If you find any mistakes, please make a comment! Thank you.

Solution to Principles of Mathematical Analysis Chapter 7 Part B


Chapter 7 Sequences and Series of Functions

Exercise 13

(By analambanomenos)

(a) Following the steps of the hint: (i) Show that some subsequence $\{f_{n_i}\}$ converges at all rational points $r$, say, to $f(r)$.

Let $\{r_m\}$ be an enumeration of the rational points of $\mathbf R$. Since $\{f_n(r_1)\}\subset[0,1]$, a compact set, there is a subsequence $\{f_{1,n_j}(r_1)\}$ which converges to $f(r_1)$, by Theorem 2.41. Similarly, the $\{f_{1,n_j}(r_2)\}$ has a subsequence $\{f_{2,n_j}(r_2)\}$ which converges to $f(r_2)$. Continuing in this fashion, we get a sequence of subsequences $\{f_{m,n_j}\}$ such that each $\{f_{m,n_j}\}$ is a subsequence of $\{f_{m-1,n_j}\}$ and $f_{m,n_j}(r_m)$ converges to $f(r_m)$. Letting $f_{n_i}=f_{i,n_i}$, we get a subsequence of $\{f_n\}$ which converges on all rational points $r$ of $\mathbf R$ to $f(r)$.

(ii) Define $f(x)$, for any $x\in\mathbf R$, to be $\sup f(r)$, the sup being taken over all rational points $r\le x$.

First we need to show that this new definition of $f(r)$ agrees with the previous one; let $f^*(r)$ be the new definition of $f(r)$. It is clear from the definition that $f^*(r)\ge f(r)$. Suppose $f^*(r)>f(r)$, that is, there is a rational point $q<r$ and a real number $a$ such that $f(q)>a>f(r)$. Since $f(q)=\lim f_{n_i}(q)$, and $f(r)=\lim f_{n_i}(r)$, there is an integer $N$ such that for all $i\ge N$ we have $f_{n_i}(q)>a>f_{n_i}(r)$, but this contradicts the fact that these functions are monotonically increasing.

It is clear from the definition that $f$ is monotonically increasing.

(iii) Show that $f_{n_i}(x)\rightarrow f(x)$ at every $x$ at which $f$ is continuous.

Suppose $f$ is continuous at $x$ and let $\varepsilon>0$. Then there are rational numbers $r_1<x<r_2$ such that $$f(x)-\frac{\varepsilon}{2}<f(r_1)\le f(x)\le f(r_2)<f(x)+\frac{\varepsilon}{2}.$$ Since $f_{n_i}(r_1)\rightarrow f(r_1)$ and $f_{n_i}(r_2)\rightarrow f(r_2)$, there is an integer $N$ such that for all $i>N$ we have $$\big|f_{n_i}(r_1)-f(r_1)\big|<\frac{\varepsilon}{2}\quad\hbox{and}\quad\big|f_{n_i}(r_2)-f(r_2)\big|<\frac{\varepsilon}{2}.$$ And since $f_{n_i}$ is monotonically increasing, we have for all $i>N$ $$f(x)-\varepsilon<f(r_1)-\frac{\varepsilon}{2}<f_{n_i}(r_1)\le f_{n_i}(x)\le f_{n_i}(r_2)<f(r_2)+\frac{\varepsilon}{2}<f(x)+\varepsilon.$$ That is, $\big|f(x)-f_{n_i}(x)\big|<\varepsilon$, so $f_{n_i}(x)\rightarrow f(x)$ as $i\rightarrow
\infty$.

(iv) Show that a subsequence of $\{f_{n_i}\}$ converges at every point of discontinuity of $f$ since there are at most countably many such points.

Since $f$ is monotonically increasing, $f$ has at most countably many simple points of discontinuity by Theorems 4.29 and 4.30. Let $\{x_m\}$ be an enumeration of the points of discontinuity of $f$. Repeating the argument of step (i), there is a subsequence $\{f_{1,n_j}(x_1)\}$ of $\{f_{n_i}(x_1)$ which converges to a value $a$ between 0 and 1. We can continue this way for all the $x_m$, getting a sequence of subsequences $\{f_{m,n_j}\}$ of $\{f_{n_i}\}$ such that each $\{f_{m,n_j}\}$ is a subsequence of $\{f_{m-1,n_j}\}$ for which $f_{m,n_j}(x_m)$ converges. Letting $f_{n_k}=f_{k,n_k}$, we get a subsequence of $\{f_n\}$ which converges at all the points $\{x_m\}$. We can redefine $f(x_i)$ at these points to be those values.

Since we have already shown that $\{f_{n_k}\}$ converges at the points where $f$ is continuous, we have $f_{n_k}(x)$ converging to $f(x)$ for all real values $x$.

(b) Let $K$ be a compact subset of $\mathbf R$ and let $\varepsilon>0$. It was shown in step (iii) of part (a) that, since $f$ is everywhere continuous, $f(x)=\lim f_{n_k}(x)$ for all points $x$. Also, we are using the initial definition of $f$, so that is it monotonically increasing. For each $x\in K$ there is a $\delta_x>0$ such that for all $y\in[x-\delta_x,x+\delta_x]$ we have
$$\big|f(y)-f(x)\big|<\frac{\varepsilon}{4}.$$Also, there is an integer $N_x$ such that for all $k>N_x$ $$\big|f_{n_k}(x-\delta_x)\big|<\frac{\varepsilon}{2}\quad\hbox{and}\quad\big|f_{n_k}(x+\delta_x)\big|<\frac{\varepsilon}{4}.$$ Since $f_{n_k}$ is monotonically increasing, we must have, for all $y\in(x-\delta_x,x+\delta_x)$ and all $k>N_x$, $$f(x)-\frac{\varepsilon}{2}<f(x-\delta_x)-\frac{\varepsilon}{4}<f_{n_k}(x-\delta_x)\le f_{n_k}(y)\le f_{n_k}(x+\delta_x)+\frac{\varepsilon}{4}<f(x)+\frac{\varepsilon}{2},$$ that is, $\big|f_{n_k}(y)-f(x)\big|<\varepsilon/2$. Hence, for $y\in(x-\delta_x,x+\delta_x)$ and $k>N_x$, $$\big|f_{n_k}(y)-f(y)\big|\le\big|f_{n_k}(y)-f(x)\big|+\big|f(x)-f(y)\big|<\varepsilon.$$ Since the open neighborhoods $(x-\delta_x,
x+\delta_x)$ cover the compact set $K$, a finite subset of these neighborhoods, centered at $x_1,\ldots,x_M$, also covers $K$. Letting $N=\max(N_{x_1},\ldots,N_{x_M})$, we have for all $y\in K$ and all $k>N$ that $\big|f_{n_k}(y)-f(y)\big|\le\varepsilon$, that is, $\{f_{n_k}\}$ converges uniformly to $f$ on $K$.


Exercise 14

(By analambanomenos) Since for each $n$ we have $\big|2^{-n}f(3^{2n-1}t)\big|\le2^{-n}$ and $\big|2^{-n}f(3^{2n}t)\big|\le2^{-n}$, the series defining $x(t)$ and $y(t)$ converge uniformly since $\sum 2^{-n}$ converges, by Theorem 7.10. And since the partial sums are continuous functions, $x(t)$ and $y(t)$ are continuous functions, by Theorem 7.12. Hence $\Phi$ is continuous.

Following the hint, let $(x_0,y_0)\in I^2$, and let $$x_0=\sum_{n=1}^\infty 2^{-n}a_{2n-1},\quad\quad y_0=\sum_{n=1}^\infty 2^{-n}a_{2n}$$ be the binary expansions of $x_0$ and $y_0$, where each $a_i$ is 0 or 1. Let $$t_0=\sum_{i=1}^\infty 3^{-i-1}(2a_i).$$ By Exercise 3.19, the set of all such $t_0$ is precisely the Cantor set.

We have for $k=1,2,3,\ldots,$
\begin{align*}
3^kt_0 &= \sum_{i=1}^\infty 3^{k-i-1}(2a_i) \\
&= 2\sum_{n=0}^{k-2}3^na_{k-1-n}+\frac{2}{3}\,a_k+\frac{2}{3}\sum_{n=1}^\infty 3^{-n}a_{k+n}
\end{align*}Note that the last term lies between 0 and $$\frac{2}{3}\sum_{n=1}^\infty 3^{-n} = \frac{2}{3}\biggl(\frac{1/3}{1-(1/3)}\biggr)=\frac{1}{3}.$$ Also, the first term is an even integer, so that, since $f(t+2)=f(t)$, we have $$f(3^kt_0)=f\Biggl(\frac{2}{3}\,a_k+\frac{2}{3}\sum_{n=1}^\infty 3^{-n}a_{k+n}\Biggr).$$ If $a_k=0$, then the expression in the argument of $f$ lies between 0 and $\frac{1}{3}$, so that $f(3^kt_0)=0$. And if $a_k=1$, then the expression in the argument of $f$ lies between $\frac{2}{3}$ and 1, so that $f(3^kt_0)=1$. In either case, we have $f(3^kt_0)=a_k$. Hence $\Phi(t_0)=\bigl(x(t_0),y(t_0)\bigr)=(x_0,y_0)$.


Exercise 15

(By analambanomenos) The function $f$ must be constant on $[0,\infty]$. Let $0\le x<y$ and let $\varepsilon>0$. Then there is a $\delta>0$ such that for all $n=1,2,3,\ldots,$ we have $\big|f_n(z)-f_n(w)\big|<
\varepsilon$ if $0\le w,z\le 1$ and $|w-z|<\delta$. For large enough $N$, we have $$\bigg|\frac{y}{N}-\frac{x}{N}\bigg|=\frac{|y-x|}{N}<\delta,\quad\quad 0\le\frac{x}{N}<\frac{y}{N}\le1.$$Hence $$\big|f(y)-f(x)\big|=\Bigg|f_N\biggl(\frac{y}{N}\biggr)-f_N\biggl(\frac{x}{N}\biggr)\Bigg|<\varepsilon.$$ Since $\varepsilon>0$ was arbitrary, this shows that $f(x)=f(y)$.

By the way, it is easy to extend this argument to show that if $\{f_n\}$ were equicontinuous on $[-1,1]$, then $f$ would be constant on all of $\mathbf R$.


Exercise 16

(By analambanomenos) Let $\varepsilon>0$ and let $\delta>0$ such that if $d(x,y)<\delta$ then $$\big|f_n(x)-f_n(y)\big|<\varepsilon/3$$ for all $n$. Let $x\in K$ and let $N_x$ be an integer large enough so that if $m>N_x$ and $n>N_x$, then $\big|f_m(x)-f_n(x)\big|<\varepsilon$. Then for all $m>N_x$ and $n>N_x$ and all $y\in K$ such that $d(x,y)<\delta$, we have
$$\big|f_n(y)-f_m(y)\big|\le\big|f_n(y)-f_n(x)\big|+\big|f_n(x)-f_m(x)\big|+\big|f_m(x)-f_m(y)\big|<\varepsilon.$$ Since $K$ is compact, there are a finite number of points $x_1,\ldots,x_M$ in $K$ such that the neighborhoods of radius $\delta$ centered at the $x_i$ cover $K$. So if we let $N=\max(N_{x_1},\ldots,N_{x_M})$, then for $m>N$ and $n>N$ and for all $y\in K$, we have $\big|f_m(y)-f_n(y)\big|<\varepsilon$. Hence by Theorem 7.8, $\{f_n\}$ converges uniformly on $K$.


Exercise 17

(By analambanomenos) The extensions of the definitions, and the statements and proofs of Theorems 7.8 through 7.11, are trivial, so I will simply copy them from the text, italicizing the changes.

Definition 7.7 We say that a sequence of mappings $\{f_n\}$, $n=1,2,3,\ldots,$ into a metric space $F$ converges uniformly on $E$ to a function $f$ if for every $\varepsilon>0$ there is an integer $N$ such that $n\ge N$ implies \begin{equation}\label{7.17.1}d\bigl(f_n(x),f(x)\bigr)\le\varepsilon \end{equation} for all $x\in E$.

Definition 7.22 A family $\mathscr F$ of mappings $f$ defined on a set $E$ in a metric space $X$ with values in a metric space $Y$ is said to be equicontinuous on $E$ if for every $\varepsilon>0$ there exists a $\delta>0$ such that $$d_Y\bigl(f(x),f(y)\bigr)<\varepsilon$$ whenever $d_X(x,y)\le\delta$, $x\in E$, $y\in E$, and $f\in\mathscr F$.

Theorem 7.8 The sequence of mappings $\{f_n\}$ into a complete metric space $F$, defined on $E$, converges uniformly on $E$ if and only if for every $\varepsilon>0$ there exists an integer $N$ such that $m\ge N$, $n\ge N$, $x\in E$ implies $d\bigl(f_n(x),f_m(x)\bigr)\le\varepsilon$.

Proof: Suppose $\{f_n\}$ converges uniformly on $E$, and let $f$ be the limit function. There there is an integer $N$ such that $n\ge E$ implies $d\bigl(f_n(x),f(x)\bigr)\le
\varepsilon/2$ so that $$d\bigl(f_n(x),f_m(x)\bigr)\le d\bigl(f_n(x),f(x)\bigr)+d\bigl(f(x),f_m(x)\bigr)\le\varepsilon$$ if $n\ge N$, $m\ge M$, $x\in E$.

Conversely, suppose the Cauchy condition holds. Since $F$ is a complete metric space, the sequence $\{f_n(x)\}$ converges, for every $x$, to a limit which we may call $f(x)$. Thus the sequence $\{f_n\}$ converges on $E$, to $f$. We have to prove that the convergence is uniform.

Let $\varepsilon>0$ be given, and choose $N$ such that \eqref{7.17.1} holds. Fix $n$, and let $m\rightarrow\infty$ in \eqref{7.17.1}. Since $f_m(x)\rightarrow f(x)$ as $m\rightarrow\infty$, this gives $d\bigl(f_n(x),f(x)\bigr)\le\varepsilon$ for every $n\ge N$ and every $x\in E$, which completes the proof.

Theorem 7.9 Suppose $\{f_n\}$ {\it is a sequence of mappings from $E$ into a metric space $F$ such that} $\lim f_n(x)=f(x)$ for $x\in E$. Put $$M_n=\sup_{x\in E}d\bigl(f_n(x),f(x)\bigr).$$ Then $f_n\rightarrow f$ uniformly on $E$ if and only if $M_n\rightarrow0$ as $n\rightarrow\infty$.

Proof: Since this is an immediate consequence of Definition 7.7, we omit the details of the proof.

Theorem 7.10 Suppose $\{f_n\}$ is a sequence of vector-valued functions with values in $\mathbf R^k$ defined on $E$, and suppose $$\big|\big|f_n(x)\big|\big|\le M_n\quad\quad
(x\in E,\, n=1,2,3,\ldots).$$ Then $\sum f_n$ converges uniformly on $E$ if $\sum M_n$ converges.

Proof: If $\sum M_n$ converges, then, for arbitrary $\varepsilon>0$, $$\Bigg|\Bigg|\sum_{i=n}^mf_i(x)\Bigg|\Bigg|\le\sum_{i=n}^m M_i\le\varepsilon\quad\quad(x\in E),$$ provided $m$ and $n$ are large enough. Uniform convergence now follows from Theorem 7.8.

Theorem 7.11 Suppose $\{f_n\}$, a sequence of mappings with values in a complete metric space $F$, converges uniformly on a set $E$ in a metric space. Let $x$ be a limit point
of $E$, and suppose that $$\lim_{t\rightarrow x}f_n(t)=A_n\quad\quad(n=1,2,3,\ldots).$$ Then $\{A_n\}$ converges and \begin{equation}\label{7.17.2}\lim_{t\rightarrow x}f(t)=\lim_{n\rightarrow\infty}A_n.\end{equation}

Proof: Let $\varepsilon>0$ be given. By the uniform convergence of $\{f_n\}$, there exists $N$ such that $n\ge N$, $m\ge N$, $t\in E$ imply \begin{equation}\label{7.17.3} d\bigl(f_n(t),f_m(t)\bigr)\le\varepsilon.\end{equation} Letting $t\rightarrow x$ in \eqref{7.17.3}, we obtain $$d(A_n,A_m)\le\varepsilon$$ for $n\ge N$, $m\ge M$, so that $\{A_n\}$ is a Cauchy sequence and therefore converges, say to $A$.

Next, \begin{equation}\label{7.17.4}d\bigl(f(t),A\bigr)\le d\bigl(f(t),f_n(t)\bigr)+d\bigl(f_n(t),A_n\bigr)+d\bigl(A_n,A\bigr)\end{equation} We first choose $n$ such that \begin{equation}\label{7.17.5} d\bigl(f(t),f_n(t)\bigr)\le\frac{\varepsilon}{3}\end{equation} for all $t\in E$ (this is possible by the uniform convergence), and such that \begin{equation}\label{7.17.6}d\bigl(A_n,A\bigr)\le\frac{\varepsilon}{3}.\end{equation} Then, for this $n$ we choose a neighborhood $V$ of $x$ such that \begin{equation}\label{7.17.7}d\bigl(f_n(t)-A_n\bigr)\le\frac{\varepsilon}{3},\end{equation} if $t\in V\cap E$, $t\ne x$.

Substituting the inequalities \eqref{7.17.5}, \eqref{7.17.6}, \eqref{7.17.7} into \eqref{7.17.4}, we see that $d\bigl(f(t),A\bigr)\le\varepsilon$, provided $t\in V\cap E$, $t\ne x$. This is equivalent to \eqref{7.17.2}.

Theorem 7.12 If $\{f_n\}$ is a sequence of continuous {\it mappings on $E$ into a metric space} $F$, and if $f_n\rightarrow f$ uniformly on $E$, then $f$ is continuous on $E$.

Proof: (In the text, Theorem 7.12 is an immediately corollary of Theorem 7.11. Here, that would require that $F$ be complete, which is not necessarily the case. So the following proof is entirely new.)

By Theorem 4.8, to show that $f$is continuous, it suffices to show that $f^{-1}(V)$ is open for any open set $V\subset F$. If $x\in f^{-1}(V)$, we need to find a neighborhood $N_x$ of $x$ which is contained in $f^{-1}(V)$, for then $f^{-1}(V)$ would be the union of the open neighborhoods $N_x$, $x\in f^{-1}(V)$, and so be open.

Let $\varepsilon>0$ be small enough so that the neighborhood $M_y$ of $y=f(x)$ of radius $\varepsilon$ is contained in $V$. Let $n$ be large enough so that $d\bigl(f_n(z),f(z)\bigr)
\le\varepsilon/3$ for all $z\in E$, which is possible by the assumption of uniform convergence. Let $\delta>0$ such that $d\bigl(f_n(z),f_n(x)\bigr)<\varepsilon/3$ for all $z$ in the neighborhood $N_x$ of $x$ of radius $\delta$, which is possible since $f_n$ is continuous. Then, for all $z\in N_x$, $$d\bigl(f(z),f(x)\bigr)\le d\bigl(f(z),f_n(z)\bigr)+d\bigl(f_n(z),f_n(x)\bigr)+
d\bigl(f_n(x),f(x)\bigr)\le\varepsilon.$$ That is, $f(z)\in M_y\subset V$, or $z\in f^{-1}(V)$, so that $N_x\subset f^{-1}(V)$.

The remaining Theorems, for mappings into $\mathbf R^k$, are largely corollaries of their scalar counterparts in the text.

Theorem 7.16 Let $\alpha$ be monotonically increasing on $[a,b]$. Suppose $\mathbf f_n=(f_{n1},\ldots,f_{nk})\in\mathscr R(\alpha)$ on $[a,b]$, for $n=1,2,3,\ldots,$ and suppose $\mathbf f_n \rightarrow\mathbf f=(f_1,\ldots,f_k)$ uniformly on $[a,b]$. Then $\mathbf f\in\mathscr R(\alpha)$ on $[a,b]$, and $$\int_a^b\mathbf f\,d\alpha=\lim_{n\rightarrow\infty}\int_a^b
\mathbf f_n\,d\alpha.$$

Theorem 7.17 Suppose $\big\{\mathbf f_n=(f_{n1},\ldots,f_{nk})\big\}$ is a sequence of vector-valued functions, differentiable on $[a,b]$ and such that $\big\{\mathbf f(x_0)\big\}$ converges for some point $x_0$ on $[a,b]$. If $\{\mathbf f_n’\}$ converges uniformly on $[a,b]$, to a function $\mathbf f=(f_1,\ldots,f_k)$, and $$\mathbf f’(x)=\lim_{n\rightarrow\infty}f_n’(x)\quad\quad(a\le x\le b).$$

Proofs: If $\mathbf x=(x_1,\ldots,x_k)\in\mathbf R^k$, then $|x_i|\le||\mathbf x||\le\sum_i^k|x_i|$. Hence $\mathbf f_n$ convergences uniformly if and only if the component functions $f_{n1},\ldots,f_{nk}$ converge uniformly. And since the integers or derivatives, and integrability or differentiability, of vector-valued functions are defined by component, the vector-valued versions of the Theorems 7.16 and 7.17 are immediate corollaries of the scalar versions in the text.

Theorem 7.24 If $K$ is a compact metric space, if $\mathbf f_n=(f_{n1},\ldots,f_{nk})\in\mathscr C(K)$ for $n=1,2,3,\ldots,$ and if $\{\mathbf f_n\}$ converges uniformly on $K$, then $\{\mathbf f_n\}$ is equicontinuous on $K$.

Proof: If $\mathbf x=(x_1,\ldots,x_k)\in\mathbf R^k$, then $|x_i|\le||\mathbf x||\le\sum_i|x_i|$. Hence $\{\mathbf f_n\}$ is equicontinuous if and only if each of the $\{f_{ni}\}$, $i=1,\ldots,k$ are equicontinuous. Hence the vector-valued version of Theorem 7.24 is an immediate corollary of the scalar version in the text.

Theorem 7.25 If $K$ is compact, if $\mathbf f_n=(f_{n1},\ldots,f_{nk})$ is pointwise bounded and equicontinuous on $K$, then

(a) $\{\mathbf f_n\}$ is uniformly bounded on $K$,

(b)$\{\mathbf f_n\}$ contains a uniformly convergent subsequence.

Proof: Since each of the sequences $\{f_{ni}\}$, $i=1,\ldots,k$, is pointwise bounded and equicontinuous on $K$, by the scalar version of Theorem 7.25 in the text each of them is uniformly bounded on $K$, so $\{\mathbf f_n\}$ is uniformly bounded on $K$. Also, $\{f_{n1}\}$ contains a uniformly convergent subsequence, $\{f_{n_j1}\}$. Since the subsequence $\{f_{n_j2}\}$ is also pointwise bounded and equicontinuous on $K$, it also has a uniformly convergent subsequence such that the corresponding subsequence of $\{f_{n_j1}\}$ also converges uniformly. Continuing in this manner, after $k$ steps we have a subsequence $\{\mathbf f_{n_j}\}$ whose component functions all converge uniformly, so $\{\mathbf f_{n_j}\}$ also converges uniformly.

Baby Rudin 数学分析原理完整第七章习题解答


Linearity

This website is supposed to help you study Linear Algebras. Please only read these solutions after thinking about the problems carefully. Do not just copy these solutions.
Close Menu