forked from wangjohn/mit-courses
-
Notifications
You must be signed in to change notification settings - Fork 0
/
PSET5-18.100B.tex
178 lines (131 loc) · 9.9 KB
/
PSET5-18.100B.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
\documentclass[psamsfonts]{amsart}
%-------Packages---------
\usepackage{amssymb,amsfonts}
\usepackage[all,arc]{xy}
\usepackage{enumerate}
\usepackage{mathrsfs}
%--------Theorem Environments--------
%theoremstyle{plain} --- default
\newtheorem{thm}{Theorem}[section]
\newtheorem{cor}[thm]{Corollary}
\newtheorem{prop}[thm]{Proposition}
\newtheorem{lem}[thm]{Lemma}
\newtheorem{conj}[thm]{Conjecture}
\newtheorem{quest}[thm]{Question}
\theoremstyle{definition}
\newtheorem{defn}[thm]{Definition}
\newtheorem{defns}[thm]{Definitions}
\newtheorem{con}[thm]{Construction}
\newtheorem{exmp}[thm]{Example}
\newtheorem{exmps}[thm]{Examples}
\newtheorem{notn}[thm]{Notation}
\newtheorem{notns}[thm]{Notations}
\newtheorem{addm}[thm]{Addendum}
\newtheorem{exer}[thm]{Exercise}
\theoremstyle{remark}
\newtheorem{rem}[thm]{Remark}
\newtheorem{rems}[thm]{Remarks}
\newtheorem{warn}[thm]{Warning}
\newtheorem{sch}[thm]{Scholium}
\makeatletter
\let\c@equation\c@thm
\makeatother
\numberwithin{equation}{section}
\bibliographystyle{plain}
\voffset = -10pt
\headheight = 0pt
\topmargin = -20pt
\textheight = 700pt
%--------Meta Data: Fill in your info------
\title{18.100B \\
Problem Set 5}
\author{John Wang}
\begin{document}
\maketitle
\section{Problem 3.1}
\begin{thm}
The convergence of $\{ s_n \}$ implies the convergence of $\{ | s_n | \}$. The converse is not true, however.
\end{thm}
\begin{proof}
We will assume that $\{ s_n \}$ is a convergent set in $\mathbb{R}$ because the norm is not defined in Rudin for any other metric space. Since $\{ s_n \}$ is a convergent sequence, it must be Cauchy. Therefore, there exists and $N$ such that $m,n \geq N$ implies that $d(s_m,s_n) < \epsilon$ for all $\epsilon > 0$. We can use the traingle inequality, then, to show the following:
\begin{eqnarray}
|s_n| &=& |(s_n + s_m) - s_m | \\
& \leq & |(s_n + s_m) - s_n| + | s_n - s_m| \nonumber \\
&=& |s_m | + d(s_n,s_m) \nonumber \\
&<& |s_m| + \epsilon \nonumber
\end{eqnarray}
This shows that $|s_n| - |s_m| < \epsilon$, which further implies that $d(|s_n|,|s_m|) < \epsilon$ for $m,n \geq N$. Thus, we see that the sequence $\{ | s_n | \}$ is Cauchy, and since it is in $\mathbb{R}$, it converges.
It is easy to show that the converse is not true, namely that the convergence of $\{ |s_n | \}$ does not imply the convergence of $\{ s_n \}$. Take the sequence defined by $s_n = (-1)^n$. Thus, we can see that $|s_n| = |(-1)^n| = 1$ is a constant sequence, and thus definitely converges. However, $s_n$ switches between $-1$ and $1$ and thus will never converge for any $n$.
\end{proof}
\section{Problem 3.2}
\begin{thm}
We show that $\lim_{ n \to \infty} \sqrt{n^2 + n} - n = \frac{1}{2}$.
\end{thm}
\begin{proof}
We multiply through by the conjugate to obtain the following:
\begin{eqnarray}
\lim_{ n \to \infty} \sqrt{n^2 + n} - n &=& \lim_{ n \to \infty} \sqrt{n^2 + n} - n \frac{\sqrt{n^2 + n} + n}{\sqrt{n^2 + n} + n} \\
&=& \lim_{n \to \infty} \frac{n^2 + n - n^2}{\sqrt{n^2 + n} + n} \nonumber \\
&=& \lim_{n \to \infty} \frac{n}{ n \sqrt{\frac{n^2 + n}{n^2} + n}} \nonumber \\
&=& \lim_{n \to \infty} \frac{1}{ \sqrt{1 + \frac{1}{n}} + 1} \nonumber \\
&=& \frac{1}{\sqrt{1} + 1} \nonumber \\
&=& \frac{1}{2} \nonumber
\end{eqnarray}
\end{proof}
\section{Problem 3.3}
\begin{thm}
If $s_1 = \sqrt{2}$ and $s_{n+1} = \sqrt{2 + \sqrt{s_n}}$ for $ n = 1,2,3, \ldots$ then $\{ s_n \}$ converges and $s_n < 2$ for all $n \in \mathbb{N}$.
\end{thm}
\begin{proof}
We know that $0 < \sqrt{2} < 2$, so we have $0 < s_1 < 2$. Moreover, if we add 2 to each side of the inequality, we obtain $2 < 2 + \sqrt{2} < 4$. Taking the square roots, we obtain $\sqrt{2} < \sqrt{2 + \sqrt{2}} < 2$. Adding 2 to each side of this, we obtain $2 + \sqrt{2} < 2 + \sqrt{2 + \sqrt{2}} < 4$. Taking square roots, we have $\sqrt{2} < \sqrt{ 2 + \sqrt{2 + \sqrt{2}}} < 2$. We can repeat the process infinitely many times and notice that $s_1 = \sqrt{2}, s_2 = \sqrt{2 + \sqrt{2}}, s_3 = \sqrt{2 + \sqrt{2 + \sqrt{2}}}, \ldots$. Thus, we see that $s_1 < s_2 < s_3 < \ldots < 2$. We have therefore shown that $\{ s_n \}$ is a monotonically increasing sequence that is bounded because $s_n < 2$ $ \forall n \in \mathbb{N}$. By a theorem in Rudin, we know that $\{ s_n \}$ must converge.
\end{proof}
\section{Problem 3.4}
\begin{thm}
Consider the sequence $\{ s_n \}$ defined by $s_1 = 0$ with $s_{2m} = \frac{s_{2m} - 1}{2}$ and $s_{2m+1} = \frac{1}{2} + s_{2m}$. The lower limit of the sequence is $\frac{1}{2}$ and the upper limit is $1$.
\end{thm}
\begin{proof}
If we compute the sequence, then we see the following, starting at $s_1 = 0$:
\begin{equation}
\{ s_n \} = 0, 0, \frac{1}{2}, \frac{1}{4}, \frac{3}{4}, \frac{3}{8}, \frac{7}{8}, \ldots
\end{equation}
We can use induction to derive the following formulas:
\begin{equation*}
\text{Even n} \rightarrow s_n = \frac{2^{\frac{n}{2}-1} - 1 }{2^{\frac{n}{2} -1}} = \frac{1}{2} - \frac{1}{2^{\frac{n}{2}-1}} \hspace{1cm} \text{Odd n} \rightarrow s_n = \frac{2^{\frac{n-1}{2}} - 1}{2^{\frac{n-1}{2}}} = 1 - \frac{1}{2^n}
\end{equation*}
We can see that for even $n$, we have $s_n < \frac{1}{2}$ and for odd $n$, we have $s_n < 1$. It is clear by inspection that they are subsequential limits. Moreover, we can show that these are the only two subsequential limits of $\{ s_{n} \}$. This is because each subsequence $\{ s_{n_k} \}$, in order to converge, must contain either a finite number of even terms or a finite number of odd terms. There must exist some $N$ such that $n_k \geq N$ implies that each subsequent $s_{n_k}$ is either all odd or all even. If this is not the case, and for $n_k \geq N$ we have $s_{n_k}$ odd (even) and $s_{n_{k+1}}$ even (odd), then $d(s_{n_k},s_{n_{k+1}}) = \frac{1}{2}$ because $s_{2m+1} - s_{2m} = \frac{1}{2}$. This would imply that the subsequence does not converge. Thus, it is clear that $\{\frac{1}{2},1 \}$ are the only two subsequential limits. Thus, we have that the upper bound $s^{*} = \sup \{ \frac{1}{2}, 1 \} = 1$ and the lower bound $s_{*} = \inf \{ \frac{1}{2},1 \} = \frac{1}{2}$.
\end{proof}
\section{Problem 3.20}
\begin{thm}
Suppose $\{ p_n \}$ is a Cauchy sequence in a metric space $X$, and some subsequence $\{ p_{n_i} \}$ converges to a point $p \in X$. Then the full sequence $\{ p_n \}$ converges to $p$.
\end{thm}
\begin{proof}
Since we know that $\{ p_{n_i} \}$ converges, we know that there exits an $N_0$ such that $n_i \geq N_0$ implies $d(p_{n_i}, p) < \epsilon$ for all $\epsilon > 0$. Since the sequence $\{ p_n \}$ is Cauchy, there exists an $M$ such that $n \geq M$ and $m \geq M$ implies $d(p_n, p_m) < \epsilon$ for all $\epsilon > 0$. By the triangle inequality, we obtain for $n, n_i \geq \max \{ N_0, M \} $ that
\begin{eqnarray}
d(p_n,p) &\leq& d(p_n, p_{n_i}) + d(p_{n_i},p) \\
&<& \epsilon + \epsilon = 2 \epsilon \nonumber
\end{eqnarray}
We know that $d(p_n, p_{n_i} ) < \epsilon$ by the fact that $\{ p_n \}$ is Cauchy, because $p_{n_i} \in \{ p_n \}$ as it $p_{n_i}$ part of a subsequence of $\{ p_n \}$. We obtain $d(p_{n_i},p) < \epsilon$ because the subsequence $\{ p_{n_i} \}$ converges to $p$. Thus, since $\epsilon$ is arbitrary, we obtain that the full sequence $\{ p_n \}$ converges to $p$.
\end{proof}
\section{Problem 3.21}
\begin{thm}
If $\{ E_n \}$ is a sequence of closed, nonempty, and bounded sets in a complete metric space $X$, if $E_n \supset E_{n+1}$, and if $\lim_{n \to \infty} \text{diam} E_n = 0$, then $\bigcap_1^\infty E_n$ consists of exactly one point.
\end{thm}
\begin{proof}
Suppose $\{ p_n \}$ is any sequence with $p_n \in E_n$. Then the assumption that $\lim_{n \to \infty} \text{diam} E_n = 0$ means that there exists an $N$ such that for all $\epsilon > 0$, we have $d(\text{diam} {E_n},0) < \epsilon$ for $n \geq N$. Thus, we have $\text{diam} E_n < \epsilon$. Moreover, since we have $E_m \supset E_n$ for $m \geq n \geq N$, we know that $ d(p_n, p_m) < \text{diam} E_n < \epsilon $ for $p_n \in E_n$ and $p_m \in E_m$. This implies that $\{ p_n \}$ is Cauchy, and since we have assumed that $X$ is complete, $\{ p_n \}$ must converge to some limit $p$ in $X$. Moreover, since $p$ is the limit of $\{ p_n \}$, it is also a limit point of $ E_n$. This is because for $n > N$, $ d(p_n, p) < \epsilon $ which implies that there exists a point $p_n \in E_n$ for every neighborhood of $p$. Since we assumed each $E_n$ is closed, we know that $p$ must be contained in each $E_n$. Thus, we know that $ p \in \bigcap_1^\infty E_n$.
It is clear that there are no more elements in $\bigcap_1^\infty E_n$. Assume by contradiction that there exists a $q \neq p$ such that $q \in \bigcap_1^\infty E_n $. Since $q \neq p$, we know that $d(p,q) > 0$ by definition of a metric space. Since both $p$ and $q$ belong to $E_1, E_2, \ldots$ we see that $ \sup d(p_n,q_n) > 0$ for $p_n, q_n \in E_n$. Thus, the diameter $\text{diam} E_n > 0$ for all $n$. Thus shows that $\lim_{n \to \infty} \text{diam} E_n \neq 0$, which is a contradiction of our assumption. Thus, the set $\bigcap_1^\infty E_n$ contains exactly one element.
\end{proof}
\section{Problem 3.23}
\begin{thm}
Suppose $\{ p_n \}$ and $\{ q_n \}$ are Cauchy sequences in a metric space $X$. Show that the sequence $\{ d(p_n, q_n) \}$ converges.
\end{thm}
\begin{proof}
Since $\{ p_n \}$ is a Cauchy sequence, we know that there exists some $N$ such that $d(p_n,p_m) < \epsilon$ for all $\epsilon > 0$ and $m,n \geq N$. Since $\{q_n \}$ is Cauchy, we know there exists some $M$ such that $d(q_n,q_m) < \epsilon$ for all $\epsilon > 0$ and $m,n \geq N$. Using the triangle inequality, we find that
\begin{eqnarray}
d(p_n,q_n) &\leq& d(p_n, p_m) + d(p_m, q_n) \\
&\leq& d(p_n,p_m) + d(p_m, q_m) + d(q_m, q_n) \nonumber \\
&<& 2 \epsilon + d(p_m, q_m) \nonumber \\
d(p_n,q_n) - d(p_m,q_m) &<& 2 \epsilon
\end{eqnarray}
This implies that $d( d(p_n,q_n), d(p_m,q_m) ) = |d(p_n,q_n) - d(p_m, q_m) | < 2 \epsilon$ for all $\epsilon >0$ and $m,n \geq N$. Thus we know that $\{ d(p_n,q_n) \}$ is a Cauchy sequence. Since $\mathbb{R}$ is complete, we know that $\{ d(p_n,q_n) \}$ converges.
\end{proof}
\end{document}