Discover millions of ebooks, audiobooks, and so much more with a free trial

Only $11.99/month after trial. Cancel anytime.

Calculus: An Intuitive and Physical Approach (Second Edition)
Calculus: An Intuitive and Physical Approach (Second Edition)
Calculus: An Intuitive and Physical Approach (Second Edition)
Ebook2,017 pages18 hours

Calculus: An Intuitive and Physical Approach (Second Edition)

Rating: 4.5 out of 5 stars

4.5/5

()

Read preview

About this ebook

Application-oriented introduction relates the subject as closely as possible to science. In-depth explorations of the derivative, the differentiation and integration of the powers of x, and theorems on differentiation and antidifferentiation lead to a definition of the chain rule and examinations of trigonometric functions, logarithmic and exponential functions, techniques of integration, polar coordinates, much more. Clear-cut explanations, numerous drills, illustrative examples. 1967 edition. Solution guide available upon request.
LanguageEnglish
Release dateMay 9, 2013
ISBN9780486134765
Calculus: An Intuitive and Physical Approach (Second Edition)

Related to Calculus

Titles in the series (100)

View More

Related ebooks

Mathematics For You

View More

Related articles

Reviews for Calculus

Rating: 4.375 out of 5 stars
4.5/5

8 ratings0 reviews

What did you think?

Tap to rate

Review must be at least 10 words

    Book preview

    Calculus - Morris Kline

    CALCULUS

    An Intuitive and Physical Approach

    Morris Kline

    SECOND EDITION

    DOVER PUBLICATIONS, INC.

    Mineola, New York

    Copyright

    Copyright © 1967, 1977 by John Wiley & Sons, Inc.

    All rights reserved.

    Bibliographical Note

    This Dover edition, first published in 1998, is an unabridged republication of the work originally published in 1977 by John Wiley and Sons, Inc., New York.

    Readers who would like to receive the solutions to the exercises may request them from the publisher at the following e-mail address:

    editors@doverpublications.com

    Library of Congress Cataloging-in-Publication Data

    Kline, Morris, 1908–

    Calculus : an intuitive and physical approach, second edition / Morris Kline.

                p. cm.

    An unabridged republication of the work originally published in 1977 by John Wiley and Sons, Inc., New York—T.p. verso.

    Includes index.

    ISBN-13: 978-0-486-40453-0

    ISBN-10: 0-486-40453-6

    1. Calculus. I. Title.

    QA303.K68 1998

    515—dc21

    98-36211

    CIP

    Manufactured in the United States by Courier Corporation

    40453609

    www.doverpublications.com

    Dedicated to the memory of

    Morris Kline

    1908–1992

    Mathematician, Educator

    PREFACE TO

    THE SECOND EDITION

    The basic features of the first edition have been retained, such as the intuitive approach and real applications. As to the approach, the last chapter introduces a rigorous treatment. Though this chapter could be used in conjunction with the opening chapters of the book, I do not recommend doing so; the rigorous presentation is difficult to grasp and obscures the understanding. Rigor undoubtedly refines the intuition but does not supplant it. The rigorous approach should be reserved for a course in advanced calculus for mathematicians.

    Most of the applications still belong with the physical sciences; however, no knowledge of physics is presupposed. Several sections on physical applications have been dropped and applications to the social and biological sciences have been added instead. These applications are vital. The theory and technique of the calculus are, in themselves, meaningless. Moreover, since most students who take calculus will be scientists or engineers, they will be highly motivated by the applications. Many calculus texts dispose of applications by asking students to calculate centers of gravity and moments of inertia. But since students have no idea of how these quantities are used, the only consequence is that the gravity of the problems produces inertia in the students.

    In addition to extending the scope of the applications, I have made a number of other improvements. More drill exercises have been added, the exercises have been carefully graded as to difficulty, and there are more illustrative examples. Students doing homework exercises acquire the habit of searching for an illustrative example which they can imitate and thereby do the exercises without thinking. To counter this many illustrative examples are incorporated into the text instead of being set apart formally. Hopefully the students will read the text instead of mechanically following the technique of the usual illustrative example.

    A chapter on differential equations and a section on such numerical methods as Simpson’s rule and the trapezoidal rule have also been added. More use has been made of vector analysis, particularly in treating velocity and acceleration.

    Chapter 9, The Definite Integral, has been moved forward so that students using the mathematics in another course taken simultaneously can utilize the concept sooner.

    Some changes in notation are unquestionably an improvement. The notation has been restricted to the customary situation where time is the independent variable; in other situations y′, dy / dx and f′(x) are used. The notation ∫y dx has been introduced early to denote antidifferentiation. The use of dy / dx from the outset is the result of numerous requests. Admittedly, a useful notation in showing the variables involved, it also suggests that the derivative is a quotient, whereas we must take great pains to convince the student that it is not.

    The pace of the first few chapters has been speeded up somewhat. However, it does seem desirable, because students are weak in algebra, to keep the algebra simple at the outset while students are acquiring the concepts of the calculus.

    Two other changes may be helpful. Since not all classes go at the same pace or have the same objectives the two volumes of the first edition have been replaced by the present single volume. Thereby instructors are freer to choose the topics they deem most appropriate. Second, since time may not permit the inclusion of all the applications, those sections that can be omitted without disrupting the continuity have been starred.

    Some figures have been improved; where the precise shape is significant the computer has been employed to achieve accuracy.

    An instructor’s manual which contains full solutions of all the exercises, suggestions for teaching, and additional material for advanced students is available to instructors on request to the publisher.

    Morris Kline

    New York, N. Y.

    August, 1976

    PREFACE TO

    THE FIRST EDITION

    Anyone who adds to the plethora of introductory calculus texts owes an explanation, if not an apology, to the mathematical community. I believe that an introductory course in the calculus should present the subject intuitively and should relate it as closely as possible to science. A text for such an approach is not available.

    In this book the justification of the theorems and techniques is consistently intuitive; that is, geometrical, physical, and heuristic arguments and generalizations from concrete cases are employed to convince. The approach is especially suitable for the calculus because the subject grew out of physical and geometrical problems. These problems tell us what functions we should take up, what concepts we want to formulate, and what techniques we should develop. In view of the fact that the human mind learns intuitively and that time does not permit both an intuitive and a rigorous presentation in elementary calculus, it seems to me that the approach adopted is the correct one.

    The intuitive approach is explained to the student so that he will know what kind of evidence is being used to support arguments. Thus he is told that a graph of a typical function may not represent all functions. On the other hand, he is also told that the elementary functions are well behaved except at isolated points and that he can usually trust his intuition. As he works with the ideas of the calculus, he will sharpen his intuition. If he continues with mathematics, he will learn the analytical foundations and proofs that guard against the failings of intuition.

    The use of an intuitive approach in the first treatment of a subject is not an innovation. Arithmetic is learned intuitively in elementary school and then the logic of it is learned gradually through the work in algebra. Geometry is learned intuitively in junior high school and then the formal deductive approach is presented in senior high school. A difficult subject such as the calculus, therefore, should certainly be introduced by an intuitive approach.

    This approach has many merits beyond that of being the only feasible one. Every pedagogue today champions discovery, but few teach it. How does one discover in mathematics? By thinking in physical and geometrical terms, by conjecturing or guessing, by formulating hypotheses and testing them, and by generalizing on specific cases. Physical problems that call for the creation of mathematics set the stage for discovery. That the intuitive approach may lead to errors is granted, but truth emerges more readily from error than from confusion. The student must be allowed to make mistakes, for if he makes no mistakes, he will not progress.

    After the basic material has been covered, the need for rigor is motivated, and the last two chapters of Part 2 of this book do offer an introduction to precise language and proof. These chapters are intended as a transition to advanced calculus.

    One alternative to an intuitive approach is a rigorous treatment. In my opinion, a rigorous first course in the calculus is ill advised for numerous reasons. First, it is too difficult for the students. Beginners are asked to learn a mass of concepts so subtle that they defied the best mathematicians for two hundred years. Even Cauchy, the founder of rigor, gave formulations that are crude compared to what the current rigorous presentations ask students to absorb. And Cauchy, despite his concern for rigor, missed the distinctions between continuity and differentiability and between convergence and uniform convergence. Before one can appreciate a precise formulation of a concept or theorem, he must know what idea is being formulated and what exceptions or pitfalls the wording is trying to avoid. Hence he must be able to call upon a wealth of experience acquired before tackling the rigorous formulation. Furthermore, having students master a polished deductive organization does not teach them how to think and how to do mathematics, for thinking and doing are not deductive processes. How can discovery take place when students are asked to work with ideas that are already overladen with sophistication and refinement? Finally, the rigorous approach is misleading. Because the introductory calculus course is the student’s first contact with higher mathematics, he obtains the impression that real mathematics is deductive and that good mathematicians think deductively.

    Rigor has its place in mathematics education. It is a check on the creations and it permits an aesthetic (as well as an anaesthetic) presentation. But it is also to some extent gilt on the lily and an interdiction against the inclusion of functions which rarely occur in practice and which must even be invented with Weierstrassian ingenuity. A rigorous first course in calculus reminds one of the words of Samuel Johnson; I have found you an argument but I am not obliged to find you an understanding. Even if the rigorous material is understood, its value is limited. As Henri Lebesgue pointed out: Logic makes us reject certain arguments but it cannot make us believe any argument.

    The reasons often given for a rigorous presentation—that students must learn what a real proof is or that students should not be asked to unlearn later what they have already been taught—are hollow. One cannot give the whole truth at once in any subject. Even we, as teachers, do not face the whole truth for, in fact, the whole question of what is rigorous in mathematics was never so much up in the air as it is now. At any rate, what may seem ideally right and efficient is pedagogically intolerable.

    There is also the alternative of compromising on rigor by offering precise definitions and proofs in some portions of the text and intuitive or pseudorigorous arguments in others. This alternative seems to me to have almost all of the disadvantages of a rigorous presentation and the additional one of confusing the student about what proof really is.

    The second essential respect in which this book differs from current ones is that the relationship of mathematics to science is taken seriously. The present trend to separate mathematics from science is tragic. There are chapters of mathematics that have value in and for themselves. However, the calculus divorced from applications is meaningless. We should also keep in mind that most of the students taking calculus will be scientists and engineers, and these students must learn how to use mathematics. But the step from mathematics to its applications is not simple and straightforward and it creates difficulties for the student from the time he is called upon to solve verbal problems in algebra. The mathematics courses fail to teach students how to formulate physical problems mathematically. The science and engineering courses, on the other hand, assume that students know how to translate physical problems into mathematical language and how to make satisfactory idealizations. The gap between mathematics and science instruction must be filled, and we can do so to our own advantage because thereby we give meaning and motivation to the calculus.

    In this book real problems are used to motivate the mathematics, and the latter, once developed, is applied to genuine physical problems—the magnificent, impressive, and even beautiful problems tendered by nature. I have selected those that do not require a background in physics. However, if the student is to think properly even about familiar concepts, such as weight, force, velocity, acceleration, light, and work, the book must say a little about them and about the physical laws that are involved. It does. To relegate physical problems only to the exercises is, in my opinion, ineffective in teaching students how to apply mathematics. If the student is not informed about the physics of a problem, all he can do in tackling an exercise is to apply a formula mechanically or guess what technical process is called for, do it, and check his answer. If the book does not supply an answer and if the student’s result is off by a factor of 1000, he will not realize it because he has no judgment about what to expect.

    I have tried to incorporate several other features that may contribute to the pedagogy. In many instances I have deliberately made false starts so as to have the student realize that correct methods and correct proofs are almost always preceded by groping and to have students appreciate why we finally take one course rather than another. I wish to dispel the impression that good mathematicians are able to proceed directly to the right conclusion because strict logic or a God-given insight guides them. I have given full details on the mathematical steps in the hope that the book will be readable. The style is informal.

    The pace of the first few chapters of Part 1 has been deliberately made slow. It is well known that calculus students are still struggling with elementary algebra. By keeping the opening algebra simple, I trust that the calculus ideas will stand forth. In later chapters the algebra becomes more difficult.

    The order of the mathematical concepts and topics was chosen so that applications could be introduced almost from the beginning. However, no artificial or complicated arrangement was necessary. Differentiation and antidifferentiation are the key concepts in the first thirteen chapters and then summation is introduced. The early introduction of applications should permit good correlation with physics courses if these are taken concurrently. The concept of differentials was deliberately delayed until Chapter 12 of Part 1. Most books do stress that the derivative is a limit of a quotient, but often present differentials early. No matter how much the student is cautioned that the derivative is not a quotient, if he is allowed to work almost at once with the quotient of the differentials, he tends to forget the true meaning of the derivative.

    Beyond the matter of the proper approach, there are a few other points that may warrant comment. Despite my own preference for a course in analytic geometry which precedes the calculus, I have included the former. This is a concession to the current fad. Some work with vectors is included, but I have not done much with the differentiation and integration of vector functions. I believe that this topic requires a new way of thinking and that we should not try to teach two major classes of techniques at the same time. The calculus of scalar functions is the basic one, and this should be mastered first. Apropos of the inclusion of vector analysis, the present trend to include this topic, linear algebra, probability, and differential equations in a three- or four-semester analytic geometry and calculus course, seems to me to be attempting the impossible.

    Any book should be adjustable to the needs of a particular course. Most of the applications are in separate sections, and therefore the number of those that need to be included is optional. Even whole chapters, such as Chapter 13 of Part 1 on further techniques of integration, Chapter 16 of Part 1 on further physical applications of the definite integral, and Chapter 19 of Part 2 on polar parametric equations, can be omitted. The last two chapters of Part 2 (Chapters 24 and 25), the introduction to rigor, can be omitted or, on the other hand, they can be read before the chapters on solid analytic geometry and functions of two or more variables.

    I wish to express my thanks to the publishers for their generous support of this approach to the calculus and for their efficient handling of the production.

    Morris Kline

    New York

    November 1966

    CONTENTS

    CHAPTER 1  WHY CALCULUS?

    1.  The Historical Motivations for the Calculus

    2.  The Creators of the Calculus

    3.  The Nature of the Calculus

    CHAPTER 2  THE DERIVATIVE

    1.  The Concept of Function

    2.  The Graph or Curve of a Function

    3.  Average and Instantaneous Speed

    4.  The Method of Increments

    5.  A Matter of Notation

    6.  The Method of Increments Applied to y = ax²

    7.  The Derived Function

    8.  The Differentiation of Simple Monomials

    9.  The Differentiation of Simple Polynomials

    10.  The Second Derivative

    CHAPTER 3  THE ANTIDERIVED FUNCTION OR THE INTEGRAL

    1.  The Integral

    2.  Straight Line Motion in One Direction

    3.  Up and Down Motion

    4.  Motion Along an Inclined Plane

    APPENDIX  The Coordinate Geometry of Straight Lines

    A1.  The Need for Geometrical Interpretation

    A2.  The Distance Formula

    A3.  The Slope of a Straight Line

    A4.  The Inclination of a Line

    A5.  Slopes of Parallel and Perpendicular Lines

    A6.  The Angle Between Two Lines

    A7.  The Equation of a Straight Line

    A8.  The Distance from a Point to a Line

    A9.  Equation and Curve

    CHAPTER 4  THE GEOMETRICAL SIGNIFICANCE OF THE DERIVATIVE

    1.  The Derivative as Slope

    2.  The Concept of Tangent to a Curve

    3.  Applications of the Derivative as the Slope

    4.  The Equation of the Parabola

    5.  Physical Applications of the Derivative as Slope

    6.  Further Discussion of the Derivative as the Slope

    CHAPTER 5  THE DIFFERENTIATION AND INTEGRATION OF POWERS OF x

    1.  Introduction

    2.  The Functions xn for Positive Integral n

    3.  A Calculus Method of Finding Roots

    4.  Differentiation and Integration of xn for Fractional Values of n

    CHAPTER 6  SOME THEOREMS ON DIFFERENTIATION AND ANTIDIFFERENTIATION

    1.  Introduction

    2.  Some Remarks about Functions

    3.  The Differentiation of Sums and Differences of Functions

    4.  The Differentiation of Products and Quotients of Functions

    5.  The Integration of Combinations of Functions

    6.  All Integrals Differ by a Constant

    7.  The Power Rule for Negative Exponents

    8.  The Concept of Work and an Application

    CHAPTER 7  THE CHAIN RULE

    1.  Introduction

    2.  The Chain Rule

    3.  Application of the Chain Rule to Differentiation

    4.  The Differentiation of Implicit Functions

    5.  Equations of the Ellipse and Hyperbola

    6.  Differentiation of the Equations of Ellipse and Hyperbola

    7.  Integration Employing the Chain Rule

    8.  The Problem of Escape Velocity

    9.  Related Rates

    APPENDIX  Transformation of Coordinates

    A1.  Introduction

    A2.  Rotation of Axes

    A3.  Translation of Axes

    A4.  Invariants

    CHAPTER 8  MAXIMA AND MINIMA

    1.  Introduction

    2.  The Geometrical Approach to Maxima and Minima

    3.  Analytical Treatment of Maxima and Minima

    4.  An Alternative Method of Determining Relative Maxima and Minima

    5.  Some Applications of the Method of Maxima and Minima

    6.  Some Applications to Economics

    7.  Curve Tracing

    CHAPTER 9  THE DEFINITE INTEGRAL

    1.  Introduction

    2.  Area as the Limit of a Sum

    3.  The Definite Integral

    4.  The Evaluation of Definite Integrals

    5.  Areas Below the x-Axis

    6.  Areas Between Curves

    7.  Some Additional Properties of the Definite Integral

    8.  Numerical Methods for Evaluating Definite Integrals

    APPENDIX  The Sum of the Squares of the First n Integers

    CHAPTER 10  THE TRIGONOMETRIC FUNCTIONS

    1.  Introduction

    2.  The Sinusoidal Functions

    3.  Some Preliminaries on Limits

    4.  Differentiation of the Trigonometric Functions

    5.  Integration of the Trigonometric Functions

    6.  Application of the Trigonometric Functions to Periodic Phenomena

    CHAPTER 11  THE INVERSE TRIGONOMETRIC FUNCTIONS

    1.  The Notion of an Inverse Function

    2.  The Inverse Trigonometric Functions

    3.  The Differentiation of the Inverse Trigonometric Functions

    4.  Integration Involving the Inverse Trigonometric Functions

    5.  Change of Variable in Integration

    6.  Time of Motion Under Gravitational Attraction

    CHAPTER 12  LOGARITHMIC AND EXPONENTIAL FUNCTIONS

    1.  Introduction

    2.  A Review of Logarithms

    3.  The Derived Functions of Logarithmic Functions

    4.  Exponential Functions and Their Derived Functions

    5.  Problems of Growth and Decay

    6.  Motion in One Direction in a Resisting Medium

    7.  Up and Down Motion in Resisting Media

    8.  Hyperbolic Functions

    9.  Logarithmic Differentiation

    CHAPTER 13  DIFFERENTIALS AND THE LAW OF THE MEAN

    1.  Differentials

    2.  The Mean Value Theorem of the Differential Calculus

    3.  Indeterminate Forms

    CHAPTER 14  FURTHER TECHNIQUES OF INTEGRATION

    1.  Introduction

    2.  Integration by Parts

    3.  Reduction Formulas

    4.  Integration by Partial Fractions

    5.  Integration by Substitution and Change of Variable

    6.  The Use of Tables

    CHAPTER 15  SOME GEOMETRIC USES OF THE DEFINITE INTEGRAL

    1.  Introduction

    2.  Volumes of Solids: The Cylindrical Element

    3.  Volumes of Solids: The Shell Game

    4.  Lengths of Arcs of Curves

    5.  Curvature

    6.  Areas of Surfaces of Revolution

    7.  Remarks on Approximating Figures

    CHAPTER 16  SOME PHYSICAL APPLICATIONS OF THE DEFINITE INTEGRAL

    1.  Introduction

    2.  The Calculation of Work

    3.  Applications to Economics

    4.  The Hanging Chain

    5.  Gravitational Attraction of Rods

    6.  Gravitational Attraction of Disks

    7.  Gravitational Attraction of Spheres

    CHAPTER 17  POLAR COORDINATES

    1.  The Polar Coordinate System

    2.  The Polar Coordinate Equations of Curves

    3.  The Polar Coordinate Equations of the Conic Sections

    4.  The Relation Between Rectangular and Polar Coordinates

    5.  The Derivative of a Polar Coordinate Function

    6.  Areas in Polar Coordinates

    7.  Arc Length in Polar Coordinates

    8.  Curvature in Polar Coordinates

    CHAPTER 18  RECTANGULAR PARAMETRIC EQUATIONS AND CURVILINEAR MOTION

    1.  Introduction

    2.  The Parametric Equations of a Curve

    3.  Some Additional Examples of Parametric Equations

    4.  Projectile Motion in a Vacuum

    5.  Slope, Area, Arc Length, and Curvature Derived from Parametric Equations

    6.  An Application of Arc Length

    7.  Velocity and Acceleration in Curvilinear Motion

    8.  Tangential and Normal Acceleration in Curvilinear Motion

    CHAPTER 19  POLAR PARAMETRIC EQUATIONS AND CURVILINEAR MOTION

    1.  Polar Parametric Equations

    2.  Velocity and Acceleration in the Polar Parametric Representation

    3.  Kepler’s Laws

    4.  Satellites and Projectiles

    CHAPTER 20  TAYLOR’S THEOREM AND INFINITE SERIES

    1.  The Need to Approximate Functions

    2.  The Approximation of Functions by Polynomials

    3.  Taylor’s Formula

    4.  Some Applications of Taylor’s Theorem

    5.  The Taylor Series

    6.  Infinite Series of Constant Terms

    7.  Tests for Convergence and Divergence

    8.  Absolute and Conditional Convergence

    9.  The Ratio Test

    10.  Power Series

    11.  Return to Taylor’s Series

    12.  Some Applications of Taylor’s Series

    13.  Series as Functions

    CHAPTER 21  FUNCTIONS OF TWO OR MORE VARIABLES AND THEIR GEOMETRIC REPRESENTATION

    1.  Functions of Two or More Variables

    2.  Basic Facts on Three-Dimensional Cartesian Coordinates

    3.  Equations of Planes

    4.  Equations of Straight Lines

    5.  Quadric or Second Degree Surfaces

    6.  Remarks on Further Work in Solid Analytic Geometry

    CHAPTER 22  PARTIAL DIFFERENTIATION

    1.  Functions of Two or More Variables

    2.  Partial Differentiation

    3.  The Geometrical Meaning of the Partial Derivatives

    4.  The Directional Derivative

    5.  The Chain Rule

    6.  Implicit Functions

    7.  Differentials

    8.  Maxima and Minima

    9.  Envelopes

    CHAPTER 23  MULTIPLE INTEGRALS

    1.  Introduction

    2.  Volume Under a Surface

    3.  Some Physical Applications of the Double Integral

    4.  The Double Integral

    5.  The Double Integral in Cylindrical Coordinates

    6.  Triple Integrals in Rectangular Coordinates

    7.  Triple Integrals in Cylindrical Coordinates

    8.  Triple Integrals in Spherical Coordinates

    9.  The Moment of Inertia of a Body

    CHAPTER 24  AN INTRODUCTION TO DIFFERENTIAL EQUATIONS

    1.  Introduction

    2.  First-Order Ordinary Differential Equations

    3.  Second-Order Linear Homogeneous Differential Equations

    4.  Second-Order Linear Non-Homogeneous Differential Equations

    CHAPTER 25  A RECONSIDERATION OF THE FOUNDATIONS

    1.  Introduction

    2.  The Concept of a Function

    3.  The Concept of the Limit of a Function

    4.  Some Theorems on Limits of Functions

    5.  Continuity and Differentiability

    6.  The Limit of a Sequence

    7.  Some Theorems on Limits of Sequences

    8.  The Definite Integral

    9.  Improper Integrals

    10.  The Fundamental Theorem of the Calculus

    11.  The Directions of Future Work

    TABLES

    INDEX

    CHAPTER

    ONE

    WHY CALCULUS?

    1. The Historical Motivations for the Calculus. Each branch of mathematics has been developed to attack a class of problems that could not be solved at all or yielded to a solution only after great efforts. Thus elementary algebra was created to find answers to simple physical problems which in mathematical form called for solving first, second, and higher degree equations with one or two unknowns. Plane and solid geometry originated in the need to find perimeters, areas, and volumes of common figures and to state conditions under which two figures, say two triangles, are congruent or have the same shape—that is, are similar. Trigonometry, introduced by astronomers, enabled man to determine the sizes and distances of heavenly bodies.

    In high school algebra and trigonometry we usually learn the fundamentals of another branch of mathematics, called coordinate or analytic geometry. Thus we learn to graph linear equations such as x + 2y = 5, to represent a circle of radius R by an equation of the form x² + y² = R², and to determine which curves correspond to such equations as y = sin x and y = cos x. The primary purpose of relating equations and curves is to enable us to use the equations in the study of such important curves as the paths of projectiles, planets, and light rays. Of course, each of the above-mentioned branches of mathematics has also helped to treat problems of the physical and social sciences which arose long after the motivating questions had been disposed of.

    During the seventeenth century, when modern science was founded and began to expand apace, a number of new problems were brought to the fore. Because the mathematicians of that century, like those of most great periods, were the very physicists and astronomers who raised the questions, they responded at once to the problems. Let us see what some of these problems were.

    Seventeenth-century scientists were very much concerned with problems of motion. The heliocentric theory created by Nicolaus Copernicus (1473–1543) and Johannes Kepler (1571–1630) introduced the concepts of the earth rotating on its axis and revolving around the sun. The earlier theory of planetary motion, dating back to Ptolemy (c. A.D. 150), which presupposed an earth absolutely fixed in space and, indeed, in the center of the universe, was discarded. The adoption of the theory involving an earth in motion invalidated the laws and explanations of motion that had been accepted since Greek times. New insights were needed into such phenomena as the motion of a projectile shot from a cannon and an answer to the question of why objects stay with the moving earth seemed called for. Furthermore, Kepler had shown on the basis of observations that the path of each planet around the sun is an ellipse, although no theoretical explanation of why the planets move on such paths had been offered. However, the notion that all bodies in the universe attract one another in accordance with the force of gravitation became prominent, and scientists decided to investigate whether the motions of planets around the sun and of moons around planets could be deduced from the proper laws of motion and gravitation. The motion of celestial bodies became the dominant scientific study.

    All of these motions—those of objects near the surface of the earth and those of the heavenly bodies—take place with variable velocity, and many involve variable acceleration. Although the difficulties in handling variable velocities and accelerations may not be apparent at the moment, the branches of mathematics that existed before the calculus was created were not adequate to treat them. We shall see later precisely what the difficulties are and how they are surmounted. In pre-calculus courses students often work on problems involving variable velocity—for example, the motion of a body falling to earth—but the intricacy is circumvented there by one dodge or another.

    The second major problem of the seventeenth century was the determination of tangents to various curves (Fig. 1-1). This question is of some interest as a matter of pure geometry, but its deeper significance is that the tangent to a curve at a point represents the direction of the curve at the point. Thus, if a projectile moves along a curve, the direction in which the projectile is headed at any point on its path is the direction of the tangent at that point. To determine whether the projectile will strike its target head on or merely at a glancing angle, we must know in which direction the projectile is moving at that point on its path at which it strikes the target. The invention of the telescope and microscope in the seventeenth century stimulated great interest in the action of lenses. To determine the course of a light ray after it strikes the surface of a lens, we must know the angle that the light ray makes with the lens, that is, the angle (Fig. 1-2) between the light ray and the tangent to the lens. Incidentally, the study of the behavior of light was, next to the study of motion, the most active scientific field in that century. It may now be apparent why the question of finding the tangent to a curve was a major one.

    Figure 1-1

    Figure 1-2

    A third class of problems besetting the seventeenth-century scientists may be described as maxima and minima problems. The motion of cannon balls was studied intensively from the sixteenth century onward. In fact, the mathematicians Nicolò Tartaglia (1500–1557) and Galileo Galilei (1564–1642) made significant progress in this investigation even before the calculus was applied to it. One of the important questions about the motion of cannon balls and other kinds of projectiles was the determination of the maximum range. As the angle of elevation of a cannon (angle A in Fig. 1-3) is varied, the range—that is, the horizontal distance from the cannon to the point at which the projectile again reaches the ground—also varies. The question is, at what angle of elevation is the range a maximum? Another maximum and minimum problem of considerable importance arises in planetary motion. As a planet moves about the sun, its distance from the sun varies. A basic question in this area is, what are the maximum and minimum distances of the planet from the sun? Some simple maxima and minima problems can be solved by the methods of elementary algebra and elementary geometry, but the most important problems are beyond the power of these branches and require the calculus.

    Still another class of problems in the seventeenth century concerned the lengths of curves and the areas and volumes of figures bounded by curves and surfaces. Elementary mathematics suffices to determine the areas and volumes of simple figures, principally figures bounded by line segments and by portions of planes. However, when curves or curved surfaces are involved, elementary geometry is almost helpless. For example, the shape of the earth is an oblate spheroid, that is, a sphere somewhat flattened on the top and bottom (Fig. 1-4). The calculation of the volume of this figure cannot be performed with elementary geometry; it can be done with the calculus. Euclidean geometry does have a method, called the method of exhaustion, for treating a very limited number of area and volume problems involving curves and surfaces, respectively. This method is difficult to apply and, moreover, involves concepts that can with considerable justification be regarded as belonging to the calculus, although the Greeks did not formulate them in modern terms. In any case, the method of exhaustion could not cope with the variety and difficulty of the area and volume problems that appeared in the seventeenth century. Closely related to these problems were those of finding the center of gravity of a body and the gravitational attraction exerted by, say, the earth on the moon. The relation may not be evident at the moment, but we shall see that the same method solves both types of problem.

    Figure 1-3

    Figure 1-4

    The efforts to treat the four classes of problem that we have thus far briefly described led mathematicians to methods which we now embrace under the term calculus. Of course, similar problems continue to be important in our time; otherwise the calculus would have only historical value. In fact, once a mathematical method or branch of any significance is created, many new uses are found for it that were not envisioned by its creators. For the calculus this has proved to be far more the case than for any other mathematical creation; we shall examine later a number of modern applications. Moreover, the most weighty developments in mathematics since the seventeenth century employ the calculus. Indeed, it is the basis of a number of branches of mathematics which now comprise its most extensive portion. The calculus has proved to be the richest lode that the mathematicians have ever struck.

    2. The Creators of the Calculus. Like almost all branches of mathematics, the calculus is the product of many men. In the seventeenth century Pierre de Fermat (1601–1665), René Descartes (1596–1650), Blaise Pascal (1623–1662), Gilles Persone de Roberval (1602–1675), Bonaventura Cavalieri (1598–1647), Isaac Barrow (1630–1677), James Gregory (1638–1675), Christian Huygens (1629–1695), John Wallis (1616–1703), and, of course, Isaac Newton (1642–1727) and Gottfried Wilhelm Leibniz (1646–1716) all contributed to it. Newton and Leibniz are most often mentioned as the creators of the calculus. This is a half-truth. Without deprecating their contributions, it is fair to say, as Newton himself put it, that they stood on the shoulders of giants. They saw more clearly than their predecessors the generality of the methods that were gradually being developed and, in addition, added many theorems and processes to the stock built up by their predecessors.

    Even Newton and Leibniz did not complete the calculus. In fact, it may be a comfort to students just beginning to work in the calculus to know that Newton and Leibniz, two of the greatest mathematicians, did not fully understand what they themselves had produced. Throughout the eighteenth century new results were obtained by, for example, James Bernoulli (1654–1705), his brother John Bernoulli (1667–1748), Michel Rolle (1652–1719), Brook Taylor (1685–1731), Colin Maclaurin (1698–1746), Leonhard Euler (1707–1783), Jean Le Rond d’Alembert (1717–1783), and Joseph-Louis Lagrange (1736–1813). However, the final clarification of the concepts of the calculus was achieved only in the nineteenth century by, among others, Bernhard Bolzano (1781–1848), Augustin-Louis Cauchy (1789–1857), and Karl Weierstrass (1815–1897). We shall find many of these great names attached to theorems that we shall be studying.*

    3. The Nature of the Calculus. The word calculus comes from the Latin word for pebble, which became associated with mathematics because the early Greek mathematicians of about 600 B.C. did arithmetic with the aid of pebbles. Today a calculus can mean a procedure or set of procedures such as division in arithmetic or solving a quadratic equation in algebra. However, most often the word means the theory and procedures we are about to study in the differential and integral calculus. Usually we say the calculus to denote the differential and integral calculus as opposed to other calculi.

    The calculus utilizes algebra, geometry, trigonometry, and some coordinate geometry (which we shall study in this book). However, it also introduces some new concepts, notably the derivative and the integral. Fundamental to both is the limit concept. We shall not attempt to describe the notion of limit and how it is used to formulate the derivative and integral, because a brief explanation may be more confusing than helpful. Nevertheless, we do wish to point out that the calculus in its introduction and utilization of the limit concept marks a new stage in the development of mathematics.

    The proper study of the calculus calls for attention to several features. The first is the theory, which leads to numerous theorems about the derivative and the integral. The second feature is the technique; to use the calculus, one must learn a fair amount of technique in differentiation and integration. The third feature is application. The calculus was created in response to scientific needs, and we should study many of the applications to gain appreciation of what can be accomplished with the subject; these applications also give insight into the mathematical ideas.

    The theory of the calculus, which depends primarily on the limit concept, is rather sophisticated. Complete proofs of all the theorems are difficult to grasp when one is beginning the study of the subject. Our approach to the theory attempts to surmount this hurdle. Many of our proofs are complete. However, other proofs are made by appealing to geometric evidence; that is, we use curves or other geometric figures to substantiate our assertions. The geometric evidence does not necessarily consist of complete geometric proofs, so that we cannot say that we are proving geometrically. Nevertheless, the arguments are quite convincing. For example, we can be certain even without geometric proof that the bisector of angle A (Fig. 1-5) divides the isosceles triangle ABC into two congruent triangles.

    Figure 1-5

    This approach, which will become clearer when we begin considering specific cases, is called the intuitive approach. It is recommended for several reasons. The first, as already suggested, is pedagogical. A thoroughly sound, deductive approach to the calculus, one which the modern mathematician would regard as logically rigorous, is meaningless before one understands the ideas and the purposes to which they are put. One should always try to understand new concepts and theorems in an intuitive manner before studying a formal and rigorous presentation of them. The logical version may dispose of any lingering doubts and may be aesthetically more satisfying to some minds, but it is not the road to understanding.

    The rigorous approach, in fact, did not become available until about 150 years after the creation of the calculus. During these years mathematicians built up not only the calculus but also differential equations, differential geometry, the calculus of variations, and many other major branches of mathematics that depend on the calculus. In achieving these results the greatest mathematicians thought in intuitive and physical terms.

    The second reason for adopting the intuitive approach is that we wish to have time for some techniques and for applications. Were we to study the rigorous formulations of limit, derivative, integral, and allied concepts, we would not have time for anything else. These concepts can be carefully and most profitably studied in more advanced courses in the calculus after one has some appreciation of what they mean in geometrical and physical terms and of what one wishes to do with them. As for the applications, the calculus more than any other branch of mathematics was created to solve major physical problems, and one should certainly learn what the calculus accomplishes in this connection.

    After we have become reasonably familiar with the ideas, techniques, and uses of the calculus, we shall consider in the last chapter how the intuitive approach may be strengthened by a more rigorous one.

    * The history of the calculus is presented very well in Carl B. Boyer, The Concepts of the Calculus, Dover Publications, New York, 1959.

    CHAPTER

    TWO

    THE DERIVATIVE

    1. The Concept of Function. Before considering any ideas of the calculus itself, we shall review a concept that is, no doubt, largely familiar—the concept of a function.

    If an object moves in a straight line—for example, a ball rolling along a floor—the time during which it moves, measured from the instant it starts its motion, is a variable. In the case of the ball, the time continually increases. The distance the object moves, measured from the point at which it starts to move, is also a variable. The two variables are related. The distance the ball travels depends on the time the ball has been in motion. By the end of 1 second it may have moved 50 feet, by the end of 2 seconds, 100 feet, and so on. The relation between distance and time is a function. Loosely stated for the moment, a function is a relation between variables.

    There are, of course, thousands of functions. If the ball were started on its journey with a different speed, the relation between distance traveled and time in motion would be different. All kinds of motions take place around us, and for each of them there is a function or relation between distance traveled and time in motion. The idea of a relation between variables is not confined to motion. The national debt of this country varies as time varies, and the relation between these variables is also a function. If money is allowed to accumulate interest in a bank, the amount in the account increases with time. Here, too, we have a function. Obviously, many other examples could be cited.

    The notion of a function as used in the calculus is more restricted than we have thus far indicated. The statement that a person’s obligation to society increases with his age expresses a function, the variables of which are obligation and age. However, it is not possible to measure the extent of this obligation in numbers. The calculus is concerned only with variables whose values can be expressed numerically. Thus time, distance, and money accumulating in a bank are variables whose various values can be measured and therefore can be expressed as numbers.

    The most effective mathematical representation of a function is what we shall call a formula. Suppose that an object is dropped and falls straight down. The time during which it falls continually increases as does the distance that it falls. The formula

    whose correctness we shall establish later, expresses the relation between the two variables if the resistance of air is neglected. In this formula t is the number of seconds the object falls measured from the instant it starts to fall and s is the number of feet it falls measured from the point at which it is dropped. (The letter s is used to denote distance because the Latin word for it is spatium.) The formula thus says that when t = 2, then s = 16 · 2² or 64. For each value that we may substitute for t there is a corresponding value of s. This example not only illustrates a formula but also possesses a property which is important for the functions that the calculus treats. For each value we may choose for t there is no more than one value of s. We say that s is a single-valued function of t.

    When written in the form (1), that is, when solved for s, formula (1) shows how s depends on t or expresses s as a function of t. Then t is called the independent variable and s, the dependent variable.

    formula (1) also tells us something about how t depends on s, for example, when s = 64, we find that t can be +2 or −2. To express more clearly the dependence of t on s, we can solve (1) algebraically for t and obtain

    formula (2) expresses t as a function of s, so that s is the independent variable and t is the dependent variable. Here we can say that t is a double-valued function of s or we can say that there are two functions,

    each of which is single-valued. Because the techniques of the calculus apply to single-valued functions, we use either one or the other of these two functions, depending on the physical problem.

    EXERCISES

    1.  The amount A of money that accumulates in n years if one dollar is invested and if the interest is compounded annually at the fixed rate of i per cent per year is A = (1 + i)n. As the formula is written, which is the independent variable? Which is the dependent variable?

    2.  Solve υ = 32t for t. Is the resulting function single-valued?

    3.  Solve y = 5.3x² for x. Is the resulting function single-valued?

    4. Write the formula for:

    (a)  the area A of a circle in terms of the radius;

    (b)  the area A of a circle in terms of the diameter.

    5.  Write the formula for the radius of a circle in terms of the area. Is the function single-valued? If not, which of the single-valued functions do you think would be more useful and why?

    6.  One arm of a right triangle is 3 units and the hypotenuse is x units. Write a formula for the length of the other arm.

    Suggestion: Use the Pythagorean theorem.

    7.  An automobile travels at 30 miles per hour. Write a formula which expresses the distance d traveled in feet as a function of the time t which represents the number of seconds of travel.

    8.  A rectangle is required to have an area of 4 square feet but its dimensions may vary. If one side has length x, express the perimeter p of the rectangle as a function of x.

    We shall be working with functions constantly. Thus we may have to deal with

    and a great variety of other functions. If we wish to refer to any of these specific functions repeatedly it is cumbersome to have to repeat the entire function each time, and yet some device is needed to avoid confusion with other functions. Moreover, there will be times when we shall want to speak about properties of all functions or all functions of some class. How should we handle these tasks without a lot of extensive repetition or verbiage?

    There is a notation which solves our problems and even has additional advantages. Suppose we wish to refer to the function in (4) repeatedly. We use the notation f(t) and f(t) then stands for the entire expression on the right side of (4). [The notation f(t) is read "f of t"; moreover the notation does not mean f times t. The symbol f(t) must be taken as a whole.] Similarly, to denote the function in (5) one writes f(x). To speak of a class of functions one can say, let f(x) be any function of such and such a class.

    The notation f(x) has many advantages. First of all, as opposed to s or y, it tells us what the independent variable is. Second, suppose one wished to speak of the value of the function f(x) = x² −9 when x = 3. We could, of course, calculate the value of the function. But, often knowing the value is not as important as is speaking about it. The notation f(3) does it. That is, f(3) means the value of f(x) when x = 3. In the case of f(x) = x² −9, f(3) = 0. Similarly, f(−2)= −5.

    The function notation has another advantage. Suppose that in some discussion we wish to talk about two different functions, say y = x² −9 and y = (x – 1)/2. To refer to both as f(x) would be confusing. What we can do is refer to x² −9 as f(x) and refer to (x − 1)/2 as g(x). In this situation f(5) is 16 and g(5) is 2.

    The function notation also lends itself to a distinction that is most often honored in the breach. Strictly speaking, a function is a relation between two variables, x and y say. The relation might be that y is the square of x. Then the relation is neither x nor y but something connecting the two. Thus when one says that John is the father of William, the relationship is fatherhood, and this relation is neither John nor William but a connection between them. Yet in speaking about these people one might use the word father to mean John, whereas the word strictly refers to the relationship. Likewise in dealing with functions we should distinguish between the relation and the value y of the function. We could use f to denote the relation between y and x and use f(x) to mean y. However, we often use y or f(x) to mean the relation. Fortunately, this use of y or f(x) in the double sense causes no confusion and avoids a lot of unnecessary words.

    Example 1. If f(x) = x² −7x + 5, f(2) = 2² −7 · 2 + 5 = −5;f(−2) = (−2)² −7(−2) + 5 = 23; f(2a) = (2a)² −7(2a) + 5 = 4a² −14a + 5; f(x + 3) = (x + 3)² −7(x + 3) + 5 = x² − x − 7.

    Example 2. If f(x) = x³ + 5, f(−2) = (−2)³ + 5 = −8 + 5 = −3; f(t) = t³ + 5; f(2t) = 2(t)³ + 5 = 8t³ + 5.

    Example 3. If f(x) = x² + 9x and g(x) = x³ −7, f(xg(x) = (x² + 9x)(x³ −7) = x⁵ + 9x⁴ −7x² −63x.

    EXERCISES

    1.  If f(x) = x² −9x, calculate f(0), f(2), f(−1), and f(9).

    2.  If f(x) = − x² −9x, calculate f(0), f(2), f( −2), f(9), and f( −9).

    3.  For the functions in Exercises 1 and 2, calculate f(a) and f(x0).

    4.  If f(x) = x − show that (a) f(−x) = −f(x),(b) f( ) = − f(x).

    5.  If f(x) = , find f(0), f(2), f( −2), and f( ).

    6.  If f(x) = x² −7x, what is f(2x)? What is f(x + h)?

    7.  If f(x) = tan x, find f(0), f(π/4), and f(− π/4).

    8.  If f(x) = x² + 5 and g(x) = x³ −7, how much is f( −2) · g( −2).

    9.  If , find f(3), f( −1), f( ).

    10.  If find f(0), f(4), f(g²).

    11.  Let g(x) = x³. Show that g(− x) = − g(x).

    12.  Let g(x) = x⁴ + 2x² + 1. Show that g(x) = g(− x).

    13.  We learn in trigonometry that sin x ≡ sin(π x). Hence f(sin x) = f(sin[π x]). Now let f(x) = x sin x. Then x sin x = (π x) sin(π x) or x = π x. Hence π = 2x and since x is any value we choose, so is π. What is wrong?

    ANSWERS AVAILABLE

    1.Ans. 0, −14, 10, 0.

    3.Ans. a² −9a; − a² −9a.

    7.Ans. 0, 1, −1.

    2. The Graph or Curve of a Function. We learn in algebra that formulas may be pictured as curves. The powerful method of interpreting formulas geometrically is known as coordinate or analytic geometry. This subject was created in the early part of the seventeenth century by René Descartes (1596–1650) and Pierre de Fermat (1601–1665) to expedite the study of functions, which were just beginning to be used, and to study geometric problems by algebraic means.

    The main idea of coordinate geometry is that a function relating two variables, such as y = x² or y = x² + 6x, can be represented geometrically by a curve and, conversely, that a curve can be represented by a function, although the function may not be single-valued. It is customary to speak of the function as the equation of the curve.

    The basic device of coordinate geometry is a coordinate system, that is, a scheme that locates points in a plane by means of a pair of numbers. The most important coordinate system is the one that utilizes two mutually perpendicular lines (Fig. 2-1), called the x- and y-axes. These lines intersect at some point O, called the origin. Along the x-axis, which is usually taken to be a horizontal line, we assign to each point its distance from the origin O. This distance is taken to be positive if the point lies to the right of O and negative if it lies to the left. Thus point T, which is 3 units to the right of O, has the number 3 or + 3 assigned to it, whereas point U, which is 3 units to the left of O, has the number −3 assigned to it. Likewise numbers that represent their distances from O are assigned to points on the y-axis, but in this case points above O are assigned positive numbers and points below O, negative numbers. It is customary and usually convenient to use the same unit of distance in assigning numbers to points on both axes.

    We are now able to locate or describe the positions of all points in the plane in relation to the axes. Consider point P in Fig. 2-1. To reach it from O we can travel 2 units to the right along the x-axis and 3 units up in the direction of or parallel to the y-axis. Point P is located by the numbers 2 and 3, which are usually written as (2, 3), the first number indicating the distance along the x-axis and the second, the distance parallel to the y-axis. The two numbers are called the coordinates of P; the first is called the abscissa and the second, the ordinate.

    Figure 2-1

    The coordinates of point Q in Fig. 2-1 are (−2, 3); the minus sign enters because we must travel 2 units to the left to reach Q. Similarly, the coordinates of R are (−2, −3) and the coordinates of S are (2, −3).

    Our method of attaching coordinates to points assigns two coordinates to each point. Accordingly, point T on the x-axis in Fig. 2-1 has the coordinates (3, 0) and point V on the y-axis has the coordinates (0, 2).

    The coordinate system we have examined is called the rectangular Cartesian system. The word Cartesian honors Descartes who introduced it, although in cruder form, and the word rectangular refers to the fact that the axes meet at right angles.

    With the help of the rectangular Cartesian coordinate system we can obtain curves that represent or picture formulas. To obtain the curve corresponding to s = 16t² we first introduce the usual horizontal and vertical axes of a rectangular coordinate system. We may as well label these axes t and s instead of x and y to remind us of the letters we are actually using for the variables in our formula. We can now graph the formula s = 16t² by making a table of values. Thus when t = 0, s = 0; when t = 1, s = 16; and so forth. The table then reads:

    Because the s-values are large, we use a smaller unit of length on the s-axis than on the t-axis. We now plot the points (0, 0), (1, 16), (2, 64),…, and join them by a smooth curve (Fig. 2-2). In the present case there is no need to include negative values of t in the table because we see that for each negative value of t, say −2, the s-value is the same as that for the corresponding positive value of t. Geometrically this means that the curve rises to the left in precisely the same way as to the right. The full curve in Fig. 2-2 is called a parabola.

    Figure 2-2

    Figure 2-3

    The graph of a formula, that is, the curve corresponding to a formula, is a useful picture of how the variables in the formula behave. Thus we see in Fig. 2-2 that s not only increases as t increases from 0 onward through positive values but does so much more rapidly than t. We must, however, be careful to distinguish the curve from the physical situation. The motion described by s = 16t² is that of an object that falls straight down. There is a correspondence between the physical happening and the curve, but the latter is usually not a picture of the physical motion. It is also important to note that both the formula and the curve differ in another essential respect from the physical motion. An object that is dropped from a point near the surface of the earth may fall for, say, 5 seconds and then hit the earth. The physical motion, then, is represented by the formula and the curve only for the values of t from 0 to 5. However, the mathematical formula and mathematical curve have meaning for every positive and negative value of t.

    The function s = 16t² is especially simple in that to each value of t that we may select there is a corresponding value of s. The concept of a function does not require this. Consider, for example, the function

    For each value of x other than 2, there is a definite value of y. For x = 2 the expression on the right side becomes 1/0, which is meaningless because division by 0 is not defined.* Hence this function is defined or, as the mathematician says, this function exists for each value of x other than 2. The graph of the function is shown in Fig. 2-3. There is no point on the graph corresponding to x = 2.

    As another example, let us consider the function

    wherein the radical sign denotes the positive square root. When x is greater than 1, the radicand is negative, and the square root of a negative number is a complex number. In elementary calculus we prefer not to deal with complex numbers. Hence we restrict the values of x to 1 and numbers algebraically less than 1; that is, the permissible values of x are the positive numbers less than or equal to 1, 0, and all negative numbers. The graph of the permissible x- and y- values is shown in Fig. 2-4. Whether we label the vertical axis y or f(x) is immaterial here.

    Thus the concept of a function does not require that there be a y – value for each value of x, but it does require a y-value for each value of x in some collection or set of x-values. The technical word for the collection of x-values for which the y-values exist is domain; the collection of corresponding y-values is called the range of the function.

    Let us consider as an example the function

    The temptation in studying this function is to use algebra at once and write

    and now one almost naturally cancels the factor x − 2 in numerator and denominator and obtains

    There seems to be nothing wrong about yielding to temptation and, in fact, one seems to gain a lot by doing so, contrary to all the moralists: the function (9) is much simpler than (8).

    Figure 2-4

    Nevertheless, there is a vital distinction between (8) and (9). To arrive at (9) we canceled the factor x − 2 in numerator and denominator. What this operation really amounts to is dividing numerator and denominator by x − 2. However, division by x − 2 is legitimate except when x = 2. Hence the steps from (8) to (9) are correct except when x = 2. Let us look at (8) when x = 2. The numerator and denominator are each 0, and what we obtain for y is 0/0. Now, division by 0 is meaningless, and therefore y in (8) has no value at x = 2. On the other hand the function in (9) has a very definite value when x = 2, namely y = 4. Thus the two functions (8) and (9) differ in their behavior at x = 2.

    Let us look at the graphs of the two functions. To graph (8) we must calculate the y-values for various x-values. For this purpose we can use the simpler function (9), which is the same as (8) except at x = 2. We find that the graph of (9), which is a first degree equation in x and y, is a straight line; this is shown in Fig. 2-5b. Then the graph of (8) must also be the same straight line except at x = 2. At x = 2 the function (8) has no value, and therefore there is no point on the graph corresponding to x = 2. The graph of (8) then consists of two half lines or rays emanating from point P of Fig. 2-5a, but point P itself is not part of the graph.

    The distinction we have drawn between the functions (8) and (9) may seem to be much ado about nothing, but let us withhold judgment and undertake a further study of (8). What can we say about the behavior of that function as x takes on values closer and closer to 2? For values of x close to 2, other than 2 itself, the function (9) has the same values as (8), and we see from (9) that as x takes on values closer and closer to 2, the y-values come closer and closer to 4. Hence the same must be true of the function (8). We say that the number

    Enjoying the preview?
    Page 1 of 1