0709.4492/arXivNewProofAnalysis4.tex
1: %
2: \documentclass[12pt, reqno]{amsart}
3: \usepackage{amsthm}
4: \usepackage{amssymb}
5: \usepackage{latexsym}
6: \usepackage{url}
7: %\usepackage{amsmath}
8: \usepackage{verbatim}
9: \usepackage{graphicx}
10: %\usepackage[font=footnotesize,labelfont=sf,textfont=sf]{subfig}
11: \usepackage{epsf}
12: \usepackage{enumerate}
13: \usepackage{geometry}
14: \geometry{top=1.1in, bottom=1.1in, left=1.38in, right=1.38in}
15: \usepackage{hyperref}
16: \hypersetup{breaklinks=true,
17: pagecolor=white,
18: colorlinks=true,%false
19: citecolor=black,%
20: filecolor=black,%
21: linkcolor=black,%
22: urlcolor=black
23: %pdftex}
24: }
25: \urlstyle{rm}
26: 
27: 
28: 
29: \newtheorem{theorem}{Theorem}[section]
30: \newtheorem{lemma}[theorem]{Lemma}
31: \newtheorem{proposition}[theorem]{Proposition}
32: \newtheorem{corollary}[theorem]{Corollary}
33: \theoremstyle{definition}
34: \newtheorem{definition}[theorem]{Definition}
35: \newtheorem{example}[theorem]{Example}
36: \newtheorem{exercise}[theorem]{Exercise}
37: \newtheorem{conclusion}[theorem]{Conclusion}
38: \newtheorem{conjecture}[theorem]{Conjecture}
39: \newtheorem{criterion}[theorem]{Criterion}
40: \newtheorem{summary}[theorem]{Summary}
41: \newtheorem{axiom}[theorem]{Axiom}
42: \newtheorem{problem}[theorem]{Problem}
43: \newtheorem{remark}[theorem]{Remark}
44: 
45: \numberwithin{equation}{section}
46: \newcommand{\Int}{\textnormal{Int}}
47: \newcommand{\Ext}{\textnormal{Ext}}
48: \newcommand{\R}{\mathbb{R}}
49: \newcommand{\Q}{\mathbb{Q}}
50: \newcommand{\C}{\mathbb{C}}
51: \newcommand{\N}{\mathbb{N}}
52: \newcommand{\sign}{\textnormal{sign}}
53: \newcommand{\ud}{\, \mathrm{d}}
54: \newcommand{\wt}{\widetilde}
55: 
56: \begin{document}
57: %\setcounter{page}{1}
58: 
59: 
60: 
61: \title[New proofs and improvements]{New proofs and improvements of classical theorems in analysis}
62: 
63: \author[Daniel Reem]{Daniel Reem}
64: 
65: \address{Department of Mathematics, The Technion - Israel Institute of Technology. Current address: Department of Mathematics, The University of Haifa, Haifa 31905, Israel}
66: 
67: \email{dream@tx.technion.ac.il}
68: 
69: 
70: %\dedicatory{This paper is dedicated to Professor ABCD}
71: 
72: \subjclass[2010]{03F99, 26A15, 26A03, 47H10, 54D05} 
73: \keywords{Constructive, fixed point, intermediate value theorem, extreme value theorem, optimal delta, uniform continuity}
74: 
75: %\date{September 7, 2010\\
76: %\hspace*{0.39cm} 2010 Mathematics Subject Classification. 03F99, 26A15, %26A03, 54D05}
77: 
78: 
79: 
80: 
81: 
82: \begin{abstract}
83: New proofs and improvements of three classical theorems in analysis are  presented. Although these theorems are well-known, and have been extensively investigated over the years, it seems that new light can be shed on them. We first present a quantitative necessarily and sufficient condition for a function to be uniformly continuous, and as a by-product we obtain explicitly the optimal delta for the given epsilon. The uniform continuity of a continuous function  defined on a compact metric space follows as a simple consequence. We  proceed with the extreme value theorem and present a ``programmer's proof'', a proof  which does not use the costume argument of proving boundedness first. We finish with the intermediate value theorem, which is generalized to a class of discontinuous functions, and, in addition, the meaning of the intermediate value property is re-examined and a fixed point theorem for (very) discontinuous functions is established. At the end we discuss briefly in which sense the proofs are constructive.
84: \end{abstract} 
85: \maketitle
86: 
87: \section{introduction}
88: In this note we present new proofs and improvements of three classical   theorems in analysis. Although these theorems are well-known, and have been extensively investigated over the years, it seems that new light can be shed on them.
89: 
90: We first present a quantitative necessarily and sufficient condition for a function between two metric spaces to be uniformly continuous, and as a by-product we obtain explicitly the optimal delta for the given epsilon. The uniform continuity of a continuous function  defined on a compact metric space follows as a simple consequence. We  proceed with the extreme value theorem and present two proofs of this theorem: the ``envelope proof'' in which the largest possible maximal point is found using an envelope function, and  the ``programmer's proof'',  which does not use the costume argument of proving boundedness first. We finish with the intermediate value theorem, which is generalized to a class of discontinuous functions and, in addition, the meaning of the intermediate value property is re-examined, and a fixed point theorem for (very) discontinuous functions is established. At the end we discuss briefly in which sense the proofs are constructive.
91: 
92: %The proofs presented here have an elementary character. However, in contrast to the well-known proofs of these theorems (see e.g., \cite{Kitchen, Nitecki, Spivak}), these proofs are probably not suitable for a first semester course in calculus.
93: 
94: \section{Uniform continuity}
95: A well-known fact is that any real continuous function defined on a compact interval/metric space $I$  is uniformly continuous, i.e., for any $\epsilon>0$ there exists $\delta>0$ such that for all $x,y\in I$, if  $|x-y|<\delta$, then $|f(x)-f(y)|<\epsilon$. Familiar proofs of this fact, e.g., the ones taken from  
96: \cite[p. 229]{Folland}, \cite[pp. 87-88]{HewittStromberg},
97: \cite[pp. 273-274]{Hille},  \cite[pp. 19-20]{Jones}, \cite[p. 193]{Kitchen}, \cite[pp. 33-34]{LangAnalysisII}, \cite[p. 168-169]{Nitecki}, \cite[pp. 48-49, 157]{Royden}, \cite[p. 91]{Rudin},  \cite[p. 114]{Sagan}, \cite[p. 143-144]{Spivak},  and \cite[pp. 238, 644-645]{ThomsonBruckner} show the existence of such positive $\delta$, but do not give any single clue for finding it explicitly. In particular, they do not give any information on how to find the largest possible such $\delta$.  Is it possible to find explicitly this optimal $\delta$? the following lemma   shows that the answer is positive. A key step is simply to reformulate the condition of uniform continuity. The uniform continuity of a continuous function defined on a compact metric space is a simple consequence of this lemma.
98: \begin{lemma}\label{lem:UniCont} 
99: Let $(X,d_X)$ and $(Y,d_Y)$ be two metric spaces, and let $f:X\to Y$. 
100: \begin{enumerate}[(a)]
101: \item $f$ is uniformly continuous if and only if for each $\epsilon>0$ there exists $\delta>0$  such that for all $x,y\in X$, if $d_Y(f(x),f(y))\geq\epsilon$, then $d_X(x,y)\geq \delta$. 
102: \item Suppose that $f$ is non-constant and let $M\in (0,\infty]$ be defined by $M:=\sup\{d_Y(f(x),f(y)): x,y\in X\}$. Then $f$  is uniformly continuous if and only if 
103: the function  $\delta:[0,M) \to [0,\infty)$ defined by
104: \begin{equation}\label{eq:delta}
105: \delta(\epsilon)=\inf \{d_X(x,y):  (x,y)\in X^2,\,d_Y(f(x),f(y))\geq \epsilon\}
106: \end{equation}
107: satisfies $\delta(\epsilon)>0$ for each $\epsilon\in (0,M)$. Moreover, the function $\delta$ defined above assigns to each $\epsilon\in (0,M)$ the largest possible delta from the definition of uniform continuity. 
108: \end{enumerate}
109: \end{lemma}
110: \begin{proof}
111: \begin{enumerate}[(a)]%$\,$
112: \item Simple.
113: \item Since the set $A_{\epsilon}:=\{(x,y)\in X^2: d_Y(f(x),f(y))\geq \epsilon\}$ is nonempty by the choice of $\epsilon$, it follows that the  function $\delta$ is well defined. Now, if $\delta(\epsilon)>0$ for each $\epsilon\in (0,M)$, then from (a small modification of) the previous part $f$ is uniformly continuous. On the other hand, suppose that $f$ is uniformly continuous, and let $\wt{\delta}$ be any positive number  with the property that  $d_X(x,y)<\wt{\delta}$ implies $d_Y(f(x),f(y))<\epsilon$. Then for each $(x,y)\in A_{\epsilon}$ we must have $\wt{\delta}\leq d_X(x,y)$, i.e.,
114:  $\wt{\delta}$ is a lower bound of the set $\{d_X(x,y):(x,y)\in A_{\epsilon}\}$. Hence $\wt{\delta}\leq \delta(\epsilon)$, and in particular $\delta(\epsilon)$ is positive and it is the largest possible delta from the definition of uniform continuity. 
115: 
116: \end{enumerate}
117: \end{proof}
118: \begin{theorem}\label{thm:UniCont}
119: Let $(X,d_X)$ and $(Y,d_Y)$ be two metric spaces, and let $f:X\to Y$ be continuous. If $X$ is compact, then $f$ is 
120: uniformly continuous. 
121: \end{theorem}
122: \begin{proof}
123: The assertion is obvious if $f$ is constant, so from now on assume it is not. By Lemma \ref{lem:UniCont} it is sufficient to show that the function $\delta$ defined in \eqref{eq:delta} satisfies $\delta(\epsilon)>0$ for each $\epsilon\in (0,M)$. Assume to the contrary that there exists some $\epsilon\in (0,M)$ for which $\delta(\epsilon)=0$. Then 
124: $0=\delta(\epsilon)=\lim_{n\to \infty}d_X(x_n,y_n)$ for some sequence $((x_n,y_n))_n$ contained
125: in $A_{\epsilon}=\{(x,y)\in X^2: d_Y(f(x),f(y))\geq \epsilon\}$, and by passing to a convergent subsequence we find some $(x,y)\in X^2$ which is the limit of this subsequence and for which $d_X(x,y)=0$. 
126: %$0=d_X(x,y)$ for some $(x,y)\in X^2$ satisfying $x$. 
127: But $f$ is continuous, and so $d_Y(f(x),f(y))\geq \epsilon$, a contradiction to $x=y$.
128: \end{proof}
129: To the best of our knowledge, the issue of the ``optimal delta'' is not treated in the literature. %/calculus courses, although it is sometimes raised by curious teachers/students/readers. 
130: We note that $\delta(\epsilon)$ from \eqref{eq:delta} is somewhat dual, but definitely different, from the modulus of (uniform) continuity
131: \begin{equation*}
132: w(\delta)=\sup\{d_Y(f(x),f(y)): x,y\in X, \,d_X(x,y)\leq \delta\},
133: \end{equation*} 
134:  which assigns to a given $\delta\geq 0$ the smallest possible $\epsilon\geq 0$ from the definition of uniform continuity, when one allows  weak inequalities. However, $\delta(\epsilon)$ can be regarded as a modulus of its own. Local versions of this modulus (as a function of $x$ too) can be defined similarly. 
135: 
136: It is of some interest to compute explicitly $\delta(\epsilon)$ in some  simple cases. For example, let $X=[0,b],\,Y=\R$ and let  
137:   $f:X\to Y$ be defined by $f(x)=x^{\alpha}$, where $\alpha, b\in (0,\infty)$. Then
138: \begin{equation*}
139: \delta(\epsilon)=\left\{
140: \begin{array}{ll}
141:  {b-(b^{\alpha}-\epsilon)}^{1/\alpha} & \alpha\geq 1,\\
142: \epsilon^{1/\alpha} & 0<\alpha\leq 1.
143: \end{array}\right.
144: \end{equation*}
145: To derive this, one simply finds the minimum of $g(x,y)=|x-y|$, or equivalently of $h(x,y)=(x-y)^2$, on the set $A_{\epsilon}=\{(x,y)\in X^2: |f(x)-f(y)|\geq \epsilon\}$. This minimum is always attained on the boundary of $A_{\epsilon}$. It is interesting whether one can compute  $\delta(\epsilon)$ using the standard definition of uniform continuity. Note that here $\epsilon<M=b^{\alpha}$, so $\delta(\epsilon)$ is indeed well defined when $\alpha\geq 1$. In fact, the above formula shows that $f$ is uniformly continuous on $[0,\infty)$ for $0<\alpha\leq 1$, but cannot be uniformly continuous there if $\alpha>1$.
146: 
147: It is tempting to conjecture, and the above example supports this, that the optimal delta is a continuous function of $\epsilon$.  Unfortunately, in general this is not true:   the ``decreasing chainsaw'' function $f:[0,1]\to \R$ defined by $f(0)=0$ and (here $ n\in\N$) by
148: \begin{equation*}
149: f(t)=\left\{\begin{array}{ll}
150: \displaystyle{\frac{1}{n+1}-(2n+1)\left(t-\frac{1}{n+1}\right)},& t\in [1/(n+1),2/(2n+1)],\\
151:  \displaystyle{(2n+1)\left(t-\frac{2}{2n+1}\right)},  & t\in [2/(2n+1),1/n],
152: \end{array}\right.
153: \end{equation*}
154:  shows that in general $\delta$ may be discontinuous at infinitely many points (here   $\delta(1/n)=1/(n(2n+1))<1/(n(2n-1))\leq\delta(\epsilon)$ whenever $1/n<\epsilon$). However, it can be easily verified that $\delta$ is always lower semicontinuous. In addition, it is increasing, so it is continuous, and actually differentiable, almost everywhere. See \cite{Enayat,Guthrie} for a related discussion about the latter issue. 
155:  
156:  
157: 
158: \section{The Extreme Value Theorem}
159: \begin{theorem}\label{thm:Weierstrass}
160: Let $(I,d)$ be a compact metric space and let $f:I\to \mathbb{R}$ be  continuous. Then $f$ has both a minimum and a maximum on $X$.
161: \end{theorem}%
162:  We present two different proofs of Theorem \ref{thm:Weierstrass}. % of this theorem. %The first proof we find
163: %a maximal point using an auxiliary ``envelope'' function $g$.
164: The first one is only for the special case where $I$ is a compact interval. It is based on a certain ``envelope'' function, and uses the costume argument of proving first that $\sup_{x\in I}f(x)$ and $\inf_{x\in I}f(x)$ are finite, and then proving that they are attained; this argument appears
165: in almost all the proofs we know, including the topological one (see, e.g., \cite[p. 89]{Rudin}). The only exceptional proof is the one of Fort \cite{Fort}. 
166: The second proof is for the general case, and it does not use the above  argument. It is in the spirit of programming, and hence we called it a 
167: ``programmer's proof''. The proof is significantly different from that of Fort, but can be thought as a dual to his one, in the sense that in his proof one proves the existence of the extreme value by starting from ``above'' (the whole space) and going ``downward'' (to smaller subsets), and in our case we start from ``below'' (a finite subset) and go ``upward'' (to a dense subset).  
168: \begin{proof}[{\bf Proof 1: An ``envelope proof''}]
169: The case where $I$ is a singleton is  obvious, so from now on assume that $I$ contains at least two points. The proof consists of two steps.\\%\textbf{first proof}\vspace{0.1cm}\\
170: \textbf{Step 1:}  We show that $f$ is bounded using the ``real induction'' argument. 
171: This part of the proof is not really new (it is a modification of \cite[p. 135]{Spivak}), but it is included for the sake of completeness. Let
172: % which is similar to induction. Let
173: \begin{equation*}
174: A=\{x\in I: f\,\, \textrm{is bounded on}\,\, [a,x]\}.
175: \end{equation*}
176: $A$ is nonempty because $a\in A$. Since $f$ is continuous, each $x\in X$ has
177:  a neighborhood in which $f$ is bounded. Hence $A$ has the property that if $x\in A$,
178: then also $[x,x+\delta]\cap I\subseteq A$ for some $\delta>0$, and in particular
179: $[a,a+\delta]\subset A$ for some $\delta>0$. Let $s=\sup A$. Because $f$ is continuous at $s$, there are $M_1,\delta\in (0,\infty)$
180: such that $|f(t)|\leq M_1$ for all $t\in (s-\delta,s+\delta)\cap I$. By the definition
181: of $s$ there exists $x\in A\cap (s-\delta,s)$, and by the definition of $A$ we know
182: that $\sup_{t\in [a,x]}|f(t)|\leq M_2$ for some $M_2\in (0,\infty)$. Hence $f$ is bounded on $[a,s+\delta)\cap I$ by $M_1+M_2$, so in particular $s\in A$. But now, by the property of $A$ described above, it must be that $s=b$, otherwise
183: $s$ is not the supremum of $A$. Thus $A=I$, and this establishes the first step.
184:   \vspace{0.2cm}\\
185: \textbf{Step 2:} 
186:  We now find explicitly a point $x_0\in I$ at which $f$ attains a maximal value. $x_0$ is in fact the largest possible such point; a similar argument
187:  can be used for showing that $f$ attains a minimal value. Let
188: \begin{equation*}
189: g(x)=\sup\{f(t): t\in [a,x]\},\quad \forall x\in I.
190: \end{equation*} By the first step the ``envelope'' function $g$
191: is well defined (and actually continuous, but we will not use this fact), and it is obviously
192: increasing, so it has a maximum at $b$. Let
193: \begin{equation*}
194: C=\{x\in I: g(x)=g(b)\},\quad x_0=\inf C.
195: \end{equation*}
196: We finish by showing that
197: $f(x_0)=g(b)=\sup_{x\in I}f(x)$. Let $\epsilon>0$. By the continuity
198: of $f$ there exists $\delta>0$ such that $f(x)<f(x_0)+\epsilon$ for each 
199: $x$ in the intersection $I\cap [x_0-\delta,x_0+\delta]$.
200: By the definition of $x_0$ there exists $x\in C\cap [x_0,x_0+\delta]$. If $x_0=a$, then
201: $g(b)=g(x)\leq f(x_0)+\epsilon$. Otherwise, assume $\delta<\min(x_0-a,b-a)$. We have
202: \begin{multline*}\label{eq:g}
203: g(b)=g(x)\leq \max(\sup\{f(t): t\in [a,x_0-\delta]\},\sup\{f(t): t\in [x_0-\delta,x_0+\delta]\})
204: \\\leq \max(g(x_0-\delta),f(x_0)+\epsilon).
205: \end{multline*}
206: If $g(b)\leq g(x_0-\delta)$, then there is equality
207: because $g$ is increasing, so we obtain a contradiction to the
208: definition of $x_0$. Hence also $a<x_0$ implies $g(b)\leq f(x_0)+\epsilon$.
209: Since $\epsilon$ was arbitrary, we conclude that $f(x_0)=g(b)$.
210: \end{proof}
211: \begin{proof}[{\bf Proof 2: A ``programmer's proof''}]
212: Consider an increasing sequence $(E_n)_n$ of finite subsets of $I$ such that $\overline{\bigcup_{n=1}^{\infty}E_n}=I$. If $I$ is a compact interval $[a,b]$, then we can take 
213: \begin{equation*}\label{eq:P_kDef}
214: P_k=a+\frac{(b-a)k}{2^n}, \quad\quad E_n=\{P_k: k=0,1,\ldots,
215: 2^n\}
216: \end{equation*}
217: for each $n\in \mathbb{N}\cup\{0\}$ and each $k=0,1,\ldots, 2^n$. In general, we can define $E_{n+1}=E_n\cup F_n,\,E_0=F_0$, where $F_n$ is the set of centers of the balls of a given $2^{-n}$-net of $I$. 
218: Let
219: \begin{equation*}
220: \begin{array}{c}
221: M_n=\max\{f(x):x\in E_n\},\\
222: m_n=\min\{f(x):x\in E_n\}.
223:  \end{array}
224: \end{equation*}
225: The function $f$ attains a maximum on the finite set $E_n$, i.e., there exists $x_n\in
226: E_n$ such that $f(x_n)=M_n$. Let $(x_{n_j})_j$ be any convergent subsequence of $(x_n)_{n}$ and let
227:  $x_{\infty}=\lim_{j\to \infty}x_{n_j}$. Then $x_{\infty}\in I$, and because
228:  $f$ is continuous,
229: \begin{equation*}
230: f(x_{\infty})=\lim_{j\to \infty}f(x_{n_j})=\lim_{j\to
231: \infty}M_{n_j}.
232: \end{equation*}
233: Actually, the whole sequence $(M_n)_{n=0}^{\infty}$ converges to
234: $f(x_{\infty})$, since it is an increasing sequence with a convergent
235: subsequence. 
236: We now show that $f$ attains its maximal value at
237: $x_{\infty}$, i.e., $f(x)\leq f(x_{\infty})$ for all $x\in I$.
238: Let $x\in I$ and let $\epsilon>0$. Since $f$ is continuous on $I$, it
239: is continuous at $x$, so there exists $\delta>0$ such that if $y\in I$
240: satisfies $d(y,x)<\delta$, then $|f(x)-f(y)|<\epsilon$. By the construction of the sequence $(E_n)_n$, for $n$ large enough there exists $t_n\in E_n$ such  that $d(t_n,x)<\delta$. Therefore
241: \begin{equation*}
242: f(x)\leq f(t_n)+\epsilon\leq
243: M_{n}+\epsilon_{\,\,\overrightarrow{n\to \infty}}\,\,
244: f(x_{\infty})+\epsilon.
245: \end{equation*}
246: But $\epsilon$ was arbitrary, so $f(x)\leq f(x_{\infty})$, and since
247: $x$ was arbitrary this means that $f$ has a maximum at $x_{\infty}$.
248: By the same way $f$ has a minimum on $I$.
249: \end{proof}
250: 
251: 
252: \section{The intermediate value theorem }
253: \begin{theorem}\label{thm:ClassicInter}
254: Let $I=[a,b]\subset \R$. If $f:I\to \mathbb{R}$ is continuous and if $c\in
255: \mathbb{R}$ is between $f(a)$ and $f(b)$, then there exists $x\in I$
256: such that $f(x)=c$.
257: \end{theorem}
258: Theorem \ref{thm:ClassicInter} is a consequence of the following
259: more general theorem (Theorem ~\ref{thm:NonClassicInter}) which  generalizes the intermediate value
260: theorem to a class of discontinuous functions, and also re-examines the meaning of the intermediate value. Before stating it, recall that a topological space $X$ is called connected if it 
261: cannot be represented as $X=A\cup B$, where $A$ and $B$ are two nonempty,
262: disjoint and open sets in $X$. A simple consequence of this definition and the completeness axiom is that every interval in $\mathbb{R}$ is a connected  space. 
263: 
264: For a subset $D$ of $X$ we denote by $\Int(D)$, $\partial D$ and $\Ext(D)=X\backslash (D\cup \partial D)=\Int(X\backslash D)$ its interior, boundary and exterior respectively. We will use the following terminology.
265: \begin{definition}\label{def:OpenContinuous}
266: Let $X,Y$ be two topological spaces. A function $f:X\to Y$ is said to be continuous with
267: respect to $D\subseteq Y$ if $f^{-1}(D)$ is an open set in $X$.
268: \end{definition}
269: For instance, $f:X\to Y$ is continuous if and only if it is continuous with respect to all open subsets of $Y$, and $f:X\to \R$ is lower semicontinuous if and only if it is continuous with respect to all the intervals of the form $(a,\infty)$.
270: \begin{theorem}\label{thm:NonClassicInter}
271: Let $X$ be a connected topological space and let $Y$ be a
272: topological space. Let $D\subseteq Y$.
273: If $f:X\to Y$ is continuous with respect to both $\Int(D)$ and $\Ext(D)$, and if there are $a,b\in X$ such
274: that $f(a)\in D$ and $f(b)\notin D$, then there exists $x\in X$ such
275: that $f(x)\in \partial D$.
276: In particular this is true if $f$ is continuous.
277: \end{theorem}
278: \begin{proof}
279: If $f(a)\in \partial D$ or $f(b)\in \partial D$, then the proof is
280: complete. Otherwise, since $f(a)\in D$ and $f(b)\notin D$, it follows
281: that $f(a)\in \Int(D)$ and $f(b)\in \Ext(D)$, so $f^{-1}(\Int(D))$ and
282: $f^{-1}(\Ext(D))$ are nonempty sets and they are open by our
283: assumption. Now, since
284: \begin{multline*}
285: X=f^{-1}(Y)=f^{-1}(\Int(D)\cup \partial D \cup
286: \Ext(D))=\\f^{-1}(\Int(D))\cup f^{-1}(\partial D)\cup
287: f^{-1}(\Ext(D)),
288: \end{multline*}
289: it follows that if $f^{-1}(\partial D)$ is empty, then $X$ is a union
290: of two open, disjoint and nonempty sets and this contradicts the
291: assumption that $X$ is connected. Hence $f^{-1}(\partial D)$ is
292: nonempty, i.e., there exists $x\in X$ such that $f(x)\in \partial D$.
293: \end{proof}
294: \emph{Proof of Theorem \ref{thm:ClassicInter}}: Denote
295: $D=(-\infty,c)$. Without loss of generality we can assume that $f(a)<c<f(b)$. Hence
296: $f(a)\in \Int(D)=D$, $f(b)\in \Ext(D)=(c,\infty)$ and
297: $f^{-1}(\Int(D)),f^{-1}(\Ext(D))$ are open because $f$ is continuous.
298: Since $I$ is connected,  by Theorem \ref{thm:NonClassicInter} there
299: is $x\in I$ such that $f(x)\in
300: \partial D=\{c\}$, i.e., $f(x)=c$. %$\Box$% \\
301: \begin{example}
302: Let $f:\R\to\R$ be defined  by $f(x)=x$ when $x$ is irrational, $f(x)=2x$ when $x\in\Q\backslash \{1/n: n\in \N\}$ and $f(1/n)=1,n\in \N$.  Obviously $f^{-1}((0,\infty))$ and $f^{-1}((-\infty,0))$ are open sets, so the conditions of Theorem \ref{thm:NonClassicInter} are satisfied and indeed  $f(x)\in \partial (0,\infty)$ for $x=0$. But $f$  is discontinuous at every point. This shows that the type of continuity expressed in Definition~\ref{def:OpenContinuous} is a very weak one. It would be interesting to find more useful examples.
303: \end{example}
304: \begin{example}
305: Here we show how Theorem  \ref{thm:NonClassicInter} can be used for deriving a fixed point theorem for (very) discontinuous functions, in contrast to the majority of fixed point theorems which are for continuous ones  \cite{GranasDugondji}. Let $I=[a,b]$ and let $f:I\to I$. Suppose that $g(x):=f(x)-x$ has the property that both  $g^{-1}((0,\infty))$ and $g^{-1}((-\infty,0))$ are open sets  in $I$. Then $f$ has a fixed point in $I$. Indeed, if $f(a)\neq a$ and $f(b)\neq b$, then %the usual trick of defining $g(x)=f(x)-x$ works here too, 
306:  $g(a)\in (0,\infty),\, g(b)\in (-\infty,0)$ and thus  Theorem  \ref{thm:NonClassicInter} implies that $g(t)=0$ for some $t\in I$, i.e., $t$ is a fixed point of $f$.   
307: % both $g^{-1}((0,\infty))$ and $g^{-1}((-\infty,0))$ are open sets  in $I$ as a simple check shows. 
308: \end{example}
309: \begin{remark} There is another theorem which generalizes the
310: intermediate value theorem \cite[p. 93]{Rudin}.
311: This theorem says that the image of a connected topological space by
312: a continuous function is a connected topological space.%\cite{s]).
313: 
314: Both Theorem \ref{thm:NonClassicInter} and the above theorem
315: generalize the classical intermediate value theorem. However, there are two main
316: differences between them. First, in Theorem \ref{thm:NonClassicInter}
317: the function $f$ is not necessarily continuous, but rather satisfies a mild condition of
318: continuity. Second, the
319: intermediate value property is expressed differently in both cases:
320: in the theorem mentioned above it is expressed in the connectivity of
321: $f(X)$, while in Theorem \ref{thm:NonClassicInter} it is expressed
322: in the fact that if $f$ passes through both $D$ and its complement $Y\backslash D$,
323: then it also passes through the boundary $\partial D$, which can be thought of as being an
324: intermediate set between $D$ and $Y\backslash D$ (or between $\Int(D)$ and $\Ext(D)$).
325: \end{remark}
326: \section{Concluding remarks}
327: The proofs given above raise some questions regarding the sense in which
328:  they are constructive.
329: 
330: In the proof of Lemma \ref{lem:UniCont}, for any $\epsilon>0$ the corresponding $\delta(\epsilon)$ was found explicitly. However, the proof presented in Theorem \ref{thm:UniCont} that $\delta(\epsilon)>0$ for each $\epsilon>0$ when the space is compact is  based on nonconstructive arguments.
331: 
332: The first proof of Theorem \ref{thm:Weierstrass} gives a representation for
333: the largest possible maximal point $x_0$, but this representation is too vague to be considered as constructive.
334: The second proof is somewhat constructive in the practical sense,
335: since the monotone sequence $(M_n)_n$ which converges to $\max_{x\in
336: I} f(x)$ can  be computed easily and explicitly (but slowly), at least when $I$ is a compact interval. The
337: sequence $(x_n)_n$ of points for which $f(x_n)=M_n$ can also be
338: easily computed.
339: 
340: On the other hand, one can argue against the sense in which this proof is constructive. First, no error estimates were given for the convergence of  $(M_n)_n$. Second, the point $x_{\infty}$ at which $f$ attains its maximal value usually cannot be found in a constructive manner, because it is a limit of a convergent subsequence, which
341: usually cannot be computed explicitly in advance (recall that the
342: existence of such subsequence follows from an infinite version of
343: the Dirichlet pigeonhole principle, so it is highly
344: nonconstructive).
345: 
346: Nevertheless, if some additional information is given about $f$, and hence about $(x_n)_n$, then we can say more about $x_{\infty}$. For instance, if it is known that $x_{\infty}$ is
347: unique, as it is the case when $f$ is strictly concave, then the proof shows that the whole sequence $(x_n)_n$ converges to $x_{\infty}$.
348: 
349: The proof of Theorem \ref{thm:NonClassicInter} can be regarded as a
350: pure existence proof, i.e., a proof without any single constructive
351: clue. However, if some additional information is known about the
352: connected space $X$, then by a repeated application of Theorem~
353: \ref{thm:NonClassicInter} one \textbf{can} compute explicitly with error estimates an
354: ``intermediate'' point $x\in X$ for which $f(x)\in \partial D$.
355: 
356: For example, Let $X,Y,$ $D\subseteq Y$ and $f:X\to Y$ satisfy the
357: conditions of Theorem \ref{thm:NonClassicInter}, where
358: $X=[a,b]\subset \mathbb{R}, f(a)\in
359: D$ and $f(b)\notin D$. Theorem ~\ref{thm:NonClassicInter} ensures that
360: there exists $x_0\in X$ such that $f(x_0)\in \partial D$. Now,
361:  for the point $P_1=a+(b-a)/2$, either
362: $f(P_1)\in D$ or $f(P_1)\notin D$. Hence there are $a_1,b_1\in \{a,P_1,b\}$ such that
363: $f(a_1)\in D,\,f(b_1)\notin D$ and $[a_1,b_1]\subset X$. Since
364: $f^{-1}(\Int(D))\cap [a_1,b_1]$ and $f^{-1}(\Ext(D))\cap[a_1,b_1]$ are
365: open sets in the connected space $[a_1,b_1]$, by
366: Theorem ~\ref{thm:NonClassicInter} there exists $x_1\in [a_1,b_1]$ such
367: that $x_1\in \partial D$, so one has a better estimate for an intermediate point.
368: Continuing in this way, one essentially gets the bisection method
369: and finds an intermediate point to within an error of $(b-a)\cdot 2^{-n}$ in the $n$-th step.\\\vspace{0.02cm}
370: 
371: {\noindent }\textbf{Acknowledgments}\vspace{0.1cm}\\ I thank Gregory Shapiro and Zbigniew H. Nitecki for helpful discussion.
372: %\newpage 
373: \bibliographystyle{amsplain}
374: \bibliography{biblio}
375: \end{document}
376: 
377: