^{\beta + \gamma} \cdot \alpha^1 = \alpha^\beta \cdot \alpha^\gamma \cdot \alpha^1 = \alpha^\beta \cdot \alpha^{\gamma + 1}$; limit: $\alpha^{\beta + \gamma} = \alpha^{\beta + \lim \gamma_n} = \lim \alpha^{\beta + \gamma_n} = \lim \alpha^\beta \cdot \alpha^{\gamma_n} = \alpha^\beta \cdot \alpha^{\lim \gamma_n} = \alpha^\beta \cdot \alpha^{\gamma}$. \\
iii) $\gamma = 0$, easy; successor: $(\alpha^\beta)^{\gamma+1} = (\alpha^\beta)^\gamma \cdot (\alpha^\beta)^1 = \alpha^{\beta \cdot \gamma} + (\alpha^\beta \cdot 1) = \alpha^{\beta \cdot (\gamma +1)}$; limit: $(\alpha^\beta)^{\gamma} = (\alpha^\beta)^{\lim \gamma_n} = \lim (\alpha^\beta)^{\gamma_n} = \lim \alpha^{\beta \cdot \gamma_n} = \alpha^{\beta \cdot (\lim \gamma_n)} = \alpha^{\beta \cdot\gamma}$.
\end{customthm}
\begin{customthm}{I.2.9} \leavevmode \\
i) $(\omega+1) \cdot 2 = \omega + 1 + \omega + 1 = \omega + \omega + 1 = \omega \cdot 2 + 1$ \\
ii) $(\omega \cdot 2)^2 = (\omega \cdot 2) \cdot (\omega \cdot 2) = (\omega \cdot \omega \cdot 2) = \omega^2 \cdot 2$
\end{customthm}
\begin{customthm}{I.2.10} \leavevmode \\
i) Let $\alpha<\beta$. It is hard not to see that $\alpha + 0 < \beta + 0$. Suppose $\alpha + \gamma < \beta + \gamma$. Observe that $(\alpha + \gamma) + 1 < (\beta + \gamma) +1$ and, if $\gamma$ is a limit ordinal, suppose the hypothesis holds for $\gamma' < \gamma$, $\alpha + \gamma = \lim \alpha + \gamma_n \leq \lim \beta + \gamma_n = \beta + \gamma$. \\
ii) Likewise, $\alpha \cdot 0 \leq \beta \cdot 0$ and $\alpha \cdot 1 \leq \beta \cdot 1$ Suppose $\alpha \cdot \gamma < \beta \cdot \gamma$. Observe $\alpha \cdot (\gamma + 1) = \alpha \cdot \gamma + \alpha \cdot 1 < \beta \cdot \gamma + \beta = \beta (\gamma + 1)$. For limits, $\alpha \cdot \gamma = \lim \alpha \cdot \gamma_n \leq \lim \beta \cdot \gamma_n = \beta \cdot \gamma$ \\
iii) Finally, $\alpha^0 < \beta^0$. Suppose $\alpha^\gamma < \beta^\gamma$. Observe $\alpha^{\gamma+1} = \alpha^\gamma \cdot \alpha \leq \beta^\gamma \cdot \beta = \beta^{\gamma+1}$. For limits, $\alpha^\gamma = \lim \alpha^{\gamma_n} \leq \lim \beta^{\gamma_n} = \beta^\gamma$.
\end{customthm}
\begin{customthm}{I.2.11}
\end{customthm}
\begin{customthm}{I.2.12}
Let $\alpha_0 = \omega$, $\alpha_{n+1} = \omega^{\alpha_n}$ and $\epsilon_0 = \lim \alpha_n$. Observe that $\omega^{\epsilon_0} = \omega^{\lim \alpha_n} = \lim \omega^{\alpha_n} = \lim \alpha_{n+1} = \epsilon_0$.
Note that the above sequence $\alpha_n$ is increasing. Suppose there is $\epsilon < \epsilon_0$ such that $\omega^\epsilon = \epsilon$. It is obvious that such $\epsilon$ must be infinite. Take the smallest $n \in \omega$ such that $\alpha_{n+1} > \epsilon$. Observe that $\omega^{\alpha_n} = \alpha_{n+1} > \epsilon = \omega^\epsilon$. But then, $\alpha_n > \epsilon$, contradiction.
\end{customthm}
\begin{customthm}{I.2.13}
\end{customthm}
\begin{customthm}{I.2.14}
If that were true, then $X = \set{a_n \colon n \in N } \subset P$ would be a nonempty set with no $E$-minimal element.
\end{customthm}
\begin{customthm}{I.2.15}
\end{customthm}
\begin{customthm}{I.2.16}
\end{customthm}
\begin{customthm}{I.3.1}
i) Suppose $X$ is finite and $Y \subset X$ is infinite. Thus, $Y$ is $T$-infinite, i.e., there is a nonempty $Z \subset P(Y)$ with no $\subset$-maximal element. But $P(Y) \subset P(X)$ and $X$ is finite (hence $T$-finite), so $Z$ must have a maximal element---contradiction. \\
ii) Let $\bigcup X_i$ be a finite union of finite sets. With each $X_i$, there is an associated bijection $f_i \colon X_i \to n_i$ for some natural number $n_i$. We define a function $f \colon \bigcup X_i \to \sum n_i$ by $f(x) = f_k(x) + \sum_{i<k} n_i$, where $k$ is minimal natural number such that $x \in X_k$. Clearly, $f$ is one-to-one. Note that $f$ is not necessarily a bijection ($X_i$ may overlap). \\
iii) Let $X$ be finite and have $n$ elements. Then $P(X)$ equinumerous with $2^X$, so has $2^n$ elements. \\
iv) Let $X$ be finite and let $f \colon X \to f(X)$ be a mapping. Since $X$ is finite, there is a bijection $g \colon X \to n$ for some natural number $n$. Consider an instance of $f^{-1}$. Then $h(x) = g \circ f^{-1}$ is in one to one correspondence with $n$ (though, again, not necessarily bijective correspondence, unless $f$ was one-to-one itself).
\end{customthm}
\begin{customthm}{I.3.2}
i) Let $X$ be countable and let $Y \subset X$. Since $X$ is countable, there is a bijection $f \colon X \to \omega$. But note that $f$ composed with the identity function $\mathrm{id} \colon Y \to X$ is in one-to-one correspondence with $\omega$. \\
ii) Let $\bigcup X_i$ be a finite union of countable sets. With each $X_i$, there is an associated bijection $f_i \colon X_i \to \omega \times \set{i}$. Take $f(x) \colon \bigcup X \to \omega \times \omega$ to be $f(x) = (f_i(x),i)$ with $i$ minimal. The desired function is $g \colon \bigcup X \to \omega$ defined by $g(x) = \Gamma \circ f(x)$, where $\Gamma$ is a bijection between $\omega \times \omega$ and $\omega$. Indeed, this proves that the finite union of countable sets is at least countable. And from i) we know that is is also at most countable. \\
iii) Let $X$ be countable and let $f \colon X \to f(X)$ be a mapping. Since $X$ is countable, there is a bijection $g \colon X \to \omega$. Consider an instance of $f^{-1}$. Then $h(x) = g \circ f^{-1}$ is in one to one correspondence with $\omega$.
\end{customthm}
\begin{customthm}{I.3.3}
It is enough to prove that $f \colon N \times N \to \omega$, $f(m,n) = 2^m(2n+1)-1$ is one-to-one. Note that $2^a(2b+1)-1 = 2^c(2d+1) - 1$ implies $2^a(2b+1) = 2^c(2d+1)$ implies $a=c, b=d$.
\end{customthm}
\begin{customthm}{I.3.4}
i) There are $\omega^n$ sequences of length $n$ in $N$. We use the axiom of choice to say that for each $n$ there is an injection $f_n \colon \text{sequences of length }n \to N$. But then, the function $f(x) = (f_n(x),n)$, where $n$ is the length of finite sequence $x$ is an injection. So the cardinality of all finite length sequences is at most countable. It is also at least countable because the number of all sequences of size $1$ is already countable. \\
ii)
\end{customthm}
\begin{customthm}{I.3.5}
We prove it by induction on $\alpha$. First, $\Gamma(0 \times 0) = 0 \leq \omega^0 = 1$. Now, $\alpha \times \alpha$ is isomorphic to $\alpha \cdot \alpha$, so $\Gamma((\alpha+1) \times (\alpha+1)) = \text{order type of } (\alpha+1) \cdot (\alpha+1) = \text{order type of } \alpha^2 + \alpha \cdot 2 + 1 = \Gamma(\alpha \times \alpha) + \alpha \cdot 2 + 1 \leq \omega^\alpha + \alpha \cdot 2 + 1 \leq \omega^{\alpha+1}$.
\end{customthm}
\begin{customthm}{I.3.7}
Let $B$ be a projection of $\omega_\alpha$, i.e., there is a surjection $f \colon \omega_\alpha \to B$. It suffices to show that there is an injection $g \colon B \to \omega_\alpha$. Let $g(x) = s(f^{-1}(x))$ where $s$ is a selector function.
\end{customthm}
\begin{customthm}{I.3.8}
We know that the set of all finite sequences in $\omega^\alpha$ is of order-type $\omega_\alpha$ and hence its cardinality is $\aleph_\alpha$. It suffices to show that there is a surjection $f: [\omega_\alpha]^{<\omega} \to \omega_\alpha^{<\omega}$. Then, by the axiom of choice we can construct an injection $g(x) = s(f^{-1}(x))$. It is enough to take $f$ to map any set in $[\omega_\alpha]^{<\omega}$ to an increasing sequence of its elements in $\omega_\alpha^{<\omega}$.
\end{customthm}
\begin{customthm}{I.3.9}
Let $B$ be a projection of $A$. That is, let there be a surjection $f \colon A \to B$. Define $g \colon P(B) \to P(A)$ as $g(X) = f^{-1}(X)$. We need to show that $g$ is one-to-one. Suppose $g(X) = g(Y)$, then $f^{-1}(X) = f^{-1}(Y)$ and $X=Y$, done.
\end{customthm}
\begin{customthm}{I.3.10}
\end{customthm}
\begin{customthm}{I.3.11}
Since $\omega_{\alpha+1}$ is a projection of $P(\alpha)$, we know that $\aleph_{\alpha+1} \leq 2^{\aleph_\alpha}$. But by the Cantor's theorem, $2^{\aleph_\alpha} < 2^{2^{\aleph_\alpha}}$. So $\aleph_{\alpha+1} < 2^{2^{\aleph_\alpha}}$.
\end{customthm}
\begin{customthm}{I.3.12}
Note that $\omega_\zeta$, $\zeta < \cof \alpha$, is a non-decreasing $\cof \alpha$ sequence of ordinals in $\omega_\alpha$ and $\lim \omega_\zeta = \omega_\alpha$, so $\cof \cof \alpha = \cof \alpha = \cof \omega_\alpha$.
\end{customthm}
\begin{customthm}{I.3.13}
We work in $\ZF$. Suppose $\omega_2 = \bigcup_{n < \omega} S_n$ and each $S_n$ is countable and let $\alpha_n$ be the order-type of $S_n$. Then, for each $n$, there is a unique isomorphism $f_n \colon \alpha_n \to S_n$. Let us denote $\sup \alpha_n$ by $\alpha$. We define a function $f \colon \omega \times \alpha \to \omega_2$ by:
\begin{equation*}
f(n, \kappa) = \begin{cases} f_n(\kappa) & \mbox{if } \kappa \in \alpha_n \\ 0 & \mbox{otherwise.} \end{cases}
\end{equation*}
Let $\gamma \in \omega_2$, then there is a minimal $n$ with $\gamma \in S_n$ and $f_n^{-1}(\gamma) \in \alpha_n$, so there is a unique inverse (the one taking $n$ to be minimal, this uses the fact that $\omega \times \omega_2$ is well-ordered under $\ZF$) $f^-1(n,\gamma) = (n,f_n^{-1}(\gamma)$. Thus, $f$ is a surjection and $\aleph_2 \leq |\omega \times \alpha| \leq \aleph_0 \cdot \aleph_1 = \aleph_1$, contradiction.
\end{customthm}
\begin{customthm}{I.3.14}
Let $S$ be $D$--infinite. So there is a one-to-one mapping $f \colon S \to X$, where $X$ is a subset of $S$. Choose $x_0 \in S \setminus X$ and define $x_{n+1} = f(x_n) \in X$. But then, the set $\set{x_n \mid n < \omega}$ is a countable subset of $S$.
For the other direction, let $X \subset S$ be countable, say, $X=\set{x_n \mid n < \omega}$. Then, define $f \colon S \to X$ by:
\begin{equation*}
f(s) = \begin{cases} x_{n+1} & \mbox{if } s = x_n \\ s & \mbox{otherwise.} \end{cases}
\end{equation*}
\end{customthm}
\begin{customthm}{I.3.15}
i) Let $A,B$ be $D$--finite sets, then $A$ and $B$ have no countable subsets. Suppose the set $A \cup B$ is $D$--infinite, and so has a countable subset $X$. But then, $X = X \cap (A \cup B) = (X \cap A) \cup (X \cap B)$. Sets $(X \cap A)$ and $(X \cap B)$ cannot be countable, hence are finite. Sum of two finite sets is finite, so $X$ is finite, hence $D$--finite.
Suppose $X \subset A \times B$ is $D$--infinite and $A,B$ are $D$--finite. Then, $|X| \leq |\set{a \mid (a,b) \in A}| \times |\set{b \mid (a,b) \in B}|$. But $|\set{a \mid (a,b) \in A}|$ and $|\set{b \mid (a,b) \in B}|$, as subsets of $D$--finite sets, cannot be countable and are finite. Hence, $|X|$ is also finite, contradiction.
\\
ii)
iii)
\end{customthm}
\begin{customthm}{I.3.16}
\end{customthm}
\begin{customthm}{I.4.1}
For the cardinality, of the set of real continuous functions, it is enough to observe that every such a function is determined by its values at rational points, i.e., let $p_n/q_n \to x$, then $f(x) = \lim f(p_n/q_n)$. So the cardinality is $\continuum^\omega = \continuum$.
For all the real functions, it is obviously $\continuum^\continuum$ (every real point can be mapper to any real point).
\end{customthm}
\begin{customthm}{I.4.2}
It is enough to prove that for two sequences $x,y \in \omega^\omega$ that are chosen arbitrarily and are distinct, the linear orderings $x_0 + \Z + x_1 + \Z + \ldots$ and $y_0 + \Z + y_1 + \Z + \ldots$ are different.
Let us first look at $x_0 + \Z$ and $y_0 + \Z$. If $x_0 \neq y_0$, say $x_0>y_0$, then these orders cannot be isomorphic---there is nowhere we can send $x_0$ to without breaking the ordering. And the same argument holds for $x_0 + \Z + \ldots + x_k + \Z$ and $y_0 + \Z + \ldots + y_k + \Z$ for any $k$. Finally, for $x_0 + \Z + x_1 + \Z + \ldots$ and $y_0 + \Z + y_1 + \Z + \ldots$, we look at the minimal $k$ such that $x_k$ and $y_k$ are different and repeat the argument.
\end{customthm}
\begin{customthm}{I.4.3}
Any polynomial has at most finitely many roots and there are countably many polynomials with integer coefficients. So there are at most $\omega^{<\omega} = \aleph_0$ algebraic reals. And of course linear polynomials with integer coefficient have different algebraic roots and form a countable set, so there are exactly countably many algebraic reals.
\end{customthm}
\begin{customthm}{I.4.4}
It suffices to prove it for $\R \times \R$, because it has the same cardinality as $\R$. Let $S$ be a countable subset of $\R \times \R$. Now, for each real $r$, the line $(\set{r} \times \R) \setminus S$ has at least one point (otherwise $S$ would be the whole line and hence of cardinality continuum). So the whole set $(\set{r} \times \R) \setminus S$ must have cardinality $1 \cdot |\R| = \continuum$.
\end{customthm}
\begin{customthm}{I.4.5}
i) The set of irrational numbers is equal to $\R \setminus \Q$.
\\
ii) The set of transcendental numbers is equal to $\R\setminus \mathbb{A}$
\\
By above and previous exercise, both sets are of cardinality continuum.
\end{customthm}
\begin{customthm}{I.4.6}
There are at least $\continuum$ many open sets of reals as each interval $(x, \infty)$ is distinct, open, and there are $\continuum$ many of them. On the other hand, each open set is a countable sum of rational intervals and there are at most $\omega^\omega = \continuum$ possible combinations of those.
\end{customthm}
\begin{customthm}{I.4.7}
Cantor set is non-empty, because it is in one-to-one correspondence with $0$-$1$ $\omega$--sequences and hence has cardinality $\continuum$. It is closed, because it is an intersection of closed sets. What is left is to prove that the Cantor set has no isolated points. Let $x$ be an arbitrary point in the Cantor set and let us pick $\epsilon > 0$. It suffices to find $y \neq x$ such that $|x-y|<\epsilon$. Recall that the Cantor set $C$ is an intersection of $2^n$ intervals $C_n$ each of length $3^{-n}$. Pick $n$ large enough so that $3^{-n}<\epsilon$. Let $[a,b]$ be one of those intervals that comprise $C_n$. Since $x \in C$, $x \in C_{n+1}$ and there is a subinterval $[a_0,b_0] \subset [a,b]$. But, by the construction of $C_{n+1}$, there is another interval $[a_1,b_1] \subset [a,b]$, which has an empty intersection with $[a_0,b_0]$. It is clear from the construction of the Cantor set that it has a nonempty intersection with the interval $[a_0,b_0]$. Hence, there is a $y \neq x$ in the Cantor set such that $|x-y| \leq 3^{-n} < \epsilon$.
\end{customthm}
\begin{customthm}{I.4.8}\label{ex:perfect-set-intersected-with-interval-cardinality-continuum}
Let $P$ be perfect and $P \cap (a,b) \neq \emptyset$. Pick an $x \in P \cap (a,b)$. Since, $P$ is perfect, there must be a sequence of elements not equal to $x$ both in $(a,x)$ and in $(x,b)$. So $P \cap (a,x) \neq \emptyset \neq P \cap (x,b)$. Now, pick $x_0 \in (a,x)$ and $x_1 \in (b,x)$. This process can be repeated and after $\omega$ many steps with have picked $\continuum$ many distinct points in $P \cap (a,b)$.
\end{customthm}
\begin{customthm}{I.4.9}\label{ex:P_1-and-P_2-perfect-and-not-subset-of-each-other-then-their-difference-is-of-cardinality-continuum}
Let $P_2 \not \subset P_1$ be perfect sets. There is a point $x \in P_2 \setminus P_1$. There must be a sufficiently small interval $(a,b)$ containing $x$ such that $(a,b) \cap P_1 = \emptyset$ (if it weren't the case we could construct a sequence of $P_1$ elements approaching $x$ and then $x \in P_1$). By the exercise above, this ends the proof.
\end{customthm}
\begin{customthm}{I.4.10}
Let $P$ be perfect. Take any $p \in P$ and any neighborhood $(a,b)$ of $p$. Then, by exercise~\ref{ex:perfect-set-intersected-with-interval-cardinality-continuum} this neighborhood contains uncountably many elements of $P$.
\end{customthm}
\begin{customthm}{I.4.11}
Let $P$ be perfect, $P \subset F$, where $F$ is closed and let $F^*$ be the set of all condensation points of $F$. By exercise above, the set of condensation points $P^*$ of $P$ is equal to $P$. But then, it must be the case that $P^* \subset F^*$, and so $P \subset F^*$.
\end{customthm}
\begin{customthm}{I.4.12}
Let $F$ be an uncountable closed set and let $P$ be the perfect set constructed in the proof of Cantor-Bendixson theorem. Take any $a \in F^*$. Each neighborhood of $a$ must have an element from $P$, because $F \setminus P$ is countable and $a$ is a condensation point. But that means $a$ is a limit of elements of $P$, hence $a \in P$. So $F^* \subset P$. This, combined with the exercise above, gives us that $F^*=P$.
\end{customthm}
\begin{customthm}{I.4.13}
By the exercise above and Cantor-Bendixson theorem, we have that $F = F^* \cup (F \setminus F^*)$, where $F^*$ is a perfect set and $F \setminus F^*$ is at most countable. Suppose there is another set $P$ such that $F = P \cup (F \setminus P)$ such that $P$ is perfect and $F \setminus P$ is countable.
Suppose $P \subset F^*$, then $F = P \cup (F \setminus F^*)$, so $P = F^*$. We proceed similarly if $F^* \subset P$. Suppose $P \not\subset F^*$. But then $\emptyset = F^* \setminus P \cup ((F \setminus F^*) \setminus (F \setminus P)) \supset F^* \setminus P$ and, by exercise~\ref{ex:P_1-and-P_2-perfect-and-not-subset-of-each-other-then-their-difference-is-of-cardinality-continuum}, $F^* \setminus P$ has cardinality continuum. This is a contradiction!
\end{customthm}
\begin{customthm}{I.4.14}
Suppose $\Q$ is the intersection of countable collection of open sets $U_n$. Since $\Q$ is dense, $U_n$ must also be dense. Let us enumerate $\Q$ by $q_n$. Consider sets $V_n = \R \setminus \set{q_n}$. Then the family of sets consisting of all $U_n$ and $V_n$ is a collection of open dense sets. By the Baire Category Theorem, the intersection of these sets is dense. But $\cap V_n = \emptyset$, so it is definitely not---contradiction.
\end{customthm}
\begin{customthm}{I.4.15}
Suppose $B$ is a Borel set and $f \colon X \to Y$ is a continuous function. We need to prove that $f^{-1}(B)$ is also Borel. It is enough to prove that $\set{f^{-1}(B) \mid B \text{ Borel}}$ forms a $\sigma$-algebra and contains all open sets:
\\
i) it contains all open sets in $f^{-1}(Y)$, because $f$ is continuous.
\\
ii) set $\set{f^{-1}(B) \mid B \text{ Borel}}$ forms a $\sigma$-algebra, because $f^(Y)$ is its element, $f^{-1}(A \cup B) = f^{-1}(A) \cup f^{-1}(B)$, and $f^{-1}(Y/A) = X/f^{-1}(A)$.
\end{customthm}
\begin{customthm}{I.4.16}
Set is $G_\delta$ if it is a countable intersection of open sets. An open set is a set such that given $x$ in that set and $\delta > 0$, whenever $|x-y| < \delta$, $y$ is an element of that set. Note that $f$ is continuous at $x$ just in case for all $n$, there is a $\delta_n$ such that for $x_1,x_2 \in B(x,\delta_n)$, we have that $|f(x_1)-f(x_2)|<1/n$. We need to prove that, for each $n$, the set
\begin{equation*}
C_n = \set{x \mid \exists \delta_n(x) \forall x_1,x_2 \in B(x, \delta) \implies |f(x_1) - f(x_2)| < 1/n}
\end{equation*}
is open. Take $x \in C_n$ and $\epsilon<\delta_n(x)/2$ (small enough to suffice, I think). Then any $y$ such that $|x-y|<\delta_n(x)/2$ is such that $y \in C_n$ (take $\delta_n(y)<\delta_n(x)/2$ or whatever). Hence, $C_n$ is open. To end the proof, note that $C = \cap C_n$ is the set of points of continuity of $f$.
\end{customthm}
\begin{customthm}{I.4.17}
i) Define a map that takes $(x,y) \mapsto x_0, y_0, x_1, y_1, \ldots$, $x,y \in \mathcal{N}$. We prove that this map, say $F$, is a homeomorphism. It is clear that it is bijective. Let us now prove that $F$ is continuous. Take any finite sequence $s$ and look at the basis $O(s) = \set{f \in \mathcal{N} \mid s \subset f}$. This set will be taken by $F^{-1}$ to a set of the form $O(s^0) \times O(s^1)$, where $s^0 = s_0, s_2, s_4,\ldots$ and $s^1 = s_1, s_3, \ldots$. And this set is clearly open. Hence, $F$ is continuous. Now, we prove that $F$ is an open map, i.e., $F$ maps open sets to open sets. But, $F(O(s^1) \times O(s^2)) = O(s)$, where $s = s^1_0, s^2_0, s^1_1, \ldots$, which is open. \\
ii) Let us define a map $F \colon \mathcal{N}^\omega \to \mathcal{N} \times \mathcal{N}$ such that $F(x) = y$ if and only if $y(n,m) = x_n(m)$. This is clearly a bijection. Take $O(s) \times O(t)$ and look at $F^{-1}(O(s) \times O(t))$. This is equal to some set $\prod_{i \in \omega} X_i$, where each $X_i$ is a subset if a copy of $\mathcal{N}$. Note that all but finitely many $X_i$ are equal to $\mathcal{N}$ and each $X_i$ is of the form $\bigcup_{i \in \omega} O(w^i)$ for some sequence of finite sequences $w_i$. Conversely, take a set $U$ that is in the basis of $\mathcal{N}^\omega$. Then $U$ is of the form $\prod_{i \in \omega} U_i$, where each $U_i$ is open in $\mathcal{N}$ and all but finitely many $U_i$ are a copy of $\mathcal{N}$. Again, the image of $U$ will be equal to some open $V \subset \mathcal{N} \times \mathcal{N}$. Indeed, it is true that for any finite sequence $s, t$ such that there are $f,g \in V$ extending $s$ and $t$ respectively, it holds that there is a finite extensions of $s,t$, say $s',t'$ such that $[s'], [t'] \in V$.
\end{customthm}
\begin{customthm}{I.4.18}
Let $s \in T_F$, then there is some $f \in F$ such that $s \subset f$. But then $s^\frown f(length(s)+1) \subset f$, so $s$ cannot be maximal.
Take a map $F \mapsto T_F$. We have just proved that $T_F$ has no maximal node, so the map is indeed between closed sets of the Baire space and sequential sequences without maximal nodes. Suppose trees $F$ and $G$ are getting mapped to the same $T$. Hence, $T_F = T_G$. Then, we have that $[T_F]=F$ and $[T_G]=G$ are equal, so $F$=$G$ and we know that our map is injective. Now, take any sequential tree without maximal nodes $T$ and note that $[T]$ is closed. So there is an $F$ that is getting mapped to $T$ via our map. Hence, the map is surjective.
\end{customthm}
\begin{customthm}{I.4.19}
Let $P$ be a perfect Polish space. And let $F \colon P \to P'$ be a homeomorphism to a separable complete metric space. We need to find a closed copy of the Cantor set within $P'$. Let $S$ be a set of all finite sequences of $0$'s and $1$'s. Let us inductively define closed sets $I_s$:
\begin{itemize}
\item[(i)] $I_s \cap P'$ is perfect;
\item[(ii)] $diam(I_s) \leq 1/length(s)$;
\item[(iii)] $I_{s^\frown 0} \subset I_s$, $I_{s^\frown 1} \subset I_s$, and the intersection $I_{s^\frown 0} \cap I_{s^\frown 1}$ is empty.
\end{itemize}
And by completeness, it follows that for each $f \in \set{0,1}^\omega$, the set $P' \cap \bigcap_{n<\omega} I_{f_{|n}}$ has exactly one element.
Now, define $F' \colon \set{0,1}^\omega \to P'$ as $F'(f) = x \in P' \cap \bigcap_{n<\omega} I_{f_{|n}}$. By the definition, $F'$ is an injection. Take any $x \in P' \cap \bigcap_{n<\omega} I_{f_{|n}}$ together with it's neighbourhood $U$. Observe that $F'^{-1}(U)$ is mapped to a set $O(s)$ in the base of the topology of the Cantor space, for some finite sequence of $0$'s and $1$'s $s$. On the other hand, take any base $O(s)$ in the Cantor space and it will be mapped by $F$ to the set of the form $U \cap \bigcup \set{P' \cap \bigcup_{n<\omega} I_{f_{|n}} \mid f \in \set{0,1}^\omega}$, where $U$ is open in $P'$. And this set is open in $P' \cap \bigcup \set{ \bigcap_{n<\omega} I_{f_{|n}} \mid f \in \set{0,1}^\omega}$. Hence, $F'$ is a homeomorphism between the Cantor space and a closed subset of $P'$. This ends the proof.
\end{customthm}
\begin{customthm}{I.4.20}
Let $x_n$ be a dense subset of a Polish space $X$ and define $f(x) = (d(x,x_n) \mid n \in N)$. Note that we can assume $0 \leq d \leq 1$ as we can always normalize the metric via map $d \mapsto 1/(1+d)$.
Suppose $f(x)=f(y)$. Let $\epsilon > 0$ and take $n$ such that $d(x,x_n) < \epsilon/2$. Then, $d(x,y) \leq d(x,x_n) + d(y,x_n = 2d(x,x_n) < \epsilon)$. But this works for any $\epsilon > 0$, so $d(x,y)=0$, hence $x=y$. So $f$ is injective.
Take any $\zeta_m \to x$, then
$f(\zeta_m) = (d(\zeta_m,x_n) \mid n \in N) = ( \lim d(\zeta_m, x_n) \mid n \in N) = f(x)$. So $f$ is continuous.
Take $f(\zeta_m) \to f(x)$. Then, $d(\zeta_m,x_n) \to d(x,x_n)$ for every $n$. But then for any $\epsilon > 0$, there is an $n$ such that for sufficiently large $m$, we have that $d(\zeta_m,x) < d(\zeta_n, x_n) + d(x_n,x) < \epsilon/2 + \epsilon/2 = \epsilon$. Hence, $\zeta_m \to x$ and $f$ is an open mapping.
We now prove that the image of $f$ is a $G_\delta$ subset of the Hilbert's cube. Observe that Hilbert's cube is a Polish space and that the image of $f$ is also a Polish space. Take the base $B_n$ of the Hilbert's cube. For any open neighborhood of $f(x)$ and $\epsilon>0$, there is a base $B_n$ contained in that neighborhood such that $\text{diam}(f(X) \cap B_N) < \epsilon$ (with respect to the induced complete metric). Define a $G_\delta$ set $Y$ as follows.
\begin{equation*}
Y = \bigcap_{m<\omega} \bigcup_{n<\omega} \set{B_n \mid \text{diam}(f(X) \cap B_n) < 1/m}.
\end{equation*}
Observe that $f(X) \subset Y$. Take $y \in Y$. For every $m$, there is an $n(m)$ such that $y \in B_{n(m)}$ with $\text{diam}(f(X) \cap B_{n(m)} < 1/m$. So $f(X)$ is dense in $Y$ and there is a Cauchy sequence $y_m \in f(X)$ converging to $y$. Hence, $y \in f(X)$ and the image of $f$ is $G_\delta$.
\end{customthm}
\end{document}