From 4a0d004d0898fd353e7adc8611c8da93f8f9704f Mon Sep 17 00:00:00 2001 From: Joshua Potter Date: Thu, 18 May 2023 12:39:36 -0600 Subject: [PATCH] Separate concept of a glossary further. Restructure fields in Apostol. --- Bookshelf/Apostol.tex | 784 ++++++++++++++++++----------------- Bookshelf/Enderton_Logic.tex | 15 +- 2 files changed, 408 insertions(+), 391 deletions(-) diff --git a/Bookshelf/Apostol.tex b/Bookshelf/Apostol.tex index 4e6a13f..c4c4176 100644 --- a/Bookshelf/Apostol.tex +++ b/Bookshelf/Apostol.tex @@ -14,11 +14,14 @@ \tableofcontents +\begingroup +\renewcommand\thechapter{G} + \chapter{Glossary}% \label{chap:glossary} \section{\defined{Characteristic Function}}% -\label{sec:def-characteristic-function} +\label{def:characteristic-function} Let $S$ be a set of points on the real line. The \textbf{characteristic function} of $S$ is the function $\mathcal{X}_S$ such @@ -32,7 +35,7 @@ The \textbf{characteristic function} of $S$ is the function $\mathcal{X}_S$ such \end{definition} \section{\defined{Infimum}}% -\label{sec:def-infimum} +\label{def:infimum} A number $B$ is called an \textbf{infimum} of a nonempty set $S$ if $B$ has the following two properties: @@ -49,36 +52,36 @@ Such a number $B$ is also known as the \textbf{greatest lower bound}. \end{definition} \section{\partial{Integrable}}% -\label{sec:def-integrable} +\label{def:integrable} Let $f$ be a function defined and bounded on $[a, b]$. $f$ is said to be \textbf{integrable} if there exists one and only one number - $I$ such that \eqref{sec:def-integral-bounded-function-eq2} holds. + $I$ such that \eqref{def:integral-bounded-function-eq2} holds. If $f$ is integrable on $[a, b]$, we say that the integral $\int_a^b f(x) \mathop{dx}$ \textbf{exists}. \section{\partial{Integral of a Bounded Function}}% -\label{sec:def-integral-bounded-function} +\label{def:integral-bounded-function} Let $f$ be a function defined and bounded on $[a, b]$. Let $s$ and $t$ denote arbitrary step functions defined on $[a, b]$ such that \begin{equation} - \label{sec:def-integral-bounded-function-eq1} + \label{def:integral-bounded-function-eq1} s(x) \leq f(x) \leq t(x) \end{equation} for every $x$ in $[a, b]$. If there is one and only one number $I$ such that \begin{equation} - \label{sec:def-integral-bounded-function-eq2} + \label{def:integral-bounded-function-eq2} \int_a^b s(x) \mathop{dx} \leq I \leq \int_a^b t(x) \mathop{dx} \end{equation} for every pair of step functions $s$ and $t$ satisfying - \eqref{sec:def-integral-bounded-function-eq1}, then this number $I$ is called + \eqref{def:integral-bounded-function-eq1}, then this number $I$ is called the \textbf{integral of $f$ from $a$ to $b$}, and is denoted by the symbol $\int_a^b f(x) \mathop{dx}$ or by $\int_a^b f$. If $a < b$, we define $\int_b^a f(x) \mathop{dx} = -\int_a^b f(x) \mathop{dx}$, - provided $f$ is \nameref{sec:def-integrable} on $[a, b]$. + provided $f$ is \nameref{def:integrable} on $[a, b]$. We also define $\int_a^a f(x) \mathop{dx} = 0$. The function $f$ is called the \textbf{integrand}, the numbers $a$ and $b$ are @@ -86,10 +89,10 @@ The function $f$ is called the \textbf{integrand}, the numbers $a$ and $b$ are \textbf{interval of integration}. \section{\partial{Integral of a Step Function}}% -\label{sec:def-integral-step-function} +\label{def:integral-step-function} -Let $s$ be a \nameref{sec:def-step-function} defined on $[a, b]$, and let - $P = \{x_0, x_1, \ldots, x_n\}$ be a \nameref{sec:def-partition} of $[a, b]$ +Let $s$ be a \nameref{def:step-function} defined on $[a, b]$, and let + $P = \{x_0, x_1, \ldots, x_n\}$ be a \nameref{def:partition} of $[a, b]$ such that $s$ is constant on the open subintervals of $P$. Denote by $s_k$ the constant value that $s$ takes in the $k$th open subinterval of $P$, so that @@ -102,17 +105,17 @@ If $a < b$, we define $\int_b^a s(x) \mathop{dx} = -\int_a^b s(x) \mathop{dx}$. We also define $\int_a^a s(x) \mathop{dx} = 0$. \section{\partial{Lower Integral}}% -\label{sec:def-lower-integral} +\label{def:lower-integral} Let $f$ be a function bounded on $[a, b]$ and $S$ denote the set of numbers $\int_a^b s(x) \mathop{dx}$ obtained as $s$ runs through all - \nameref{sec:def-step-function}s below $f$. + \nameref{def:step-function}s below $f$. That is, let $$S = \left\{ \int_a^b s(x) \mathop{dx} : s \leq f \right\}.$$ The number $\sup{S}$ is called the \textbf{lower integral of $f$}. It is denoted as $\ubar{I}(f)$. \section{\partial{Monotonic}}% -\label{sec:def-monotonic} +\label{def:monotonic} A function $f$ is called \textbf{monotonic} on set $S$ if it is increasing on $S$ or if it is decreasing on $S$. @@ -122,11 +125,11 @@ $f$ is said to be \textbf{strictly monotonic} if it is strictly increasing on A function $f$ is said to be \textbf{piecewise monotonic} on an interval if its graph consists of a finite number of monotonic pieces. In other words, $f$ is piecewise monotonic on $[a, b]$ if there is a - \nameref{sec:def-partition} of $[a, b]$ such that $f$ is monotonic on each of + \nameref{def:partition} of $[a, b]$ such that $f$ is monotonic on each of the open subintervals of $P$. \section{\defined{Partition}}% -\label{sec:def-partition} +\label{def:partition} Let $[a, b]$ be a closed interval decomposed into $n$ subintervals by inserting $n - 1$ points of subdivision, say $x_1$, $x_2$, $\ldots$, $x_{n-1}$, subject @@ -148,9 +151,9 @@ A collection of points satisfying \eqref{sec:partition-eq1} is called a \end{definition} \section{\partial{Refinement}}% -\label{sec:def-refinement} +\label{def:refinement} -Let $P$ be a \nameref{sec:def-partition} of closed interval $[a, b]$. +Let $P$ be a \nameref{def:partition} of closed interval $[a, b]$. A \textbf{refinement} $P'$ of $P$ is a partition formed by adjoining more subdivision points to those already in $P$. @@ -158,10 +161,10 @@ $P'$ is said to be \textbf{finer than} $P$. The union of two partitions $P_1$ and $P_2$ is called the \textbf{common refinement} of $P_1$ and $P_2$. \section{\defined{Step Function}}% -\label{sec:def-step-function} +\label{def:step-function} A function $s$, whose domain is a closed interval $[a, b]$, is called a - \textbf{step function} if there is a \nameref{sec:def-partition} + \textbf{step function} if there is a \nameref{def:partition} $P = \{x_0, x_1, \ldots, x_n\}$ of $[a, b]$ such that $s$ is constant on each open subinterval of $P$. That is to say, for each $k = 1, 2, \ldots, n$, there is a real number $s_k$ @@ -180,7 +183,7 @@ That is to say, for each $k = 1, 2, \ldots, n$, there is a real number $s_k$ \end{definition} \section{\defined{Supremum}}% -\label{sec:def-supremum} +\label{def:supremum} A number $B$ is called a \textbf{supremum} of a nonempty set $S$ if $B$ has the following two properties: @@ -197,15 +200,17 @@ Such a number $B$ is also known as the \textbf{least upper bound}. \end{definition} \section{\partial{Upper Integral}}% -\label{sec:def-upper-integral} +\label{def:upper-integral} Let $f$ be a function bounded on $[a, b]$ and $T$ denote the set of numbers $\int_a^b t(x) \mathop{dx}$ obtained as $t$ runs through all - \nameref{sec:def-step-function}s above $f$. + \nameref{def:step-function}s above $f$. That is, let $$T = \left\{ \int_a^b t(x) \mathop{dx} : f \leq t \right\}.$$ The number $\inf{T}$ is called the \textbf{upper integral of $f$}. It is denoted as $\bar{I}(f)$. +\endgroup + \chapter{A Set of Axioms for the Real-Number System}% \label{chap:set-axioms-real-number-system} @@ -236,7 +241,7 @@ Every nonempty set $S$ of real numbers which is bounded above has a supremum; {Apostol.Chapter\_I\_03.is\_lub\_neg\_set\_iff\_is\_glb\_set\_neg} Suppose $L = \sup{S}$ and fix $x \in S$. - By definition of the \nameref{sec:def-supremum}, $x \leq L$ and $L$ is the + By definition of the \nameref{def:supremum}, $x \leq L$ and $L$ is the smallest value satisfying this inequality. Negating both sides of the inequality yields $-x \geq -L$. Furthermore, $-L$ must be the largest value satisfying this inequality. @@ -244,7 +249,8 @@ Every nonempty set $S$ of real numbers which is bounded above has a supremum; \end{proof} -\section{\verified{Theorem I.27}}% +\section{\verified{Existence of a Greatest Lower Bound}} +\label{sec:existence-greatest-lower-bound} \label{sec:theorem-i.27} \begin{theorem}[I.27] @@ -262,13 +268,14 @@ Every nonempty set $S$ of real numbers which is bounded above has a supremum; Let $S$ be a nonempty set bounded below by $x$. Then $-S$ is nonempty and bounded above by $x$. By the \nameref{sec:completeness-axiom}, there exists a - \nameref{sec:def-supremum} $L$ of $-S$. + \nameref{def:supremum} $L$ of $-S$. By \nameref{sec:lemma-1}, $L$ is a supremum of $-S$ if and only if $-L$ is an infimum of $S$. \end{proof} -\section{\verified{Theorem I.29}}% +\section{\verified{Positive Integers Unbounded Above}}% +\label{sec:positive-integers-unbounded-above} \label{sec:theorem-i.29} \begin{theorem}[I.29] @@ -407,7 +414,7 @@ Every nonempty set $S$ of real numbers which is bounded above has a supremum; \end{proof} -\section{Theorem I.32}% +\section{\verified{Theorem I.32}}% \label{sec:theorem-i.32} Let $h$ be a given positive number and let $S$ be a set of real numbers. @@ -426,7 +433,7 @@ Let $h$ be a given positive number and let $S$ be a set of real numbers. \lean{Bookshelf/Apostol/Chapter\_I\_03} {Apostol.Chapter\_I\_03.sup\_imp\_exists\_gt\_sup\_sub\_delta} - By definition of a \nameref{sec:def-supremum}, $\sup{S}$ is the least upper + By definition of a \nameref{def:supremum}, $\sup{S}$ is the least upper bound of $S$. For the sake of contradiction, suppose for all $x \in S$, $x \leq \sup{S} - h$. @@ -452,7 +459,7 @@ Let $h$ be a given positive number and let $S$ be a set of real numbers. \lean{Bookshelf/Apostol/Chapter\_I\_03} {Apostol.Chapter\_I\_03.inf\_imp\_exists\_lt\_inf\_add\_delta} - By definition of an \nameref{sec:def-infimum}, $\inf{S}$ is the greatest lower + By definition of an \nameref{def:infimum}, $\inf{S}$ is the greatest lower bound of $S$. For the sake of contradiction, suppose for all $x \in S$, $x \geq \inf{S} + h$. @@ -464,7 +471,8 @@ Let $h$ be a given positive number and let $S$ be a set of real numbers. \end{proof} -\section{Theorem I.33}% +\section{\verified{Additive Property of Supremums and Infimums}}% +\label{sec:additive-property-supremums-infimums} \label{sec:theorem-i.33} Given nonempty subsets $A$ and $B$ of $\mathbb{R}$, let $C$ denote the set @@ -496,7 +504,7 @@ Given nonempty subsets $A$ and $B$ of $\mathbb{R}$, let $C$ denote the set Let $x \in C$. By definition of $C$, there exist elements $a' \in A$ and $b' \in B$ such that $x = a' + b'$. - By definition of a \nameref{sec:def-supremum}, $a' \leq \sup{A}$. + By definition of a \nameref{def:supremum}, $a' \leq \sup{A}$. Likewise, $b' \leq \sup{B}$. Therefore $a' + b' \leq \sup{A} + \sup{B}$. Since $x = a' + b'$ was arbitrarily chosen, it follows $\sup{A} + \sup{B}$ @@ -567,7 +575,7 @@ Given nonempty subsets $A$ and $B$ of $\mathbb{R}$, let $C$ denote the set Let $x \in C$. By definition of $C$, there exist elements $a' \in A$ and $b' \in B$ such that $x = a' + b'$. - By definition of an \nameref{sec:def-infimum}, $a' \geq \inf{A}$. + By definition of an \nameref{def:infimum}, $a' \geq \inf{A}$. Likewise, $b' \geq \inf{B}$. Therefore $a' + b' \geq \inf{A} + \inf{B}$. Since $x = a' + b'$ was arbitrarily chosen, it follows $\inf{A} + \inf{B}$ @@ -649,15 +657,18 @@ Given nonempty subsets $A$ and $B$ of $\mathbb{R}$, let $C$ denote the set \end{proof} -\chapter{The Concept of Area as a Set Function}% -\label{chap:concept-area-set-function} +\chapter{The Concepts of Integral Calculus}% +\label{chap:concepts-integral-calculus} + +\section{The Concept of Area as a Set Function}% +\label{sec:concept-area-set-function} We assume there exists a class $\mathscr{M}$ of measurable sets in the plane and a set function $a$, whose domain is $\mathscr{M}$, with the following properties: -\section{\defined{Nonnegative Property}}% -\label{sec:nonnegative-property} +\subsection{\defined{Nonnegative Property}}% +\label{sub:nonnegative-property} For each set $S$ in $\mathscr{M}$, we have $a(S) \geq 0$. @@ -668,8 +679,8 @@ For each set $S$ in $\mathscr{M}$, we have $a(S) \geq 0$. \end{axiom} -\section{\defined{Additive Property}}% -\label{sec:area-additive-property} +\subsection{\defined{Additive Property}}% +\label{sub:area-additive-property} If $S$ and $T$ are in $\mathscr{M}$, then $S \cup T$ and $S \cap T$ are in $\mathscr{M}$, and we have $a(S \cup T) = a(S) + a(T) - a(S \cap T)$. @@ -681,8 +692,8 @@ If $S$ and $T$ are in $\mathscr{M}$, then $S \cup T$ and $S \cap T$ are in \end{axiom} -\section{\defined{Difference Property}}% -\label{sec:area-difference-property} +\subsection{\defined{Difference Property}}% +\label{sub:area-difference-property} If $S$ and $T$ are in $\mathscr{M}$ with $S \subseteq T$, then $T - S$ is in $\mathscr{M}$, and we have $a(T - S) = a(T) - a(S)$. @@ -694,8 +705,8 @@ If $S$ and $T$ are in $\mathscr{M}$ with $S \subseteq T$, then $T - S$ is in \end{axiom} -\section{\defined{Invariance Under Congruence}}% -\label{sec:area-invariance-under-congruence} +\subsection{\defined{Invariance Under Congruence}}% +\label{sub:area-invariance-under-congruence} If a set $S$ is in $\mathscr{M}$ and if $T$ is congruent to $S$, then $T$ is also in $\mathscr{M}$ and we have $a(S) = a(T)$. @@ -707,8 +718,8 @@ If a set $S$ is in $\mathscr{M}$ and if $T$ is congruent to $S$, then $T$ is \end{axiom} -\section{\defined{Choice of Scale}}% -\label{sec:area-choice-scale} +\subsection{\defined{Choice of Scale}}% +\label{sub:area-choice-scale} Every rectangle $R$ is in $\mathscr{M}$. If the edges of $R$ have lengths $h$ and $k$, then $a(R) = hk$. @@ -720,18 +731,18 @@ If the edges of $R$ have lengths $h$ and $k$, then $a(R) = hk$. \end{axiom} -\section{\partial{Exhaustion Property}}% -\label{sec:area-exhaustion-property} +\subsection{\partial{Exhaustion Property}}% +\label{sub:area-exhaustion-property} Let $Q$ be a set that can be enclosed between two step regions $S$ and $T$, so that \begin{equation} - \label{sec:exhaustion-property-eq1} + \label{sub:exhaustion-property-eq1} S \subseteq Q \subseteq T. \end{equation} If there is one and only one number $c$ which satisfies the inequalities $$a(S) \leq c \leq a(T)$$ for all step regions $S$ and $T$ satisfying - \eqref{sec:exhaustion-property-eq1}, then $Q$ is measurable and $a(Q) = c$. + \eqref{sub:exhaustion-property-eq1}, then $Q$ is measurable and $a(Q) = c$. \begin{axiom} @@ -740,16 +751,16 @@ If there is one and only one number $c$ which satisfies the inequalities \end{axiom} -\chapter{Exercises 1.7}% -\label{chap:exercises-1.7} +\section{Exercises 1.7}% +\label{sec:exercises-1.7} -\section{Exercise 1.7.1}% -\label{sec:exercise-1.7.1} +\subsection{\partial{Exercise 1.7.1}}% +\label{sub:exercise-1.7.1} Prove that each of the following sets is measurable and has zero area: -\subsection{\partial{Exercise 1.7.1a}}% -\label{sub:exercise-1.7.1a} +\subsubsection{\partial{Exercise 1.7.1a}}% +\label{ssub:exercise-1.7.1a} A set consisting of a single point. @@ -757,15 +768,15 @@ A set consisting of a single point. Let $S$ be a set consisting of a single point. By definition of a point, $S$ is a rectangle in which all vertices coincide. - By \nameref{sec:area-choice-scale}, $S$ is measurable with area its width + By \nameref{sub:area-choice-scale}, $S$ is measurable with area its width times its height. The width and height of $S$ is trivially zero. Therefore $a(S) = (0)(0) = 0$. \end{proof} -\subsection{\partial{Exercise 1.7.1b}}% -\label{sub:exercise-1.7.1b} +\subsubsection{\partial{Exercise 1.7.1b}}% +\label{ssub:exercise-1.7.1b} A set consisting of a finite number of points in a plane. @@ -778,7 +789,7 @@ A set consisting of a finite number of points in a plane. \paragraph{Base Case}% Consider a set $S$ consisting of a single point in a plane. - By \nameref{sub:exercise-1.7.1a}, $S$ is measurable with area $0$. + By \nameref{ssub:exercise-1.7.1a}, $S$ is measurable with area $0$. Thus $P(1)$ holds. \paragraph{Induction Step}% @@ -790,14 +801,14 @@ A set consisting of a finite number of points in a plane. Denote the remaining set of points as $S_k$. By construction, $S_{k+1} = S_k \cup T$. By the induction hypothesis, $S_k$ is measurable with area $0$. - By \nameref{sub:exercise-1.7.1a}, $T$ is measurable with area $0$. - By the \nameref{sec:area-additive-property}, $S_k \cup T$ is + By \nameref{ssub:exercise-1.7.1a}, $T$ is measurable with area $0$. + By the \nameref{sub:area-additive-property}, $S_k \cup T$ is measurable, $S_k \cap T$ is measurable, and \begin{align} a(S_{k+1}) & = a(S_k \cup T) \nonumber \\ & = a(S_k) + a(T) - a(S_k \cap T) \nonumber \\ - & = 0 + 0 - a(S_k \cap T). \label{sub:exercise-1.7.1b-eq1} + & = 0 + 0 - a(S_k \cap T). \label{ssub:exercise-1.7.1b-eq1} \end{align} There are two cases to consider: @@ -810,11 +821,11 @@ A set consisting of a finite number of points in a plane. $S_k \cap T \neq \emptyset$. Since $T$ consists of a single point, $S_k \cap T = T$. - By \nameref{sub:exercise-1.7.1a}, $a(S_k \cap T) = a(T) = 0$. + By \nameref{ssub:exercise-1.7.1a}, $a(S_k \cap T) = a(T) = 0$. \vspace{8pt} \noindent - In both cases, \eqref{sub:exercise-1.7.1b-eq1} evaluates to $0$, implying + In both cases, \eqref{ssub:exercise-1.7.1b-eq1} evaluates to $0$, implying $P(k + 1)$ as expected. \paragraph{Conclusion}% @@ -823,8 +834,8 @@ A set consisting of a finite number of points in a plane. \end{proof} -\subsection{\partial{Exercise 1.7.1c}}% -\label{sub:exercise-1.7.1c} +\subsubsection{\partial{Exercise 1.7.1c}}% +\label{ssub:exercise-1.7.1c} The union of a finite collection of line segments in a plane. @@ -839,7 +850,7 @@ The union of a finite collection of line segments in a plane. Consider a set $S$ consisting of a single line segment in a plane. By definition of a line segment, $S$ is a rectangle in which one side has dimension $0$. - By \nameref{sec:area-choice-scale}, $S$ is measurable with area its width + By \nameref{sub:area-choice-scale}, $S$ is measurable with area its width $w$ times its height $h$. Therefore $a(S) = wh = 0$. Thus $P(1)$ holds. @@ -854,13 +865,13 @@ The union of a finite collection of line segments in a plane. By construction, $S_{k+1} = S_k \cup T$. By the induction hypothesis, $S_k$ is measurable with area $0$. By the base case, $T$ is measurable with area $0$. - By the \nameref{sec:area-additive-property}, $S_k \cup T$ is measurable, + By the \nameref{sub:area-additive-property}, $S_k \cup T$ is measurable, $S_k \cap T$ is measurable, and \begin{align} a(S_{k+1}) & = a(S_k \cup T) \nonumber \\ & = a(S_k) + a(T) - a(S_k \cap T) \nonumber \\ - & = 0 + 0 - a(S_k \cap T). \label{sub:exercise-1.7.1c-eq1} + & = 0 + 0 - a(S_k \cap T). \label{ssub:exercise-1.7.1c-eq1} \end{align} There are two cases to consider: @@ -877,7 +888,7 @@ The union of a finite collection of line segments in a plane. \vspace{8pt} \noindent - In both cases, \eqref{sub:exercise-1.7.1c-eq1} evaluates to $0$, implying + In both cases, \eqref{ssub:exercise-1.7.1c-eq1} evaluates to $0$, implying $P(k + 1)$ as expected. \paragraph{Conclusion}% @@ -886,8 +897,8 @@ The union of a finite collection of line segments in a plane. \end{proof} -\section{\partial{Exercise 1.7.2}}% -\label{sec:exercise-1.7.2} +\subsection{\partial{Exercise 1.7.2}}% +\label{sub:exercise-1.7.2} Every right triangular region is measurable because it can be obtained as the intersection of two rectangles. @@ -916,9 +927,9 @@ Prove that every triangular region is measurable and that its area is one half \centering \end{figure} - By \nameref{sec:area-choice-scale}, both $R$ and $S$ are measurable. + By \nameref{sub:area-choice-scale}, both $R$ and $S$ are measurable. By this same axiom, $a(R) = ab$ and $a(S) = ca\sin{\theta}$. - By the \nameref{sec:area-additive-property}, $R \cup S$ and $R \cap S$ are + By the \nameref{sub:area-additive-property}, $R \cup S$ and $R \cap S$ are both measurable. $a(R \cap S) = a(T)$ and $a(R \cup S)$ can be determined by noting that $R$'s construction implies identity $a(R) = 2a(T)$. @@ -932,13 +943,13 @@ Prove that every triangular region is measurable and that its area is one half & = ab + ca\sin{\theta} - ca\sin{\theta} - a(T). \end{align*} Solving for $a(T)$ gives the desired identity: $$a(T) = \frac{1}{2}ab.$$ - By \nameref{sec:area-invariance-under-congruence}, $a(T') = a(T)$, concluding + By \nameref{sub:area-invariance-under-congruence}, $a(T') = a(T)$, concluding our proof. \end{proof} -\section{\partial{Exercise 1.7.3}}% -\label{sec:exercise-1.7.3} +\subsection{\partial{Exercise 1.7.3}}% +\label{sub:exercise-1.7.3} Prove that every trapezoid and every parallelogram is measurable and derive the usual formulas for their areas. @@ -959,9 +970,9 @@ Prove that every trapezoid and every parallelogram is measurable and derive the Suppose $S$ is a right trapezoid. Then $S$ is the union of non-overlapping rectangle $R$ of width $b_1$ and height $h$ with right triangle $T$ of base $b_2 - b_1$ and height $h$. - By \nameref{sec:area-choice-scale}, $R$ is measurable. - By \nameref{sec:exercise-1.7.2}, $T$ is measurable. - By the \nameref{sec:area-additive-property}, $R \cup T$ and $R \cap T$ are + By \nameref{sub:area-choice-scale}, $R$ is measurable. + By \nameref{sub:exercise-1.7.2}, $T$ is measurable. + By the \nameref{sub:area-additive-property}, $R \cup T$ and $R \cap T$ are both measurable and \begin{align*} a(S) @@ -969,7 +980,7 @@ Prove that every trapezoid and every parallelogram is measurable and derive the & = a(R) + a(T) - a(R \cap T) \\ & = a(R) + a(T) & \text{by construction} \\ & = b_1h + a(T) & \text{Choice of Scale} \\ - & = b_1h + \frac{1}{2}(b_2 - b_1)h & \textref{sec:exercise-1.7.2} \\ + & = b_1h + \frac{1}{2}(b_2 - b_1)h & \textref{sub:exercise-1.7.2} \\ & = \frac{b_1 + b_2}{2}h. \end{align*} @@ -979,15 +990,15 @@ Prove that every trapezoid and every parallelogram is measurable and derive the Then $S$ is the union of non-overlapping triangle $T$ and right trapezoid $R$. Let $c$ denote the length of base $T$. Then $R$ has longer base edge of length $b_2 - c$. - By \nameref{sec:exercise-1.7.2}, $T$ is measurable. + By \nameref{sub:exercise-1.7.2}, $T$ is measurable. By Case 1, $R$ is measurable. - By the \nameref{sec:area-additive-property}, $R \cup T$ and $R \cap T$ are + By the \nameref{sub:area-additive-property}, $R \cup T$ and $R \cap T$ are both measurable and \begin{align*} a(S) & = a(T) + a(R) - a(R \cap T) \\ & = a(T) + a(R) & \text{by construction} \\ - & = \frac{1}{2}ch + a(R) & \textref{sec:exercise-1.7.2} \\ + & = \frac{1}{2}ch + a(R) & \textref{sub:exercise-1.7.2} \\ & = \frac{1}{2}ch + \frac{b_1 + b_2 - c}{2}h & \text{Case 1} \\ & = \frac{b_1 + b_2}{2}h. \end{align*} @@ -999,9 +1010,9 @@ Prove that every trapezoid and every parallelogram is measurable and derive the Let $c$ denote the length of base $T$. Reflect $T$ vertically to form another right triangle, say $T'$. Then $T' \cup R$ is an acute trapezoid. - By \nameref{sec:area-invariance-under-congruence}, + By \nameref{sub:area-invariance-under-congruence}, \begin{equation} - \label{sec:exercise-1.7.3-eq1} + \label{sub:exercise-1.7.3-eq1} \tag{3.1} a(T' \cup R) = a(T \cup R). \end{equation} @@ -1009,7 +1020,7 @@ Prove that every trapezoid and every parallelogram is measurable and derive the meaning \begin{align*} a(T \cup R) - & = a(T' \cup R) & \eqref{sec:exercise-1.7.3-eq1} \\ + & = a(T' \cup R) & \eqref{sub:exercise-1.7.3-eq1} \\ & = \frac{b_1 - c + b_2 + c}{2}h & \text{Case 2} \\ & = \frac{b_1 + b_2}{2}h. \end{align*} @@ -1026,32 +1037,31 @@ Prove that every trapezoid and every parallelogram is measurable and derive the Let $c$ denote the length of base $T$. Reflect $T$ vertically to form another right triangle, say $T'$. Then $T' \cup R$ is an acute trapezoid. - By \nameref{sec:area-invariance-under-congruence}, + By \nameref{sub:area-invariance-under-congruence}, \begin{equation} - \label{sec:exercise-1.7.3-eq2} - \tag{3.2} + \label{sub:exercise-1.7.3-eq2} a(T' \cup R) = a(T \cup R). \end{equation} By construction, $T' \cup R$ has height $h$ and bases $b - c$ and $b + c$ meaning \begin{align*} a(T \cup R) - & = a(T' \cup R) & \eqref{sec:exercise-1.7.3-eq2} \\ + & = a(T' \cup R) & \eqref{sub:exercise-1.7.3-eq2} \\ & = \frac{b - c + b + c}{2}h & \text{Area of Trapezoid} \\ & = bh. \end{align*} \end{proof} -\section{Exercise 1.7.4}% -\label{sec:exercise-1.7.4} +\subsection{\partial{Exercise 1.7.4}}% +\label{sub:exercise-1.7.4} Let $P$ be a polygon whose vertices are lattice points. The area of $P$ is $I + \frac{1}{2}B - 1$, where $I$ denotes the number of lattice points inside the polygon and $B$ denotes the number on the boundary. -\subsection{\partial{Exercise 1.7.4a}}% -\label{sub:exercise-1.7.4a} +\subsubsection{\partial{Exercise 1.7.4a}}% +\label{ssub:exercise-1.7.4a} Prove that the formula is valid for rectangles with sides parallel to the coordinate axes. @@ -1063,7 +1073,7 @@ Prove that the formula is valid for rectangles with sides parallel to the We assume $P$ has three non-collinear points, ruling out any instances of points or line segments. - By \nameref{sec:area-choice-scale}, $P$ is measurable with area $a(P) = wh$. + By \nameref{sub:area-choice-scale}, $P$ is measurable with area $a(P) = wh$. By construction, $P$ has $I = (w - 1)(h - 1)$ interior lattice points and $B = 2(w + h)$ lattice points on its boundary. The following shows the lattice point area formula is in agreement with @@ -1078,8 +1088,8 @@ Prove that the formula is valid for rectangles with sides parallel to the \end{proof} -\subsection{\partial{Exercise 1.7.4b}}% -\label{sub:exercise-1.7.4b} +\subsubsection{\partial{Exercise 1.7.4b}}% +\label{ssub:exercise-1.7.4b} Prove that the formula is valid for right triangles and parallelograms. @@ -1101,12 +1111,12 @@ Prove that the formula is valid for right triangles and parallelograms. By construction, $T$ shares two sides with $R$. Therefore \begin{equation} - \label{sub:exercise-1.7.4b-eq1} + \label{ssub:exercise-1.7.4b-eq1} B_T = \frac{1}{2}B_R - 1 + H_L. \end{equation} Likewise, \begin{equation} - \label{sub:exercise-1.7.4b-eq2} + \label{ssub:exercise-1.7.4b-eq2} I_T = \frac{1}{2}(I_R - (H_L - 2)). \end{equation} The following shows the lattice point area formula is in agreement with @@ -1114,21 +1124,21 @@ Prove that the formula is valid for right triangles and parallelograms. \begin{align*} I_T + \frac{1}{2}B_T - 1 & = \frac{1}{2}(I_R - (H_L - 2)) + \frac{1}{2}B_T - 1 - & \eqref{sub:exercise-1.7.4b-eq2} \\ + & \eqref{ssub:exercise-1.7.4b-eq2} \\ & = \frac{1}{2}\left[ I_R - H_L + B_T \right] \\ & = \frac{1}{2}\left[ I_R - H_L + \frac{1}{2}B_R - 1 + H_L \right] - & \eqref{sub:exercise-1.7.4b-eq1} \\ + & \eqref{ssub:exercise-1.7.4b-eq1} \\ & = \frac{1}{2}\left[ I_R + \frac{1}{2}B_R - 1 \right] \\ - & = \frac{1}{2}\left[ wh \right] & \textref{sub:exercise-1.7.4a}. + & = \frac{1}{2}\left[ wh \right] & \textref{ssub:exercise-1.7.4a}. \end{align*} We do not prove this formula is valid for parallelograms here. - Instead, refer to \nameref{sub:exercise-1.7.4c} below. + Instead, refer to \nameref{ssub:exercise-1.7.4c} below. \end{proof} -\subsection{\partial{Exercise 1.7.4c}}% -\label{sub:exercise-1.7.4c} +\subsubsection{\partial{Exercise 1.7.4c}}% +\label{ssub:exercise-1.7.4c} Use induction on the number of edges to construct a proof for general polygons. @@ -1141,7 +1151,7 @@ Use induction on the number of edges to construct a proof for general polygons. \paragraph{Base Case}% A $3$-polygon is a triangle. - By \nameref{sub:exercise-1.7.4b}, the lattice point area formula holds. + By \nameref{ssub:exercise-1.7.4b}, the lattice point area formula holds. Thus $P(3)$ holds. \paragraph{Induction Step}% @@ -1175,7 +1185,7 @@ Use induction on the number of edges to construct a proof for general polygons. & = a(S) + (I_T + \frac{1}{2}B_T - 1) & \text{induction hypothesis} \\ & = a(S) + a(T). & \text{base case} \end{align*} - By the \nameref{sec:area-additive-property}, $S \cup T$ is measurable, + By the \nameref{sub:area-additive-property}, $S \cup T$ is measurable, $S \cap T$ is measurable, and \begin{align*} a(P) @@ -1193,8 +1203,8 @@ Use induction on the number of edges to construct a proof for general polygons. \end{proof} -\section{\partial{Exercise 1.7.5}}% -\label{sec:exercise-1.7.5} +\subsection{\partial{Exercise 1.7.5}}% +\label{sub:exercise-1.7.5} Prove that a triangle whose vertices are lattice points cannot be equilateral. @@ -1207,7 +1217,7 @@ ways, using Exercises 2 and 4.] Let $T$ be an equilateral triangle whose vertices are lattice points. Assume each side of $T$ has length $a$. Then $T$ has height $h = (a\sqrt{3}) / 2$. - By \nameref{sec:exercise-1.7.2}, + By \nameref{sub:exercise-1.7.2}, \begin{equation} \label{sub:exercise-1.7.5-eq1} \tag{5.1} @@ -1215,7 +1225,7 @@ ways, using Exercises 2 and 4.] \end{equation} Let $I$ and $B$ denote the number of interior and boundary lattice points of $T$ respectively. - By \nameref{sec:exercise-1.7.4}, + By \nameref{sub:exercise-1.7.4}, \begin{equation} \label{sub:exercise-1.7.5-eq2} \tag{5.2} @@ -1229,8 +1239,8 @@ ways, using Exercises 2 and 4.] \end{proof} -\section{\partial{Exercise 1.7.6}}% -\label{sec:exercise-1.7.6} +\subsection{\partial{Exercise 1.7.6}}% +\label{sub:exercise-1.7.6} Let $A = \{1, 2, 3, 4, 5\}$, and let $\mathscr{M}$ denote the class of all subsets of $A$. @@ -1311,16 +1321,16 @@ Prove that the set function $n$ satisfies the first three axioms for area. \end{proof} -\chapter{Exercises 1.11}% -\label{chap:exercises-1-11} +\section{Exercises 1.11}% +\label{sec:exercises-1-11} -\section{Exercise 1.11.4}% -\label{sec:exercise-1.11.4} +\subsection{\partial{Exercise 1.11.4}}% +\label{sub:exercise-1.11.4} Prove that the greatest-integer function has the properties indicated: -\subsection{\verified{Exercise 1.11.4a}}% -\label{sub:exercise-1.11.4a} +\subsubsection{\verified{Exercise 1.11.4a}}% +\label{ssub:exercise-1.11.4a} $\floor{x + n} = \floor{x} + n$ for every integer $n$. @@ -1339,8 +1349,8 @@ $\floor{x + n} = \floor{x} + n$ for every integer $n$. \end{proof} -\subsection{\verified{Exercise 1.11.4b}}% -\label{sub:exercise-1.11.4b} +\subsubsection{\verified{Exercise 1.11.4b}}% +\label{ssub:exercise-1.11.4b} $\floor{-x} = \begin{cases} @@ -1388,8 +1398,8 @@ $\floor{-x} = \end{proof} -\subsection{\verified{Exercise 1.11.4c}}% -\label{sub:exercise-1.11.4c} +\subsubsection{\verified{Exercise 1.11.4c}}% +\label{ssub:exercise-1.11.4c} $\floor{x + y} = \floor{x} + \floor{y}$ or $\floor{x} + \floor{y} + 1$. @@ -1406,7 +1416,7 @@ $\floor{x + y} = \floor{x} + \floor{y}$ or $\floor{x} + \floor{y} + 1$. & = \floor{\floor{x} + \{x\} + \floor{y} + \{y\}} \nonumber \\ & = \floor{\floor{x} + \floor{y} + \{x\} + \{y\}} \nonumber \\ & = \floor{x} + \floor{y} + \floor{\{x\} + \{y\}} - & \textref{sub:exercise-1.11.4a} \label{sub:exercise-1.11.4c-eq1} + & \textref{ssub:exercise-1.11.4a} \label{ssub:exercise-1.11.4c-eq1} \end{align} There are two cases to consider: @@ -1414,7 +1424,7 @@ $\floor{x + y} = \floor{x} + \floor{y}$ or $\floor{x} + \floor{y} + 1$. Suppose $\{x\} + \{y\} < 1$. Then $\floor{\{x\} + \{y\}} = 0$. - Substituting this value into \eqref{sub:exercise-1.11.4c-eq1} yields + Substituting this value into \eqref{ssub:exercise-1.11.4c-eq1} yields $$\floor{x + y} = \floor{x} + \floor{y}.$$ \paragraph{Case 2}% @@ -1422,7 +1432,7 @@ $\floor{x + y} = \floor{x} + \floor{y}$ or $\floor{x} + \floor{y} + 1$. Suppose $\{x\} + \{y\} \geq 1$. Because $\{x\}$ and $\{y\}$ are both less than $1$, $\{x\} + \{y\} < 2$. Thus $\floor{\{x\} + \{y\}} = 1$. - Substituting this value into \eqref{sub:exercise-1.11.4c-eq1} yields + Substituting this value into \eqref{ssub:exercise-1.11.4c-eq1} yields $$\floor{x + y} = \floor{x} + \floor{y} + 1.$$ \paragraph{Conclusion}% @@ -1432,8 +1442,8 @@ $\floor{x + y} = \floor{x} + \floor{y}$ or $\floor{x} + \floor{y} + 1$. \end{proof} -\subsection{\partial{Exercise 1.11.4d}}% -\label{sub:exercise-1.11.4d} +\subsubsection{\partial{Exercise 1.11.4d}}% +\label{ssub:exercise-1.11.4d} $\floor{2x} = \floor{x} + \floor{x + \frac{1}{2}}.$ @@ -1442,12 +1452,12 @@ $\floor{2x} = \floor{x} + \floor{x + \frac{1}{2}}.$ \lean{Bookshelf/Apostol/Chapter\_1\_11} {Apostol.Chapter\_1\_11.exercise\_4d} - This is immediately proven by applying \nameref{sec:hermites-identity}. + This is immediately proven by applying \nameref{sub:hermites-identity}. \end{proof} -\subsection{\partial{Exercise 1.11.4e}}% -\label{sub:exercise-1.11.4e} +\subsubsection{\partial{Exercise 1.11.4e}}% +\label{ssub:exercise-1.11.4e} $\floor{3x} = \floor{x} + \floor{x + \frac{1}{3}} + \floor{x + \frac{2}{3}}.$ @@ -1456,12 +1466,13 @@ $\floor{3x} = \floor{x} + \floor{x + \frac{1}{3}} + \floor{x + \frac{2}{3}}.$ \lean{Bookshelf/Apostol/Chapter\_1\_11} {Apostol.Chapter\_1\_11.exercise\_4e} - This is immediately proven by applying \nameref{sec:hermites-identity}. + This is immediately proven by applying \nameref{sub:hermites-identity}. \end{proof} -\section{\partial{Hermite's Identity}}% -\label{sec:hermites-identity} +\subsection{\partial{Hermite's Identity}}% +\label{sub:hermites-identity} +\label{sub:exercise-1.11.5} The formulas in Exercises 4(d) and 4(e) suggest a generalization for $\floor{nx}$. @@ -1475,7 +1486,7 @@ State and prove such a generalization. We prove that for all natural numbers $n$ and real numbers $x$, the following identity holds: \begin{equation} - \label{sec:exercise-1.11.5-eq1} + \label{sub:exercise-1.11.5-eq1} \floor{nx} = \sum_{i=0}^{n-1} \floor{x + \frac{i}{n}} \end{equation} By definition of the floor function, $x = \floor{x} + r$ for some @@ -1486,16 +1497,16 @@ State and prove such a generalization. By construction, $\cup\; S = \ico{0}{1}$. Therefore there exists some $j \in \mathbb{N}$ such that \begin{equation} - \label{sec:exercise-1.11.5-eq2} + \label{sub:exercise-1.11.5-eq2} r \in \ico{\frac{j}{n}}{\frac{j+1}{n}}. \end{equation} With these definitions established, we now show the left- and right-hand sides - of \eqref{sec:exercise-1.11.5-eq1} evaluate to the same number. + of \eqref{sub:exercise-1.11.5-eq1} evaluate to the same number. \paragraph{Left-Hand Side}% - Consider the left-hand side of identity \eqref{sec:exercise-1.11.5-eq1}. - By \eqref{sec:exercise-1.11.5-eq2}, $nr \in \ico{j}{j + 1}$. + Consider the left-hand side of identity \eqref{sub:exercise-1.11.5-eq1}. + By \eqref{sub:exercise-1.11.5-eq2}, $nr \in \ico{j}{j + 1}$. Therefore $\floor{nr} = j$. Thus \begin{align} @@ -1503,15 +1514,15 @@ State and prove such a generalization. & = \floor{n(\floor{x} + r)} \nonumber \\ & = \floor{n\floor{x} + nr} \nonumber \\ & = \floor{n\floor{x}} + \floor{nr}. \nonumber - & \textref{sub:exercise-1.11.4a} \\ + & \textref{ssub:exercise-1.11.4a} \\ & = \floor{n\floor{x}} + j \nonumber \\ - & = n\floor{x} + j. \label{sec:exercise-1.11.5-eq3} + & = n\floor{x} + j. \label{sub:exercise-1.11.5-eq3} \end{align} \paragraph{Right-Hand Side}% Now consider the right-hand side of identity - \eqref{sec:exercise-1.11.5-eq1}. + \eqref{sub:exercise-1.11.5-eq1}. We note each summand, by construction, is the floor of $x$ added to a nonnegative number less than one. Therefore each summand contributes either $\floor{x}$ or $\floor{x} + 1$ to @@ -1519,12 +1530,12 @@ State and prove such a generalization. Letting $z$ denote the number of summands that contribute $\floor{x} + 1$, we have \begin{equation} - \label{sec:exercise-1.11.5-eq4} + \label{sub:exercise-1.11.5-eq4} \sum_{i=0}^{n-1} \floor{x + \frac{i}{n}} = n\floor{x} + z. \end{equation} The value of $z$ corresponds to the number of indices $i$ that satisfy $$\frac{i}{n} \geq 1 - r.$$ - By \eqref{sec:exercise-1.11.5-eq2}, it follows + By \eqref{sub:exercise-1.11.5-eq2}, it follows \begin{align*} 1 - r & \in \ioc{1 - \frac{j+1}{n}}{1-\frac{j}{n}} \\ @@ -1534,22 +1545,22 @@ State and prove such a generalization. indices $i$ that satisfy $$\frac{i}{n} \geq \frac{n - j}{n}.$$ Rearranging terms, we see that $i \geq n - j$ holds for $z = (n - 1) - (n - j) + 1 = j$ of the $n$ summands. - Substituting the value of $z$ into \eqref{sec:exercise-1.11.5-eq4} yields + Substituting the value of $z$ into \eqref{sub:exercise-1.11.5-eq4} yields \begin{equation} - \label{sec:exercise-1.11.5-eq5} + \label{sub:exercise-1.11.5-eq5} \sum_{i=0}^{n-1} \floor{x + \frac{i}{n}} = n\floor{x} + j. \end{equation} \paragraph{Conclusion}% - Since \eqref{sec:exercise-1.11.5-eq3} and \eqref{sec:exercise-1.11.5-eq5} + Since \eqref{sub:exercise-1.11.5-eq3} and \eqref{sub:exercise-1.11.5-eq5} agree with one another, it follows identity - \eqref{sec:exercise-1.11.5-eq1} holds. + \eqref{sub:exercise-1.11.5-eq1} holds. \end{proof} -\section{\partial{Exercise 1.11.6}}% -\label{sec:exercise-1.11.6} +\subsection{\partial{Exercise 1.11.6}}% +\label{sub:exercise-1.11.6} Recall that a lattice point $(x, y)$ in the plane is one whose coordinates are integers. @@ -1565,7 +1576,7 @@ Prove that the number of lattice points in $S$ is equal to the sum Let $i = a, \ldots, b$ and define $S_i = \mathbb{N} \cap \ioc{0}{f(i)}$. By construction, the number of lattice points in $S$ is \begin{equation} - \label{sec:exercise-1.11.6-eq1} + \label{sub:exercise-1.11.6-eq1} \sum_{n = a}^b \abs{S_n}. \end{equation} All that remains is to show $\abs{S_i} = \floor{f(i)}$. @@ -1586,13 +1597,13 @@ Prove that the number of lattice points in $S$ is equal to the sum \paragraph{Conclusion}% By cases 1 and 2, $\abs{S_i} = \floor{f(i)}$. - Substituting this identity into \eqref{sec:exercise-1.11.6-eq1} finishes the + Substituting this identity into \eqref{sub:exercise-1.11.6-eq1} finishes the proof. \end{proof} -\section{Exercise 1.11.7}% -\label{sec:exercise-1.11.7} +\subsection{\partial{Exercise 1.11.7}}% +\label{sub:exercise-1.11.7} If $a$ and $b$ are positive integers with no common factor, we have the formula $$\sum_{n=1}^{b-1} \floor{\frac{na}{b}} = \frac{(a - 1)(b - 1)}{2}.$$ @@ -1601,8 +1612,8 @@ When $b = 1$, the sum on the left is understood to be $0$. \note{When $b = 1$, the proofs of (a) and (b) are trivial. We continue under the assumption $b > 1$.} -\subsection{\partial{Exercise 1.11.7a}}% -\label{sub:exercise-1.11.7a} +\subsubsection{\partial{Exercise 1.11.7a}}% +\label{ssub:exercise-1.11.7a} Derive this result by a geometric argument, counting lattice points in a right triangle. @@ -1612,10 +1623,10 @@ Derive this result by a geometric argument, counting lattice points in a right Let $f \colon [1, b - 1] \rightarrow \mathbb{R}$ be given by $f(x) = ax / b$. Let $S$ denote the set of points $(x, y)$ satisfying $1 \leq x \leq b - 1$, $0 < y \leq f(x)$. - By \nameref{sec:exercise-1.11.6}, the number of lattice points of $S$ is equal + By \nameref{sub:exercise-1.11.6}, the number of lattice points of $S$ is equal to the sum \begin{equation} - \label{sub:exercise-1.11.7a-eq1} + \label{ssub:exercise-1.11.7a-eq1} \sum_{n=1}^{b-1} \floor{f(n)} = \sum_{n=1}^{b-1} \floor{\frac{na}{b}}. \end{equation} Define $T$ to be the triangle of width $w = b$ and height $h = f(b) = a$ @@ -1648,43 +1659,43 @@ Derive this result by a geometric argument, counting lattice points in a right Let $I_R$ denote the number of interior lattice points of $R$. Let $I_T$ and $B_T$ denote the interior and boundary lattice points of $T$ respectively. - By \nameref{sub:exercise-1.7.4b-eq2}, + By \nameref{ssub:exercise-1.7.4b-eq2}, \begin{align} I_T & = \frac{1}{2}(I_R - (H_L - 2)) \nonumber \\ & = \frac{1}{2}(I_R - (2 - 2)) & \textref{par:exercise-1.11.7a-i} \nonumber \\ - & = \frac{1}{2}I_R. & \label{sub:exercise-1.11.7a-eq2} + & = \frac{1}{2}I_R. & \label{ssub:exercise-1.11.7a-eq2} \end{align} Furthermore, since both the adjacent and opposite side of $T$ are not included in $T$ and there exist no lattice points on $T$'s hypotenuse besides the endpoints, it follows \begin{equation} - \label{sub:exercise-1.11.7a-eq3} + \label{ssub:exercise-1.11.7a-eq3} B_T = 0. \end{equation} Thus the number of lattice points of $T$ equals \begin{align} I_T + B_T - & = I_T & \eqref{sub:exercise-1.11.7a-eq3} \nonumber \\ - & = \frac{1}{2}I_R & \eqref{sub:exercise-1.11.7a-eq2} \nonumber \\ + & = I_T & \eqref{ssub:exercise-1.11.7a-eq3} \nonumber \\ + & = \frac{1}{2}I_R & \eqref{ssub:exercise-1.11.7a-eq2} \nonumber \\ & = \frac{(b - 1)(a - 1)}{2}. - & \textref{sub:exercise-1.7.4a} \label{sub:exercise-1.11.7a-eq4} + & \textref{ssub:exercise-1.7.4a} \label{ssub:exercise-1.11.7a-eq4} \end{align} \paragraph{Conclusion}% - By \eqref{sub:exercise-1.11.7a-eq1} the number of lattice points of $S$ is + By \eqref{ssub:exercise-1.11.7a-eq1} the number of lattice points of $S$ is equal to the sum $$\sum_{n=1}^{b-1} \floor{\frac{na}{b}}.$$ But the number of lattice points of $S$ is the same as that of $T$. - By \eqref{sub:exercise-1.11.7a-eq4}, the number of lattice points in $T$ is + By \eqref{ssub:exercise-1.11.7a-eq4}, the number of lattice points in $T$ is equal to $$\frac{(b - 1)(a - 1)}{2}.$$ Thus $$\sum_{n=1}^{b-1} \floor{\frac{na}{b}} = \frac{(a - 1)(b - 1)}{2}.$$ \end{proof} -\subsection{\partial{Exercise 1.11.7b}}% -\label{sub:exercise-1.11.7b} +\subsubsection{\partial{Exercise 1.11.7b}}% +\label{ssub:exercise-1.11.7b} Derive the result analytically as follows: By changing the index of summation, note that @@ -1700,9 +1711,9 @@ Now apply Exercises 4(a) and (b) to the bracket on the right. By hypothesis, $a$ and $b$ are coprime. Furthermore, $n < b$ for all values of $n$. Thus $an / b$ is not an integer. - By \nameref{sub:exercise-1.11.4b}, + By \nameref{ssub:exercise-1.11.4b}, \begin{equation} - \label{sub:exercise-1.11.7b-eq1} + \label{ssub:exercise-1.11.7b-eq1} \floor{-\frac{an}{b}} = -\floor{\frac{an}{b}} - 1. \end{equation} Consider the following: @@ -1712,9 +1723,9 @@ Now apply Exercises 4(a) and (b) to the bracket on the right. & = \sum_{n=1}^{b-1} \floor{\frac{ab - an}{b}} \\ & = \sum_{n=1}^{b-1} \floor{-\frac{an}{b} + a} \\ & = \sum_{n=1}^{b-1} \floor{-\frac{an}{b}} + a. - & \textref{sub:exercise-1.11.4a} \\ + & \textref{ssub:exercise-1.11.4a} \\ & = \sum_{n=1}^{b-1} -\floor{\frac{an}{b}} - 1 + a - & \eqref{sub:exercise-1.11.7b-eq1} \\ + & \eqref{ssub:exercise-1.11.7b-eq1} \\ & = -\sum_{n=1}^{b-1} \floor{\frac{an}{b}} - \sum_{n=1}^{b-1} 1 + \sum_{n=1}^{b-1} a \\ & = -\sum_{n=1}^{b-1} \floor{\frac{an}{b}} - (b - 1) + a(b - 1). @@ -1725,12 +1736,12 @@ Now apply Exercises 4(a) and (b) to the bracket on the right. \end{proof} -\section{\partial{Exercise 1.11.8}}% -\label{sec:exercise-1.11.8} +\subsection{\partial{Exercise 1.11.8}}% +\label{sub:exercise-1.11.8} Let $S$ be a set of points on the real line. -Let $\mathcal{X}_S$ denote the \nameref{sec:def-characteristic-function} of $S$. -Let $f$ be a \nameref{sec:def-step-function} which takes the constant value +Let $\mathcal{X}_S$ denote the \nameref{def:characteristic-function} of $S$. +Let $f$ be a \nameref{def:step-function} which takes the constant value $c_k$ on the $k$th open subinterval $I_k$ of some partition of an interval $[a, b]$. Prove that for each $x$ in the union $I_1 \cup I_2 \cup \cdots \cup I_n$ we have @@ -1743,7 +1754,7 @@ This property is described by saying that every step function is a linear Let $x \in I_1 \cup I_2 \cup \cdots \cup I_n$ and $N = \{1, \ldots, n\}$. Let $k \in N$ such that $x \in I_k$. Consider an arbitrary $j \in N - \{k\}$. - By definition of a nameref{sec:def-partition}, $I_j \cap I_k = \emptyset$. + By definition of a nameref{def:partition}, $I_j \cap I_k = \emptyset$. That is, $I_j$ and $I_k$ are disjoint for all $j \in N - \{k\}$. Therefore, by definition of the characteristic function, $\mathcal{X}_{I_k}(x) = 1$ and $\mathcal{X}_{I_j}(x) = 0$ for all @@ -1760,16 +1771,16 @@ This property is described by saying that every step function is a linear \end{proof} -\chapter{Properties of the Integral of a Step Function}% -\label{chap:properties-integral-step-function} +\section{Properties of the Integral of a Step Function}% +\label{sec:properties-integral-step-function} -\section{\partial{Additive Property}}% -\label{sec:step-additive-property} -\label{sec:theorem-1.2} +\subsection{\partial{Additive Property}}% +\label{sub:step-additive-property} +\label{sub:theorem-1.2} \begin{theorem}[1.2] - Let $s$ and $t$ be \nameref{sec:def-step-function}s on closed interval + Let $s$ and $t$ be \nameref{def:step-function}s on closed interval $[a, b]$. Then $$\int_a^b \left[ s(x) + t(x) \right] \mathop{dx} = @@ -1780,7 +1791,7 @@ This property is described by saying that every step function is a linear \begin{proof} Let $s$ and $t$ be step functions on closed interval $[a, b]$. - By definition of a step function, there exists a \nameref{sec:def-partition} + By definition of a step function, there exists a \nameref{def:partition} $P_s$ such that $s$ is constant on each open subinterval of $P_s$. Likewise, there exists a partition $P_t$ such that $t$ is constant on each open subinterval of $P_t$. @@ -1793,7 +1804,7 @@ This property is described by saying that every step function is a linear $P$. Let $t_k$ denote the constant value of $t$ on the $k$th open subinterval of $P$. - By definition of the \nameref{sec:def-integral-step-function}, + By definition of the \nameref{def:integral-step-function}, \begin{align*} \int_a^b \left[ s(x) + t(x) \right] \mathop{dx} & = \sum_{k=1}^n (s_k + t_k) \cdot (x_k - x_{k-1}) \\ @@ -1806,13 +1817,13 @@ This property is described by saying that every step function is a linear \end{proof} -\section{\partial{Homogeneous Property}}% -\label{sec:step-homogeneous-property} -\label{sec:theorem-1.3} +\subsection{\partial{Homogeneous Property}}% +\label{sub:step-homogeneous-property} +\label{sub:theorem-1.3} \begin{theorem}[1.3] - Let $s$ be a \nameref{sec:def-step-function} on closed interval $[a, b]$. + Let $s$ be a \nameref{def:step-function} on closed interval $[a, b]$. For every real number $c$, we have $$\int_a^b c \cdot s(x) \mathop{dx} = c\int_a^b s(x) \mathop{dx}.$$ @@ -1821,13 +1832,13 @@ This property is described by saying that every step function is a linear \begin{proof} Let $s$ be a step function on closed interval $[a, b]$. - By definition of a step function, there exists a \nameref{sec:def-partition} + By definition of a step function, there exists a \nameref{def:partition} $P = \{x_0, x_1, \ldots, x_n\}$ such that $s$ is constant on each open subinterval of $P$. Let $s_k$ denote the constant value of $s$ on the $k$th open subinterval of $P$. Then $c \cdot s$ is a step function with step partition $P$. - By definition of the \nameref{sec:def-integral-step-function}, + By definition of the \nameref{def:integral-step-function}, \begin{align*} \int_a^b c \cdot s(x) \mathop{dx} & = \sum_{k=1}^n c \cdot s_k \cdot (x_k - x_{k-1}) \\ @@ -1837,13 +1848,13 @@ This property is described by saying that every step function is a linear \end{proof} -\section{\partial{Linearity Property}}% -\label{sec:step-linearity-property} -\label{sec:theorem-1.4} +\subsection{\partial{Linearity Property}}% +\label{sub:step-linearity-property} +\label{sub:theorem-1.4} \begin{theorem}[1.4] - Let $s$ and $t$ be \nameref{sec:def-step-function}s on closed interval + Let $s$ and $t$ be \nameref{def:step-function}s on closed interval $[a, b]$. For every real $c_1$ and $c_2$, we have $$\int_a^b \left[ c_1s(x) + c_2t(x) \right] \mathop{dx} = @@ -1860,20 +1871,20 @@ This property is described by saying that every step function is a linear \begin{align*} & \int_a^b \left[ c_1s(x) + c_2t(x) \right] \mathop{dx} \\ & = \int_a^b c_1s(x) \mathop{dx} + \int_a^b c_2t(x) \mathop{dx} - & \textref{sec:step-additive-property} \\ + & \textref{sub:step-additive-property} \\ & = c_1\int_a^b s(x) \mathop{dx} + c_2\int_a^b t(x). \mathop{dx} - & \textref{sec:step-homogeneous-property} + & \textref{sub:step-homogeneous-property} \end{align*} \end{proof} -\section{\partial{Comparison Theorem}}% -\label{sec:step-comparison-theorem} -\label{sec:theorem-1.5} +\subsection{\partial{Comparison Theorem}}% +\label{sub:step-comparison-theorem} +\label{sub:theorem-1.5} \begin{theorem}[1.5] - Let $s$ and $t$ be \nameref{sec:def-step-function}s on closed interval + Let $s$ and $t$ be \nameref{def:step-function}s on closed interval $[a, b]$. If $s(x) < t(x)$ for every $x$ in $[a, b]$, then $$\int_a^b s(x) \mathop{dx} < \int_a^b t(x) \mathop{dx}.$$ @@ -1883,7 +1894,7 @@ This property is described by saying that every step function is a linear \begin{proof} Let $s$ and $t$ be step functions on closed interval $[a, b]$. - By definition of a step function, there exists a \nameref{sec:def-partition} + By definition of a step function, there exists a \nameref{def:partition} $P_s$ such that $s$ is constant on each open subinterval of $P_s$. Likewise, there exists a partition $P_t$ such that $t$ is constant on each open subinterval of $P_t$. @@ -1896,7 +1907,7 @@ This property is described by saying that every step function is a linear $P$. Let $t_k$ denote the constant value of $t$ on the $k$th open subinterval of $P$. - By definition of the \nameref{sec:def-integral-step-function}, + By definition of the \nameref{def:integral-step-function}, \begin{align*} \int_a^b s(x) \mathop{dx} & = \sum_{k=1}^n s_k \cdot (x_k - x_{k-1}) \\ @@ -1906,13 +1917,13 @@ This property is described by saying that every step function is a linear \end{proof} -\section{\partial{Additivity With Respect to the Interval of Integration}}% -\label{sec:step-additivity-with-respect-interval-integration} -\label{sec:theorem-1.6} +\subsection{\partial{Additivity With Respect to the Interval of Integration}}% +\label{sub:step-additivity-with-respect-interval-integration} +\label{sub:theorem-1.6} \begin{theorem}[1.6] - Let $a, b, c \in \mathbb{R}$ and $s$ a \nameref{sec:def-step-function} on the + Let $a, b, c \in \mathbb{R}$ and $s$ a \nameref{def:step-function} on the smallest closed interval containing them. Then $$\int_a^c s(x) \mathop{dx} + \int_c^b s(x) \mathop{dx} + @@ -1924,7 +1935,7 @@ This property is described by saying that every step function is a linear WLOG, suppose $a < c < b$ and $s$ be a step function on closed interval $[a, b]$. - By definition of a step function, there exists a \nameref{sec:def-partition} + By definition of a step function, there exists a \nameref{def:partition} $P$ such that $s$ is constant on each open subinterval of $P$. Let $Q = \{x_0, x_1, \ldots, x_n\}$ be a refinement of $P$ that includes $c$ @@ -1933,7 +1944,7 @@ This property is described by saying that every step function is a linear that $x_i = c$. Let $s_k$ denote the constant value of $s$ on the $k$th open subinterval of $Q$. - By definition of the \nameref{sec:def-integral-step-function}, + By definition of the \nameref{def:integral-step-function}, \begin{align*} \int_a^b s(x) \mathop{dx} & = \sum_{k=1}^n s_k \cdot (x_k - x_{k - 1}) \\ @@ -1952,9 +1963,9 @@ This property is described by saying that every step function is a linear \end{proof} -\section{\partial{Invariance Under Translation}}% -\label{sec:step-invariance-under-translation} -\label{sec:theorem-1.7} +\subsection{\partial{Invariance Under Translation}}% +\label{sub:step-invariance-under-translation} +\label{sub:theorem-1.7} \begin{theorem}[1.7] @@ -1968,7 +1979,7 @@ This property is described by saying that every step function is a linear \begin{proof} Let $s$ be a step function on closed interval $[a, b]$. - By definition of a step function, there exists a \nameref{sec:def-partition} + By definition of a step function, there exists a \nameref{def:partition} $P = \{x_0, x_1, \ldots, x_n\}$ such that $s$ is constant on each open subinterval of $P$. Let $s_k$ denote the constant value of $s$ on the $k$th open subinterval of @@ -1981,7 +1992,7 @@ This property is described by saying that every step function is a linear Let $t_k$ denote the value of $t$ on the $k$th open subinterval of $Q$. By construction, $t_k = s_k$. - By definition of the \nameref{sec:def-integral-step-function}, + By definition of the \nameref{def:integral-step-function}, \begin{align*} \int_{a+c}^{b+c} s(x - c) \mathop{dx} & = \int_{a+c}^{b+c} t(x) \mathop{dx} \\ @@ -1993,9 +2004,9 @@ This property is described by saying that every step function is a linear \end{proof} -\section{\partial{Expansion or Contraction of the Interval of Integration}}% -\label{sec:step-expansion-contraction-interval-integration} -\label{sec:theorem-1.8} +\subsection{\partial{Expansion or Contraction of the Interval of Integration}}% +\label{sub:step-expansion-contraction-interval-integration} +\label{sub:theorem-1.8} \begin{theorem}[1.8] @@ -2009,7 +2020,7 @@ This property is described by saying that every step function is a linear \begin{proof} Let $s$ be a step function on closed interval $[a, b]$. - By definition of a step function, there exists a \nameref{sec:def-partition} + By definition of a step function, there exists a \nameref{def:partition} $P = \{x_0, x_1, \ldots, x_n\}$ such that $s$ is constant on each open subinterval of $P$. Let $s_i$ denote the value of $s$ on the $i$th open subinterval of $P$. @@ -2023,7 +2034,7 @@ This property is described by saying that every step function is a linear Then $t(x) = s(x / k)$ is a step function on closed interval $[ka, kb]$ with partition $Q = \{kx_0, kx_1, \ldots, kx_n\}$. Furthermore $t_i = s_i$. - By definition of the \nameref{sec:def-integral-step-function}, + By definition of the \nameref{def:integral-step-function}, \begin{align*} \int_{ka}^{kb} s(x / k) \mathop{dx} & = \int_{ka}^{kb} t(x) \mathop{dx} \\ @@ -2039,7 +2050,7 @@ This property is described by saying that every step function is a linear Then $t(x) = s(x / k)$ is a step function on closed interval $[kb, ka]$ with partition $Q = \{kx_n, kx_{n-1}, \ldots, kx_0\}$. Furthermore $t_i = s_i$. - By definition of the \nameref{sec:def-integral-step-function}, + By definition of the \nameref{def:integral-step-function}, \begin{align*} \int_{ka}^{kb} s(x / k) \mathop{dx} & = -\int_{kb}^{ka} s(x / k) \mathop{dx} \\ @@ -2053,8 +2064,8 @@ This property is described by saying that every step function is a linear \end{proof} -\section{\partial{Reflection Property}}% -\label{sec:step-reflection-property} +\subsection{\partial{Reflection Property}}% +\label{sub:step-reflection-property} Let $s$ be a step function on closed interval $[a, b]$. Then @@ -2063,7 +2074,7 @@ Then \begin{proof} Let $k = -1$. - By \nameref{sec:step-expansion-contraction-interval-integration}, + By \nameref{sub:step-expansion-contraction-interval-integration}, $$\int_{-a}^{-b} s \left( \frac{x}{-1} \right) = -\int_a^b s(x) \mathop{dx}.$$ Simplifying the left-hand side of the above identity, and multiplying both @@ -2071,16 +2082,16 @@ Then \end{proof} -\chapter{Exercises 1.15}% -\label{chap:exercises-1.15} +\section{Exercises 1.15}% +\label{sec:exercises-1.15} -\section{Exercise 1.15.1}% -\label{sec:exercise-1.15.1} +\subsection{\partial{Exercise 1.15.1}}% +\label{sub:exercise-1.15.1} Compute the value of each of the following integrals. -\subsection{\partial{Exercise 1.15.1a}}% -\label{sub:exercise-1.15.1a} +\subsubsection{\partial{Exercise 1.15.1a}}% +\label{ssub:exercise-1.15.1a} $\int_{-1}^3 \floor{x} \mathop{dx}$. @@ -2091,7 +2102,7 @@ $\int_{-1}^3 \floor{x} \mathop{dx}$. $P = \{-1, 0, 1, 2, 3\} = \{x_0, x_1, x_2, x_3, x_4\}$. Let $s_k$ denote the constant value $s$ takes on the $k$th open subinterval of $P$. - By definition of the \nameref{sec:def-integral-step-function}, + By definition of the \nameref{def:integral-step-function}, \begin{align*} \int_{-1}^3 \floor{x} \mathop{dx} & = \sum_{k=1}^4 s_k \cdot (x_k - x_{k-1}) \\ @@ -2101,8 +2112,8 @@ $\int_{-1}^3 \floor{x} \mathop{dx}$. \end{proof} -\subsection{\partial{Exercise 1.15.1c}}% -\label{sub:exercise-1.15.1c} +\subsubsection{\partial{Exercise 1.15.1c}}% +\label{ssub:exercise-1.15.1c} $\int_{-1}^3 \left(\floor{x} + \floor{x + \frac{1}{2}}\right) \mathop{dx}$. @@ -2118,7 +2129,7 @@ $\int_{-1}^3 \left(\floor{x} + \floor{x + \frac{1}{2}}\right) \mathop{dx}$. \end{align*} Let $s_k$ denote the constant value $s$ takes on the $k$th open subinterval of $P$. - By definition of the \nameref{sec:def-integral-step-function}, + By definition of the \nameref{def:integral-step-function}, \begin{align*} \int_{-1}^3 \floor{x} + \floor{x + \frac{1}{2}} & = \sum_{k=1}^8 s_k \cdot (x_k - x_{k-1}) \\ @@ -2129,23 +2140,23 @@ $\int_{-1}^3 \left(\floor{x} + \floor{x + \frac{1}{2}}\right) \mathop{dx}$. \end{proof} -\subsection{\partial{Exericse 1.15.1e}}% -\label{sub:exercise-1.15.1e} +\subsubsection{\partial{Exericse 1.15.1e}}% +\label{ssub:exercise-1.15.1e} $\int_{-1}^3 \floor{2x} \mathop{dx}$. \begin{proof} Let $s(x) = \floor{2x}$. - By \nameref{sec:hermites-identity}, + By \nameref{sub:hermites-identity}, $s(x) = \floor{x} + \floor{x + \frac{1}{2}}$. - Thus, by \nameref{sub:exercise-1.15.1c}, + Thus, by \nameref{ssub:exercise-1.15.1c}, $$\int_{-1}^3 \floor{2x} \mathop{dx} = 6.$$ \end{proof} -\section{\partial{Exercise 1.15.3}}% -\label{sec:exercise-1.15.3} +\subsection{\partial{Exercise 1.15.3}}% +\label{sub:exercise-1.15.3} Show that $\int_a^b \floor{x} \mathop{dx} + \int_a^b \floor{-x} \mathop{dx} = a - b$. @@ -2155,13 +2166,13 @@ Show that Let $s(x) = \floor{x}$ and $t(x) = \floor{-x}$, both with domain $[a, b]$. Let $x_1$, $\ldots$, $x_{n-1}$ denote the integers found in interval $(a, b)$. Then $P = \{x_0, x_1, \ldots, x_n\}$, $x_0 = a$ and $x_n = b$, is a step - \nameref{sec:def-partition} of both $s$ and $t$. + \nameref{def:partition} of both $s$ and $t$. Let $s_k$ and $t_k$ denote the constant values $s$ and $t$ take on the $k$th open subinterval of $P$ respectively. - By \nameref{sub:exercise-1.11.4b}, $\floor{-x} = -\floor{x} - 1$ for all $x$ + By \nameref{ssub:exercise-1.11.4b}, $\floor{-x} = -\floor{x} - 1$ for all $x$ in every open subinterval of $P$. That is, $s_k = -t_k - 1$. - By definition of the \nameref{sec:def-integral-step-function}, + By definition of the \nameref{def:integral-step-function}, \begin{align*} \int_a^b \floor{x} \mathop{dx} + \int_a^b \floor{-x} \mathop{dx} & = \sum_{k=1}^n s_k (x_k - x_{k-1}) + @@ -2175,22 +2186,22 @@ Show that \end{proof} -\section{Exercise 1.15.5}% -\label{sec:exercise-1.15.5} +\subsection{\partial{Exercise 1.15.5}}% +\label{sub:exercise-1.15.5} -\subsection{\partial{Exercise 1.15.5a}}% -\label{sub:exercise-1.15.5a} +\subsubsection{\partial{Exercise 1.15.5a}}% +\label{ssub:exercise-1.15.5a} Prove that $\int_0^2 \floor{t^2} \mathop{dt} = 5 - \sqrt{2} - \sqrt{3}$. \begin{proof} Let $s(t) = \floor{t^2}$ with domain $[0, 2]$. - Then $s$ is a \nameref{sec:def-step-function} with partition + Then $s$ is a \nameref{def:step-function} with partition $P = \{0, 1, \sqrt{2}, \sqrt{3}, 2\} = \{x_0, x_1, \ldots, x_4\}$. Let $s_k$ denote the constant value that $s$ takes in the $k$th open subinterval of $P$. - By the \nameref{sec:def-integral-step-function}, + By the \nameref{def:integral-step-function}, \begin{align*} \int_0^2 \floor{t^2} \mathop{dt} & = \sum_{k=1}^4 s_k \cdot (x_k - x_{k-1}) \\ @@ -2201,15 +2212,15 @@ Prove that $\int_0^2 \floor{t^2} \mathop{dt} = 5 - \sqrt{2} - \sqrt{3}$. \end{proof} -\subsection{\partial{Exercise 1.15.5b}}% -\label{sub:exercise-1.15.5b} +\subsubsection{\partial{Exercise 1.15.5b}}% +\label{ssub:exercise-1.15.5b} Compute $\int_{-3}^3 \floor{t^2} \mathop{dt}$. \begin{proof} Let $s(t) = \floor{t^2}$ with domain $[0, 3]$. - Then $s$ is a \nameref{sec:def-step-function} with \nameref{sec:def-partition} + Then $s$ is a \nameref{def:step-function} with \nameref{def:partition} \begin{align*} P & = \{\sqrt{0}, \sqrt{1}, \sqrt{2}, \sqrt{3}, \sqrt{4}, \sqrt{5}, @@ -2218,7 +2229,7 @@ Compute $\int_{-3}^3 \floor{t^2} \mathop{dt}$. \end{align*} Let $s_k$ denote the constant value that $s$ takes in the $k$th open subinterval of $P$. - By the \nameref{sec:def-integral-step-function}, + By the \nameref{def:integral-step-function}, \begin{align} \int_0^3 \floor{t^2} \mathop{dt} & = \sum_{k=1}^9 s_k \cdot (x_k - x_{k-1}) @@ -2232,7 +2243,7 @@ Compute $\int_{-3}^3 \floor{t^2} \mathop{dt}$. \label{sub:exercise-1.15.5b-eq2} \int_{-3}^0 \floor{t^2} \mathop{dt} = \int_0^3 \floor{t^2}. \end{equation} - By \nameref{sec:step-additivity-with-respect-interval-integration}, + By \nameref{sub:step-additivity-with-respect-interval-integration}, \begin{align*} \int_{-3}^3 \floor{t^2} \mathop{dt} & = \int_{-3}^0 \floor{t^2} \mathop{dt} + @@ -2245,22 +2256,22 @@ Compute $\int_{-3}^3 \floor{t^2} \mathop{dt}$. \end{proof} -\section{Exercise 1.15.7}% -\label{sec:exercise-1.15.7} +\subsection{\partial{Exercise 1.15.7}}% +\label{sub:exercise-1.15.7} -\subsection{\partial{Exercise 1.15.7a}}% -\label{sub:exercise-1.15.7a} +\subsubsection{\partial{Exercise 1.15.7a}}% +\label{ssub:exercise-1.15.7a} Compute $\int_0^9 \floor{\sqrt{t}} \mathop{dt}$. \begin{proof} Let $s(t) = \floor{\sqrt{t}}$ with domain $[0, 9]$. - Then $s$ is a \nameref{sec:def-step-function} with \nameref{sec:def-partition} + Then $s$ is a \nameref{def:step-function} with \nameref{def:partition} $P = \{0, 1, 4, 9\} = \{x_0, x_1, x_2, x_3\}$. Let $s_k$ denote the constant value that $s$ takes in the $k$th open subinterval of $P$. - By the \nameref{sec:def-integral-step-function}, + By the \nameref{def:integral-step-function}, \begin{align*} \int_0^9 \floor{\sqrt{t}} \mathop{dt} & = \sum_{k=1}^3 s_k \cdot (x_k - x_{k-1}) \\ @@ -2270,8 +2281,8 @@ Compute $\int_0^9 \floor{\sqrt{t}} \mathop{dt}$. \end{proof} -\subsection{\partial{Exercise 1.15.7b}}% -\label{sub:exercise-1.15.7b} +\subsubsection{\partial{Exercise 1.15.7b}}% +\label{ssub:exercise-1.15.7b} If $n$ is a positive integer, prove that $$\int_0^{n^2} \floor{\sqrt{t}} \mathop{dt} = n(n - 1)(4n + 1) / 6.$$ @@ -2289,11 +2300,11 @@ If $n$ is a positive integer, prove that Let $n = 1$. Define $s(t) = \floor{\sqrt{t}}$ with domain $[0, 1]$. - Then $s$ is a \nameref{sec:def-step-function} with - \nameref{sec:def-partition} $P = \{0, 1\} = \{x_0, x_1\}$. + Then $s$ is a \nameref{def:step-function} with \nameref{def:partition} + $P = \{0, 1\} = \{x_0, x_1\}$. Let $s_k$ denote the constant value of $s$ on the $k$th open subinterval of $P$. - By definition of the \nameref{sec:def-integral-step-function}, the left-hand + By definition of the \nameref{def:integral-step-function}, the left-hand side of \eqref{sub:exercise-1.15.7b-eq1} evaluates to \begin{align*} \int_0^{n^2} \floor{\sqrt{t}} \mathop{dt} @@ -2309,8 +2320,7 @@ If $n$ is a positive integer, prove that Let $n > 0$ be a positive integer and suppose $P(n)$ is true. Define $s(t) = \floor{\sqrt{t}}$ with domain $[0, (n + 1)^2]$. - Then $s$ is a \nameref{sec:def-step-function} with - \nameref{sec:def-partition} + Then $s$ is a \nameref{def:step-function} with \nameref{def:partition} \begin{align*} P & = \{0, 1, 4, \ldots, n^2, (n + 1)^2\} \\ @@ -2318,7 +2328,7 @@ If $n$ is a positive integer, prove that \end{align*} Let $s_k$ denote the constant value of $s$ on the $k$th open subinterval of $P$. - By definition of the \nameref{sec:def-integral-step-function}, it follows + By definition of the \nameref{def:integral-step-function}, it follows that \begin{align*} & \int_0^{(n + 1)^2} s(x) \mathop{dx} \\ @@ -2346,21 +2356,21 @@ If $n$ is a positive integer, prove that \end{proof} -\section{\partial{Exercise 1.15.9}}% -\label{sec:exercise-1.15.9} +\subsection{\partial{Exercise 1.15.9}}% +\label{sub:exercise-1.15.9} Show that the following property is equivalent to - \nameref{sec:step-expansion-contraction-interval-integration}: + \nameref{sub:step-expansion-contraction-interval-integration}: \begin{equation} - \label{sec:exercise-1.15.9-eq1} + \label{sub:exercise-1.15.9-eq1} \int_{ka}^{kb} f(x) \mathop{dx} = k \int_a^b f(kx) \mathop{dx}. \end{equation} \begin{proof} Let $f$ be a step function on closed interval $[a, b]$ and $k \neq 0$. - Applying \nameref{sec:step-expansion-contraction-interval-integration} to the - right-hand side of \eqref{sec:exercise-1.15.9-eq1} yields + Applying \nameref{sub:step-expansion-contraction-interval-integration} to the + right-hand side of \eqref{sub:exercise-1.15.9-eq1} yields $$k\int_{ka}^{kb} f(kx / k) \mathop{dx} = k\left[k\int_a^b f(kx) \mathop{dx}\right].$$ Simplifying the left-hand side and dividing both sides by $k$ immediately @@ -2368,24 +2378,24 @@ Show that the following property is equivalent to \end{proof} -\section{Exercise 1.15.11}% -\label{sec:exercise-1.15.11} +\subsection{\partial{Exercise 1.15.11}}% +\label{sub:exercise-1.15.11} If we instead defined the integral of step functions as \begin{equation*} - \label{sec:exercise-1.15.11-eq1} + \label{sub:exercise-1.15.11-eq1} \int_a^b s(x) \mathop{dx} = \sum_{k=1}^n s_k^3 \cdot (x_k - x_{k-1}), \end{equation*} a new and different theory of integration would result. Which of the following properties would remain valid in this new theory? -\subsection{\partial{Exercise 1.15.11a}}% -\label{sub:exercise-1.15.11a} +\subsubsection{\partial{Exercise 1.15.11a}}% +\label{ssub:exercise-1.15.11a} $\int_a^b s + \int_b^c s = \int_a^c s$. \note{This property mirrors - \nameref{sec:step-additivity-with-respect-interval-integration}.} + \nameref{sub:step-additivity-with-respect-interval-integration}.} \begin{proof} @@ -2395,8 +2405,8 @@ $\int_a^b s + \int_b^c s = \int_a^c s$. WLOG, suppose $a < b < c$. Let $s$ be a step function defined on closed interval $[a, c]$. - By definition of a \nameref{sec:def-step-function}, there exists a - \nameref{sec:def-partition} such that $s$ is constant on each open + By definition of a \nameref{def:step-function}, there exists a + \nameref{def:partition} such that $s$ is constant on each open subinterval of $P$. Let $Q = \{x_0, x_1, \ldots, x_n\}$ be a refinement of $P$ that includes $b$ as a subdivision point. @@ -2404,7 +2414,7 @@ $\int_a^b s + \int_b^c s = \int_a^c s$. that $x_i = c$. Let $s_k$ denote the constant value of $s$ on the $k$th open subinterval of $Q$. - By \eqref{sec:exercise-1.15.11-eq1}, + By \eqref{sub:exercise-1.15.11-eq1}, \begin{align*} \int_a^c s & = \sum_{k=1}^n s_k^3 \cdot (x_k - x_{k-1}) \\ @@ -2415,12 +2425,12 @@ $\int_a^b s + \int_b^c s = \int_a^c s$. \end{proof} -\subsection{\partial{Exercise 1.15.11b}}% -\label{sub:exercise-1.15.11b} +\subsubsection{\partial{Exercise 1.15.11b}}% +\label{ssub:exercise-1.15.11b} $\int_a^b (s + t) = \int_a^b s + \int_a^b t$. -\note{This property mirrors the \nameref{sec:step-additive-property}.} +\note{This property mirrors the \nameref{sub:step-additive-property}.} \begin{proof} @@ -2429,7 +2439,7 @@ $\int_a^b (s + t) = \int_a^b s + \int_a^b t$. \vspace{6pt} Let $s$ and $t$ be step functions on closed interval $[a, b]$. - By definition of a step function, there exists a \nameref{sec:def-partition} + By definition of a step function, there exists a \nameref{def:partition} $P_s$ such that $s$ is constant on each open subinterval of $P_s$. Likewise, there exists a partition $P_t$ such that $t$ is constant on each open subinterval of $P_t$. @@ -2443,7 +2453,7 @@ $\int_a^b (s + t) = \int_a^b s + \int_a^b t$. $P_s$. Let $t_k$ denote the constant value of $t$ on the $k$th open subinterval of $P_t$. - By \eqref{sec:exercise-1.15.11-eq1}, + By \eqref{sub:exercise-1.15.11-eq1}, \begin{align*} \int_a^b s + t & = \sum_{k=1}^n (s_k + t_k)^3 \cdot (x_k - x_{k-1}) \\ @@ -2462,12 +2472,12 @@ $\int_a^b (s + t) = \int_a^b s + \int_a^b t$. \end{proof} -\subsection{\partial{Exercise 1.15.11c}}% -\label{sub:exercise-1.15.11c} +\subsubsection{\partial{Exercise 1.15.11c}}% +\label{ssub:exercise-1.15.11c} $\int_a^b c \cdot s = c \int_a^b s$. -\note{This property mirrors the \nameref{sec:step-homogeneous-property}.} +\note{This property mirrors the \nameref{sub:step-homogeneous-property}.} \begin{proof} @@ -2476,13 +2486,13 @@ $\int_a^b c \cdot s = c \int_a^b s$. \vspace{6pt} Let $s$ be a step function on closed interval $[a, b]$. - By definition of a step function, there exists a \nameref{sec:def-partition} + By definition of a step function, there exists a \nameref{def:partition} $P = \{x_0, x_1, \ldots, x_n\}$ such that $s$ is constant on each open subinterval of $P$. Let $s_k$ denote the constant value of $s$ on the $k$th open subinterval of $P$. Then $c \cdot s$ is a step function with step partition $P$. - By \eqref{sec:exercise-1.15.11-eq1}, + By \eqref{sub:exercise-1.15.11-eq1}, \begin{align*} \int_a^b c \cdot s & = \sum_{k=1}^n (c \cdot s_k)^3 \cdot (x_k - x_{k-1}) \\ @@ -2494,12 +2504,12 @@ $\int_a^b c \cdot s = c \int_a^b s$. \end{proof} -\subsection{\partial{Exercise 1.15.11d}}% -\label{sub:exercise-1.15.11d} +\subsubsection{\partial{Exercise 1.15.11d}}% +\label{ssub:exercise-1.15.11d} $\int_{a+c}^{b+c} s(x) \mathop{dx} = \int_a^b s(x + c) \mathop{dx}$. -\note{This property mirrors \nameref{sec:step-invariance-under-translation}.} +\note{This property mirrors \nameref{sub:step-invariance-under-translation}.} \begin{proof} @@ -2508,15 +2518,19 @@ $\int_{a+c}^{b+c} s(x) \mathop{dx} = \int_a^b s(x + c) \mathop{dx}$. \vspace{6pt} Let $s$ be a step function on closed interval $[a + c, b + c]$. - By definition of a \nameref{sec:def-step-function}, there exists a \nameref{sec:def-partition} $P = \{x_0, x_1, \ldots, x_n\}$ such that $s$ is constant on each open subinterval of $P$. - Let $s_k$ denote the constant value of $s$ on the $k$th open subinterval of $P$. + By definition of a \nameref{def:step-function}, there exists a + \nameref{def:partition} $P = \{x_0, x_1, \ldots, x_n\}$ such that $s$ is + constant on each open subinterval of $P$. + Let $s_k$ denote the constant value of $s$ on the $k$th open subinterval of + $P$. Let $c$ be a real number. - Then $t(x) = s(x + c)$ is a step function on closed interval $[a, b]$ with partition $Q = \{x_0 - c, x_1 - c, \ldots, x_n - c\}$. + Then $t(x) = s(x + c)$ is a step function on closed interval $[a, b]$ with + partition $Q = \{x_0 - c, x_1 - c, \ldots, x_n - c\}$. Furthermore, $t$ is constant on each open subinterval of $Q$. Let $t_k$ denote the value of $t$ on the $k$th open subinterval of $Q$. By construction, $t_k = s_k$. - By \eqref{sec:exercise-1.15.11-eq1}, + By \eqref{sub:exercise-1.15.11-eq1}, \begin{align*} \int_{a+c}^{b+c} s(x) \mathop{dx} & = \sum_{k=1}^n s_k^3 \cdot (x_k - x_{k-1}) \\ @@ -2528,12 +2542,12 @@ $\int_{a+c}^{b+c} s(x) \mathop{dx} = \int_a^b s(x + c) \mathop{dx}$. \end{proof} -\subsection{\partial{Exercise 1.15.11e}}% -\label{sub:exercise-1.15.11e} +\subsubsection{\partial{Exercise 1.15.11e}}% +\label{ssub:exercise-1.15.11e} If $s(x) < t(x)$ for each $x$ in $[a, b]$, then $\int_a^b s < \int_a^b t$. -\note{This property mirrors the \nameref{sec:step-comparison-theorem}.} +\note{This property mirrors the \nameref{sub:step-comparison-theorem}.} \begin{proof} @@ -2542,8 +2556,9 @@ If $s(x) < t(x)$ for each $x$ in $[a, b]$, then $\int_a^b s < \int_a^b t$. \vspace{6pt} Let $s$ and $t$ be step functions on closed interval $[a, b]$. - By definition of a \nameref{sec:def-step-function}, there exists a \nameref{sec:def-partition} - $P_s$ such that $s$ is constant on each open subinterval of $P_s$. + By definition of a \nameref{def:step-function}, there exists a + \nameref{def:partition} $P_s$ such that $s$ is constant on each open + subinterval of $P_s$. Likewise, there exists a partition $P_t$ such that $t$ is constant on each open subinterval of $P_t$. Let $$P = P_s \cup P_t = \{x_0, x_1, \ldots, x_n\}$$ be the common refinement @@ -2555,7 +2570,7 @@ If $s(x) < t(x)$ for each $x$ in $[a, b]$, then $\int_a^b s < \int_a^b t$. $P$. Let $t_k$ denote the constant value of $t$ on the $k$th open subinterval of $P$. - By \eqref{sec:exercise-1.15.11-eq1}, + By \eqref{sub:exercise-1.15.11-eq1}, \begin{align*} \int_a^b s & = \sum_{k=1}^n s_k^3 \cdot (x_k - x_{k-1}) \\ @@ -2565,11 +2580,11 @@ If $s(x) < t(x)$ for each $x$ in $[a, b]$, then $\int_a^b s < \int_a^b t$. \end{proof} -\chapter{Theorey of Integrability}% -\label{chap:theory-integrability} +\section{Upper and Lower Integrals}% +\label{sec:upper-lower-integrals} -\section{\partial{Theorem 1.9}}% -\label{sec:theorem-1.9} +\subsection{\partial{Theorem 1.9}}% +\label{sub:theorem-1.9} \begin{theorem}[1.9] @@ -2577,12 +2592,12 @@ If $s(x) < t(x)$ for each $x$ in $[a, b]$, then $\int_a^b s < \int_a^b t$. $\ubar{I}(f)$ and an upper integral $\overline{I}(f)$ satisfying the inequalities \begin{equation} - \label{sec:theorem-1.9-eq1} + \label{sub:theorem-1.9-eq1} \int_a^b s(x) \mathop{dx} \leq \ubar{I}(f) \leq \bar{I}(f) \leq \int_a^b t(x) \mathop{dx} \end{equation} - for all \nameref{sec:def-step-function}s $s$ and $t$ with $s \leq f \leq t$. - The function $f$ is \nameref{sec:def-integrable} on $[a, b]$ if and only if + for all \nameref{def:step-function}s $s$ and $t$ with $s \leq f \leq t$. + The function $f$ is \nameref{def:integrable} on $[a, b]$ if and only if its upper and lower integrals are equal, in which case we have $$\int_a^b f(x) \mathop{dx} = \ubar{I}(f) = \bar{I}(f).$$ @@ -2592,7 +2607,7 @@ If $s(x) < t(x)$ for each $x$ in $[a, b]$, then $\int_a^b s < \int_a^b t$. Let $f$ be a function bounded on $[a, b]$. We prove that (i) $f$ has a lower and upper integral satisfying - \eqref{sec:theorem-1.9-eq1} and (ii) that $f$ is integrable on $[a, b]$ if + \eqref{sub:theorem-1.9-eq1} and (ii) that $f$ is integrable on $[a, b]$ if and only if its lower and upper integrals are equal. \paragraph{(i)}% @@ -2612,13 +2627,11 @@ If $s(x) < t(x)$ for each $x$ in $[a, b]$, then $\int_a^b s < \int_a^b t$. By construction, $s \leq t$ for every $s$ in $S$ and $t$ in $T$. Therefore \nameref{sec:theorem-i.34} tells us $S$ has a - \nameref{sec:def-supremum}, $T$ has an \nameref{sec:def-infimum}, and + \nameref{def:supremum}, $T$ has an \nameref{def:infimum}, and $\sup{S} \leq \inf{T}$. - By definition of the \nameref{sec:def-lower-integral}, - $\ubar{I}(f) = \sup{S}$. - By definition of the \nameref{sec:def-upper-integral}, - $\bar{I}(f) = \inf{S}$. - Thus \eqref{sec:theorem-1.9-eq1} holds. + By definition of the \nameref{def:lower-integral}, $\ubar{I}(f) = \sup{S}$. + By definition of the \nameref{def:upper-integral}, $\bar{I}(f) = \inf{S}$. + Thus \eqref{sub:theorem-1.9-eq1} holds. \paragraph{(ii)}% @@ -2626,20 +2639,22 @@ If $s(x) < t(x)$ for each $x$ in $[a, b]$, then $\int_a^b s < \int_a^b t$. there exists one and only one number $I$ such that $$\int_a^b s(x) \mathop{dx} \leq I \leq \int_a^b t(x) \mathop{dx}$$ for every pair of step functions $s$ and $t$ satisfying - \eqref{sec:def-integral-bounded-function-eq1}. - By \eqref{sec:theorem-1.9-eq1} and the definition of the supremum/infimum, + \eqref{def:integral-bounded-function-eq1}. + By \eqref{sub:theorem-1.9-eq1} and the definition of the supremum/infimum, this holds if and only if $\ubar{I}(f) = \bar{I}(f)$, concluding the proof. \end{proof} -\section{\partial{Measurability of Ordinate Sets}}% -\label{sec:measurability-ordinate-sets} -\label{sec:theorem-1.10} +\section{The Area of an Ordinate Set Expressed as an Integral}% +\label{sec:area-ordinate-set-expressed-integral} + +\subsection{\partial{Theorem 1.10}}% +\label{sub:theorem-1.10} \begin{theorem}[1.10] - Let $f$ be a nonnegative function, \nameref{sec:def-integrable} on an interval + Let $f$ be a nonnegative function, \nameref{def:integrable} on an interval $[a, b]$, and let $Q$ denote the ordinate set of $f$ over $[a, b]$. Then $Q$ is measurable and its area is equal to the integral $\int_a^b f(x) \mathop{dx}$. @@ -2648,29 +2663,28 @@ If $s(x) < t(x)$ for each $x$ in $[a, b]$, then $\int_a^b s < \int_a^b t$. \begin{proof} - Let $f$ be a nonnegative function, \nameref{sec:def-integrable} on $[a, b]$. + Let $f$ be a nonnegative function, \nameref{def:integrable} on $[a, b]$. By definition of integrability, there exists one and only one number $I$ such that $$\int_a^b s(x) \mathop{dx} \leq I \leq \int_a^b t(x) \mathop{dx}$$ for every pair of step functions $s$ and $t$ satisfying - \eqref{sec:def-integral-bounded-function-eq1}. + \eqref{def:integral-bounded-function-eq1}. In other words, $I$ is the one and only number that satisfies $$a(S) \leq I \leq a(T)$$ for every pair of step regions $S \subseteq Q \subseteq T$. - By the \nameref{sec:area-exhaustion-property}, $Q$ is measurable and its area + By the \nameref{sub:area-exhaustion-property}, $Q$ is measurable and its area is equal to $I = \int_a^b f(x) \mathop{dx}$. \end{proof} -\section{\partial{Measurability of the Graph of a Nonnegative Function}}% -\label{sec:measurability-graph-nonnegative-function} -\label{sec:theorem-1.11} +\subsection{\partial{Theorem 1.11}}% +\label{sub:theorem-1.11} \begin{theorem}[1.11] Let $f$ be a nonnegative function, integrable on an interval $[a, b]$. Then the graph of $f$, that is, the set \begin{equation} - \label{sec:theorem-1.11-eq1} + \label{sub:theorem-1.11-eq1} \{(x, y) \mid a \leq x \leq b, y = f(x)\}, \end{equation} is measurable and has area equal to $0$. @@ -2692,38 +2706,40 @@ If $s(x) < t(x)$ for each $x$ in $[a, b]$, then $\int_a^b s < \int_a^b t$. such that $$\int_a^b s(x) \mathop{dx} \leq I \leq \int_a^b t(x) \mathop{dx}$$ for every pair of step functions $s$ and $t$ satisfying - \eqref{sec:def-integral-bounded-function-eq1}. + \eqref{def:integral-bounded-function-eq1}. In other words, $I$ is the one and only number that satisfies $$a(S) \leq I \leq a(T)$$ for every pair of step regions $S \subseteq Q' \subseteq T$. - By the \nameref{sec:area-exhaustion-property}, $Q'$ is measurable and its + By the \nameref{sub:area-exhaustion-property}, $Q'$ is measurable and its area is equal to $I = \int_a^b f(x) \mathop{dx}$. \paragraph{(ii)}% Let $Q$ denote the ordinate set of $f$. - By \nameref{sec:measurability-ordinate-sets}, $Q$ is measurable with area - equal to the integral $I = \int_a^b f(x) \mathop{dx}$. + By \nameref{sub:theorem-1.10}, $Q$ is measurable with area equal to the + integral $I = \int_a^b f(x) \mathop{dx}$. By \nameref{par:theorem-1.11-i}, $Q'$ is measurable with area also equal to $I$. - We note the graph of $f$, \eqref{sec:theorem-1.11-eq1}, is equal to set + We note the graph of $f$, \eqref{sub:theorem-1.11-eq1}, is equal to set $Q - Q'$. - By the \nameref{sec:area-difference-property}, $Q - Q'$ is measurable and + By the \nameref{sub:area-difference-property}, $Q - Q'$ is measurable and $$a(Q - Q') = a(Q) - a(Q') = I - I = 0.$$ Thus the graph of $f$ is measurable and has area equal to $0$. \end{proof} \section - [\partial{Integrability of Bounded Monotonic Functions}] - {\partial{Integrability of Bounded Monotonic \texorpdfstring{\\}{}Functions}} + [Integrability of Bounded Monotonic Functions] + {Integrability of Bounded Monotonic \texorpdfstring{\\}{}Functions} \label{sec:integrability-bounded-monotonic-functions} -\label{sec:theorem-1.12} + +\subsection{\partial{Theorem 1.12}}% +\label{sub:theorem-1.12} \begin{theorem}[1.12] - If $f$ is \nameref{sec:def-monotonic} on a closed interval $[a, b]$, then $f$ - is \nameref{sec:def-integrable} on $[a, b]$. + If $f$ is \nameref{def:monotonic} on a closed interval $[a, b]$, then $f$ + is \nameref{def:integrable} on $[a, b]$. \end{theorem} @@ -2733,8 +2749,8 @@ If $s(x) < t(x)$ for each $x$ in $[a, b]$, then $\int_a^b s < \int_a^b t$. That is to say, either $f$ is increasing on $[a, b]$ or $f$ is decreasing on $[a, b]$. Because $f$ is on a closed interval, it is bounded. - By \nameref{sec:theorem-1.9}, $f$ has a \nameref{sec:def-lower-integral} - $\ubar{I}(f)$, $f$ has an \nameref{sec:def-upper-integral} $\bar{I}(f)$, + By \nameref{sub:theorem-1.9}, $f$ has a \nameref{def:lower-integral} + $\ubar{I}(f)$, $f$ has an \nameref{def:upper-integral} $\bar{I}(f)$, and $f$ is integrable if and only if $\ubar{I}(f) = \bar{I}(f)$. Consider a partition $P = \{x_0, x_1, \ldots, x_n\}$ of $[a, b]$ in which @@ -2748,13 +2764,13 @@ If $s(x) < t(x)$ for each $x$ in $[a, b]$, then $\int_a^b s < \int_a^b t$. on every $k$th open subinterval of $P$. Let $t$ be the step function above $f$ with constant value $f(x_k)$ on every $k$th open subinterval of $P$. - Then, by \eqref{sec:theorem-1.9-eq1}, it follows + Then, by \eqref{sub:theorem-1.9-eq1}, it follows \begin{equation} - \label{sec:theorem-1.12-eq1} + \label{sub:theorem-1.12-eq1} \int_a^b s(x) \mathop{dx} \leq \ubar{I}(f) \leq \bar{I}(f) \leq \int_a^b t(x) \mathop{dx}. \end{equation} - By definition of the \nameref{sec:def-integral-step-function}, + By definition of the \nameref{def:integral-step-function}, \begin{align*} \int_a^b s(x) \mathop{dx} & = \sum_{k=1}^n f(x_{k-1})\left[\frac{b - a}{n}\right] \\ @@ -2769,7 +2785,7 @@ If $s(x) < t(x)$ for each $x$ in $[a, b]$, then $\int_a^b s < \int_a^b t$. & = \left[\frac{b - a}{n}\right] \sum_{k=1}^n f(x_k) - f(x_{k-1}) \\ & = \frac{(b - a)(f(b) - f(a))}{n}. \end{align*} - By \eqref{sec:theorem-1.12-eq1}, + By \eqref{sub:theorem-1.12-eq1}, \begin{align*} \ubar{I}(f) & \leq \bar{I}(f) \\ @@ -2787,13 +2803,13 @@ If $s(x) < t(x)$ for each $x$ in $[a, b]$, then $\int_a^b s < \int_a^b t$. on every $k$th open subinterval of $P$. Let $t$ be the step function above $f$ with constant value $f(x_{k-1})$ on every $k$th open subinterval of $P$. - Then, by \eqref{sec:theorem-1.9-eq1}, it follows + Then, by \eqref{sub:theorem-1.9-eq1}, it follows \begin{equation} - \label{sec:theorem-1.12-eq2} + \label{sub:theorem-1.12-eq2} \int_a^b s(x) \mathop{dx} \leq \ubar{I}(f) \leq \bar{I}(f) \leq \int_a^b t(x) \mathop{dx}. \end{equation} - By definition of the \nameref{sec:def-integral-step-function}, + By definition of the \nameref{def:integral-step-function}, \begin{align*} \int_a^b s(x) \mathop{dx} & = \sum_{k=1}^n f(x_k)\left[\frac{b - a}{n}\right] \\ @@ -2808,7 +2824,7 @@ If $s(x) < t(x)$ for each $x$ in $[a, b]$, then $\int_a^b s < \int_a^b t$. & = \left[\frac{b - a}{n}\right] \sum_{k=1}^n f(x_{k-1}) - f(x_k) \\ & = \frac{(b - a)(f(a) - f(b))}{n}. \end{align*} - By \eqref{sec:theorem-1.12-eq2}, + By \eqref{sub:theorem-1.12-eq2}, \begin{align*} \ubar{I}(f) & \leq \bar{I}(f) \\ @@ -2821,10 +2837,8 @@ If $s(x) < t(x)$ for each $x$ in $[a, b]$, then $\int_a^b s < \int_a^b t$. \end{proof} -\section{\partial{% - Calculation of the Integral of a Bounded Increasing Function}}% -\label{sec:calculation-integral-bounded-increasing-function} -\label{sec:theorem-1.13} +\subsection{\partial{Theorem 1.13}}% +\label{sub:theorem-1.13} \begin{theorem}[1.13] @@ -2832,7 +2846,7 @@ If $s(x) < t(x)$ for each $x$ in $[a, b]$, then $\int_a^b s < \int_a^b t$. Let $x_k = a + k(b - a) / n$ for $k = 0, 1, \ldots, n$. If $I$ is any number which satisfies the inequalities \begin{equation} - \label{sec:theorem-1.13-eq1} + \label{sub:theorem-1.13-eq1} \frac{b - a}{n} \sum_{k=0}^{n-1} f(x_k) \leq I \leq \frac{b - a}{n} \sum_{k=1}^n f(x_k) @@ -2844,12 +2858,12 @@ If $s(x) < t(x)$ for each $x$ in $[a, b]$, then $\int_a^b s < \int_a^b t$. \begin{proof} Let $f$ be increasing on a closed interval $[a, b]$ and $I$ be a number - satisfying \eqref{sec:theorem-1.13-eq1}. + satisfying \eqref{sub:theorem-1.13-eq1}. Let $s$ be the step function below $f$ with constant value $f(x_{k-1})$ on every $k$th open subinterval of $P$. Let $t$ be the step function above $f$ with constant value $f(x_k)$ on every $k$th open subinterval of $P$. - By definition of the \nameref{sec:def-integral-step-function}, + By definition of the \nameref{def:integral-step-function}, \begin{align*} \int_a^b s(x) \mathop{dx} & = \sum_{k=1}^n f(x_{k-1})\left[\frac{b - a}{n}\right] \\ @@ -2857,20 +2871,20 @@ If $s(x) < t(x)$ for each $x$ in $[a, b]$, then $\int_a^b s < \int_a^b t$. \int_a^b t(x) \mathop{dx} & = \sum_{k=1}^n f(x_k)\left[\frac{b - a}{n}\right]. \end{align*} - Therefore \eqref{sec:theorem-1.13-eq1} can alternatively be written as + Therefore \eqref{sub:theorem-1.13-eq1} can alternatively be written as \begin{equation} - \label{sec:theorem-1.13-eq2} + \label{sub:theorem-1.13-eq2} \int_a^b s(x) \mathop{dx} \leq I \leq \int_a^b t(x) \mathop{dx}. \end{equation} - By \nameref{sec:theorem-1.12}, $f$ is integrable. - Therefore \nameref{sec:theorem-1.9} indicates $f$ satisfies + By \nameref{sub:theorem-1.12}, $f$ is integrable. + Therefore \nameref{sub:theorem-1.9} indicates $f$ satisfies \begin{equation} - \label{sec:theorem-1.13-eq3} + \label{sub:theorem-1.13-eq3} \int_a^b s(x) \mathop{dx} \leq \int_a^b f(x) \mathop{dx} \leq \int_a^b t(x) \mathop{dx}. \end{equation} - Manipulating \eqref{sec:theorem-1.13-eq2} and \eqref{sec:theorem-1.13-eq3} + Manipulating \eqref{sub:theorem-1.13-eq2} and \eqref{sub:theorem-1.13-eq3} together yields \begin{align*} I - \int_a^b f(x) \mathop{dx} @@ -2895,10 +2909,8 @@ If $s(x) < t(x)$ for each $x$ in $[a, b]$, then $\int_a^b s < \int_a^b t$. \end{proof} -\section{\partial{% - Calculation of the Integral of a Bounded Decreasing Function}}% -\label{sec:calculation-integral-bounded-decreasing-function} -\label{sec:theorem-1.14} +\subsection{\partial{Theorem 1.14}}% +\label{sub:theorem-1.14} \begin{theorem}[1.14] @@ -2906,7 +2918,7 @@ If $s(x) < t(x)$ for each $x$ in $[a, b]$, then $\int_a^b s < \int_a^b t$. Let $x_k = a + k(b - a) / n$ for $k = 0, 1, \ldots, n$. If $I$ is any number which satisfies the inequalities \begin{equation} - \label{sec:theorem-1.14-eq1} + \label{sub:theorem-1.14-eq1} \frac{b - a}{n} \sum_{k=1}^n f(x_k) \leq I \leq \frac{b - a}{n} \sum_{k=0}^{n-1} f(x_k) @@ -2918,12 +2930,12 @@ If $s(x) < t(x)$ for each $x$ in $[a, b]$, then $\int_a^b s < \int_a^b t$. \begin{proof} Let $f$ be decreasing on a closed interval $[a, b]$ and $I$ be a number - satisfying \eqref{sec:theorem-1.14-eq1}. + satisfying \eqref{sub:theorem-1.14-eq1}. Let $s$ be the step function below $f$ with constant value $f(x_k)$ on every $k$th open subinterval of $P$. Let $t$ be the step function above $f$ with constant value $f(x_{k-1})$ on every $k$th open subinterval of $P$. - By definition of the \nameref{sec:def-integral-step-function}, + By definition of the \nameref{def:integral-step-function}, \begin{align*} \int_a^b s(x) \mathop{dx} & = \sum_{k=1}^n f(x_k) \left[\frac{b - a}{n}\right] \\ @@ -2931,20 +2943,20 @@ If $s(x) < t(x)$ for each $x$ in $[a, b]$, then $\int_a^b s < \int_a^b t$. & = \sum_{k=1}^n f(x_{k-1}) \left[\frac{b - a}{n}\right] \\ & = \sum_{k=0}^{n-1} f(x) \left[\frac{b - a}{n}\right]. \end{align*} - Therefore \eqref{sec:theorem-1.14-eq1} can alternatively be written as + Therefore \eqref{sub:theorem-1.14-eq1} can alternatively be written as \begin{equation} - \label{sec:theorem-1.14-eq2} + \label{sub:theorem-1.14-eq2} \int_a^b s(x) \mathop{dx} \leq I \leq \int_a^b t(x) \mathop{dx}. \end{equation} - By \nameref{sec:theorem-1.12}, $f$ is integrable. - Therefore \nameref{sec:theorem-1.9} indicates $f$ satisfies + By \nameref{sub:theorem-1.12}, $f$ is integrable. + Therefore \nameref{sub:theorem-1.9} indicates $f$ satisfies \begin{equation} - \label{sec:theorem-1.14-eq3} + \label{sub:theorem-1.14-eq3} \int_a^b s(x) \mathop{dx} \leq \int_a^b f(x) \mathop{dx} \leq \int_a^b t(x) \mathop{dx}. \end{equation} - Manipulating \eqref{sec:theorem-1.14-eq2} and \eqref{sec:theorem-1.14-eq3} + Manipulating \eqref{sub:theorem-1.14-eq2} and \eqref{sub:theorem-1.14-eq3} together yields \begin{align*} I - \int_a^b f(x) \mathop{dx} @@ -2969,11 +2981,11 @@ If $s(x) < t(x)$ for each $x$ in $[a, b]$, then $\int_a^b s < \int_a^b t$. \end{proof} -\section{\unverified{% - Calculation of the Integral \texorpdfstring{$\int_0^b x^p \mathop{dx}$}{int-x-p} - when \texorpdfstring{$p$}{p} is a Positive Integer}}% -\label{sec:calculation-integral-int-x-p-p-positive-integer} -\label{sec:theorem-1.15} +\subsection{\unverified{% + Integral of \texorpdfstring{$\int_0^b x^p \mathop{dx}$}{int-x-p} when + \texorpdfstring{$p$}{p} is a Positive Integer}}% +\label{sub:calculation-integral-int-x-p-p-positive-integer} +\label{sub:theorem-1.15} \begin{theorem}[1.15] diff --git a/Bookshelf/Enderton_Logic.tex b/Bookshelf/Enderton_Logic.tex index 243aeb4..37c3384 100644 --- a/Bookshelf/Enderton_Logic.tex +++ b/Bookshelf/Enderton_Logic.tex @@ -5,16 +5,21 @@ \begin{document} -\header - {A Mathematical Introduction to Logic} - {Herbert B. Enderton} +\header{A Mathematical Introduction to Logic}{Herbert B. Enderton} \tableofcontents -% Sets first chapter to `0` to match Enderton book. +\begingroup +\renewcommand\thechapter{G} + +\chapter{Glossary}% +\label{chap:glossary} + +\endgroup + +% Reset counter to mirror Enderton's book. \setcounter{chapter}{0} \addtocounter{chapter}{-1} - \chapter{Useful Facts About Sets}% \label{chap:useful-facts-about-sets}