calculus an introductory approach

181
CALC-ULUS An Introductory Approach

Upload: sergecheshut

Post on 22-Dec-2015

65 views

Category:

Documents


12 download

DESCRIPTION

Calculus

TRANSCRIPT

Page 1: Calculus an Introductory Approach

CALC-ULUS

An Introductory Approach

Page 2: Calculus an Introductory Approach

Editors

John L. Kelley, University of California

Paul R. Halmos, University of Chicago

PATRICK SuPPEs-Introduction to Logic

PAUL R. HALMOs-Finite-Dimensional Vector Spaces, 2nd Ed.

EDWARD J. McSHANE AND TRUMAN A. BOTTS-Real Analysis

JoHN G. KEMENY AND J. LAURIE SNELL--Finite Markov Chains

PAo'IUCK SuPPEs-Axiomatic Set Theory

PAUL R. HALMOS-Naive Set Theory

JoHN L. KELLEY-lntroduction to Modern Algebra

R. DUBISCH-Student Manual to Modern Algebra

lvAN NrvEN-Calculus: An Introductory Approach

Additional titles will be listed and announced as published. A series of distinguished texts for undergraduate mathematics.

Page 3: Calculus an Introductory Approach

CALCULUS

An Introductory Approach

by

IVAN NIVEN Professor of Mathematics The University of Oregon

D. V AN NOSTRAND COMPANY, INC. PRINCETON,. NEW JERSEY

TORONTO LONDON

NEW YORK

Page 4: Calculus an Introductory Approach

D. VAN NOSTRAND COMPANY, INC.

120 Alexander St., Princeton, New Jersey (Principal office) 24 West 40 Street, New York 18, New York

D. V AN NosTRAND CoMPANY, LTD.

358, Kensington High Street, London, W.14, England

D. VAN NosTRAND CoMPANY (Canada), LTD.

25 Hollinger Road, Toronto 16, Canada

COPYRIGHT@ 1961, BY

D. VAN NOSTRAND COMPANY, !Ne.

Published simultaneously in Canada by D. VAN NosTRAND CoMPANY (Canada), llrn.

No reproduction in any form of this book, in whole or in part (except for brief quotation in critical articles or reviews), may be made without written authorization from the publishers.

PRINTED IN THE UNITED STATES OF AMERICA

Page 5: Calculus an Introductory Approach

PREFACE

My purpose is to present the ideas that lie at the heart of calculus, along with the necessary background material from analytic geometry. I re­strict attention to a small collection of central concepts, and hence many of the topics offered in the larger books on calculus are omitted. I have aimed at a balance between theory and applications.

This book grew out of lectures given at Stanford University in the summers of 1957 and 1958 in a special program for high school teachers sponsored by the General Electric Company. It is hoped that the book will be of use in similar situations where a brief course in ·calculus is wanted. For example, it should be suitable in a calculus course for liberal arts students, a course designed to set forth the nature of the subject with an economy of time. The reader should have a reasonably good knowl­edge of basic algebra and trigonometry, although discussions of a few topics-inequalities and radian measure, for example-have been included for convenience.

The section headings in the table of contents suffice to describe the topics considered. Special features of the book are as follows: an immediate discussion of actual problems in calculus, with a minimum of prelimi­naries; a simplicity of theory, relatively speaking, obtained on the one hand by restricting sharply the class of functions in the discourse, and on the other hand by not attempting to state results in their most general form, or for that matter anywhere near their most general form; a devel­opment of the series expansions of the trigonometric, logarithmic, and exponential functions without the elaborate preparation that is ordinarily used; the postponing of the proof of the existence of the definite integral to an appendix, partly to break up the theory into smaller portions, and partly to keep more difficult ideas later in the presentation.

One of the most troublesome problems facing a writer of a beginning book on calculus is the matter of rigor. The difficulty is twofold; first that accuracy of statement can lead to overlengthy statements in which the reader may lose sight of the central idea in a welter of detail; second that since theorems in calculus are propositions about real numbers, ideally there should be a preamble on the logical foundations of the real number

V

Page 6: Calculus an Introductory Approach

~'i~----------------------~P~R,EFACE

system. I have tried to avoid the first of these difficulties by a restric­tion, mentioned already, on the scope of the material. As to the second difficulty, I take two fundamental propositions about real numbers as axiomatic. It is assumed that a bounded sequence of increasing real numbers has a limit. The mean value theorem is also assumed without a strict proof, although a heuristic argument is given to show the plausi­bility of this result. These results cannot be established without a rather thorough analysis of the real number system, which is precluded by the self-imposed limitation on the length of this book.

The more difficult problems are starred. Answers to about one third of the problems are given at the end of the book.

I was fortunate in having the manuscript read by two friendly critics of quite different backgrounds, one a university student, the other an experienced mathematician. First, I am indebted to my son Scott Niven who, in addition to making helpful suggestions about details, drew my attention to certain obscure passages. Second, I am grateful to Professor Herbert S. Zuckcrman for pointing out possible arrangements of the material; among other things, the final chapter was expanded considerably along lines he suggested.

Eugene, Oregon Jan'Uary 1961

I. N.

Page 7: Calculus an Introductory Approach

TABLE OF CONTENTS

CHAPTER PAGE

PREFACE V

1. WHAT IS CALCULUS? 1 1.1 Slope 1 1.2 An Example 4 1.3 The Slope of a Special Curve 5 1.4 The Notation for a Sum 7 1.5 Some Special Sums 9 1.6 An Example from Integral Calculus 12 1.7 A Refinement 15

2. Lll\IIITS 17 2.1 Inequalities 18 2.2 Absolute Values 21 2.3 Sequences and Limits 24 2.4 Theorems on Limits 29 2.5 Functions 33 2.6 Radian Measure 35 2.7 The Limit of a Function 36 2.8 The Definition of the Limit of a Function 39 2.9 Continuous Functions 46

3. INTEGRATION 48 3.1 The Limit of a Sum 48 3.2 The General Definition of an Integral 52 3.3 The Evaluation of Certain Integrals 56 3.4 The Interpretation of an Integral as an Area 58 3.5 Further Evaluation of Integrals 60 3.6 The Integral of sin x 62 3.7 The Volume of a Sphere 65

4. DIFFERENTIATION 67 4.1 The Derivative 68 4.2 The Definition of the Derivative 70 4.3 Simple Derivatives 73

vii

Page 8: Calculus an Introductory Approach

I, I

viii TABLE OF CONTENTS

CH:\PTF..R PAGE

4.4 Trigonometric Functions 79 4.5 The Chain Rule 82 4.6 Inverse Functions 84 4.7 The lvlcan Value Theorem 90 4.8 Maxima and Minima 92 4.9 The Problem of Minimum Surface Area of a Cylinder

of Fixed V olumc 96 4.10 The Refraction of Light 97

5. THE FUNDAMENTAL THEOREM 100 5.1 Derivatives and Integrals 100 5.2 Upper and Lower Sums 101 5.3 The Fundamental Theorem of Calculus 103 5.4 The Indefinite Integral 106 5.5 Inequalities for Integrals 108

6. THE TRIGONOMETRIC FUNCTIONS 112 6.1 Two Important Limits 112 6.2 Infinite Series Expansions for sih x and cos x 114 6.3 Remarks on Infinite Series 119

7. 'l'HE LOGARITHMIC AND EXPONENTIAL FUNCTIONS

7.1 A Function Defined 7.2 Properties of L(x) 7.3 The Exponential Function 7.4 The Series Expansion for ex 7.5 The Number e 7.6 The Series Expansion for the Logarithmic Function 7.7 The Computation of Logarithms

8. FURTHER APPLICATIONS

8.1 The Series Expansion for the Arc Tangent Function 8.2 The Computation of " 8.3 The Velocity of Escape 8.4 Radioactive Decay 8.5 A Mixing Problem 8.6 The Focusing Property of the Parabola 8.7 Newton's Method for Solving Equations

APPENDIX A. On the Existence of the Definite Integral

ANSWERS TO SOME PROBLEMS

INDEX

121 121 125 128 131 136 138 141

143 143 144 146 150 152 154 157

161

166

171

Page 9: Calculus an Introductory Approach

CHAPTER 1

WHAT IS CALCULUS?

1.0. A short definition of calculus, in a sentence or two for example, is likely to be meaningless except to persons already familiar with the subject. The reason for this is that such a definition necessarily refers to certain mathematical operations whose nature can be known thoroughly only by prolonged examination and study. Even this book in its entirety is only a partial answer to the question "What is cal­culus?", because a complete answer cannot be given in a short book. A brief volume cannot encompass the elegant general results of calculus. For example the advanced topics that are treated in the later chapters of this book are approached in special ways as individual results, where­as larger books on calculus ordinarily get these results as by-products of broad sweeping theories.

In this first chapter we pose a few simple problems of calculus, with solutions given at once in some cases, but solutions postponed until later chapters in others. In diving right in, we assume that the reader has a rudimentary knowledge of inequalities and functions; these topics are elaborated in Chapter 2, but in more detail than necessary for the present chapter. Prior to the study of actual problems from calculus, there is one preliminary discussion on the concept of slope.

1.1 Slope.* The slope of a straight line is simply a measure of the steepness of rise or fall of the line, viewed from left to right. Suppose that the line makes an angle rt. with the x-axis, the orientation of rt.

being from the positive end of the x-axis counterclockwise around to the line. Then the slope of the line is defined as tan rt.. If the line is falling from left to right, as in the case AB in Figure l.l, then the slope is negative, whereas if the line is rising from left to right, as in the case CD in Figure l.l, then the slope is positive. These cases correspond to

* Any reader with a ~now ledge of the concept of slope should bypass this section.

I

Page 10: Calculus an Introductory Approach

2 CALCULUS: AN INTRODUCTORY APPROACH CHAP. l

the angle " being obtuse and acute, respectively. If any two points with coordinates (x1, YI) and (x,, y,) are selected on the line, then the definition of the tangent function in trigonometry gives the well­known formula for the slope m,

(1)

Any line parallel to the x-axis has " = 0° and y2 = y1, and the slope of such a line is zero. Any line parallel to the y-axis has " = 90°, and so such a line has no slope because tan 90° does not exist. Another way of seeing that a vertical line has no slope is to observe that in such a case x2 = x1 0, and so equation (1) involves division by zero.

y-axis y-axis

A D

x-axis

x-axis

FIG, 1.1 FIG. 1.2

The slope of a curve at any point Pis defined as the slope of the tan­gent line to the curve at P. The tangent line at P can be described as the limiting position of the chord PQ as the point Q is moved along the curve towards the point P. In case the curve is a circle for example, a well-known theorem from elementary geometry states that the tangent line at any point is perpendicular to the radius drawn to the point of tangency. For curves other than the circle there is no corresponding theorem, but we shall be able with the use of calculus to determine the tangent lines of many types of curves. From these considerations we can observe that whereas a straight line has one slope, a curve has a slope at each point, and the slope is generally different from point to point.

Page 11: Calculus an Introductory Approach

SEC. 1 WHAT IS CALCULUS~

Q

FIG. 1.3

Problems

l. Prove that the formula (l ) can be written

Yl - Y2 ?n = ---.

Xl-X2

2. Find the slopes of the following lines: (a) through (0, 0) o.nd (5, 7); (b) through ( -1, 2) and (4, -6);

FIG. 1.4

(c) through (l, 2) and (a, b), presuming that a =f: l ; (d) through (2, -3) o.nd ( -3, -3).

3

3. Find the numerical value of x so that the line joining (x, -4) and (6, l) sho.ll have slope 2.

4. Prove tho.t the line joining A(5, -7) and B(6, 1) has the same slope as the line joining B(6, l ) o.nd 0(8, 17), and thus establish that the three points A, B, C are collinear.

5. Prove that the points (4, -9), (-1, - 6), and (- 16,3) are collineo.r. 6. Find the numerical value of y so that the point (7, y) shall lie on the

line joining (1, 3) and (9, 19). 7. Prove that the line through (4, 2) and (7, 1) is parallel to t he line

through (5, 6) o.nd ( - 1, 8). Suggestion: two lines (not parallel to the y-axis) are parallel if and only if their slopes are equal.

8. Find the numerical value of y required so that the line through ( - 2, 3) and (8, l) shall be parallel to the line through (1, -4) and ( -4, y).

9. Prove that t he four points (1, -2), (3, - 5), (8, 4), and (6, 7} are the vertices of a parallelogram.

10. Prove that the points (1, -2), (3, -5), (8, 4), and (10, l ) are the vertices of a parallelogram.

*ll . Find a point, other than (6, 7) and (10, 1}, which forms a parallelo­gram along with (1, -2), (3, -5), and (8, 4).

* The more difficult problems are starred.

Page 12: Calculus an Introductory Approach

I

4 CALCULUS: AN INTRODUCTORY APPROACH CHAP. I

12. Prove that the line through (7, 0) and (!0, 3) is inclined to the x-axis at an angle of 30°. (Recall that tan 30° ~ Ifv3.)

!3. Prove that the line through (!, 5) and (8, 12) is inclined to the x-axis at an angle of 45°. (Recall than tan 45° ~ !.)

14. Find the slope of the line joining (a, b) and (c, d), presuming that a is not equal to c.

!5. Find the slope of the line joining (a, b) and (a+c, b+d), presuming that c is not equal to zero.

1.2 An Example. We now give a sample problem from differential calculus. What should be the dimensions of a cylindrical tin can of fixed volume ISrr cubic inches, so that the surface area is a minimum?. Since the surface area is roughly proportional to the total amount of metal in the can, the solution of the problem will give almost the dimen­sions for minimum cost of the metal. The figure lSrr for the volume was chosen so that the arithmetic calculations would work out simply. Actually !Srr cubic inches is very close to a volume of one quart.

Let, r, h, S, and V denote, respectively, the radius, the height, the total surface area, and the volume of a circular cylinder. Then it will be recalled that

V ~ rrr"h and S ~ 2m2+ 2nrh.

The symbol V can be replaced by the constant lSrr, and so by simple algebra we can eliminate h in the formula forS,

!Srr ~ nr2h, h ~ 18jr2, S ~ 2nr2+2nrh,

S ~ 2rr(r2+ 1r8

).

Thus we have obtained S as a function of r. (It may be noted that we could have eliminated r in the algebraic process, and arrived at a for­mula giving S as a function of h. But square roots enter in, and so the above procedure is preferred.)

The graph of the functionS ~ 2rr(r2 + 18/r) is shown in Figure 1.5, for a reasonable range of positive values of r. To find the minimum value of S, we need some technique for locating the low point on the curve, the point labeled Pin Figure 1.5. It is the point where the slope of the curve is zero. To the right of the pointP the slope of the curve is positive, whereas to the left of the point P the slope of the curve is negative. The question of determining the coordinates of the point P can be settled easily by the methods of differential calculus. We are not at present in a position to finish the above problem of minimizing S, and we will return to the question later. But the example shows that

Page 13: Calculus an Introductory Approach

r

SEC. 3 WHAT IS CALCULUS ? 5

some questions of maxima and minima could be handled if we knew how to calculate the slope of a curve, and so find where the slope is zero.

S-axis

30

1 r-axis

FIG. 1.5 Graph of S = 27T[r2+18/r]. (Xote the different units of length on the axes.)

Problem

In the above analysis, we derived a formula for S as a function of r. Obtain a formula for S as a function of h.

1.3 The Slope of a Special Curve. We turn to a much simpler equation than S = 211(r2 + 18/r) to begin our treatment of the slope of a curve. The problem now is to find the slope of the curve y = x2

at the point (3, 9). The parabolic equation y = x 2 is about the sim­plest among non-linear equations, and so provides a good starting point. As in Figure 1.6, let us de­note the point (3, 9) by A, and let P be any nearbypointon the curve. Since the point P, unlike A, is not a fixed point, we give it coordin­ates (x, x'). This is nothing but the general coordinates (x, y) with y replaced by x2 in accordance with the presumption that P is a point on the curve y = x2• The slope of the chord AP, by formula (1) of § 1.1, is (x'- 9)/(x- 3). The tangent

A (3, 9)

Fw. 1.6 Graph of y = xz.

Page 14: Calculus an Introductory Approach

i.'

I'· !

6 CALCULUS: AN INTRODUCTORY APPROACH CHAP. 1

line to the curve y = x2 at A is the limiting position of the chord AP as the point P approaches the point A. Thus the slope of the tangent line at A, which by definition is the same as the slope of the curve at A, is the limiting value of (x2-9)j(x-3) as x approaches 3. We cannot simply substitute x = 3, because (i) geometrically the points PandA coincide and there is no chord AP, and (ii) algebraically the expression (x2- 9)/(x- 3) becomes OJO, which has no meaning.

Now by elementary algebra we have

(x+ 3)(x- 3) -- = _:._ _ _:_;__ _ _:_ = X+ 3, x-3 x-3

x2 -9

and (x2-9)j(x-3) is indeed equal to x+3 for all values of x except one, namely x = 3. The function f(x) = x+ 3 has a straight-line graph, whereas the functionf(x) = (x2-9)/(x-3) has the same graph

/ '-Gap at the point (3, 6)

FIG. 1.7 Graph ofj(x) = x+3. FIG. 1.8 Graph ofj(x) = (x2-9)/(x-3).

with a point missing, a gap at (3, 6). It is clear intuitively that the limit of (x2-9)/(x-3) as x approaches 3 is the same as the limit of x+ 3 as x approaches 3. We shall spell out a precise definition of limit later. In the meantime we write the standard notation for limits

x2 -9 lim -- = lim (x+3) = 6, x-)3 x _ 3 zry3

where ''Iim'' stands for ''limit'', and ''x --+3'' is short for ''as x approaches 3". We notice that while (x2-9)j(x-3) has no value at x = 3, never­theless it has a limit as x approaches 3. This limiting value 6 is thus the slope of the curve y = x2 at the point (3, 9).

Page 15: Calculus an Introductory Approach

-SEd.4 WHAT IS CALCULUS 1 7

It is instructive also to consider a set of values of (x2-9)/(x-3) as X-> 3:.

X

3.1 3.01 3.001 3.0001 3+ l0-10

(x'-9)/(x-3)

6.1 6.01 6.001 6.0001 6+ l0-10

This table of values suggests the germ of the idea of limit, namely that

x2-9 lim-- = 6 X--+3 X- 3

means that (x2-9)f(x-3) gets closer and closer to 6 as x approaches 3. We shall formulate this technically in the next chapter.

Problems

l. Find the slope of the curve y = 2x' at the point (3, 18). 2. Find the slope of the curve y = x2 at the point (2, 4). 3. Find the slope of the curve y = x2 at the point (l, l). 4. Find the slope of the curve y = x2 at the point ( -3, 9). 5. Solve the preceding problem by using the information that the slope

of the curve y = x' is 6 at the point (3, 9), and the fact that the graph of the curve is symmetric about the y-axis.

1.4 The Notation for a Sum. Before giving an example in integral calculus, we introduce the mathematical notation for a sum. Consider the sum of the integers from l to lOO,

l +2+ 3 +4+ ... +98+ 99+ lOO.

This is written

lOO "i,j, j=l

meaning the sum of integers j ranging from j = l to j = lOO. Notice that j plays a dummy role here, and we could just as well write

100

L k or k~l

lOO

L i or i=l

100

"i,h. h=l

Page 16: Calculus an Introductory Approach

8 CALCULUS: AN INTRODUCTORY APPROACH CHAP. l

If we wanted to indicate the sum of the integers from 7 to lOO, we would write

100 7+8+9+ ... +98+99+100 = 2j or

j-7

100

2 k or k~?

100

2 i. i=7

If the sum of the square of the integers from 7 to l 00 were desired, then we would write

100 72+82+92+ ... +982+992+1002 = 2P·

j=7

In general, if m and n are any two integers with m < nand ifj(x) is a function which is defined (i.e., has meaning) for all the integers from m ton, then

n 2 j(j)

j=m

is an abbreviated way of writing the sum

f(m) + f(m+ l) + j(m+ 2) + ... + j(n-1)+ f(n).

Here are some more examples:

15

2 j3 = 103+113+123+133+143+153; j=lO

90

2 j(i+l) = 3. 4+4. 5+5. 6+ ... +89. 90+90. 91; j-3

70

2 k2 = 20'+212+222+ ... +69'+702; k'-20

70

2 (k2+l) = 202+1+212+1+222+1+ ... +692+1+702+1; k=20 ., n

2P = l2+22+3'+ ... +(n-l)2+n2. j=l

Problems

\Vhich of the following are correct, and which incorrect? Give reasons for each conclusion.

70 70

l. 2 (k2+l) =51+ 2 k2. k=20 k=ZO

Page 17: Calculus an Introductory Approach

SEc: 5 WHAT IS CALCULUS 1 9

100 99

2. 2:P = lOO+ 2_)2. j=1 j=1

100 100

3. 2_j'= I+ 2: '2 J . j~1 j~2

100 99

4. 2_j' = 10001+ 2: '2 J . j=l j-2

n n-1

5. 2_j' = n2+ L ·z J . j-1 i=1

100 100

6. 2: (2+j) = 2+ 2j. j=1 i'""l

100

7. 2: 2 = 200. j~1

n 8. 2: I = n.

j=1

n n 9. 2 3j' = 3 2 j'.

.1-1 j-1

n n 10. 2j= 2i·

j=O j=l

n n 11. 2 (j+l) = 2 (j + 1).

i'-0 j~1

n j I n 12. 2'- 2' --- J,

j=1 c c j=l

where c is a constant, not zero. 10 10 10

13. 2 .i' = 2 j . 2 j'. j=1 j=l j~l

1.5 Some Special Sums. It is well known that the sum of the positive integers from I to n is equal to n(n +I )/2, thus

(2) ij=n(n+l). j-1 2

Although this is commonly proved in elementary mathematics by the

Page 18: Calculus an Introductory Approach

10 CALCULUS: AN INTRODUCTORY APPROACH CHAP. 1

use of arithmetic progressions, we give here a proof by the method of differences. The equation

(x+ 1)2-x2 = 2x+ 1

is an identity in x, that is, it is true for every value of x. Replace x successively by x = 1, 2, 3, ... , n to get then equations

22- 12 = 2 . 1 + 1

32-22 = 2. 2+ 1

42-32 = 2 · 3 +I

n2-(n-1)2 = 2(n-1)+1

(n+1)2-n2 = 2n+l. We add these equations, noting the extensive cancellation on the left, to get

n n n (n+1)"-12 = _L2j+ ,L 1 = ,L 2j+n.

j=! j=l j=!

We note that n n

(n+1)"-12 = n2+2n, and ,L 2j = 2 _Lj, j=l j=l

and substituting these we get n

n2+2n = 2 _Lj+n, l"l

n 2 _Lj = n2+n = n(n+l).

j~l

Dividing by 2, we obtain formula (2). Next we prove by ~ similar method that

n 1 I I (3) ,LP= -ns+ -nz+ -n.

j~l 3 2 6

To establish this, we begin by noting that a simple expansion of (x +I )3 leads to the identity

(x+ 1)"-x3 = 3x2+3x+ I.

Substituting x = I, 2, 3, 4, ... , n we get n equations

23-13 = 3. 12+3 ·1+1

32-23 = 3 . 22 + 3 . 2 +I

n3- (n-1)3 = 3(n-1)2+3(n-1)+ I

(n+1)3-n3 = 3n2+3n+l.

Page 19: Calculus an Introductory Approach

SEC. 5 WHAT IS CALCULUS?

Adding, and noting the cancellation of terms on the left, we get

n n n (n+lJ3-13 ~ 3 2:P+3 2:j+ 2: 1.

j=l j=l j=l

From elementary algebra, and from formula (2), we get

(n+I)3 ~ n3+3n"+3n+I,

and hence 3n(n+I)

n3+3n'+3n ~ ---- +n+3 2

This reduces to formula (3), by some such steps as

n L I~ n,

j=l

3n(n+ I) 3n2 n 3 'J'" ~ n"+3n"+3n- -n ~ n3+- +-~ 2 2 2'

I I I 2:P ~ -ns+ -nz+ -n.

3 2 6

Problems

I. Find the value of l+2+3+4+ ... +I000. 2. Find the value of l"+2'+32 +4'+ ... +I002. 3. Find the value of 50'+ 5J2+ 522+ ... +lOO'. 4. Evaluate the sum

5. Prove that

6. Prove that

7. Evaluate the sum

8. Evaluate the sum

n-1 . n(n-1) 2-J ~ . j~-1 2

n n 2_ j2 ~ - (n+ l)(2n+ l). j=l 6

n-1 ''2 ~J. j=l

n-1

" '2 ~J. j=O

11

Page 20: Calculus an Introductory Approach

12 CALCULUS: AN INTRODUCTORY APPROACH CHAP. 1

9. Prove that n n-1

2: U-1)2= 2:P-j=! j=l

10. Evaluate the sum n

2: (~i+3j2). J~l

ll. Verify the identity (x+1)'-x' = 4x3+6x'+4x+l. *12. Use the identity of the preceding question to find a formula for

n '·s "'"J . J=l

1.6 An Example from Integral Calculus. We shall now use the algebraic preamble above to calculate the area bounded by a parabolic arc and straight lines. The problem we solve was settled by the Greeks in ancient times, but their methods were by comparison more difficult and less suited to generalization to other problems.

Consider the parabola y = x2, y- axis that is, th<> collection of points

whose coordinates satisfy this Y=X

2 equation. In particular, we are

x-axis

interested in the portion of the curve between (0, 0) and (1, 1). The problem is to find the area A of the configuration bounded above by the curve, below by the x-axis from (0, 0) to ( 1, 0), and on the right by the line joining (1, 0) and (1, 1). Calling the line segment from (0, 0) to (1, 0) the

FIG. 1.9 baseline, we divide it into a whole number n of equal parts, each of

length 1/n. The division points are

(0,0), (~.o). (~.o) ..... c: 1.o). G·o). Drawing vertical lines from these points on the baseline, we construct a pattern of rectangles as in Figure 1.10. On each segment of the base­line there stand two rectangles. For example, on B1B2 there stand the lower rectangle B,B2C2C1, and the upper rectangle B1B2D2D1. Since

Page 21: Calculus an Introductory Approach

SEC. 6 WHAT IS CALCULUS 1 I3

the points Cr and D2 are on the curve y = x2, their coordinates are

. (j- I (.i- I )2 ) . c,. --, , n n 2

D2: (LJ2 )·

n n2

y-axis

x-axis

FIG. 1.10

Because the base segment B1B2 has length !Jn, we have the area for­mulas

and

area BrB2 C2 Cr I (j-I)2

n n 2

(j- I )2

n"

I j2 j2 area BrB2D2D1 = - · - = -

n n2 n3

The sum of the areas of the lower rectangles is less than the area A under the parabola, whereas the sum of the upper rectangles is greater than A. In symbols, we have*

n (j -I)2 n j2 2 <A< 2-. J~l n3 J=l n3

(4)

* It is presumed that the reader is familiar with the symbol <,which is used in this chapter in its simplest aspect. In later chapters more elaborate use is made of in­equalities, and so a treatment of this topic is given in § 2.1.

Page 22: Calculus an Introductory Approach

14 CALCULUS: AN INTRODUCTORY APPROACH CHAP. l

The n" can be factored out of each sum, and applying formula (3) of § 1.5 we get

n j2 1 n .L- =- .LP i=l n3 n3 J=l

n

.L (j- I)'

j=l

I I I = -+-+-,

3 2n 6n'

l n 1 n-1 I ( n ) _L (j-1)' =- _L k' =- _L lc'-n'

n3 J=l n3 k=l n3 k=l

= _J_(~n3+ ~n'+ ~n-n') = ~- _2 + -1-.

n3 3 2 6 3 2n 6n2

Thus the above inequalities can be written

I I I l l l (5) ---+-<A<-+-+-,

3 2n 6n' 3 2n 6n'

and we shall see that A is thus trapped between values which get arbitrarily close to l/3 as n increases indefinitely.

The expression on the left of ( 5) can be written in the form

~-__1_+_1 __ ~-_1_(3n-l) 3 2n 6n2 3 2n 3n

and since l/2n and (3n-l)j3n are positive, this shows that

I I I l ---+- < -. 32n6n' 3

But though the expression on the left of (5) is less than 1/3, we see that we can make it as close to 1/3 as we please, because as n gets very large, lj2n and lj6n2 get very small. We can state this in another way by saying that as n tends to infinity, lj2n and lj6n2 tend to zero, in symbols thus

l lim- = 0, n-KD 2n

l Iim- = 0, n-)co 6n2

The meaning of these statements will be made more precise in the next chapter. For the present, we see that also

lim (~+__J_+-1-) = ~. n-•oo 3 2n 6n2 3

and so from (5) we can conclude that A = 1/3.

Page 23: Calculus an Introductory Approach

SEC. 7 WHAT IS CALCULUS 1 15

We have used limiting processes to compute the area A. It should not be concluded that the only application of integral calculus is to the calculation of areas. The principal concept is not the area, but the limit of a sum-in our example the limit of a sum of areas of lower rectangles and the limit of a sum of areas of upper rectangles. The idea of a limit of a sum occurs in many forms throughout advanced mathe­matics.

Problems

!. Find the area bounded by the parabola y = x' from (0, 0) to (1, 1), the y-axis from (0, 0) to (0, 1), and the straight line from (0, 1) to (1, 1). Suggestion: use the answer A = 1/3 from § !.6.

2. Find the area bounded by the parabola y = x' and the straight line joining the points (1, I) and (-1, 1).

3. Find the area enclosed between the parabola y = x2 and the straight line y = x.

4. Find the area bounded by the parabola y = I +x2 from (0, 1) to (I, 2), and by the three straight-line segments from (I, 2) to (I, 0), from (I, 0) to (0, 0), and from (0, 0) to (0, 1).

*5. Find the area bounded by the parabola y = x' from (0, 0) to (2, 4), the straight line from (0, 0) to (2, 0), and the straight line from (2, 0) to (2, 4).

*6. Find the area bounded by the parabola y = 3x' from (0, 0) to (1, 3), the straight line from (0, 0) to (I, 0), and the straight line from (I, 0) to (I, 3).

1.7 A Refinement. We took the position in the preceding section that we knew in some sense what is meant by the "area'' of a configura­tion bounded partly by a curve. However, that was a little naive, as the following argument demonstrates. The idea of area is first intro­duced in terms of a square unit, that is, the area of a square each of whose sides is of unit length. For example, if the unit of length is an inch, we are talking about one square inch. From this unit of area we extend the concept to the area of any rectangle, and thence to any

D Unit area Rectangle Triangle

Fw.l.ll

Polygon with 5 sides

triangle, and from there to any figure that can be triangulated, namely any polygon. We do not propose to go through the steps of this develop­ment, but we draw attention to the essential difficulty that arises at

Page 24: Calculus an Introductory Approach

16 CALCULUS: AN INTRODUCTORY APPROACH CHAP. I

the next step, when we introduce figures bounded, at least in part, by curved lines. What is meant by the area of such a figure? This question is not answered in this book, but we do point out the direction the answer takes. The point is this: whereas in § 1.6 we talked simply as though we had a computation to perform, the whole process should have been separated into (i) a discussion of the definition of area in terms of the limit of a sum of rectangles, as in Figure 1.9, and (ii) the subsequent computation based on such a definition. This separation of the problem of area into two parts is a relatively modern development in mathe­matics.

Page 25: Calculus an Introductory Approach

CHAPTER 2

LIMITS

2.0. From the sample problems of the preceding chapter, it is clear that calculus is based on limiting processes, and so we now discuss the question of limits at some length. The theory of limits, in turn, is based on inequalities between real numbers. Whereas a lengthy treatise on calculus will ordinarily begin with an analysis of the real number system, the following observations must suffice for us.

Let a straight line be marked off with coordinates, after the fashion of the x-axis in coordinate geometry. That is, a zero point and a unit length are chosen, and this unit length is used to determine the location of the points marked I, 2, 3, 4, ... on one side of the zero point, and the points marked -I, -2, -3, -4, ... on the other side.

-4 -3 -2 -I 0 I 3 4

Then to each point on the line there belongs a number which apart from sign denotes the distance of that point from the origin. The collection of all snch numbers constitutes the system of real numbers, and when we refer to a "number" in this book, the referent is one of these real numbers.

Continuing our intuitive description of real numbers, we observe that all numbers belonging to points on the right side (the same side as the point I) are positive, and all on the left side are negative. By a non-negative number is meant one which is positive or zero. We shall take as axiomatic-without proof that is-the following properties of real numbers: both the sum and the product of two positive numbers are positive numbers; the sum of two negative numbers is negative, but their product is positive; the product of a positive number and a negative number is negative; for any real number a, a+ 0 = a and

17

Page 26: Calculus an Introductory Approach

18 CALCULUS: AN INTRODUCTORY APPROACH CHAP. 2

a · 0 = 0; the sum and product of two non-negative numbers are non­negative; the sum of a positive number and a non-negative number is positive.

If two numbers a and b are equal we write a = b, and if they are not equal we write a of b. If two or more mathematical statements imply one another, we say that they are equivalent. For example, the four equations x = y, y = x, x- y = 0 and y- x = 0 are equivalent. If a of 0 then a-1 or Ija is called the inverse or the reciprocal of a.

2.1 Inequalities. The verbal statement that 2 is less than 4 is written mathematically in the form 2 < 4. This can also be written 4 > 2, i.e., 4 is greater than 2. These inequality signs are also used with variables; for example x > 2 states that x is greater than 2. The notation x ~ 2 states that x is greater than or equal to 2.

DEFINITION. The inequalities x > y and y < x are equivalent, and they mean x-y is positive, or, what is the same thing, y-x is negative. Thus the following four inequalities are equivalent:

x > y, y < x, x-y > 0, y,__x < 0.

Similady x ~ y means that x- y is positive or zero, and the following four inequalities are equivalent:

x ~ y, y ~ x, x-y i 0, y-x ~ 0.

That the sum and product of two positive numbers are positive can be written mathematically as follows:

If a > 0 and b > 0, then a+ b > 0 and ab > 0.

Other statements in the preamble preceding this section can be written as inequalities, notably these:

If a ~ 0 and b ~ 0, then a+ b ~ 0 and ab ~ 0; if a ~ 0 and b > 0, then a+ b > 0.

THEOREM I. If a > b and b > c, then a > c.

PROOF. From the definition we see that a-b and b- c arc positive, so their sum (a-b)+(b-c) is also positive. Thus a-c is positive, and hence a > c.

THEOREM 2. If a> b then a+c > b+c and a-c > b-e for any number c. If a > b, and c is positive, then ac > be.

PRoOF. Since a+c > b+c amounts to (a+c)-(b+c) > 0, and so to a-b.> 0, we see that it follows from a > b. A similar argument

l

Page 27: Calculus an Introductory Approach

SEC. 1 LIMITS 19

applies to the second conclusion. To see that ac > be, we note that a-b and c are both positive, and so is their product ac- be. Thus all parts of the theorem are established.

THEOREM 3. Inequalities can be added. If a > b and c > d, then a+c > b+d.

PROOF. Since a-b and c- d are positive, so is their sum (a-b) + (c-d). Hence we have

(a+c)-(b+d) > 0, a+c > b+d.

THEOREi>I 4. Inequalities can be multiplied sometimes. If a > b and c > d, and if b and d m·e non-negative, then ac > bd.

PROOF. Since a > band b is non-negative, it follows that a is positive. Similarly c is positive. We separate the proof into two cases, b ~ 0 and b > 0. In case b ~ 0 then bd ~ 0, but ac is positive, so ac > bd.

In case b > 0, we apply Theorem 2 to c > d and b to get be > bd. We also apply Theorem 2 to a > b and c, to get ac > be. Hence by Theorem 1, ac > bd.

THEOREM 5. The inequalities a > b and -a < - b are equivalent.

PROOF. Each inequality means that a-b is positive. Note that this theorem says in effect that an inequality is reversed

if both sides are multiplied by - 1. A similar effect is produced (some­times) by taking reciprocals:

THEOREM 6. If a and b are positive, then the inequalities

are equivalent.

1 1 a> b and - <­

a b

PROOF. Applying the last part of Theorem 2 with (ab)-1 playing the role of c, we get

1 1

b a a(ab)-1 > b(ab)-1, - >

Conversely, if we assume that (1/b) > (1/a), we can multiply both sides of this inequality by ab to get a > b.

It may be noted that Theorems 1 to 6 remain valid if each inequality sign > or < is replaced by ~ or ;;; respectively. The reason for this is that the only principles used in the proofs of these theorems are that the sum and the product of two positive numbers are positive. But

Page 28: Calculus an Introductory Approach

20 CALCULUS: AN INTRODUCTORY APPROACH CHAP. 2

we can replace these by the principles that the sum and the product of two non-negative numbers are non-negative. This is all that we need to observe, because a ~ b is equivalent to a-b ~ 0, and it means that a-b is non-negative.

One of the principal uses that we will have for inequalities is to designate sets of values. If we want to indicate that x is positive we can write x > 0. If we want to say that y is non-negative, we can write y ;oo 0, or what is the same thing, 0 ;;; y. Again, if we want to say that x lies strictly between 3 and 7, we can write 3 < x < 7. If x lies between 3 and 7 with the end values included, we write 3 ;;; x ;;; 7. Another useful device for indicating sets of numbers is the notion of absolute value, which we discuss in the next section.

Problems

I. Give a numerical example to show that the second sentence of Theorem 2 does not hold if the restriction " c is positive" is removed.

2. Give an example to show that "if a > b and c > d then ac > bd" does not hold under all circumstances. (Compare this with Theorem 4).

3. Give an example to show that Theorem 6 does not hold if the restriction "a and b are positive" is removed.

4. Prove that if a ~ b and b > c then a > c. 5. Prove that if a-;oo b and c > d then a+ c > b +d. If furthermore b

and d are positive, prove that ac > bd. 6. Prove that ljn2 < ljn holds for every positive integer n, with one

exception. 7. If the temperature rp in the last 24 hours ranged from 50 to 80 degrees,

show that this can be written mathematically in the form 50 ~ T ~ 80. Show that an alternative way of writing this is -15 ~ T- 65 ~ 15.

8. Given that '1' satisfies the inequalities -20 ~ '1'- 62 ;;;::; 20, what are the smallest and largest possible values of T?

*9. Let a and b be two different numbers. (This is written mathematically a# b.) Prove that a'+b' > 2ab. Suggestion: a-b is not zero, so (a-b)' is positive.

*10. Let c and d be positive numbers. Their arithmetic mean is (c+d)j2, and their geom({tric mean is V cd. Prove that if c ¥=- d, then their arithmetic mean exceeds their geometric mean. Suggestion: VC and Vd are unequal, so ('Vc- V d)' is positive.

11. Which of the following propositions are true for all real numbers a, b, c, d, and which are false? Give reasons for your answers.

(i) If a# band c # d then a+c # b+d. (ii) If a# b then c+a # c+b.

(iii) If a # b, a # 0, b # 0, then !fa # ljb. (iv) If a # b and c # d, then ac # bd. (v) If a # b then a 2 # b'.

(vi) If a # b and c # 0, then ac # be.

Page 29: Calculus an Introductory Approach

SEC. 2 LIMITS 21

12. Which of the following propositions are true, and which false? (Remark: A mathematical proposition is true only if it is universally true, not just in some cases or some circumstances.)

(a) The sum of two positive numbers is non.negative. (b) The sum of two non-negative numbers is positive. (c) The sum of a positive number and a non-negative number is positive. (d) The product of a positive number and a non-negative number is

positive. (e) The product of two positive numbers is non-negative. (f) The product of two non-negative numbers is positive.

2.2 Absolute Values. Let x be any real nnmber. The notation lxl stands for "the absolute value of x", and it is defined as follows:

if xis positive or zero, jxj = x; if x is negative, lxl = - x.

We give some eX'\l.mples:

131 = 3, 1- SI = s, l-141 = 14, IOI = 0, 161 = 6.

There is another way of defining lxl that amounts to the same thing, namely lxl = V x2. This is the same as the earlier definition because of the convention in mathematics that the v symbol always denotes a non-negative value in the real number system. For example we see that

v(- S)2 = v64 = S, so 1- SI = v(- s)2.

Notice that ISI = 1- SI, and in generallxl = 1-xl-

THEOREM 7. The absolute value of a product of two real numbers equals the product of their absolute values, i.e., lxyl = lxl · IYI·

PROOF. First we give a direct proof, and then a short, more sophisti­cated proof. In case x or y is zero, then xy is zero and we see that

lxyl = 0, lxl · IYI = 0, so lxyl = lxl · IYI·

There remain the cases where neither x nor y is zero. If both are posi­tive, then their product xy is also positive, and we see that

lxyl = xy, lxl = x, IYI = y, so lxyl = lxl · IYI·

If both x and y are negative, then xy is positive, and we see that

lxyl = xy, lxl = -x, IYI = -y, so lxyl = lxl · IYI·

Finally we must consider the case where either x or y is positive and the

Page 30: Calculus an Introductory Approach

22 CALCULUS: AN INTRODUCTORY APPROACH CHAP. 2

other negative. There is no loss of generality in taking x positive and y negative. Now xy is negative, and it follows that

lxyl = -xy, lxl = x, IYI = -y, so lxyl = lxl . IYI· Now we give a shorter proof of Theorem 7. By using the definition

lxl = V x2 we see that

lxl = vx', IYI = yy2, lxl . IYI = yx2 vY2 = yX2Y2 =I xyl.

THEOREM 8. Let x, y, u, v be any real numbers satisfying lxl < u, IYI < v. Then these inequalities can be multiplied to give lxyl < uv.

PROOF. Because lxl and IYI are non-negative whatever may be the values of x and y, we can apply Theorem 4 with a, b, c, d replaced by u lxl, v, IYI, respectively. Thus we conclude that lxl · IYI < uv, and from this the present theorem follows by the use of Theorem 7.

A moment's reflection will reveal that such a statement as "x lies strictly between -3 and + 3" can be written in two ways in mathe­matical symbolism:

-3 <X< 3 and lxl < 3. This can be conceived pictorially on the real line:

<- set of values of x .....

I I I I I I I I I I I I I I I I I I

-4 -3 -2 -l 0 l 2 3 4

Similarly if we wanted to say that x lies strictly between - lQ-5 and lQ-5, we could write lxl < J0-5. Thus x would be very small numeri­cally. However, degrees of smallness are a relative matter, and so we shall rely on the mathematical symbols to tell the story.

How can we write mathematically that x is (strictly) within J0-5, plus or minus, of the value 31 Using inequalities we could write

3-JQ-5 <X< 3+ JQ-5.

By theorem 2 we can subtract 3 across the series of inequalities to get

-lQ-5 < x-3 < J0-5.

Now this can be written lx-31 < J0-5. Since lx-31 and 13-xl are the same, this can also be written 13- xl < J0-5.

Page 31: Calculus an Introductory Approach

SEC. 2 LIMITS 23

Thus the symbol lx- 31 is a measure of how near xis to 3. Similarly lx- ti ~s a measure of how near x is to t, when these are regarded as numbers on the real line. More precisely,

X

I I I I I I I I I I I I I I I I

-4 -3 -2 -l 0 l 2 3 4

lx- tl is the distance between the points represented by x and t. Here distance is defined in terms of the unit length, which in turn is defined as the length of the line segment from the point 0 to the point l.

ExAMPLE: What values of x are described by the inequality lx- 51 :£ 21 Stated in terms of the geometry of the real line, this says that the distance from the point x to the point 5 is at most 2. Hence the inequality denotes the set of all real numbers from x = 3 to x = 7. This set can also be described by the inequalities 3 :£ x <; 7.

Problems I. Given that a variable x stays strictly between 3 and 7, show that this

can be written mathematica1ly in any one of the following ways:

(a) 3 < x < 7 (c) lx-51 < 2

(b) 7 > x > 3 (d) 2 > lx-51 2. Given that a variable x stays strictly between 4 and 10, write this

mathematically in four ways, analogous to those in Question 1. 3. Given that a variable x stays strictly between lOO and 200, write this

mathematically in four ways. 4. Given that a variable x stays strictly between f and - t, write this

mathematically in four ways. 5. When a certain thermostat is set at 72, it does not allow the temperature

T to vary from this by more than 2 degrees. Write this mathematically in four ways.

6. Given that two real numbers x and y differ by at most 3, write this mathematically in as many ways as you can (not more than eight). Remark: Actually, any inequalitity can be written in infinitely many ways. For example x > 4 can be written 2x > 8, 3x > 12, 4x > 16, 5x > 20, etc.

7. Given that two real numbers x and y differ by at most E, where E is positive, write thi.s mathematically in as many ways as you can, up to eight.

8. Write each of the following inequalities without absolute value signs, given that E is positive:

(a) lxl < 6 (e) lx-31 "'

7

(b) lxl < < (f) lx-31 ~ €

(c) lx-31 <7 (g) lx-71 < 3

(d) lx-31 < < (h) lx-3! :£ 7

Page 32: Calculus an Introductory Approach

24 CALCULUS: AN INTRODUCTORY APPROACH ~--

CHAP. 2

9. Write each of the following inequalities with absolute value signs:

(a) -9 < X < 9 (d) 12 < X < 20

(b) -9;;; x;;; 9 (e) -4 < x-3 < 4

(c) 14 < x < 30 (f) -4 < x-3 < 0

2.3 Sequences and Limits. In our examination of sample problems from calculus in Chapter 1, we were led in each case into a problem in limits or limiting processes. We now study the problem systematically, confining our attention, however, to cases that will arise in subsequent chapters.

A sequence is an ordered collection of terms, as for example

2 3 4 5 6 7

l' 2' 3' 4' 5' 6' ... ' 111111

or 2' 4' 8' 16' 32' 64' ... ,

or 0, l, 0, 1, 0, I, 0, 1, ... .

Some sequences tend to limits, others do not. In the three examples of sequences just cited, the first sequence tends to the limit 1, the second sequence to the limit 0, and the third sequence has no limit. Another example is the well-known sequence

.3, .33, .333, .3333, .33333, ... '

and this has limit }. The sequence

1 1 1 1 1

l' 2' 3' 4' 5' ... is tending towards the value 0, so the limit of the sequence is 0. Since the nth term of the sequence is 1/n, we write

1 lim- = 0. (1) n-)oo n

In words, the limit of 1/n as n tends to infinity equals zero. In the examples given it may have been noted that the symbol n

is used to denote a discrete variable, that is, one of the positive in­tegers n = 1, 2, 3, 4, .... On the other hand, letters like x and y are customarily used to denote continuous variables. For example, if we write

(2) l

lim- = 0, X-iCO X

Page 33: Calculus an Introductory Approach

SEC. 3 LIMITS 25

we mean that as x increases indefinitely through real values the limit is zero. Thus in (2) we think of x moving to the right over the positive end of the real line,

l 2 3 4 5

I I I I I

whereas in (l) the variable n jumps from positive integer to positive integer along the real line. Although the reasoning underlying ( l) and (2) is very similar, we make a distinction between these two. The first is called the limit of a sequence, the second the limit oL a function.

What we have done thus far is purely descriptive. We now turn to an exact formulation of the idea of a limit. Consider a general infinite sequence of real numbers

(3)

whose nth term is an. We designate such a sequence briefly by {an), using braces around the nth term. Now if this sequence is to have limit a, what this means is that the difference an- a is tending to zero as n gets larger and larger. Thinking of n changing from l to 2 to 3 to 4 and so on, the difference an- a gets as small as we please, and stays small, for all values of n that are large enough. For instance, we require that an- a shall be less than l/10,000,000 for all n beyond some parti­cular value. Furthermore, this must be true not only for l/10,000,000 but for any positive numerical value, as small as we please. To make this precise, we state it as follows.

DEFINITION. The sequence (3) has limit a provided that for any given positive number, < no matter how small, the inequality

(4) )an-a\<£, Le., -e <an-a< e

holds for all but a finite number of terms of the sequence (3).

When this is the case we write

lim an = a, or an -+ a as n -> oo; n~oo

in words, the limit of an as n tends to infinity is a, or an tends to a as n tends to infinity.

Note that the effect of the requirement (4) is to make certain that an is numerically close to a, at least for large values of n-for example,

Page 34: Calculus an Introductory Approach

26 CALCULUS: AN INTRODUCTORY APPROACH CHAP. 2

if E is very small, say E = 10-20, then an must be within 10-20 of a, ex­cept for a finite number of values of n. Furthermore, in order that a sequence may have a limit, the inequalities (4) must hold for < = lQ-50, • = 10-1oo, in fact for any given positive <, not for all values I, 2, 3, ... of n, but for all but a finite number.

Let us apply the definition to the simple sequence 1/1, 1/2, 1/3, 1/4, ... , and let us test whether our claim that this has limit 0 meets the re­quirement (4). To be specific, let us take< = 1/100. Then (4) becomes

I ~ - ol < _I_ i.e., n wo'

I I I -- <- < -.

100 n !00

It is clear that these inequalities are satisfied by all terms of the sequence {1/n} except the first lOO terms, namely the terms

I I I I -,-,-, ... ,-. I 2 3 100

Let us take another value for <, say • = 10-20. Then ( 4) becomes

,I I

1

1 i - -01 < lQ-20, ,n

i.e., I

--- <- < --. 1020 n 1020

These inequalities hold except for n = 1, n = 2, n = 3, ... , n = 1020, and this is a finite number of exceptions.

The only trouble with taking < = 1/100 and • = 10-20 is that we have verified (4) for just two values of •· Nevertheless these two cases suggest how we can deal with the general value of <. For any given positive< the inequality (4) for the sequence 1/1, 1/2, 1/3, ... takes the form

I (5) -€ < - < €.

n

Now the inequality- • < (1/n) certainly holds since nand< are positive. Furthermore, (1/n) < • holds for all values 0f n satisfying n > _-1, because these two inequalities are equivalent by Theorem 6 of § 2.1. Thus we have established that the sequence {1/n} has limit zero.

Not every sequence has a limit. For example, consider the sequence -I, I,.-!, 1, -I, I, ... whose nth term is ( -J)n. It is intuitively clear that this sequence has no limit, but we can establish this conclusion analytically in terms of the definition. For suppose there were a limit, say a. Then let the given number • be taken as }. By (4) the inequalities

(6)

Page 35: Calculus an Introductory Approach

SEC. 3 LIMITS 27

would have to hold for all but a finite number of values of n. Now either a ;:"; 0 or a < 0, and we treat these two possibilities separately.

In case a ;:-; 0, then for n = 1, 3, 5, 7, ... the absolute value in (6) becomes I( -1)- al. But this is a measure of how near -1 is to a, so we see that I( -1) -a I ;:"; 1, and hence (6) fails to hold for all odd values ofn.

In case a < 0 we look to the even values of n to establish the result. For n = 2, 4, 6, 8, ... the absolute value in (6) becomes 11- al. Think­ing of this as a measure of how near I is to a, we have [I- a[ > I, and so (6) fails to hold for even values of n. Hence the sequence {(- 1)"} has no limit.

The definition does not tell us what the limit of a sequence is. That is determined by educated guessing. The definition can be used to test whether the guess, or conjecture, is right, as we tested the limit 0 for the sequence {1/n}. However, we shall not ordinarily take each sequence to the definition as a test, but rather we shall use some con­sequences of the definition as our criteria. That is, from the definition we now prove some theorems* which enable us, by starting from certain simple sequences like {1/n}, to conclude that other more complicated sequences have limits, and also to determine what these limits are.

THEOREM 9. Let {an} be a sequence of non-negative numbers such that an -+ 0 as n -+ w, and let {bn} be a sequence of non-negative num­bers satisfying b, ~ a,, b, ~ a,, ba ~ a3, etc. Then bn -> 0 as n -+ oo.

PROOF. Since the sequence {an} satisfies {4) with a replaced by 0, then for any positive number < the inequalities

-E<an<e

hold for all but a finite number of terms of the sequence. But - < < an < e implies - e < bn < e because bn is non-negative, and because

bn ;;;; an. Hence the theorem is proved.

CoROLLARY 10. As n -+ w, (1/n) -> 0, (1/n') -> 0 and in general (ljnk) -> 0 for any positive integer le.

PROOF. We have already proved that lim 1/n = 0, and since by Theorem 6 (1/n2) ~ (1/n) we conclude that (1jn2) -+ 0 as n -> w, by Theorem 9. A similar argument applies to 1jnk.

* Some of the proofs of the theorems on limits in § § 2.4 and 2.8 might well be omitted at a first reading, especially if these proofs seem obscure or difficult. However, the reader should certainly study the statements of the theorems with care in order to comprehend their meanings.

Page 36: Calculus an Introductory Approach

28 CALCULUS: AN INTRODUCTORY APPROACH CHAP. 2

We conclude this section with an observation on our definition of limit of a sequence. An alternative definition which is commonly given in books on mathematics (but which we shall not use) is that the sequence (3) has limit a provided that for any given positive number <there is a fixed integer N such that - < < an -a < <for all values of n exceeding N. Note that this amounts to saying that the inequalities on an- a hold for all but a finite number of values of n. i

Problems

I. Find the limits (if any) of the following sequences:

(a)

(b)

(c)

Ill! I I -, -, -, -, -, ~, ... ; 2 4 6 8 10 12

1 1 I z' 4'- 6'

I 1 3'- 10'

3 4 5 6 7 8 9 10 -, -, -, -, -, -· -, -, ... ; 1234567 8

I -, ... ; 12

(d) I I 1 I I I

1, -, l, -, l, -, l, -, I,-, 1, -, ... ; 2 3 4 5 6 7

(e) I 2 3 4 5 6

1, -, l, -, l, -, l, -, l, -, 1, -, ... ; 2 3 4 5 6 7

I I I I I (f)

2. What is the limit of the sequence 2, 2, 2, 2, ... ? of the sequence c, c, c, c, ... 1 3. Given that the sequence a1, a2, aa, a4, ... has limit a, what would you

expect to be the limits of the following sequences:

(ii) I +a1, 1 +a2, I +as, I +a4, ... ; (iii) ia1, ia,, ias, ia4, ... ; (iv) a1-a, a2-a, aa-a, a4-a, ....

Give proofs in cases (iii) and (iv).

4. If the nth term of a sequence is 1/(n'+n), what are the first three terms? What is the limit of this sequence?

t

Page 37: Calculus an Introductory Approach

SEC. 4 LIMITS 29

5. Given that the sequence a1, a2, a3, a4, ... has limit a, what are the limits of the sequences

(l) a4, as, a6, a7, as, ag, ... ;

(2) a2, a4, a6, as, a10, a12, ... ?

Give a proof in each case. *6. Prove that the sequence .3, .33, .333, .3333, ... has limit}. Suggestion:

use the formula a(J-rn)/(1-r) for the sum of n terms of the geometric . 2 3 progressiOn a, ar, ar , ar , ....

2.4 Theorems on Limits. In many cases it is a lot of bother to test a sequence for a limit by using the definition. For example, one of the above problems is to show that .3, .33, .333, ... has limit}, and, with this done, there is no need to use the < process to establish that the sequences

- .3, - .33, - .333, ...

and .6, .66, .666, ...

have limits --} and !i respectively, These can be obtained by multi­plying the original sequence by - 1 and 2, respectively. The general result is as follows.

THEOREM 11. If c is any constant, and if the sequence a,, a2, aa, has limit a, then the sequence ea,, ca2, ca3, ... has limit ea. Stated more briefly, if lim an = a, then lim can = ea.

PROOF. First consider the special case c = 0. In this case the se­quence {can) consists of 0, 0, 0, ... , and the limit is 0, which is the value of ea.

Second, we consider the situation when c is not equal to zero, that is c # 0. Then [c[ is positive, and for any positive < the number </[c[ is also positive. We apply the definition of limit with •/[c[ as the "given positive number". Thus by (4) the inequality

€ [an-a[ <­

[c[ holds for all but a finite number of values of n. Multiplying this in­equality by [c[, and using Theorems 2 and 7, we see that

[can-ea[ < <

holds for all but a finite number of values of n. Hence the limit of can is c.a.

Page 38: Calculus an Introductory Approach

30 CALCULUS: AN INTRODUCTORY APPROACH CHAP. 2

THEOREM 12. Iflim an= a and lim bn = b, then lim(an+bn) = a+b, lim(an-bn) =a-b, and lim(anbn) =ab. ALTERNATIVE STATEMENT: Jj lim an and lim bn exist, then

and

lim(an + bn) = lim an+ lim bn,

lim(an-bn) = liman-limbn,

lim(anbn) = (lim an)(lim bn)·

PROOF. Part 1. Let < be any positive number. We use •/2 as the "given positive number" in the definition of limi~ for the sequences {an} and {bn}· Thus by (4) each of the inequalities

-(</2) <an-a< (</2) and -(</2) < bn-b < (</2)

holds for all but a finite number of values of n. Hence these inequalities on an- a and bn- b hold simultaneously for all but a finite number of values of n. For example, if the inequalities on an- a hold for all values of n except n = 1, 2, 3, 4, 5, 8, 9, and if the inequalities on bn- b hold for all values of n except n = 1, 2, 3, 5, 9, 12, then the inequalities hold simultaneously for all values of n except n = 1, 2, 3, 4, 5, 8, 9, 12. By Theorem 3 we can add these inequalities to get

-< < (an+bn)-(a+b) < E,

which holds for all but a finite number of values of n, and so lim(an+bn) = a+b.

Pad 2. By Theorem 11 applied to the sequence {bn} with c = -1, we conclude that lim( -bn) = -b. Then we have, using Part 1,

lim(an-bn) = lim[an+(-bn)] = liman+lim(-bn) =a-b.

Part 3. To prove that the limit of a product of two sequences is equal to the product of the limits of the sequences, we begin with the special case where the limits are zero. That is to say, we prove that if two sequences { o:n} and {f3n} have the properties lim "'n = 0 and lim f3n = 0, then lim(o:nf3n) = 0. By (4) we conclude that for any positive< the inequalities [o:n[ < E and [f3n[ < <hold for all but a finite number of values of n. Using Theorems 4 and 7 we can multiply these inequalities to get [o:nf3n[ < .z, aud this too holds for all but a finite number of values of n. Now since the whole question of limits centers on small values of<, we may certainly presume that E < 1. Multiplying this inequality by E, we get E2 < E, and so we have !cxn~nl < E2 < E.

It follows that lim( o:nf3n) = 0.

1 I

Page 39: Calculus an Introductory Approach

SEc. 4 LIMITS 31

Having proved this special case, we return to the general question of the value of lim(anbn), given that lim an = a and lim bn = b. Noting that lim(a11 -a) = 0 and lim(bn-b) = 0, we identify an-a and an, and similarly bn- b and f3n· Then we have

an= a+an, bn = b+f3n, anbn = ab+af3n+b(/.n+(/.nf3n·

The last equation suggests that we can study lim(anbn) in the following way:

lim(ab + af3n + b(l.n + (l.nf3n)

= lim(ab) + lim(af3n) + lim(b(/.n) + lim('·nf3n)

= ab+a limf3n+b lime<n+O

= ab+0+0+0 =ab.

In these steps we have used part of Theorem 12 already proved, and also Theorem 11. Thus we have proved that lim(anbn) = ab.

CoROLLARY 13. If lim an = a and lim(an -b11 ) = 0, then lim bn = a.

PROOF. We apply Theorem 12 in the following fashion:

limbn = lim[an-(an-bn)] = lima11 -lim(a11 -bn) = a-0 =a.

THEOREM 14. Let c be a constant and let {an) be a sequence with no term less than c, and {bn) a sequence with no term greater than c. In other wm·ds an ;:; c and bn ~ c for all valtws of n. If lim(an- bn) = 0 then lim an = c and lim bn = c.

PROOF. Let < be a given positive quantity. Then the inequalities

-e<an-bn<E

hold for all but a finite number of values of n, by the definition ( 4) applied to the sequence {an- bn). The inequality an- bn < < can be written in the form

(an-c)+(c-b11 ) < <.

But this implies an - c < e and c- bn < E since both an- c and c- bn are non-negative. Hence \Ve get

0 ~ an- c < e, and - E < an- c < E.

Since this holds for all but a finite number of values of n, we conclude

Page 40: Calculus an Introductory Approach

32 CALCULUS: AN INTRODUCTORY APPROACH CHAP. 2

that lim an = c. Applying Corollary 13, with c now playing the role of a, we also conclude that lim bn = c.

THEOREM 15. If no term of the seq"ence {en) is positive, and if lim Cn = c, then c ~ 0.

PROOF. We use an indirect argument. Suppose that c > 0. We apply the definition of limit to the sequence { cn), using c/2 as the "given positive number". Thus by (4),

- (c/2) < Cn- c < (c/2)

holds for all but a finite number of terms of the sequence { cn). But to the inequality - (c/2) < Cn- c we can add c by Theorem 2 to get (c/2) < Cn. However, c > 0, and so (c/2) > 0, and thus Cn > 0. But this is a contradiction, and hence we conclude that c ::'i 0.

THEOREM 16. If lim an = a and lim bn = b, and if an ::'i bn holds for n = 1, 2, 3, ... , then a ::'i b.

PROOF. We see that lim(an-bn) =a-b by Theorem 12. Applying Theorem 15 with Cn replaced by an- bn, we conclude that a-b ::'i 0, or a ::'i b.

THEOREM 17. S"ppose there are seq"ences { Sn ), { Sn} and {an} satisfying lim 8n = c, lim Sn = c and

for n = 1, 2, 3, 4, .... Then lim an = c.

PROOF. By Theorem 12 we can write

lim(Sn-Sn) = limSn-limsn =c-c= 0,

Next we subtract Sn from the given inequalities to get, by Theorem 2,

Applying Theorem 9 with an replaced by Sn- Bn, and bn by an- Sn, we conclude that

lim(an-Sn) = 0.

This implies that lim(sn- an) = 0, by Theorem 11 with c replaced by -l. Finally we use Corollary 13 to conclude that lim an = c.

Page 41: Calculus an Introductory Approach

Silo. 5 LIMITS

Problems

1. Evaluate the following limits:

(a) lim (2+ 1/n+ 1/n') (e) n7oo

4n'+n-5 (b) lim (f)

n~oo n'

100+200n (c) lim (g)

n7oo n2

(d) lim (n-l)(n-2)(n-3)

n-)XJ n3

33

0 en-] r hm --n~co n

lim -+-+-c I I ) n-)oo 3 2n 6n2

c I I ) lim ---+-n-)(o 3 2n 6n2

2. Prove that if all terms of a sequence {an} are positive, then the limit (if it exists} of the sequence must be positive or zero. Suggestion: Apply Theorem 15 to the sequence {-an)·

3. Given that the sequence {an) has limit a, and that the sequence {bn) has limit b, find the limits of the sequences:

(i) 3a,, 3az, 3a,, 3a4, 3as, 3a6, ... ;

(ii) b,-2a,, b,-2az, bs-2as, b4-2a4, ... ;

(iii) a,-az, az-as, as-a<, a4-a5, a,-a,, ... ;

(iv) a1+bz, az+bs, as+b<, a4+bs, as+b,, .. ..

*4. Given that a sequence {an} does have a limit, and that a1, aa, as, ... are all positive but az, a4, ao, ... are all negative, what must the limit be? Exhibit such a sequence.

5. Prove that iflim an = a then lim(c+an) = c+a where c is any constant.

2.5 Functions. In §§ 1.3 and 1.6 we discussed the equation y = x2

and its graph. This equation can be thought of as a rule of correspon­dence: to every real value assigned to x there corresponds a real value of y. For example, if X = 5, then y = 25; if X = - 8, then y = 64. This correspondence is an example of a function. In functional nota­tion the equation could be written j(x) = xz, where the symbol j(x) is read ''j of x." The numerical examples above can then be written as /(5) = 25 and j(- 8) = 64.

We now drop this special case j(x) = x2 to make some general observations. The word "function" is used in a very restricted sense in this book. In order to avoid a lengthy discussion of the theory of sets, we do not give a general definition of the term. Indeed, for our purposes it is sufficient to say that y is a function of x if whenever x is

Page 42: Calculus an Introductory Approach

34 CALCULUS: AN INTl~ODUCTORY APPROACH CHAP. 2

assigned a numerical value, y is determined uniquely. If we use f(x) in place of y, we can say that a function f is a correspondence which determines exactly one real value f(x) corresponding to each value of x in a certain set of real numbers X. The set X of values of xis called the domain of the function; the set of values of j(x) is called the range of the function. Notice that we are restricting the domain and the range to be sets of real numbers.

Furthermore, we shall confine our attention to functions for which there is a specific mathematical equation or formula relating x and

f(x), for example,f(x) = x2. Also, unless otherwise stated, this equation will specify both the domain and the range of the function: the domain X will consist of all real numbers for which the equation provides a unique real value for j(x); the range is simply the set of all these values of f(x).

For example in the case f(x) = x2, the domain of values of xis the set of all real numbers, and the range of values of f(x) is the set of non­negative real numbers. On the other hand, in the case f(x) = .Yx, the domain is the set of non-negative real nurrbers, because Vx is not real when xis negative. The range of this function is also the set of non-negative real numbers, because of the mathematical convention that vx is positive or zero, never negative. To take a third example, in the case f(x) = Vl-x2, the domain is the set of x values satisfying -l ::o; x ::o; l, and the range is the set of real numbers between 0 and l; thus 0 ::o; j(x) ::o; I.

Because of the restricted class of functions with which we deal, we can specify the function by its equation, and so speak of the function j(x) = x2, or what is the same thing in different notation, the function y = x2. In fact there is need for neither y nor f(x) here; we can speak of the function x2.

In the same way, we shall speak variously of the graph of the func­tion xz, or the graph of the function f(x) = x2, of the graph of y = x2, and in each case we mean the same thing, namely the graph illustrated in Figure 6 of § 1.3.

The word "function" is properly attached to the entire correspon­dence system between the domain and the range, but we will sometimes use the word loosely to refer to the range alone. The sense will be indi­cated by the context, so that there will be no need to make any fine distinction between the function, f, and the values of the function,f(x).

If we wish to discuss two or more functions simultaneously, we would not call them both f(x), of course. Other notation for functions would be used, such as g(x), h(x), F(x), G(x), etc.

Page 43: Calculus an Introductory Approach

SEC. 6 LIMITS 35

Problems

1. If the domain of values of xis the set of all real numbers) we can write this symbolically as - oo < x < oo. Write similar symbolic inequalities to _indicate that the domain of x is (a) the set of all positive real numbers; (b) the set of all non-negative real numbers; (c) the set of all negative real numbers.

2. Assuming that the equation specifies the domain and range of the function, give the domain and range of each of the following:

(a) j(x) = y4-x' (e) j(x) = 2-x

(b) j(x) = y4-x4 (f) f(x) = yx-1

(c) J(x) = v2+ v4-x' (g) f(x) = y1-x

(d) j(x) = 2x (h) j(x) = -y1-x

3. Given j(x) = 2x2-3x+4, find the values of j(1), f(2), f(3), f(4), j(O) andf(-1).

4. Verify thatj(x) = x+3 and F(x) = (x'-9)/(x-3) are almost identical functions, the difference being that whereas the domain of J(x) is - w < x < oo, the domain of F(x) is -eo < x < eo except for the value x = 3.

2.6 Radian Measure. Just as there are various units such as foot, centimeter, and mile for the measurement of length, so there are various units for the measurement of angles. The best-known measure of angles is the degree, and its subparts, minutes and seconds. This unit of measure, the degree, is obviously an arbitrary measure. A right angle contains 90°, but why not 80° or 100°? A more natural measure of an angle is the radian, which we now define. Given any angle e, we locate it in a circular sector of radius 1,sothatCA =CB = 1 in the diagram. Then the radian measure of the angle e is the length of the curved arc AB. The length of the circumference of

FIG. 2.1

a circle of radius 1 is 2rr, and since this corresponds to an angle of 360°, we have

27T radians = 360°, 7T radians = 180°.

These equations give the relation between radian measure and the common degree measure. It is customary to use no sign or symbol for radians, so that such a trigonometric equation as sin 90° = l is

Page 44: Calculus an Introductory Approach

36 CALCULUS: AN INTRODUCTORY APPROACH CHAP. 2

written in the form sin 7Tj2 ~ l. Similarly, sin 30° ~ l/2 and cos 0° ~ l are written sin 7Tj6 ~ l/2 and cos 0 ~ l.

Problems

l. Prove that 1 radian is approximately 57°18'. Suggestion: Begin with n radians = 180° and divide by tr. If -rr is taken as 22/7, which is too large, then the result will be too small.

2. Find the values of the following:

(a) COS7Tj2 (f) sin7T/4

(b) eOS7Tj3 (g) sin7T/3

(c) COS7Tj4 (h) sin7T

(d) COS7Tj6 (i) sinO

(c) tan7Tj4 ( j ) tanO

3. How many degrees are there in 1 radian? in 5 radians? in x radians? (Leave answers in terms of 1r).

4. How many radians are there in 1 degree? in 5 degrees? in y degrees? (Leave answers in terms of 1r). .

5. Prove that the radian measure of a given angle is equal to twice the area of the sector of a unit circle having the given angle at the center. (A unit circle is a circle with radius unity.)

2.7 The Limit of a Function. In § 1.3 we discussed the limit of a special function, namely

(7) x2 -9

lim-- ~ 6. x--)3 x-3

Before giving a precise definition of the limit of a function, we inquire a little further from an intuitive standpoint. Note that the function in (7) above is not defined at x ~ 3; the domain of the function (x2- 9)/(x- 3) is the set of all real numbers except x ~ 3. In case a function is defined for a specific value of x, the limit of the function and the value of the function are frequently the same, as in the cases

lim(x2 + 4) ~ 4, x~o

x2 -9 lim-- = 8, x-)5 X- 3

lim(x2-6x) ~ 16. x~s

Next we turn to a limit which will be of great importance in Chapter 4, namely

(8) sine

lim--· 0--'>0 () ,

in words, the limit of (sin 8)j8 as e approaches 0. The use of 8 in place

'i I !

Page 45: Calculus an Introductory Approach

SEC. 7 LIMITS 37

of the variable x is of no significance here. We might just as well inquire about

sinx lim--. x->0 x

Now sin 0 ~ 0, so at 8 ~ 0 the function (sin 8)/8 has the "value" OJO. But this is no value at all, and so the function (sin 8)j8 is not defined at 8 ~ 0. The domain of the function is the entire set of real numbers except 8 ~ 0.

Why do we say that OJO has no meaning? One way to answer this is to recall the relation of a fraction to an equation. The fraction 54/6 denotes the solution of the equation 6x ~ 54, namely, the solution x ~ 9. Similarly 35/7 is the solution of the equation 7x ~ 35, namely, x ~ 5. But if in a similar way we try to evaluate OJO as the solution of Ox ~ 0, we run into a difficulty. For the product Ox has the value 0 no matter what number is substituted for x. In other words every real number is a solution of Ox ~ 0. Hence 0/0 can have no meaning.

Let us return to the limit problem (8), and make a brief table of values of (sin 8)/8 as 8 tends toward 0:

sin .I -- ~ .998334166

.l

sin .01 -----:-c- ~ . 999 983 333

.01

sin .001 --- ~ .999999833

.001

sin .0001

.0001 ~ . 999 999 998

(accuracy to 9 places of decimals). These cases suggest that

lim(sin O)JO ~ I, 0->0

and this we now confirm. Con­sider a sector of a circle of radius I, with angle 8 at the center C. Let 8 subtend an arc AB on the circumference. As in the diagram, Figure 2.2, we draw a perpendi­cular from B to CA, intersecting CA at D. Also we draw a

E

B

8 90° C D A

FIG. 2.2

Page 46: Calculus an Introductory Approach

38 CALCULUS: AN INTRODUCTORY APPROACH CnAP. 2

tangent to the circle at A, intersecting the extension of CB at E. The following calculations are made readily:

radius of circle ~ CA ~ CB ~ I,

BD BD sin8 ~- ~ -, BD~ sin8,

BC I

CD CD cos8 ~- ~ -- CD~ cos8,

BC l '

EA EA tan8 ~- ~ -, EA~ tan8,

CA l

area of triangle CBD ~ j(CD)(BD) ~ f cos8 sin8,

area of triangle CAE ~ {(CA)(EA) ~ i(l)(tan8) ~ ttan8.

The area of the entire circle (of which only the arc AB is shown in Figure 2.2) is 7f · ]2 or 7f. The area of the sector CAB is in proportion to the angle at the center. That is, the sector subtends an angle 8 at the center, whereas the entire circle subtends an angle 27f. Hence the area of the sector CAB is to the area of the whole circle as 8 is to 27T, and so

8 8 area of sector CAB ~ - (7r)

21T 2. Next we see that ·

area of triangle CBD < area of sector CAB < area of triangle CAE,

or I 8 l -cos8sin8 <- < -tan8 2 2 2

Now we are considering small positive values of 8, so 2/(sin 8) is posi­tive, and we can use it as a multiplier in accordance with Theorem 2 to get

8 I cosB < -- < --.

sin 8 cos e If we take reciprocals and use Theorem 6, we obtain

l sin8 -- > -- > cos8. cos8 8

(9)

Now if 8 is a small angle, cos 8 < l and !/(cos 8) > l, and as 8 approaches 0, both cos 8 and its reciprocal tend towards l. Thus, as

Page 47: Calculus an Introductory Approach

SEC; 8 LIMITS 39

8 tends to 0, we see by formula (9) that (sin 8)/8 is caught between two values that are tending to 1. A similar argument applies to 8/(sin 8) in the inequalities preceding (9), and so we conclude that

sin 8 8 (10) lim -- = I, lim -- = I.

1)--->0 e 0--->0 sin 8

It is unfortunate in a sense that this limit should equal I, because it suggests that any limit of the form 0/0 is equal to I. If we set 8 = 0 in the fraction sin 8J8 we get the meaningless symbol OJO, which suggests the value l by analogy to such cases as 7j7, 5/5, 2/2. The point here is that

aja = I for all values of a except a = 0.

In the set of problems at the end of § 2.8, there are examples of func­tions which, although they have the form 0/0 at 8 = 0, have limits that are different from I as 8 tends to 0.

2.8 The Definition of the Limit of a Function. The result (10) that we have just obtained rests on rather shaky ground, because we have not yet given any definition of the limit of a function. The defini­

. tion is not very different from that of the limit of a sequence in § 2.3. In analogy to (4), for example, we would expect a functionf(x) to be near the limit, say a, and so we 'vould require

(11) jf(x)-aj < e, i.e., -e < f(x)-a <e.

DEFINITION. A function f(x) is said to have the limit a as x tends to 0

provided that to any given positive number e, there corresponds a positive number o such that the inequalities ( 11) hold for all values of x in the domain of f(x) scttisfying 0 < jxj < o.

Stated informally, this says that f(x) lies as close to a as we please, say within e of a, provided x is close to 0, specifically within o of 0. For any positive E, however small, a positive number S must exist for which the conditions are satisfied. And when the conditions are satisfied we write

limf(x) = a. x->0

Another way of writing this is

f(x) -+a as x -+ 0;

in words,j(x) tends to a as x tends to 0. Figure 2.3 provides a pictorial representation of the definition.

Page 48: Calculus an Introductory Approach

40 CALCULUS: AN INTRODUCTORY APPROACH CHAP. 2

Height a+E

Height a-E

y-axis

FIG. 2.3

y= f(x)

x-axis

Given any horizontal band of width 2e, bounded by x ~ a+ E and x ~ a- e, there exists a vertical band of width 23 (with the y-axis in the center of the band) so that the graph of the function f(x) falls entirely inside the 2e band for values of x in the 23 band (except x ~ 0). One reason for making an exception of the x ~ 0 value can be seen from the example of § 2.7, Iim(sin x)fx.

This definition has been confined to the case in which the variable x tends to zero because this is the form in which limits arise in dif­ferential calculus. In the next section we give the definition of the limit of a function as x tends to any numerical value.

Although Figure 2.3 suggests that the domain of f(x) is presumed to include all values of x near zero on both sides of zero, and although this will in general be the case for the functions we deal with, nevertheless we also want to admit functions like j(x) ~ y~; whose domains include values of x on one side of zero only. So in the case of the function v; we confine atterl.tion to positive values of x.

The definition does not require that (ll) hold in case x ~ 0. This omission of x ~ 0 is not to be interpreted as meaning that (ll) is false when x ~ 0. It means that what happens to f(x) at x ~ 0 is irrelevant; (ll) may or may not be satisfied when x ~ 0; indeedf(x) need not be defined at x ~ 0. The latter situation will be the most common occur­rence in our use of the limit of a function; that is to say, in most cases the domain of f(x) will not include x ~ 0. )

Let us take another look at the result (10) in § 2.7 in the light of the above definition. By rewriting (9) with the inequalities reversed

Page 49: Calculus an Introductory Approach

SEC. 8 LIMITS 41

(a < b < c instead of c > b > a) and with 1 subtracted from each expression, we have

(12) cosB-1 sinB

< ---1 B

1 <---1.

cosB

But since we are considering values of B near 0 but not equal to 0 we know that cos B is near 1, and so we can presume that cos B > 1/2. Applying Theorem 6 of§ 2.1 we get

1 --< 2, cos8

and so

1 1-cosB 1

cosB 1 = -----;,--- = --(l-cosB) < 2(1-cosB).

cos B cosB

Hence using Theorem 1 of § 2.1 we can rewrite (12) as

sinB (13) -(1-cosB) < ---1 < 2(1-cosB).

B

This form is beginning to fit the pattern of inequalities (ll) where we now have B in place of x, (sin B)/ B in place of f(x), and 1 in place of a. The question now is, can 1-cos B and 2(1- cos B) be made smaller than any given • if we take 8 sufficiently close to 0? This follows from the graph of cos B in the vicinity of 8 = 0, as in Figure 2.4.

There is one final point to be made. The entire analysis lead­ing up to inequalities (9) was based on positive values of B. But in the FIG. 2.4 Portion of the graph of cos 8.

definition of the limit of a function, the inequalities (ll) must be satisfied for both positive and negative values of x (or B, or whatever symbol is appropriate) whenever the domain of the function extends to both positive and negative values. This point is easily taken care of, because not only is cos 8 symmetric about B = 0 as shown in Figure 2.4, but also (sin B)/B has the same symmetry property:

sin(-B) -sinB sinB _ B _ B = -B-, analogous to cos(- 8) = cos B.

Consequently the inequalities (13) hold for both positive and negative values of B in the vicinity of 0.

Page 50: Calculus an Introductory Approach

42 CALCULUS: AN INTRODUCTORY APPROACH CHAP. 2

The definition of the limit of a function, like the definition of the limit of a sequence in § 2.3, is not used to find limits nor even very often to verify conjectured limits. Rather it is used to make the limit concept mathematically precise and to prove theorems that can be effectively used to compute various limits. We turn now to theorems of this kind.

THEOREM 18. If limf(x) = a as x tends to 0, then lim cf(x) = c · limf(x) = ea for any constant c.

PROOF. In case c = 0, then cf(x) = 0 and ea = 0, and the result is obvious.

Next we treat the case c = -1. Since the central idea in the mean­ing of limf(x) = a is concerned with [ f(x)- a[, and since

[f(x)- a[ = [- f(x)- (-a) [

it follows that lim(- f(x)) = -a. Next let c be positive. For any given positive<, we use the definition

of limit of a function with •fc now playing the To le of <. Thus f(x) is within </C of a, provided x is sufficiently close to 0, say 0 < [x[ < o1;

E E - - < f(x)- a < - for 0 < [x[ < 31.

c c

Multiplying the inequalities by c (permissible by Theorem 2 of § 2.1) we get

-< < cf(x)-ca < < for 0 < [x[ < 31,

and so the theorem holds for any positive value of c. Finally, for any negative value of c, we write c = -le!, and we

apply the previous cases to get

limcf(x) = lim[ -[c[f(x)] = -lim[[c[f(x)]

= -[c[limf(x) = -[c[a = ea.

All cases have been handled, and the theorem is proved. Now we prove that the limit of a sum, difference, or product of two

functions is the sum, difference, or product of their limits.

THEOREM 19. If limf(x) = a and lim g(x) = b as x tends to 0, then lim[f(x)+g(x)] = a+b and lim[f(x)-g(x)] =a-b.

PROOF. Let< be any given positive number. We apply the definition

Page 51: Calculus an Introductory Approach

SEC. 8 LIMITS 43

of limit to bothj(x) and g(x) with •/2 now playing the role of<. Hence there exist positive numbers 81 and 82 such that

- •/2 < j(x)- a < </2 for 0 < lxl < 8,, and

-</2 < g(x)-b < </2 for 0 < lxl < 8,.

We can add these inequalities, by Theorem 3 of§ 2.1, and so we get

-• <j(x)+g(x)-(a+b) < < for 0 < lxl < 8,

provided 8 is the smaller of 81 and oz. Hence lim(j(x)+g(x)) = a+b. In the case j(x)- g(x) we use what we have already proved along

with Theorem IS in the case c = -I, thus

lim[f(x)-g(x)] = lim[J(x)+(-g(x))J = limj(x)+lim(-g(x))

= limf(x)-limg(x) =a-b.

THEOREM 20. If limf(x) = a and lim g(x) = b as x tends to 0, then lim[J(x)g(x)] = Iimj(x) · lim g(x) = ab.

PROOF. First we prove the theorem in the special case when a = 0 and b = 0. So let there be two functions F(x) and G(x) such that lim F(x) = 0 and Iim G(x) = 0 as x tends to 0. For any given posi­tive < we apply the definition oflimit with V~ playing the role of<:

and IF(x)l < y€ for 0 < lxl < 8,,

I G(x)l < v€ for 0 < lxl < 8,.

We can multiply the inequalities by Theorem 8, and so we have

/F(x) G(x) I < < for 0 < lxl < 8,

provided 8 is the smaller of 81 and 8,. Hence lim[F(x)G(x)] = 0. Turning now to the general case as described in the theorem, we

define F(x) and G(x) as

F(x) =f(x)-a, G(x) = g(x)-b,

so that

limF(x) = lim[J(x)-a] = Iimj(x)-lima = a-a = 0,

and

lim G(x) = lim[g(x)-b] = limg(x)-Iimb = b-b = 0.

Page 52: Calculus an Introductory Approach

44 CALCULUS: AN INTRODUCTORY APPROACH CHAP. 2

Hence by the first part of the proof we can conclude that as x tends to zero,

lim[f(x)g(x)] = lim[{F(x)+a}{G(x)+b}J

= !im[F(x)G(x)+aG(x)+bF(x)+ab]

= !imF(x)G(x) + !imaG(x) +limbF(x) +limab

= O+a ·lim G(x)+b ·limF(x)+ab

= 0+0+0+ab =ab.

THEOREM 21. If !im g(x) = a as x tends to 0, and if a i" 0, then lim 1Jg(x) = 1Ja.

PROOF. First we establish this in case a = l. Consider a function F(x) such that lim F(x) = 1 as x tends to 0. We apply the definition of limit with •/2 in place of E, so that

(14) IF(x)-11 < •/2 for 0 < lxl < 3.

Since ( 14) is a technical way of stating that F(x) is close to 1 when x is close to 0, we may presume that F(x) >. 1/2 for the values of x under discussion. By Theorem 6 of§ 2.1, this implies that {F(x)}-1 < 2, and by Theorem 4 of § 2.1 we can multiply this into (14) to get

{F(x1}-1 · IF(x)-11 < •, 11- -1-l < E, for 0 < lxl < 3.

F(x),

Hence we conclude that lim 1/F(x) = l. To prove Theorem 21 in the general case, we note that lim g(x)Ja = 1

by Theorem 18. Hence we can apply the result of the preceding paragraph, with F(x) replaced by g(x)Ja, and we conclude that

/ g(x) )-1 a lim\ -a- = 1, lim g(x) = l.

By Theorem 18 we can multiply this by 1/a to get lim 1/g(x) = 1ja, and the proof of Theorem 21 is complete.

l. Evaluate

2. Evaluate

Problems

sin28 lim--. 8->0 28

sin28 lim--. e->o (}

Page 53: Calculus an Introductory Approach

SEC. 8

Suggestion: Write

LIMITS

sin28 sin28 --~2--

8 28 ,

or use the identity sin 28 = 2 sin e cos e. 3. Evaluate

sin38 lim-- and

sin38 lim--.

{HQ 3(;1

4. Evaluate

5. Evaluate

Suggestion: Multiply by

6. Evaluate

7. Prove that

from the definition of limit. 8. Prove that

0->0 ()

sinx/2 lim---. x-;.O X

1-cosx lim--­x->0 X

l +cosx

l+cosx

tanv lim--. v->0 v

lim x ~ 0 x~o

lim xz = 0. x~o

Suggestion: Use Theorem 20 and the result of the preceding problem. 9. Evaluate the following limits:

x2-x sinZx (a) lim-- (d) lim--

x~o X x->0 5x2

sin2x x2-2x (b) lim--- (e) lim---

x->0 X COSX x~o X

x3+4x 7x2 +4 sinx (c) lim--- (f) lim

x->D 5x x~o 3x

45

*!0. Given that as x tends to 0, lim F(x) ~ a and lim G(x) ~ b # 0, prove that lim F(x)JG(x) ~ ajb.

Page 54: Calculus an Introductory Approach

!I 1: i: I I

46 CALCULUS: AN INTRODUCTORY APPROACH CHAP. 2

2. 9 Continuous Functions. The definition of limit of a function in the preceding section was restricted to the case "x tends to 0". This admits the rather obvious generalization:

DEFINITION. A function f(x) is said to have the limit a as x tends to b, written

limf(x) = a x->b

if for any given positive E there eXiStS a COrresponding positive a SUCh that the inequality

lf(x)-aj < e, i.e., -e < f(x)-a < e

holds for all X in the domain of f(x) Satisfying 0 < jx-bl < a. In short, f(x) stays within e of a as long as xis within a of b. Now we

can define the notion of a continuous function.

DEFINITION. A function f(x) is continuous at x = b if as x tends to b, f(x) tends to f(b), i.e.,

!imf(x) = f(b). X->b

A junction is continuous over an interval if it is continuous at each point in the interval.

What this says is that the function f(x) has a value at x = b which is in harmony with the trend of values of the function as x approaches b. Another informal way of thinking of a continuous functionf(x) is this: that a small change in x produces only a small change in f(x). The definition gives precision to. such vague statements. One can think intuitively of a continuous function as one whose graph has no jumps and no breaks. But such a rough formulation, whatever its descriptive value, has no mathematical value because its very imprecision prevents any careful analysis. "No jumps, no breaks"~no theorems!

In order that a function f(x) be continuous at x = b, it is neces­sary that the domain of definition of the function include the value x = b; in other words it is necessary that f(b) exist. However, it is not necessary that the function be defined both for values of x on both sides of b: one side will do. For example, the function y = -vlx with domain x ;;; 0 is continuous at x = 0 even though the function is not defined for negative values of x. We write

lim -ylx = o, x->0

with the understanding that the limit is taken with positive values of x only.

Page 55: Calculus an Introductory Approach

. SEC. 9 LIMITS 47

Problems

l. Prove that the function j(x) = xis continuous at every real number. 2. Given that j(x) and g(x) are continuous functions at x = 0, prove that

j(x)+g(x), j(x)-g(x), cf(x), andj(x)g(x) are continuous functions at x = 0. Suggestion: Use the theorems of § 2.8.

Page 56: Calculus an Introductory Approach

CHAPTER 3

INTEGRATION

3.0. We have now developed enough of the theory of limits to con­tinue the study of integral calculus, begun in § 1.6. Our program is to define the integral and to evaluate some specific examples. The defi­nition is not logically completed in this chapter; in order to avoid too much theory too soon, we postpone the question of the existence of the integral in general to Chapter 5 and Appendix A. Our present pur­pose is to discuss the idea of an integral; existenc.e must wait!

In § 3. 6 we shall use the trigonometric identity

2 sin a sin ,8 = cos( a- ,8)- cos( a+ ,8),

which can be obtained by subtracting the well-known identities

cos(a-,8) =cos a cos,B+sina sin,B, and

cos(a+,B) =cos a cos,B-sina sin ,B.

3.1 The Limit of a Sum. Having studied the limit process in some measure of detail, we return to our examination of the concept of an integral. In Chapter l we computed the area A between the curve y = x2 and the x-axis, from x = 0 to x = 1. This was done by showing geometrically that A lies between the sums of the areas of two sets of rectangles. Thus we had the inequalities in formula (4) of§ 1.6,

n (j-!)2 n j2 L <A< L -. J=I n3 i=l n3

The sum on the left can be written as

ni' j2, i=l n3

and thus the two sums differ by a single term n2jn3 or !jn. But lim !Jn = 0 as n tends to infinity, so that the above inequalities show

48

Page 57: Calculus an Introductory Approach

SEC. I INTEGRATION 49

that A is caught between two values whose difference is tending to zero. Thus we can obtain A by using the limit of either sum,

n {j-1)2 n j2 A = lim L -- or A = lim L -.

n-.oo J=l n3 n->co j=l n3

The fact that the difference between these two sums tends to zero tells us that if one limit exists so does the other, by Corollary 3 of § 2.4. To the second of the above limits we apply formula (3) of § 1.5 (and also Theorem 12 of § 2.4 and Corollary 10 of§ 2.3). Thus we get

n j2 1 n A = lim L- = lim- L;P

n->ro J=l n3 n3 1=1

I I I = lim- +lim- +lim-

3 2n 6n2

I I =- +0+0 = -.

3 3

The notation used in calculus for the limit of the sum just calculated is

J>2 dx, "the integral of x2 from x = 0 to x = I". In general the notation for the integral of x2 from x = a to x = b, where a < b, is

J:X2dx, and it represents the following limit of a sum,

(I) n ( b-a)2b-a

lim L a+j-- --. n-)oo J=l n n

We next explain how this formula (1) arises. Geometrically, this limit can be interpreted as the area bounded

above by y = x2, below by the x-axis, and at the sides by the vertical lines x = a and x = b (Figure 3.1). The portion of the x-axis from x =a to x = b is divided into n parts each of length (b-a)Jn, thus giving us n +I points

(xo, 0), (x,, 0), (x2, 0), ... , (xu, 0),

Page 58: Calculus an Introductory Approach

50 CALCULUS: AN INTRODUCTORY APPROACH CHAP. 3

where Xo is just another way of writing a, Xn another way of writing b, and in general

(2) b-a

x; = a+j--. n

V erticallines (called ordinates) are constructed from these n + I points up to the curve y = x', as illustrated in Figure 3.2 for (x;-1, 0) and

y-axis

x -axis ----~~~*-~~

(b, 0)

(x2 ,0)

Fw. 3.1

y-axis

FIG. 3.2

(x;, 0). The rectangle in Figure 3.2 has base (b-a)jn and height x;2,

and so has area

b-a ~ .b-a)'b-a x;2 • -- or a+y-- --

n n n

by use of (2). Thus the sum in (I) represents the total area of the rectangles illustrated in Figure 3.3. For purposes of illustration we have used n = 10, and it is intuitively clear that as n tends to infinity the sum of the areas of the rectangles tends to the area under the curve.

We note that (I) could be written, in view of (2),

f• • b-a • x' dx = lim .Z: x;' · -- = lim .Z: xJ'(x;- x;-1).

a n-•co 1=1 n n--.co 1=1

Page 59: Calculus an Introductory Approach

SEC. 1 INTEGRATION 51

In general, if we replace x2 by a function f(x) we can write

(3) Ib n b-a n f(x) dx = lim L f(x;)~ = lim L (x;-x;-l)f(xJ).

a n-K.o j=l n n->co j=l

The first expression is read "the integral of f(x) from a to b". The geometric interpretation of this is analogous to Figure 3.3 with y = x2

replaced by the curve y = f(x). Thus the integral (3) can be inter­preted geometrically as an area, as in Figure 3.4, bounded by y = j(x) above, the x-axis below, the line x = a on the left, and the line x = b on the right.

y=f(x)

X=a

FIG. 3.3 FIG. 3.4

Although an integral is understood most readily as an area, it is not so defined. By defining an integral more formally as a limit of a certain kind of sum, we will have a more general concept than area. The idea of a limit of a sum, or an integral, is adaptable in many places in mathe­matics. Thus area is merely one interpretation of an integral.

Problems

I. Express the integral J~ x dx as a limit of a sum, and evaluate this limit by use of equation (2) of § 1.5.

2. Verify the answer to the preceding question by noting that the integral represents tbe area of the triangle with vertices (0, 0), (1, 0) and (1, 1).

3. Express the integral J~ x dx as a limit of a sum, and evaluate this limit. 4. Verify the answer to the preceding question by observing that the

Page 60: Calculus an Introductory Approach

52 CALCULUS: AN INTRODUCTORY APPROACH CHAP. 3

integral represents the area of the triangle with vertices (0, 0), (2, 0) and (2, 2).

5. Express the integral J~ 1 dx, usually written more simply as J~ dx, as a limit of a sum. Evaluate this limit, and then verify the answer by noting that the integral represents the area of the rectangle bounded by (0, 0), (0, 1), (1, 1), and (1, 0).

6. Express the integral J~ 3x dx as a limit of a sum, and evaluate the limit. Verify the answer by calculating the area of the appropriate triangle.

7. Express the integral J~ 3x2 dx as a limit of a sum, and evaluate it.

3.2 The General Definition of an Integral. In equation (3) the base line from x = a to x = b was divided into n equal parts, each of length (b- a)Jn. While this is correct as far as it goes, it is not suffi­ciently general. It turns out, as we shall see, that the value of the in­tegral is not changed if we divide the x-axis from x = a to x = b into n parts which are not required to be equal. Thus the value x0 is still identified with a, and Xn with b, but the intermediate points x 1, x2 , ... , Xn-1 are spaced between a and b without the subintervals being necessarily of equal length. Thus the leqgths x1 - x0 ,

a

xo Xg

b

I

Xn-1 Xn

x 2 -x1, x3-x2, ... , Xn-Xn-1 may be all different. However, a condition must be imposed, namely, that the largest of these lengths must tend to zero as n tends to infinity.

Furthermore, there is no need, as in equation (3), to use j(x1) in the sum: We can replace it by j(!;;), where /;; is any value between XJ-!

and x1, thus XJ-1 ;;; fJ ;;; XJ. So the general definition of an integral is given by

(4)

where it is understood that the largest one of the numbers x1 - xo, x 2 -x1, ... , Xn-Xn-1 tends to zero as n tends to infinity. This is illu~ strated in the accompanying Figure 3.5.

This definition of an integral implies at once that

(5) J>(x) dx = J:f(x) dx+ S: j(x) dx, if a < c < b.

Page 61: Calculus an Introductory Approach

S1lc. 3 INTEGRATION 53 ~~~~~----------------=

FIG. 3.5

Equation (5) is illustrated in Figure 3.6; the total area is equal to the sum of the areas of the two parts into which it is divided by the line PQ.

From an analytic point of view, equation (5) can be seen from the definition {4) in this way: sup-pose that the interval from x = a to x = b on the x-axis is divided into n parts, with the point x = c always one of the division points. Then the sum in ( 4) can be separ­ated into two parts around the division point x = c, and these separate parts of the sum give the two integrals on the right side of equation (5).

x=a

X=C

p x-axis

FIG. 3.6

Before discussing further properties of the integral, let us consider whether the definition {4) really defines something. For after all, the integral is defined as a limit, and as we have seen, limits do not neces­sarily exist. Moreover, the limit in equation ( 4) is not as simple as those discussed in Chapter 2; the limit is now to be taken over many possible points of division xr, x2, ... , Xn-I, and many possible inter­mediate points {',, 6, ... , /;n. Thus the definition (4) is not a simple limit of a single sequence. It is the limit of a whole class of sequences, and hence the definition (4) makes sense only if the limit exists and has the same value for all possible choices of the sets of division points xr, x2, ... , Xn-1 and all possible intermediate points {',, /;2, ... , tn.

Thus we see that there is much more involved in the limit {4) than in the limits discussed in Chapter 2. The limit (4) does not in fact exist for all conceivable functions f(x), but it does exist for the fairly simple types of functions with which we deal, such as x, x2, x3, sin x, and cos x.

Page 62: Calculus an Introductory Approach

54 CALCULUS: AN INTRODUCTORY APPROACH CHAP. 3

Proving the existence of the integral is one of the most basic parts of the theory underlying calculus. We postpone this proof in order to develop first some of the needed background, and to permit the reader to gain more familiarity with the notation. In Chapter 5 we prove the existence of the integral in a restricted setting which covers virtually all cases in this book. Then in Appendix A we remove one of the re­strictions to get a more general theory covering all the situations that we treat. These proofs will not rely on any intuitive ideas about geometry.

In the present chapter we shall presume that the integrand f(x) in the integral s: f(x) dx is an integrable function, meaning that it is a function for which the limit (4) exists. Since the limit given in equation (3) is just a special case of (4), we are presuming a fortiori that (3) exists.

In all developments of the theory we shall use (4) as the definition of the integral. But in evaluating integrals, it is usually easier to employ the special case (3). The full justification for this procedure will be given in Chapter 5 and Appendix A.

The discussion of equation (4), and the illustration of its meaning in Figure 3.5, have suggested that the number a is always to be taken less than the number b, that is a < b. There is no necessity for this restriction in the formulation of an integral. The numbers a and b in s: f(x) dx, called the lower and upper limits of the integral, can satisfy any one of the three possibilities a > b, a < b, or a ~ b. (Notice that the word "limit" is being used in a different way here; a and b are not limits of the type discussed in Chapter 2.) We now extend the definition of an integral to cover all these cases.

The extended definitions are suggested by either (4) or (5). If for example we want (5) to hold for all possible sets of values a, b, c, then the replacement of c by a suggests the definition

J: f(x) dx ~ 0.

This would also be suggested by the interpretation of this integral as an area between x ~ a and x = a. Next, if we replace b by a in (5) we get

I: f(x) dx =I: f(x) dx+ f f(x) dx,

which, in view of the preceding equation, suggests that the integral from a to c is the negative of the integral from c to a. Hence we assert

1

Page 63: Calculus an Introductory Approach

SEC. 2 INTEGRATION 55

the following equation as a definition, reverting to the symbols a and b as the limits of integration,

(6) J: j(x) dx = - J: j(x) dx.

With these definitions it turns out that (5) holds whatever the re­lation between a, b, and c, provided of course that f(x) is an integrable function. We give no proof of this, but the reader is asked to prove it in a few special cases in the problem set of this section.

The case c = 0 of equation ( 5) is especially useful, namely

J: j(x) dx = J: f(x) dx + J: j(x) dx.

Applying (6) to the second of these three integrals we get

(7) J: f(x) dx = J: f(x) dx- J: j(x) dx.

Finally, we point out that there is no requirement in the definition (4) of an integral that the integrand f(x) be positive. In case f(x) is negative for some values of x, the sum of terms (x1 -XJ-r)f(~J) is ob­tained nevertheless, and then the limit is taken as n tends to infinity. This more general meaning of an integral has some marked conse­quences for the interpretation as an area; we shall discuss this matter in detail in § 3.4.

Problems

l. Taking equation (5) for granted for all numbers a, b, c such that a ;:; c ;:; b, and presuming the definition (6), establish the following results:

.(a) J>x) dx =I: f(x) dx- J>x) dx;

(b) J>x) dx = J>(x) dx- J>(x) dx;

(c) J>(x) dx+ J>x) dx- J>(x) dx = 0;

(d) J>(x) dx+ J>(x) dx+ J>x) dx+ J:f(x) dx = 0;

Page 64: Calculus an Introductory Approach

56 CALCULUS: AN INTRODUCTORY APPROACH CHAP. 3

2. Combine the sum

57 J'" /(x) dx+ 7

f(x) dx

into a single integral. 3. Combine the collection

into a single integral.

3.3 The Evaluation of Certain Integrals. Although the general definition ( 4) of an integral as a limit of a sum does not require equal divisions of the base line as in the earlier, less general, form (3) of the definition, nevertheless we shall use (3) for our calculations. For example, to evaluate the integral J: x 2 dx, we shall use the form ( l) because it is the special case of (3) for the particular functionf(x) = x2. We begin with the case where a = 0, thus

fb n (jb )2b n j2b3 x2dx=lim 2- -=lim 2-.

o n--;ooo 1=1 n n n-)oo J=l n3

Since the sum is over values of j, we can factor out b3fn3 and then use formula (3) of § 1.5, thus

fb [ b3 n ] x2dx = lim -2 j2 o n3 J=l

= lim[~: (:3 + :2 +~)] =lim[ba(~+2_+-1 )]·

3 2n 6n2

Next we apply Theorems ll and 12 of § 2.4 and Corollary lO of § 2.3 to get

{8) fb x2 dx = b3(Jim(~ + 2_ + _l )) o 3 2n 6n2

= b3(lim ~ + lim 2_ + Jim _l_ ) 3 2n 6n2

THEOREM 1. For any numbers a and b,

fb b3 a3 ax2dx = 3 - 3 .

T

Page 65: Calculus an Introductory Approach

SEC. 3 INTEGRATION 57

PROOF. We use (8) and (7) to get

fb x2 dx ~fb x2 dx-Ja x2 dx ~ b3

- a3

a 0 o 3 3

THEOREM 2. For any numbers a and b,

fb fb b' a' dx ~ b-a and xdx ~ ---. 9 9 a a ... ""'

PROOF. As in equations (8) preceding Theorem l, we begin with a ~ 0 and prove that

(9) fb dx ~ b and fb x dx ~ ~2

• 0 0 2

These results can be obtained by the use of the definition (3), with j(x) replaced by l and x respectively. First, if f(x) ~ l we have

fb n b (b b b) dx ~ lim L - · l ~ lim - + - + ... + - ~ lim{b) ~ b. o n-->oo j=l n n n n

In the second case withj(x) ~ x, we use (2) to observe that

and so

bj XJ = -,

n

bj j(xJ) ~ XJ ~ -,

n

fb •bbj _b2j b2

xdx= lim L -·- ~ lim L- = lim- L;j. o n--)oo j=l n n n2 n2

Using formula (2) of § 1.5 we get

fb b' n(n+l) b' ( l) x dx = lim- · ~ lim- l +-

o n2 2 2 n = b

2lim(l + ~) = b'.

2 n 2

Hence we have established the results (9), and to finish the proof of the theorem, we use (7) again. So we get

J>x ~ J>x- J: dx = b~a, and

fb fb fa b' a' xdx= xdx- xdx=---. a 0 0 2 2

Page 66: Calculus an Introductory Approach

58 CALCULUS: AN INTRODUCTORY APPROACH CHAP. 3

Equations (9) can be interpreted quite simply in terms of areas. The integral J: dx denotes the area of the rectangle in Figure 3. 7, and J: x dx the area of the triangle in Figure 3.8.

y-axis y-axis (b, b)

(0, 1) (b, 1)

(0, 0) x-axis x-ax1s (b, 0) (0, 0) (b, 0)

FIG. 3.7 j(x) ~ I. FIG. 3.8 j(x) ~ x.

Problems l. Evaluate the integrals

s: x' dx, r x' dx, [ x' dx .

by use of Theorem 1. 2. Evaluate the following integrals:

(a) J:Xax

(b) J: xdx

(c) J: dx

(d) J: x2 dx

*3. Evaluate the integral g x3 dx by expressing it as a limit of a sum, and also by using the answer to Problem 12 of § 1.5.

*4. Evaluate the integrals

J: x3 dx and J: x3 dx.

5. Use Theorem 1 to write algebraic expressions for the values of

J: x2 dx, J: x' dx, J: x' dx.

Check that these algebraic expressions satisfy equation (5) of § 3.2 with j(x) replaced by x'.

3.4. The Interpretation of an Integral as an Area. The inter­pretation of J:f(x) dx as an area, as illustrated in Figure 3.4, needs

T

Page 67: Calculus an Introductory Approach

SEC. 4 INTEGRATION 59

some reviSIOn in the face of the possibilities that a may not be less than Q, andj(x) may not be positive. Now in case a = b, the integral has value zero, and this harmonizes with the idea of an area bounded by x = a and x = b. Furthermore, in view of formula (6), there is no need to consider any cases other than a < b, because (6) enables us to write any integral with the lower and upper limits satisfying this inequality.

So we presume that every integral s:J(x) dx under discussion has a < b. It follows that (b -a)Jn or x;-XJ-1 in (3) is positive, and con­sequently the value of the integral will have the same sign, positive or negative, as j(x). But this is on the assumption that the integrand j(x) has the same sign over the range of integration from x = a to X= b.

Let us turn to the situation where j(x) is positive for part of the range of integration, and negative for another part. In such a case the value of the integral will be positive or negative depending on the dominance of the positive or negative portions of j(x). For example, by Theorem 2 we have

53

x dx = ~ - ~ = 4, -1 2 2

-4,

and these integrals correspond to Figures 3.9 and 3.10. The individual areas in Figure 3. 9 correspond to the integrals

fo xdx= 1

-1 -2,

and in Figure 3.10

fo xdx = 9

- 2' -3

y -axis

x -axis

FIG. 3.9

J"xdx = ~, 0 2

r 1 x dx = -. 0 2

y-axis

1 ~Area2

x-axis

FIG. 3.10

Page 68: Calculus an Introductory Approach

60 CALCULUS: AN INTRODUCTORY APPROACH CHAP. 3

Thus the integral of f(x) from a to b, with a < b, has a positive or negative value depending on whether the areas above the x-axis dominate or those below.

x=a[+\ ~ x-axis

~Gx=b FIG. 3.11

Problems

l. Given that

J1 1

x4dx = -, 0 5 J

2 21 J2 31 x5dx = -, x4dx = -,

1 2 1 5

evaluate the following by area interpretations:

(a) [ x4dx (e) s-1 x5 dx -2

(b) J1

x' dx (f) fz x4dx -1 (c) J: x• dx (g) f

2 x'dx

s-2

x4 dx r (d) (h) 0

sinxdx

-1 2. For any odd positive integer n, evaluate

3. Show that

4. Show that

Jl xn dx. -1

f" f"/2 0

sinxdx = 2 0

sinxdx.

f"/2 J"/2 0

sinxdx = 0

cosxdx.

3.5 Further Evaluation of Integrals. In § 3.3 we evaluated some integrals by means of the equation (3). This is the hard way to

T

Page 69: Calculus an Introductory Approach

SEC. 5 INTEGRATION 61

do the calculation, as we shall see in Chapter 5 where the fundamental theorem of calculus is obtained. But even within the present restricted theory we can derive some techniques for the evaluation of an integral.

THEOREM 3. If j(x) and g(x) are integrable junctions from x = a to x=bthen

J: [j(x) ± g(x)] dx = J>x) dx ± J:g(x) dx,

where the plus signs go together, and the minus signs go together. Further­more, for any constant c

J: cj(x) dx = c J: j(x) dx.

PROOF. By (4) we have

J: [j(x) ± g(x)] dx = ~i~ 1~ (x; -X;-l)[f(t;) ± g(t;)].

But by simple algebra we can write

(x;-XJ-l)[f(t;) ±g(t;)] = (x;-X;-t}f{gj) ± (Xj-Xj-l)g(t;), n n n L (x; -x;-l)[f(t;) ± g(t;)] = L (x; -x;-l)](t;) ± L (x; -xH)g(tJ).

1=1 j=! j=l

Taking limits as n tends to infinity, and using Theorem 12 of Chapter 2, we get the first part of the theorem.

Turning to the second part of the theorem, we see that

J: cj(x)dx = H~ 1~ (x;-x;-l)cf(t;).

But c, a constant factor, can be factored out of the sum to give

fb n

cj(x) dx = lim c L (x;-x;-l)](t;). a n~co j=l

Next we apply Theorem ll of Chapter 2, and the proof of the theorem is complete.

ExAMPLE. Evaluate

J: (x2+x) dx.

Page 70: Calculus an Introductory Approach

62 CALCULUS: AN INTRODUCTORY APPROACH CHAP. 3

Applying Theorems 3, 1 and 2, we get

J5 J5 J5 53 23 52 22 99

(x2+x)dx= x2dx+ xdx=---+---=-. 2 2 2 33222

1. Evaluate the integrals

(a) J: (xL 1) dx

(b) J: (xL2x)dx

Problems

(c) J: (5x'+Gx) dx

(d) J" (x2-x-1)dx -3

2. Using the data given in Problem 1 of § 3.4, evaluate the integrals

(a) J: 5x' dx (c) f (x'-x') dx

(b) J: 10x'dx (d) I: (2x'-5x4)dx

3.6 The Integral of sin x. We will need the identity

. . . . cos(x/2)- cos{[(2n + 1)x ]/2} (10) smx+sm2x+sm3x+ ... +smnx = ,

2 sin(x/2)

which holds for all values of x for which sin(x/2) is not zero. Equation (10) can be established in the following way. Taking the well-known trigonometric identity

2sino:sin;'l = cos(a.-;'l)-cos(o:+;'l),

we apply it n times to get X X 3x

2 sinx sin-= cos- -cos-2 2 2 x 3x 5x

2 sin2xsin- =cos- -cos-2 2 2

x 5x ?x 2 sin3xsin- =cos- -cos-

2 2 2

x (2n-3)x (2n-1)x 2 sin(n-1)x sin-= cos cos-'------'--

2 2 2 . . x (2n-1)x (2n+ 1)x

2 s1nnx s1n- =cos -cos . 2 2 2

T I

Page 71: Calculus an Introductory Approach

SEc. 6 INTEGRATION

Adding these equations we have

X 2 sin- [sinx+sin2x+sin3x+ ... +sinnx] =

2

Dividing by 2 sin(x/2), we get ( !0). Now we use this result to evaluate

J: sinxdx

63

x (2n+ l)x cos- -cos

2 2

for any positive b. From definition (3) it follows that this integral has the meaning

n [ jb] b b jb lim 2 sin- ·- = Jim- · 2 sin-. n~)co J=l n n n n

The last sum can be evaluated by means of ( !0) with x replaced by bjn, thus

I sin jb = cos(bj2n)- cos{[(2n + l )b ]j2n).

1-1 n 2sm(bj2n)

Hence we have

fb . . b cos(b/2n)-cos{[(2n+l)b]J2n) sm x dx = hm - · -~_.__.__~_._ __ :__:::___o

0 n-•oo n 2 sin(bj2n)

=lim[cos}'__-cos(Zn+l)b]· (bjZn). 2n 2n sin(b j2n)

Now as n tends to eo, bj2n tends to 0, and so by equation (!0) of§ 2.7,

(bj2n) e lim ~-c-c-- = Jim -- = l. n~oo sin(bj2n) 8-•0 sin&

Furthermore as n tends to oo, we observe that

and

b limcos- = cosO = 1,

2n

Jimcos (Zn;l)b = limcos[b(l+ 2~)] = cosb.

Combining these results, we obtain

(11) J: sinxdx = 1-cosb.

Page 72: Calculus an Introductory Approach

64 CALCULUS: AN INTRODUCTORY APPROACH CHAP. 3

THEOREM 4. For any numbers a and b,

J: sinxdx = cosa-cosb.

PROOF. The result follows at once from (ll) and (7). The integral in Theorem 4 is the last that we will evaluate from the

definition. As we go from the simpler functions, like x, x2, and sin x, in the integrand to the more complex functions, the evaluation of the limit of the sum by the methods of this chapter becomes more and more difficult, and certainly impractical even if not impossible. For example, although we have just integrated sin x, there is no comparable tech­nique for evaluating*

J: tanxdx,

because there is no simple formula for

tanx+tan2x+tan3x+ ... +tannx

such as we had for the sine series in formula (10). In Chapter 5 we shall discuss the fundamental theorem of calculus, and it will enable us to bypass such lengthy evaluations as in the present section by getting at the integration process through the back door, differentiation.

Problems 1. Evaluate the integrals

fn/6

(a) 0

sinxdx (c) J: sinxdx

Jn/2

(b) 0

sinxdx Jn/3

(d) 4 sinxdx n/6

2. Evaluate

Jn/2

0

cosxdx.

Suggestion: See Problem 4 of§ 3.4. 3. Evaluate the integrals

(a) Jn/2

0 (sinx+cosx) dx

(b) fn/4

0 sinxdx

r, (c) cosxdx n/4

(d) J: (x+sinx) dx

* The evaluation of this integral is Problem 11 in § 7 .3.

Page 73: Calculus an Introductory Approach

SEC. 7 INTEGRATION 65

3.7 The Volume of a Sphere. The integral is defined as a limit of a sum, not as an area. The reason for this is that the concept of a limit of a sum has much wider ramifications than area. Although many of the applications of the integral concept are beyond the scope of this book, we shall treat some results other than areas bounded by curves. In this section we derive the formula for the volume of a sphere, assuming that we are given the formula for the volume of a circular cylinder.

Let us take, as in Figure 3.12, a semicircle of radius rand center at (0, 0). If (x, y) is any point on the semicircle, then the right triangle formed by the points (0, 0), (x, 0) and (x, y) has sides of length jxj,

y-axis (0, r)

(x,y)

( -r, 0) (0, 0)

FIG. 3.13

y, and r. By Pythagoras' theorem it follows that x' +y' = r', and this is the equation of the circle. Solving this equation for y we get

y = yr2 -x2 and y = -yr2 -x2.

Our interest is in the first of these formulas, because it is the equation of the semicircle shown in Figure 3.12. The second equation pertains to the other half of the circle below the x-axis, not shown in Figure 3.12. If the semi-circle y = Vr2 -x2 is revolved about the x-aJ!:iS, it sweeps out a sphere in 3-dimensional space, and this is the sphere whose volume we will compute. In actuality we will confine attention to the quarter circle from the point (0, r) on the y-axis down to the point (r, 0) on the x-axis, and so when this is revolved about the x-axis we get exactly half a sphere.

Let us divide the portion of the x-axis from (0, 0) to (r, 0) into n equal parts, giving us the points (0, 0), (x1, 0), (x2, 0), ... , (xn, 0) where Xn is the same as r. Thus Xj = (jr)fn for j = 1, 2, ... , n. On each subdivision we construct a rectangle as in Figure 3.13.

Since Xj -Xj-1 = rfn, the rectangle has base rfn and height Vr' -xi'·

The volume of a half-sphere will be obtained by conceiving of the whole

Page 74: Calculus an Introductory Approach

66 CALCULUS: AN INTRODUCTORY APPROACH CHAP. 3

graph of Figure 3.13 as being revolved in space about the x-axis, so that the quarter circle generates a half sphere of radius r. The rec­tangle in Figure 3.13 generates a cylinder of radius V r2 - x;2 and height rjn, as illustrated in Figure 3.14. The volume of this cylinder is

FIG. 3.14

1r( ylr'- x;')'rfn = ,(,·2- x1')rjn,

and if we sum these volumes from Xt to Xn and then take the limit as n tends to infinity, we will obtain the volume of the half sphere. Thus if V denotes the volume of a sphere of radius r, we have

n V = 2 lim L 1r(r2 - XJ2 )rjn.

n-'>oo J=l

This limit can be compared with the right side of equation (3), where j(x1) is interpreted as ,(,-'-xJ'), and (b-a)jn has the value rjn if we take b = r and a = 0. Hence we can write

V = 2 J: 1r(r'- x') dx,

and this can be evaluated by use of Theorems 3, 1, and 2, to give

V= 2 J: m 2 dx-2J "x'dx

2"r' J: dx- 27T J: x' dx

,.s 4 2"r'(r)- 27T- = -m·3.

3 3

Problems

I. If in the above analysis we had worked with the entire portion of the x-axis from ( -r, 0) to (r, 0) instead of just half of this, we would have arrived at the equation

V= J~,. 7T(r'-x') dx.

Verify that this integral leads to the same answer, V = (4/3)-rrrs. *2. Consider the two parts into which a sphere of radius r is divided by a

plane which passes at a distance k from the center of the sphere. Assume of course that 0 < le < ,.. Find the volume of the smaller of these two parts.

*3. By revolving the triangle formed by (0, 0), (h, 0) and (h, r) around the x~axis, derive the formula for the volume of a right circular cone of radius r and height h. Suggestion: The equation of the straight line through (0, 0) and (h, ?') is y = (rjh)x.

Page 75: Calculus an Introductory Approach

CHAPTER 4

DIFFERENTIATION

4.0. Two problems of differential calculus, or differentiation, were introduced in §§ 1.2 and 1.3. In this chapter we treat this topic in greater detail, starting with the definition of a derivative, going on to techniques of differentiation, and concluding with applications to problems of maxima and minima.

Let us review a few identities from trigonometry that we shall use. The identity sin2x + cos2x = l, which is Pythagoras' theorem written in trigonometric form, is sufficiently well-known to need no explanation. If this is divided throughout by cos2x and then revised slightly, there results another basic identity, sec2x = l +tan2x. In § 4.4 we use the identity

A-B A+B sinA-sinB = 2 sin-- cos--.

2 2

In case the reader is not familiar with this identity, we note that it can be o_btained from the well-known identities

and sin(oo+f3) =sin"' cosf3+sinf3 cosx,

sin(oo-{3) = sinO< cosf3-sinf3 cosO<.

For, subtracting these we get

sin(Ct.+/3) -sin(oo-{3) = 2 sinf3 cos a.

If now we write A for ~- + f3 and B for a- {3, thus

A+B oo+f3 =A, oo-{3 = B, so "'= -- and

2

A-B f3 = -2-,

then we get the identity we wanted. The identity

A-B A+B cosA-cosB = -sin-- sin-~,

2 2

67

Page 76: Calculus an Introductory Approach

68 CALCULUS: AN INTRODUCTORY APPROACH CHAP. 4

which is also used in § 4.4, can be obtained in a similar way from

cos(o<+/l) = coso: cos!J-sin<X sin!J, and

cos(o.-/l) =cos <X cos!J+sino: sin!J.

4.1 The Derivative. In § 1.3 we solved the problem of finding the slope of the curve y = x2 at the point (3, 9). We now run through the

(3

+b. same solution using notation that P g+t.~) is m?re common to _calculus. As A (3 9) m F1gure 4.1, cons1der a fixed

' point A with coordinates (3,9) on the parabola y = x2, and a near­by point P with coordinates (3 + b.x, 9 + b.y). Thus b.x is the change in the x coordinate, and b.y the change in the y coordinate, measured from the point A to the point P.

The symbol b. is not a multiplier Fxo. 4.1 Graph of y = x2. or a factor when written as part of

the notation b.x or b.y. Rather the symbol b. denotes a difference, so that b.x is the difference in the x coordinates from the point A to the point P, and similarly b.y is the difference in the y coordinates.

The point P, being on the curve, has coordinates which satisfy the equation y = x2; thus

9+b.y = (3+b.x)2 = 9+Mx+(b.x)2,

b.y = Mx + (b.x)2•

The slope of the line AP is, by equation (1) of § 1.1,

9+b.y-9 b.y

3+b.x-3 b.x'

and from the preceding equations we have

b.y - = 6+b.x. b.x

The tangent line at the point A is the limiting position of the line AP as the point P moves toward A, and so the slope of the tangent line is obtained by letting b.x approach zero; thus

lim b.y = lim (6+b.x) = 6. 6.X -> 0 6_x 6.X -) 0

Page 77: Calculus an Introductory Approach

SEC. I DIFFERENTIATION 69

It should be noted that 1\.x is not necessarily a positive quantity. The limit above is 6, whether 1\.x tends to zero through positive values or through negative values. In Figure 4.1, the point P would be "down" the curve from A in case ~x were negative.

Let us generalize from the point (3, 9) to any point A with coordinates (x, y) on the curve y = x'. In this general case the point P would have coordinates (x + 1\.x, y + Ll.y). The points A and P satisfy the equation of the parabola, so we have

y = x' and y+LI.y = (x+LI.x)2 = x'+2xLI.x+(LI.x)2 •

Subtracting these two equations and dividing by 1\.x, we get

1\.y = 2xLI.x+ (LI.x)2 ,

1\.y - = 2x+LI.x 1\.x ,

1\.y lim - = lim (2x + 1\.x) = 2x. ~X->O~X L\X-70

It should be emphasized that 1\.x designates the change in the x co­ordinate from the point A to the point P, and so 1\.x is a variable in its own right, independent of x. Thus 1\.x does not mean some quantity 11. multiplied by some quantity x, but a single quantity.

We have established that the slope of the curve y = x' at any point on the curve has value 2x, and the usual notation for this in calculus is

dy - = 2x dx

Thus dyfdx, called the derivative of y with respect to x, is defined in general by the equation

(I) dy

dx

1\.y lim -.

L\x -->O 6-x

The process of obtaining dyfdx = 2x from y = x' is called differentia­tion, and this is the central process in that part of our subject known as differential calculus. In contrast, the subject matter of Chapter 3 is known as integral calculus.

ExAMPLE. Differentiate y = x3.

Page 78: Calculus an Introductory Approach

70 CALCULUS: AN INTRODUCTORY APPROACH CnAP. 4

Solution. Using analogous steps to the ones above, we write

y = x3 for the point (x, y),

y+t.y = (x+t.x)3 for the point (x+t.x, y+t.y),

y+t.y = x"+3x2t.x+3x(t.x)2+(t.x)3,

dy

dx

t.y = 3x2t.x + 3x(t.x)2 + (t.x)3,

t.y - = 3x2+3xt.x+(t.x)2, t.x

lim t.y = lim[3x"+3xt.x+(t.x)'J = 3x2. 6-X ->0 ~X

Thus the slope of the curve y = x3 at any point has value 3x2•

Problems

1. Differentiate the follO\ving

(a) y = x2 +x

(b) y = 2x2

(c) y = 2x"

(d) y = x"-x

2. Find the slope of each of the curves in Question l at the point where X= 2.

3. Find the derivative of y = x4. 4. At what point or points on the curve y = x' is the slope equal (a) to

the x coordinate 1 (b) to the y coordinate 1 5. At what point or points on the curve y = x3 is the slope equal {a) to

the x coordinate1 (b) to they coordinate1 6. Prove that no two tangent lines to the curve y = x2 are parallel. 7. Is there a tangent line to the curve y = x3 which is parallel to the

tangent line at the point (4, 64)1 8. If A and B are two distinct points on the curve y = x3, under what

circumstances is the slope of the curve at A equal to the slope of the curve at B1

9. Prove that the X·axis is a tangent line to the curve y = x3 at the point (0, 0). (This is an example of a tangent line crossing the curve at the point of tangency. Such a point is called a point of inflection.)

4.2 The Definition of the Derivative. It is desirable to have a definition of derivative in a different form than equation (!), a form completely independent of geometric concepts. This we can arrive at by treating a general equation y = J(x) by the process used on y = x2

Page 79: Calculus an Introductory Approach

SEC. 2 DIFFERENTIATION 71

and y = x3 in § 4.1. We write the equations pertaining toy = f(x) in parallel to those of y = x', for clarity:

y = x' y+D.y = (x+D.x)'

D.y = (x+b.x)2 -x2

D.y (x + D.x)'- x'

D.x D.x

dy . (x+b.x)2 -x2

-= hm dx fl.x-->o .6..x

y = f(x)

y+D.y = f(x+D.x)

D.y = f(x+ D.x)- f(x)

D.y f(x + D.x)- f(x)

D.x D.x

dy . f(x + D.x)- f(x) -= hm . dx "x~o D.x

Of these two final equations, the one on the left yields a final simplifi­cation

dy - = 2x, dx

but the one on the right cannot be simplified without specific know­ledge about the function f(x). One of the central problems of this chapter is to evaluate dyfdx for various functions f(x). The above equations enable us to reformulate the definition (I).

DEFINITION. The derivative of a function f(x) is '

(2) . f(x + D.x)- f(x)

hm , .a.x-->0 .6..x,

provided this limit exists.

This limit may exist for some values of x and not for others. So we can say that the derivative at x = c is

(3) . f(c+b.x)-f(c)

lnn , a.x -•o .6..x

provided the limit exists. If the limit (3) exists, we say that the function f(x) is differentiable at x = c. If the limit (2) exists for every real value of x, we say that the function f(x) is differentiable everywhere. If (2) exists for all x satisfying a ;£ x ;£ b, we say that f(x) is differentiable for this set of values.

In order that a function may be differentiable at x = c, the limit (3) must exist. Now as D.x tends to zero, the numerator f(c+D.x)-f(c) in (3) must also tend to zero if f(x) is differentiable at x = c. Thus the

Page 80: Calculus an Introductory Approach

72 CALCULUS: AN INTRODUCTORY APPROACH CHAP. 4

existence of the derivative at x ~ c implies that

lim f(c+L'1x) ~ f(c), L\x ->0 '

and by the definition of a continuous function in § 2.9, this says that f(x) is continuous at x ~ c. Passing from a single point x ~ c to all the points of an interval we see that we have proved the following result.

THEOREM 1. If a function f(x) is differentiable at all points of an interval, then f(x) is continuous in that interval.

We did not specify, because there is no need, whether the interval is of the type a ;;; x ;;; b (a so-called closed interval), or a < x < b (an open interval), or a ;;; x < b or a < x ;;; b (half-open).

x-axis

FIG. 4.2 Graph of f(x) ~ lxl.

The converse of Theorem 1 is false: a function need not be differentiable at a point just be­cause it is continuous. For ex-ample thefunctionf(x) ~ lxl, with graph as in Figure 4.2, is not differentiable at x ~ 0 even though it is a continuous function for every real value of x.

The notation dyjdx has advantages and disadvantages. The principle disadvantage is that it looks like a fraction, which it is not. It is the limit of a fraction, which is quite a different thing. It is the limit of the ratio

L'1y f(x + L'1x)- f(x) or -------

!:,x

as !:,x tends to zero. Notice that L'1x never actually becomes zero. It is the essential nature of the limit process, as we analyzed it throughout Chapter 2, that a limit has to do with the trend of values of a function up to but not including the critical point.

The advantages of the notation dyjdx fm a derivative are its flexi­bility and convenience. We shall draw attention later to specific instances of what is meant by this.

A REMARK ABOUT NOTATION: although we have used dyjdx to indicate the derivative of any function y ~ f(x), other notation for the same thing is often used, for example,

d Y', f'(x), (y)

dx ' d -f(x). dx

Page 81: Calculus an Introductory Approach

SEC. 3 DIFFERENTIATION 73

Thus, the statement "if y = x2 then dyjdx = 2x" can be written in the abbreviated form (djdx)x" = 2x. There is no reason for restricting this to the variable x; for example

d -(v2) = 2v. dv

In the same way the derivative (2) can occur in any variable:

, . j(x+ !1x)- f(x) f (x) = IIm ,

t:..x-->0 ~X

j'(n) = . j(u+!1u)-j(u)

hm , t:..u --o fl.u

. j(t+!1t)-j(t) hm .

t:..t--> 0 /j.t j'(t) =

Since the derivative is the limit of the ratio of !1y to !1x, it is a measure of the rate of change of y with respect to x. For example, consider the function y = x 2 with derivative dxjdy = 2x. If we think of x as in­creasing steadily from x = 0 to x = !0, then y increases at the same time from y = 0 to y = 100. As x passes through the value 3, y is increasing six times as fast as x, because the derivative is 6. As x passes throu~h the value 7, y is increasing fourteen times as fast as x. Thus the derivative can be thought of as a measure of the rate of change of y compared with the rate of change of x.

Problems

l. Prove that the function y = x2 or j(x) = x2 is continuous for every value of x.

2. Prove that the function f(x) = x3 is continuous for every value of x. 3. Write the derivative of sin x as a limit.

4.3 Simple Derivatives. The procedure in § 4.1, by which it was established that the derivative of y = x2 is dyjdx = 2x, is too primitive for repeated application to function after function. We now outline some theorems that shorten the work.

To illustrate the next theorem, let us differentiate y = 5x2 by the technique of § 4.1.

y = 5x2,

y+!1y = 5(x+!1x)2 = 5x2+ IOx!1x+5(!1x)2,

!1y = !Ox!1x+ 5(!1x)2,

Page 82: Calculus an Introductory Approach

74 CALCULUS: AN INTRODUCTORY APPROACH CHAP. 4

6.y

6.x

dy

dx

lOx+ 56.x,

lim (!Ox+ Mx) = lOx. dX -7 0

The constant factor 5 in y = 5x2 stayed in as a multiplier throughout the process.

THEOREM 2. If y = cu where c is a constant, and u is a differentiable function of x, then

dy du -=c-. dx dx

For example, if u = x2 and c = 5, then y = 5x2, and from the deriva­tive dufdx = 2x we obtain dyfdx = 5(2x) = !Ox.

PROOF. Let a change 6.x in x produce changes 6.u and 6.y in u and y respectively. Then we have

y = cu,

y+6.y = c(u+6.u) = cu+c6.u,

6.y = c6.u,

6.y !J.u -=C-, !J.x !J.x

!J.y [ !J.u] lim - = Iim c- , t..x->0 ~x t..x--;0 l\x

dy du ' -=C-dx dx

by Theorem 18 of § 2.8.

THEOREM 3. If y = c, then dyjdx = 0. If y = x, then dyjdx = 1.

PROOF. In the first case we have

y = c, y+!J.y = c, !J.y = 0, 6.y dy -=0, !J.x dx

0,

and in the second,

y = x, y+!J.y = x+!J.x, !J.y = !J.x, !J.y dy -=I, !J.x dx

I.

The geometric interpretation of this theorem is that the line y = c has slope 0, the line y = x has slope I, as illustrated in Figure 4.3.

Page 83: Calculus an Introductory Approach

SEC. 3 DIFFERENTIATION 75

In the next three theorems we shall assume that u and v are differ­entiable functions of x.

THEOREM 4. Ify ~ u+v, then dyjdx ~ dufdx+dvfdx. Ify ~ u-v, then dyfdx ~ dufdx-dvfdx. y-axis

y=c PROOF. Let us prove both parts at once by use of the sign ± , de­noting plus or minus as the case may be. We can write

y=x

y~u±v, y+fly

~ (u+flu) ± (v+flv),

fly~ flu±flv,

fly flu flv -~-+-. flx flx - flx

x -axis

FIG. 4.3

The theorem is obtained from these equations by taking limits as flx approaches zero, and using Theorem 19 of § 2.8.

ExAMPLE. Differentiate y ~ x2+x3. Solution. Define u and v as x2 and x3 respectively, so that

du dv - = 2x, - = 3x2, dx dx

dy du dv - ~ -+- ~ 2x+3x2. dx dx dx

Theorem 4 states in effect that the derivative of a sum equals the sum of the derivatives, and that the derivative of a difference equals the difference of the derivatives. The analog for products does not hold. It is not true in general that the derivative of a product equals the product of the derivatives. For example, the functions x and x2

have derivatives l and 2x, but their product x" does not have derivative l · 2x. However, there is a rule, as the following result shows.

THEOREM 5. If y ~ uv then

dy dv du - = -u.-+v-. d:r dx dx

PROOF. Again let a change flx in the independent variable x cause changes flu, flv, fly in the dependent variables u, v, y. Then we have

y ~ uv

y+fly ~(u+flu)(v+flv) ~ uv+uflv+vflu+(flu)(flv),

fly~ uflv+vflu+(flu)(flv),

Page 84: Calculus an Introductory Approach

76 CALCULUS: AN INTRODUCTORY APPROACH CHAP. 4

!J.y i'J. V !J.u tJ. V - = u-+v-+!J.u-. tJ.x tJ.x tJ.x tJ.x

As tJ.x tends to zero, so also does tJ.u. Hence taking limits, we get, by Theorems 19 and 20 of § 2.8,

dy dv du dv - = u-+v-+0-dx dx dx dx

dv du = 'lt-+v-.

dx dx

THEOREM 6. If y = ujv then

dy v(dufdx) -u(dvfdx)

dx v2

except fM values of x for which v = 0.

PRoOF. As in the proof of the preceding theorem we suppose that a change tJ.x in x produces changes tJ.u, tJ.v, tJ.y in u, .v' y. Then we have

u y = -,

V

u+!J.u u+!J.u y + tJ.y = --, tJ.y = ---

v+!J.v v+!J.v

v(u + tJ.u) -u( v + !J.v) !J.y = ---:--;-:--­

v(v + tJ.v)

vtJ.u -utJ.v

Dividing by tJ.x we get

tJ.y v(tJ.ujtJ.x) -u(tJ.vftJ.x)

tJ.x v( v + tJ.v)

' V

Finally, we take the limit as tJ.x tends to zero, and the theorem follows because v + tJ.v tends to v.

THEOREM 7. If y = xn, where n is any constant, then

dy - = nxn~l dx

PARTIAL PROOF. Here we establish this for any integer n, and for n = !J2. In § 4.6 we extend the proof to any rational number n. Thus we eventually treat all cases except the irrational values of n, for which

Page 85: Calculus an Introductory Approach

SEC. 3 DIFFERENTIATION 77

the result also holds-although we make no use of such cases in this book. Of course the cases n = 0, l, 2, 3 have been treated earlier in this chapter.

Part I. Let n be a positive integer. If n = 4, y = x•, we can apply Theorem 5 with u = x and v = x3; thus

dy dv du - = u-+v- = (x)(3x2)+(x3)(1) = 4x3. dx dx dx

Just as this gives the derivative of x4 from the known derivative of x3, so we can get the step from xn-1 to xn. If we presume that the derivative of xn-1 is (n -l)x•-2, then we can apply Theorem 5 with u = x and v = xn-l; thus

du

dx

dy

dx

1, dv

dx (n-l)xn-2,

dy dv du - = u-+v-, dx dx dx

(x)(n-1)x•-2+ (x•-1)(1) = (n-1)x•-1+xn-1 = nxn-1.

Thus if the theorem holds for any positive integer n- 1, it holds also for the next integer. Since we have already seen that the theorem holds for n = 1, 2, 3 and 4, it follows that the result holds for any positive integer n. (The proof technique that we have used is called "mathe­matical induction".)

Part 2. Let n be a negative integer, say n = -m, where m is a positive integer. Then y = xn can be written in the form

1 y = -,

xm

and we can apply Part 1 above and Theorem 6 to get

since the derivatives replacing m by - n,

dy

dx

dy xm(O) -1(mxm-1)

de (xm)2

of I and xm are 0 and mxm-1. Hence we have,

-mxm-1 - ( -n)x-n-1 --:::-- = = nxn-1,

x2m x 2n

Part 3. n = !· In this case the equation under discussion is y = x1/2 or y = Vx, whose graph is the upper half of a parabola, as illustrated in Figure 4.4. Notice that we are making a distinction between y = x112 and the full parabola y 2 = x. The other half of the parabola, not shown

Page 86: Calculus an Introductory Approach

78 CALCULUS: AN INTRODUCTORY APPROACH CnAP. 4

in Figure 4.4, has equation y = - xl/2; its graph is symmetric to the graph of y = ,;;;; with respect to the x-axis.

For the equation y = ,;;;; we have

y+!:J.y = yx+!:J.x,

!:J.y = vx+!:J.x- y'x,

-x -axis

FIG. 4.4 Graph of y = V;;, !:J.y

!:J.x !:J.x

To evaluate the limit as !:J.x tends to zero, we multiply the fraction on the right, numerator and denominator, by

thus

!:J.y

!:J.x !:J.y

!:J.x

vx+!:J.x+y'x;

,;x:t~:,.~.,- -v:r !:J.x

(x+!:J.x) -x

!:J.x[ vx + t,-;, + vx J

v~x+vx yx+!:J.x-1- yx'

I

As ~x ->- 0, x+~x ->- x, and hence we have

dy I I _ = ___ = -x-112, dx 2yx 2

It should be noted that the theorem is to be interpreted for appro­priate values of x. For example, in case n = -1, \Ve are dealing with the function y = x-l which is not defined at x = 0. So of course the derivative has no meaning at x = 0. To take another example, let us consider n = f. The domain of the function y = xl/2 is x ~ 0, since xl/2

is not real-valued for negative values of x. Thus the derivative has no meaning for negative values of x. In fact the derivative has no meaning for x = 0, so that the function y = x112 is differentiable only for x > 0.

Problems I. Differentiate the follo·wing:

(a) y = 7x2 (e) y = x-3

(b) y = 7x2+I (f) y = Ifx4

(c) y = 7x2 -4x+I (g) y = 4x

(d) y = 5x6-x4 (h) y =-I

Page 87: Calculus an Introductory Approach

SEC. 4 DIFFERENTIATION 79

2. Extend the proof of Theorem 7 to the case n = i by writing x312 as the product x · x¥, and using Theorem 5.

3. Extend the proof of Theorem 7 to the case n = ~. 4. Extend the proof of Theorem 7 to the case n = - 1· 5. Find the derivatives of the following:

X (a) y = -- (d) xy-x2-7 = 0.

3+x

2+x vx (b) y=-- (e) y=--

x' x+3

x' (c) y=-- (f) y = 4ylx(x+7)-3

x'+3

6. Given that u, v, and w are differentiable functions of x, find formulas for dyjdx in terms of u, v, w, dvjdx, dvjdx, dwjdx in the following cases:

(a) y = u+v+w (d) y = uvw

uv (b) y = u+v-w (e) y =-

w

(c) y = uv-w (f) u+v

y=--w

7. Obtain the formula of Theorem 6 by writing y = ujv in the form u = yv and then applying Theorem 5 with u and y interchanged. (Remark: This procedure requires the assumption that y and v be differentiable functions of x. In contrast, the proof given in the text requires, as we noted just prior to Theorem 4, the assumption that u and v are differentiable, and the differentiability of y is a conclusion. In our use of Theorem 6, it will be important to have this property of y as a conclusion, not as an assumption.)

4.4 Trigonometric Functions. We shall find the derivatives of sin x, cos x, and tan x, where in all cases xis in radian measure.

THEOREM 8. Jj y = sin x, then dyfdx = COS X.

PROOF. Applying the definition of a derivative, we get

y+l:!.y = sin(x+l:!.x),

l:!.y = sin(x+l:!.x)-sinx,

l:!.y sin(x+l:!.x) -sinx

l:!.x l:!.x

dy

dx

sin(x+l:!.x)-sinx lim _...:...__..,.:... __ _ AX-->0 ~X

Page 88: Calculus an Introductory Approach

80 CALCULUS: AN INTRODUCTORY APPROACH CHAP. 4

The problem is to evaluate this limit. Recalling the trigonometric identity

A-B A+B sinA-sinB = 2 sin-- cos--,

2 2

we replace A by x + D.x and R by x to get

sin(x+i'!.x)-sinx ~ 2sin ~x cos(x+ ~x),

(5) sin(x+b.x)-sinx sin(l'!.x/2) ( D.x) ----,----- ~ COS X+- .

D.x D.xf2 2

In equation (10) of§ 2.7 we saw that

sinB lim-- =I, 0--?0 8

and so it follows by replacing 8 by D.xf2, that

. sin(l'!.x/2) hm ~ 1.

<>x-•0 D.xf2 (6)

Combining (4), (5), and (6) and applying Theorem 20 of§ 2.8, we find that

dy ( D.x) - = lim cos x+- = cosx. dx 6-x-)0 2

THEOREM 9. Jj y ~ cos X, then dyfdx ~ -sin x.

PROOF. The steps are very similar to those in the preceding proof. In the present theorem we need the identity

A-B A+B cosA-cosB = -2 sin--sin--.

2 2 Thus '"e can write

y ~ cosx, y+b.y ~ cos(x+b.x),

D.x ( D.x) D.y ~ cos(x+i'!.x)-cosx ~ -2sin 2 sin x+T,

dy

dx

D.v_ ~ _ sin(l'!.x/2) sin(x+ D.x), D.x D.xf2 2

D.y lim- ~

6.X->0 ~X -lim ·lim sin x+-sin(l'!.x/2) ( D.x)

D.x/2 2

~ - (l)(sinx) ~ -sinx.

T

Page 89: Calculus an Introductory Approach

SEC. 4 DIFFERENTIATION 81

THEOREM 10. If y = tan x, then dyjdx = sec2x.

REMARK: This result is not valid, of course, for those values of x for which tan x is not defined, namely, x = ± 7Tj2, ± 37Tj2, ± 57Tj2, ....

PROOF. By Theorems 6, 8, and 9 we have

sinx y=--,

cosx

dy (cos x)(cos x)- (sinx)(- sin x)

dx cos2x

cos2x+sin2x

cos2x

1 = -- = sec2x,

cos2x

where we have used the identity sin'x+ cos2x = 1.

Problems 1. Differentiate the following:

(a) y = cosecx (f) y = sinx cosx

(b) y = secx (g) y = sin2x

(c) y = cotx (h) y = sin3x

(d) y = -5 sinx (i) y = cos2x

(e) xy = sinx+x ( j) y = sin2x+cos2x

2. Use derivatives to find the following limits:

(a)

(b)

(c)

. tan(x+£,.x)-tanx hm --'--.,----'----

Llx_,.o Llx

. tan(u+f,.u)-tanu hm ---.,------

t:.u--Joo 6.u

. cos(v+f,.v)-cosv hm-------t:.v->o Llv

. cos(x+2£,.x)-cosx (d) hm------

t..x->0 ~X

3. Find the derivatives of the following:

(a) y = xsinx (c)

(b) y = x2-cosx (d)

sinx y=-

X

tanx y=--

(1+x)2

Page 90: Calculus an Introductory Approach

82 CALCULUS: AN INTRODUCTORY APPROACH CHAP. 4

4.5 The Chain Rule. Consider the problem of differentiating y = (l +x3)5. With our results so far in this chapter, we would handle this problem by expanding the fifth power of the binomial l + x3; thus

y = l + 5x"+ IOxB+ !Ox• + 5xl2 +x15,

and using our previous results, we get

dy (7) - = l5x' + 60x5 + 90x8 + 60x11 + l5xl4.

dx

It is easier to introduce an intermediate variable u between y and x as follows: y =us, u = l+x". Then we have

dy du - = 5u4, - = 3x2 . du dx

Thus we have separated the original function into two simpler functions whose derivatives are obtained easily, and the next theorem indicates how to combine these derivatives to get dyfdx.

THEOREM ll. If y is a differentiable function of u, and u is a differenti­able function of x, then

dy dy du -=-·-dx du dx·

Before turning to the proof of this, let us see how it applies to the problem under discussion just prior to the theorem, namely y = u•, u = l + x". The theorem gives

dy - = (5u4)(3x') = l5x2u4 = l5x'(l +x3)4. dx

It is perhaps not immediately apparent that this is the same result as in equation (7). but a little simple algebra will establish the equivalence.

PROOF. Any change Llx in the variable x produces a change !'!.u in the variable u and a change !'!.y in the variable y. Now by simple algebra we have

(8) !'!.x !'!.u !'!.x -=-·-.

Taking limits as !'!.x tends to zero, we get the equation of the theorem, by using Theorem 20 of § 2.8.

Unfortunately it is not quite so simple. This proof is valid if !'!.u is not 0, corresponding to any ~x value under consideration. The proof

l I

Page 91: Calculus an Introductory Approach

SEC. 5 DIFFERENTIATION 83

is valid,* for example, for all the problems given at the end of this section, because in each case we can work with sufficiently small values of L'lx that L'lu # 0. However, in case L'lu = 0, the equation (8) becomes meaningless. Now the theorem is correct even in case ~u = 0, as the following argument shows.

First whenever L'lu = 0 we observe that L'ly = 0, because y is a function of u; so a zero change in 'U corresponds to a zero change in y. Now as L'lx tends to zero, the corresponding values of L'lu may be zero or may not be zero. Define v as the following function of L'lu, and so in turn as a function of L'lx:

v = 0 in case !J.u = 0;

L'ly dy . v = - - - m case L'lu ,P 0.

L'lu du

As L'lx tends to zero, so also v tends to zero. Notice that in either of the cases L'lu = 0 or L'lu # 0 we have

dy vL'lu = L'ly- L'lu­

du

dy or L'ly = L'lu- + vL'lu.

du

Dividing by L'lx we get

L'ly L'lu dy L'lu -=-·-+v-. L'lx L'lx du L'lx

Now let L'lx tend to zero, and since v tends to zero, the theorem is proved.

EXAMPLE. Differentiate y = sin 5x.

Solution. Write y = sin u, u = 5x, so that by Theorem 8

dy

du cosu,

du

dx 5,

dy

dx 5 cosu = 5 cos 5x.

Notice that it is not proper to leave the answer in the form 5 cos u, because the variable u was introduced for convenience in solving the problem, but u is not a variable in the original statement of the problem.

*Indeed, it is valid in every case in this book where Theorem 11 is applied. However, if we wanted to terminate the proof of Theorem ll with this simple analysis of equation (8), it would be necessary to add another hypothesis to the theorem along these lines: "and if D.u o;i; 0 for sufficiently small non-zero values of 6.x."

Page 92: Calculus an Introductory Approach

84 CALCULUS: AN INTRODUCTORY APPROACH CHAP. 4

Problems

1. Differentiate

(a) y = cos7x (f) y = cos2x

(b) y = tan3x (g) y = cos2x

(c) y = vx+l (h) y = sin7x

(d) y = yx2 +1 (i) y = va'+x'

(e) y = (2+x')' ( j) y = yb'+(c-x)2

2. Differentiate y = 2 cos2x -1, and prove that the answer is the same as in Question I (f) above.

3. Prove Theorem 9 from Theorem 8 by the use of Theorem I! and the trigonometric identity

cosx =sin(~ -x )·

Suggestion: Write y = sinu, u = (rr/2)-x. "'4. Find the value of

sin(2x + 2L'>x) -sin 2x lim . t~.x->0 6.x

*5. Evaluate

vx+L'>x+3- vx+3 lim .

t!.x->0 6.x

6. Differentiate y = sin 2x and y = 2 sin x cos x, and show that the derivatives are the same (as they should be since sin 2x = 2 sin x cos x is a basic identity of trigonometry).

7. Verify that equation (7) gives the same result for the derivative as the expression 15x2 (! +x3)4 obtained from Theorem 1!.

8. Given the identity sin 3x = 3 sin x-4 sin3x, differentiate both sides to get an identity relating cos 3x, sin2x and cos x.

4.6 Inverse Functions. Consider the function y = x', with graph as in Figure 4.5. The domain of this function is the set of all real numbers, so that to each real number x there corresponds one real value of y. This function has the property that given two different values of x, the corresponding values of y are also different. On the graph this means that any straight line drawn parallel to the x-axis

Page 93: Calculus an Introductory Approach

SEC. 6 DIFFERENTIATION 85

intersects the curve at not more than one point. Because of this pro­perty, we can reverse the correspondence and think of x as a function of y: to each value of y there corresponds a value of x. This is called the inverse junction, with equation x ~ y113. The graph of this equation is the same as the graph of y ~ x3, as in Figure 4.5. y-axis

Since the graphs of y ~ x3 and x = yll3 are the same, we can ob­tain a simple direct relationship between the derivatives of these functions. With respect to x ~ y113 we are seeking the derivative with the roles of x and y interchanged; thus

dx 1\.x lim-

dy "-Y~o 1\.y

. (y + L\.y)ll3- y113 ~ hm -'----'----'----

L\y~o 6.y FIG. 4.5 Graph of y = ,x:3 or x = yl/3.

However, 1\.x and 1\.y here have exactly the same meaning as for the function y ~ x3, and as either of 1\.x or 1\.y tends to zero, so does the other. Hence we can write, by Theorem 21 of§ 2.8,

(9) dy dx 1\.y 1\.x -·- ~ lim-·lim- ~ l. dx dy 1\.x 1\.y

For y ~ x3 we know that dyfdx ~ 3x2, and so from (9) we get

dx l l l - ~- ~ -x-2 ~ -:Y-2/3 for x ~ yli3. dy3x2 3 3

A direct interchange of x and y now gives:

dy l __ -x-213 dx- 3 '

If y ~ x113, then

and so we have established another case of Theorem 7 of§ 4.3. To generalize what we have done, we will use the concept of a mono­

tonic function, i.e., roughly speaking one whose graph is rising from left to right, or falling from left to right, as in Figures 4. 6 and 4.7.

Page 94: Calculus an Introductory Approach

86 CALCULUS: AN INTRODUCTORY APPROACH ~----~~~~--~~~~~~----~~-

CHAP.4

FIG. 4.6 Monotonic increasing function. FIG. 4. 7 Mono tonic decreasing function.

DEFINITION. A function f(x) on the interval a < x < b is said to be monotonic increasing if for any two values of x in the interval, say Xr and xz with xr < xz, the functional values satisfy the inequality f(xr) < f(a·2). A functionf(x) is said to be monotonic decreasing if under the same ciTcum­stances f(xr) > f(xz). A function is said to be monotonic if it is either

/''·" (a, a)

Fw. 4.8 Graph of y = f(x) or x = g(y).

monotonic increasing or monotonic decreasing.

If y ~ f(x) is monotonic on an interval, it is clear that to differ­ent values of x in the interval there correspond different values of f(x). Hence we can reverse the correspondence and think of x as a function of y, say x ~ g(y). This is the inverse function.

THEOREM 12. Let y ~ f(x) be monotonic and differentiable, with j'(x) # 0, on the interval a < x < b. Write (/. and f3 for f(a) and f(b) respectively, and x ~ g(y) for the inverse function. Then x ~ g(y) is a differentiable function on the interval (/. < y < f3 or (/. > y > f3 according as f(x) is monotonic increasing or decreasing, and

' dy dx -.- ~ f'(x) . g'(y) ~ 1. dx dy

(The interval a ~ x ;S b, a so-called "closed" interval was used in the definition of a monotonic function, whereas we use the "open" interval a < x < b in the present theorem. It could be either way in the definition, but the theorem is more useful with x over an open interval.)

1 I

Page 95: Calculus an Introductory Approach

SEC. 6 DIFFERENTIATION 87

PROal!'. The assumption that f(x) is monotonic assures us that D.y of 0 as long as D.x of 0, and D.x of 0 as long as D.y of 0. Hence we can apply Theorem 21 of § 2.8 to the simple algebraic relation

D.y D.x -·--=I, D.x D.y

and the theorem follo·ws. It should be noted that the interval a < x < b need not be finite.

For example, if we want to restrict x to positive real values, we can write 0 < x < oo. This would be an appropriate restriction for y = x" as discussed at the opening of this section, because the derivative of this function, 3x2, equals zero at the point (0, 0). Hence Theorem 12, while applicable to y = x 3 over 0 < x < oo, does not apply to y = x3

over say -2 < x < oo. This is not surprising, since the inverse function x = y113 has no derivative at (0, 0).

Theorems 11 and 12 illustrate one advantage of the notation dyjdx for a derivative. The equations in these theorems would be obvious if the derivative were a fraction with numerator dy and denominator dx. In many cases it is permissible to treat the derivative in just this way, as these theorems establish. Furthermore this notation makes it easy to remember what these theorems assert.

Next we use Theorems 11 and 12 to extend the proof of Theorem 7 of § 4.3 to any rational number n. Theorem 7 has already been estab­lished for any integer n. Next let n be the reciprocal of a positive integer, say n = Ijk. We want to differentiate the function y = xltk.

In order that we may apply Theorem 12, we want to avoid the point (0, 0), so we shall take 0 < x < oo as the domain of the function. For this domain the function is monotonic increasing, and the inverse function is x = yk. The derivative of this is, by Theorem 7,

and by Theorem 12 we get

dx - = kyk-1, dy

I dy 1 - = -- = = nxn-1. dx kyk-1 kx1-l!k

Now we let n be any rational number, say n = rnjk, where rn and le are integers. Although n may be positive or negative, we may presume that k is positive, absorbing any negative sign into rn. We want to differentiate y = xmlk, and we write

y = um, u = xllk,

Page 96: Calculus an Introductory Approach

I I ·'I ,,

I. i

. '

88 CALCULUS: AN INTRODUCTORY APPROACH CnAP. 4

and so

dy

du

du 1 _ = -x(llk)-1. dx le

By Theorem ll we conclude that

dy (1 ) - = (mum-1) -x(lik)-1 dx le

m = -x(m-1)/k. x(l-k)lk

le

= nx(m-1)/k = nxn-1,

Hence we have proved Theorem 7 for all rational exponents. Next we turn to the equation y = tan x. To get an inverse function

from this it is desirable to restrict severely the domain of values of x.

I I

"I X= --1 21

I I I I I

y -axis

I I I

I I

x-axis

I 1r iX=--1 2 I I I I I

Fw. 4.9 Graph of y = tan x, -'Tr/2 <X< 7Tj2.

dy - = sec2x, dx

When we discussed the equation y = tan x in Theorem 10, the domain was - oo < x < oo except for x ~ ± Trj2, ± 3Trj2, ± 5Trj2, .... We now consider what is in effect a different function, namely, y =

tan x with the domain -Tr/2 < x < Tr/2, with graph as in Figure 4.9. Inasmuch as this function differs from the one in Theorem 10 only in the narrowing of the domain of values of x, we observe that the conclusion of Theorem 10 is still valid. As illustrated by Figure 4.9, the function under consideration is monotonic in­creasing, and so it has an inverse function x = arc tan y, this being the standard notation from trigo­nometry. By Theorems 10 and 12 we can write

dx 1

dy sec2x

From elementary trigonometry we use the well-known identity

sec2x = 1 +tan2x,

Page 97: Calculus an Introductory Approach

r--· f

SEC. 6 DIFFERENTIATION 89

whence dx 1

sec2x = 1 +y2 and - = ~-dy 1 +y2

Thus we have established the proposition:

dx 1 If x = arctany, then-=~-.

dy 1 +y2

Now if we interchange x and y so as to put this in the more standard form where y is a function of x, we obtain the following result.

THEOREM 13. The derivative of y = arc tan x is

dx 1 -=--dx 1 +x2

It should be kept in mind that the domain of the function y = arc tan x is - oo < x < oo and the range is -7T/2 < y < 7Tj2, as in Figure 4.10.

Y= L 2

y -axis

-------------

x-axis

Y=-f

FIG. 4.10 Graph of y = arc tan x.

Problems

1. Prove that the tangent line to the graph of y = arc tan x at the point (0, 0) has slope 1.

2. Find the x coordinates of the points, if any, on the graph of y = arc tan x where the curve has slope i.

3. Find the x coordinates of the points, if any, on the graph of y = arc tan x where the curve has slope 5.

4. Find the derivatives of the following:

(a) y = arctan2x (c) y = (arc tanx)'

(b) y=2+arctanx (d) y = arc tany'x

Page 98: Calculus an Introductory Approach

90 CALCULUS: AN INTRODUCTORY APPROACH CnAP. 4

5. (a) In Theorem 11 let us write y ~ g(u) and u ~ J(x). Show that the conclusion of the theorem can now be written in the form

dy - ~ g'(u)J'(x) dx

(b) Next eliminate u and show that the conclusion of Theorem ll can be stated in the form: If y ~ g(f(x)) then

dy ~ g'(J(x))J'(x). dx

(c) Next remove the y and show that the conclusion of Theorem ll can be stated thus: The derivative of g(f(x)) is g'(f(x))j'(x).

6. Prove Theorem 12 by using part (c) of the preceding problem. Sug­gestion: x ~ g(y) and y ~ J(x) imply that x ~ g(J(x)). This equation gives two forms of the same function, and differentiating \Ve get

l ~ g'(f(x))j'(x) ~ u'(y)J'(x).

4.7 The Mean Value Theorem. We now discuss one of the most important results in calculus.

THEOREM 14. Suppose that a junction F(x) has· a derivative j(x) for all x satisfying a < x < b. Then the.·e is a value of x strictly between a and b, say x ~ c with a < c < b, such that

F(b)-F(a) j(c) ~ or F(b) -F(a) ~ (b -a)j(c).

b-a

To give a plausible argument for the truth of this, let us look at the matter geometrically. The quotient [F(b) -F(a)]j(b- a) is the slope of

s Q

p~ FIG. 4.11

the straight line joining the two points (a, F(a)) and (b, F(b)) on the graph of y ~ F(x). Calling these points P and Q respectively, we see that the theorem states that there is at least one point on the graph between P and Q where the tangent line is parallel to PQ. In Figure 4.11 there are two such points.

To continue our intuitive argument, let R be any point on the graph y ~ F(x) between P and Q with the following property: The perpen­dicular distance from R to PQ is a maximum. It is intuitively clear that such a point exists. Next let a straight line RS be drawn through R parallel to PQ. There will be no points of the graph of y ~ F(x) lying

Page 99: Calculus an Introductory Approach

SEC. 7 DIFFERENTIATION 91

on the far side of RS, that is, the side away from PQ, at least not in the range x = a to x = b. Thus RS is a tangent line to the curve, and RS is parallel to PQ.

Such an intuitive appeal to the geometry of the situation is not to be regarded as an adequate proof. A rigorous proof must be analytic in nature, that is, it must be based only on the foundations oflogic and the real number system, not on geometric intuition. Such a proof is beyond the scope of this book. Our viewpoint is that the mean value theorem is taken as an axiom, and we build the further structure of calculus on it. As a first example of the use of the mean value theorem, we establish a connection between the sign of the derivative and the increase or decrease of a function.

THEOREM 15. Let F(x) be a differentiable function on a <;; x <;; b, such that F'(x) > 0 for all x on the open interval a < x < b. Then the function F(x) is monotonic increasing on the closed interval a <;; x <;; b. In a similar setting, if P'(x) < 0 then P(x) is monotonic decreasing.

PROOF. We prove the first part only, since the second part is analog­ous. Let xr and x2 be any two numbers in the closed interval such that xr < x2, and so a ;" xr < x2 ;" b. We apply Theorem 14 to F(x2) -F(xr). Thus there is a number c strictly between xr and x2 such that

F(x2)-F(xr) = (x,-x,)F'(c).

Now F'(c) is positive by hypothesis, and so is x,-x,. Hence F(x2) -F(xr) is positive, and the theorem is established.

The theorem could have been stated more readily if F'(x) > 0 had been assumed on the closed interval a ~ x ~ b. However the result so stated would have been weaker, and would not have applied, for example, to the function F(x) = x2 with a = 0 and b = 1, even though this function is monotonic increasing over the interval 0 ~ x ~ l. Such an example is not as special as it may seem; it occurs in more general form in the proof of Theorem 17.

Problems

l. Find the c of the mean values theorem (Theorem 14) in each of the following cases:

(i) F(x) = x(x-1), a= 0, b = 2;

(ii) F(x) = x(x-1), a= 0, b = 1 ;

(iii) F(x) = x(x-1), a= 1, b = 2;

(iv) F(x) = sinx+2x, a = 0, b = 1r;

(v) F(x) = x'-x', a= 0, b = l.

Page 100: Calculus an Introductory Approach

92 CALCULUS: AN INTRODUCTORY APPROACH CHAP.4

*2. If rt., (3, y are real numbers satisfying 2rt. + 3(3 + 6y = 0, prove that r~.xZ+ (3x+y = 0 has at least one root between 0 and l. Suggestion: Apply the mean value theorem to F(x) = 2a.x3+3(3x2+6yx with a= 0, b = l.

3. Prove that the function F(x) = x(x-1) is monotonic decreasing on 0 ~ x ;;; }, but monotonic increasing on t ;;; x ~ 1.

4. Prove that x- sin x is monotonic increasing on 0 ~ x :;;; k where k is any positive number. Hence prove that sin x < x for any positive value of x.

5. Prove that tan x > x for any x satisfying 0 < x < 7Tj2. 6. For which positive values of x is y = x-ljx a monotonic increasing

function1 7. Prove that arc tan x is a monotonic increasing function of x for all

values of x. 8. Prove that y = x2- 4x is monotonic increasing for x ;,;; 2, but mono.

tonic decreasing for x ~ 2.

4.8 Maxima and Minima. The differentiation process can be used in many problems to find the maximum and minimum values of a function. In Figure 4.12 it is intuitively clear that at points A and B,

D

~y=F(x) .t( :' f' 1 I I

I I

I '

a b c FIG. 4.12 y = F(x). Fra. 4.13

where the function F(x) has maximum and minimum values, the derivative F'(x) is zero. Note that we are discussing so-called local maximum and minimum values, that is to say, maximum and minimum points compared with nearby points on the graph. For example, in Figure 4.12 the values of F(x) in the vicinity of the point C are smaller than at B, and likewise the values of F(x) in the vicinity of the point D are greater than at A. To be specific, we shall say thatF(x) is a maximum at x = g if there is an interval containing g, say a < g < b, such that F(g) > F(x) for all values of x satisfying a ;;; x ;;; b except x = g itself. The definition of a minimum is the same with F(g) > F(x) re­placed by F(g) < F(x). First we establish that F'(g) = 0 is a necessary condition for a maximum or minimum.

THEOREM 16. Let F(x) be a function having a derivative for all x satisfying a ;;; x ;;; b. Suppose that F(x) is a maximum at a specific value x = g,

Page 101: Calculus an Introductory Approach

SEC. 8 DIFFERENTIATION 93

where a < ~ < b; i.e., suppose that F(~) > F(x) for all x in the interval a ;:;; X ;:;; b except X ~ r ThenF'(~) ~ 0. Similarly if F(x) is a minimum at x ~ ~. then F'(~) ~ 0.

PROOF. We restrict our attention to the maximum case, since the minimum case is analogous. The proof is indirect; we show that it is impossible that F'(~) > 0 or that F'(g) < 0. First we suppose that F'(g) > 0. Then by (3) of § 4.2 we see that

. F(~+L'.x) -F(~) hm > 0.

tlx->O 6-x

Consider positive values of L'.x. Since the limit is positive, the difference F(~ + L'.x) -F(~) must be positive for L'.x sufficiently small. That is, F(~+L'.x) > F(~) for small positive values of L'.x. But this contradicts our assumption that F(~) is a maximum.

On the other hand, suppose that F'(~) < 0, so that

. F(~ + L'.x) -F(~) hm < 0.

tlx->0 6-x

Consider now negative values of L'.x. Since the limit is negative, the difference F(~ + L'.x) -F(g) must be positive for L'.x sufficiently close to zero. Again this gives us the inequalityF(~+L'.x) > F(~), which contra­dicts the assumption that F(~) is a maximum.

Since F'(x) ~ 0 at both A and Bin Figure 4.12, how can we distin­guish a maximum point from a minimum point1 To answer this, we recall that the derivative F'(x) can be thought of as the slope of the curve. Thus F' (x) is positive at points on the curve immediately to the left of A, and negative at points immediately to the right. In the vicinity of the point Bit is the other way around; F'(x) is negative at points immediately to the left of B, and positive at points immediately to the right. These considerations suggest the following result.

THEOREM 17. Suppose that a junction F(x) has a derivative for all x satisfying a ;:;; x ;:;; b. Suppose furthermore that for a specific value of x between a and b, say x ~ ~with a < ~ < b, the derivative F'(x) changes from positive to zero to negative as x passes through ~ from left to right:

F'(x) > 0 for a ;:;; x < ~; F'(x) ~ 0 for x ~ ~;

F' (x) < 0 for ~ < x ;:;; b.

Then F(x) is a maximum at x ~ ~in the sense that F(~) > F(x) for every value of x satisfying a ;:;; x < ~ or f < x ;:;; b. Similarly F(x) is a

Page 102: Calculus an Introductory Approach

94 CALCULUS: AN INTRODUCTORY APPROACH CHAP. 4

m'mmum at x ~ £if F'(x) changes from negative to zero to positive as x passes through £ from left to right.

PROOF. We prove this in the case of a maximum; the case of a mini­mum is analogous. By Theorem 15 the function F(x) is monotonic in­creasing on the closed interval a ;;; x ;;; ~, and monotonic decreasing on the closed interval£ ;'i x ;o; b. Hence F(x) is a maximum at x ~ g.

It can occur that F'(£) ~ 0, but that F'(x) does not have different signs to the left and right of x ~ g. For example, consider the function F(x) ~ x3. The derivative is F'(x) ~ 3x2, so that F'(O) ~ 0. But F'(x) is positive both for x < 0 and for x > 0. This function assumes neither a maximum nor a minimum value at x = 0, as illustrated in Figure 4.5 in § 4.6.

In summary then, we have established in Theorem 16 that in order that a differentiable function F(x) have a maximum or minimum strictly inside an interval a ;'i x ;'i b it is necessary that F'(x) ~ 0 for some value of x in the interval. However, knowing that F'(x) ~ 0 for a certain value of x is not sufficient to guarantee a maximum or minimum. Additional conditions are set forth in Theorem 17 to assure the presence of a maximum or n1inimum. Consequently, given a function F(x) we must take two steps to determine its maxima and minima. First we must find all values £ such that F'(£) ~ 0. Second we must apply Theorem 17 to each such value in order to determine whether F(x) is actually a maximum, a minimum, or neither. (In more extended works ou calculus, alternatives to Theorem 17 are given, but we limit ourselves to this one result.)

EXAMPLE. For what values of x is F(x) ~ x3- 12x2 + 45x- 40 a maximum or a minimum?

Solution. By Theorem 16 all maxima and minima occur at values of x where F'(x) ~ 0. We note that

F'(x)

F'(x)

3x2-24x+45;

0 if 3x2-24x+45 ~ 0.

The solutions of this equation are x = 3 and x = 5. Furthermore, since F'(x) can be written in the form 3(x- 3)(x- 5), we observe that

F'(x) > 0 for x < 3, F'(x) < 0 for 3 < x < 5, F'(x) > 0 for x > 5.

Hence F(x) is a maximum at x ~ 3, a minimum at x ~ 5, by Theorem I 7.

NOTE. Our discussion of maxima and minima has been restricted to situations where the maximum or minimum occurs at an interior point of the domain of definition of the function. For example in Theorem 16

Page 103: Calculus an Introductory Approach

SEC. 8 DIFFERENTIATION 95

the maximum or minimum was presumed to be located at x = g, where a < g < b. We did not allow a ;;; g ;;; b. A consequence of this is that our precedure will not necessarily locate maximum or minimum points at the end of the domain of the function. As an illustration of this, consider the function y = V1-x". The graph of this is a semi­circle, actually the upper half of a circle with radius 1, center at (0, 0). Our procedure, applied to this function, will locate the maximum point at x = 0, y = 1, but not the minimum points at x = 1, y = 0 and x = -1, y = 0. These minimum points are end points of the domain of the function, the domain being - l <; x ;;; 1.

Problems

1. Examine the following functions for maximum and minimum values of y:

(a) y = x2-4x-4 (f) y = x'-3x'+3x+4

x' (b) y = 6+7x-x2 (g) y=--

x+1

9 (c) y = 2x'+3x2 -36x+45 (h) y = x'+-

X

18 (d) y = 2x'+3x2-36x+42 (i) y = 2x"+-

X

(e) y = x'-6x"+12x+l0

2. For any fixed numbers k1, kz, ... , kn determine x so that each of the following shall be a minimum:

(a) (x-k1)2

(b) (x-k1)2 +(x-k,)2

(c) (x-kl)2 +(x-kz)'+(x-ks)2

(d) (x-kl)2 +(x-/cz)2 +(x-ks)2+ ... +(x-kn)2.

*3. Among all rectangles of fixed perimeter p, find the dimensions of the one with largest area.

4. Find two numbers whose sum is 12 and whose product is a maximum. 5. Let c be a positive constant. Find two numbers whose sum is c and

whose product is a maximum. 6. Find three positive numbers whose sum is 12 and whose product is a

maximum. Suggestion: Use the result of the preceding problem. If the three numbers are x, y, and 12-x-y, then x = y since otherwise

x+y x+y -- · -- · (12-x-y) > xy(l2-x-y).

2 2 Similarly x = 12-x-y and y = 12-x-y.

Page 104: Calculus an Introductory Approach

96 CALCULUS: AN INTRODUCTORY APPROACH CHAP. 4.

7. Find two positive numbers x and y whose smn is 10, (a) such that x2y is a maximum, (b) such that x3y is a maximmn.

8. Find the maximum value of sin x+ cos x for all values of x between 0 and -rr/2.

4.9 The Problem of Minimum Surface Area of a Cylinder of Fixed Volume. ~Te return to the problem posed in § 1.2, to find the dimensions of a cylinder of volume 1877 with minimum surface areaS. The equations were, without repetition of all the details,

367T S = 2-rrr2 +- = 27Tr2 + 367Tr-1,

r

where rand h denote the radius and height of the cylinder. The deriva­tive is, by Theorem 7,

dS

dr 41Tr- 361Tr-2,

and this equals zero in case

47Tr = 367T:r-2, r = 9r-2, r3 = 9, r = ~9.

For positive values of,. less than vi) we see that

r3 < 9, 9

I<­r3 , dS = 47Tr(l- ~) < 0. dr r3

For values of r greater than -\1'9 we can write

9 dS r3>9 1>-

, r3' dr

Hence Theorem 17 assures us that S is a mm1mum when r = viJ. The other dimension of the cylinder would have the value

18 18 -h =- =- = 2-\1'9.

:r2 9213

The actual minimum value of S is

S = 27Tr2+27Trh = 277(-\Y9)2+27T(-\Y9)(2-{Y9) = 6"9213.

The use of Theorem 17 could have been dispensed with here. The geometry of the situation assures us that there is a minimum for some positive value of r. Theorem 16 tells us that the derivative is zero at this

Page 105: Calculus an Introductory Approach

SEC. lO DIFFERENTIATION 97

location. Then our calculation showed that the derivative is zero for a single value of r, and so this must provide the location of the minimum value ofS.

Problems

1. Find the dimensions of a cylinder of volume 250 cubic units, if the surface area is to be a minimum.

2. Find the dimensions of a cylinder of surface area 547T square units, if the volume is to be a maximum.

3. Prove that a cylinder of fixed volume has minimum surface area in case h = 2r, where h and r denote the height and radius of the cylinder.

*4. A rectangular dog run is to be built, using one side of a house and 60 linear feet of fencing. What is the maximum area that can be fenced off?

*5. A page of a book is to have area 48 square inches, with l-inch margins at the top and the bottom of the page, and i-inch margins at each side. What should be the dimensions of the page so that the space available for printed matter is a maximum?

*6. A right triangle has sides 6, 8, and 10 units long. What are the dimen­-sions of the rectangle of maximum area that can be inscribed, with one side of the rectangle lying along the longest side of the triangle 1

*7. What are the dimensions of the cylinder of maximum volume that can be inscribed in a sphere of radius 11

*8. Given that one base and the two non-parallel sides of a trapezoid are each 6 units long, what should be the length of the fourth side for maximum area1

*9. Two points A and B lie in the same plane as, and on the same side of, a certain fixed line. The perpendicular distances from A and B to the line are 16 and 4 units, respectively. The points A and B are 13 units apart, thus AB = 13. A point Pis chosen on the line in such a way that AP+PB is a minimum. What is the length PB?

4.10 The Refraction of Light. As light travels from one medium into another, say from water into air, it is refracted in the sense that its path is bent or deflected from travel in a straight line. It is this phenomenon that makes a rod appear bent when it is thrust into water. The law of the index of refraction (Snell's law) is

Water Path of light ray

FIG. 4.14

Surface

'-;_'"',Appearance ..... .:::-::.:::::;..::::.

Reality

Fw. 4.15

Page 106: Calculus an Introductory Approach

98 CALCULUS: AN INTRODUCTORY APPROACH CHAP. 4

Slllr.t. (10)

sin,B

where v, is the velocity of light in the upper medium (say air), v2 is the velocity of light in the lower medium (say water), a is the angle of incidence (see Figure 4.16), ,8 is the angle of refraction (see Figure 4.16).

The ratio of velocities v1jv2 is approximately 4/3 in the air-to-water case, so that a is the larger angle, as illustrated.

A

Velocity v1

Velocity v2

B

Fw. 4.16

A

a

c

FIG. 4.17

D

b ,8

B

There is a general law, the minimum principle, that says that light travels from B to A in such a way as to minimize the elapsed time. We will show that (10) can be deduced from this law. It will also be assumed that light follows a straight line in a homogeneous medium such as air or water. We regard A and B as fixed points, C and D as the points of intersection of the perpendiculars drawn from A and B to the surface separating the media, and P as the point on the surface where the path of light changes direction. Denote the lengths AC, BD, and CD by the constants a, b, and c. Since we do not know at the outset the location of the point P, except that it is on the line CD, we denote the length CP by x, and so PD by e-x. Hence we have

AP = ya2+x2, PB = yb2+(c-x)2,

and so the total time t of passage of the ray of light from A to B via P is

ya2+x2 yb2+(c-x)2 t = + '

V1 V2

1 1 t = -(a2+x2)112+ -(b2+c2-2cx+x2)112,

Vt V2

Page 107: Calculus an Introductory Approach

SEC. lO DIFFERENTIATION 99

The only variables in this equation are x and t, although we should keep in mind that e< and flare also variables in Figure 4.17. We use the above equation to determine the values of x that make t a minimum. Differentiating we get

dt x -e+x (11) - = -(a'+x')-11'+ ----(b'+e'-2ex+x2)-112.

dx VI vz

It is not easy to solve for x when this derivative is equated to zero, so we proceed indirectly. The derivative (ll) is zero in case

X e-x -(a'+x')-112 --(b' + e'- 2ex+x')-112 V1 v, 1 X 1 e-x

(12) ya2 +x2 yb2+(e-x)2 ' VI v,

1 1 -sin a - sinfl. VI v,

To prove that this result gives a minimnm value oft, we must proceed slightly differently than in preceding sections, because we have not computed the actual value of x for which dtfdx = 0. In view of equations (12) we see that (ll) can be written in the form

dt l l (13) -=-sina--sinfl.

dx VI V2

Let x = q be the value of x for which dxfdt = 0. Then from Figure 4.17 we observe that a decrease in x from x = q causes a decrease in a and an increase in f3. Since a and f3 are acute angles, this causes a decrease in sin a. and an increase in sin fl. From (13) we see that this produces a negative value for dtfdx. Hence for x < q we have dtfdx < 0.

On the other hand an increase in x from the value x = q causes an increase in a aud a decrease in {3, as seen in Figure 4.17. This results in an increase in sin a and a decrease in sin {3, and so from (13) we see that dxfdt is positive. We have shown that dxjdt passes from negative to zero to positive as x passes through x = q, and hence by Theorem 17 the timet is a minimum when equations (12) are satisfied.

Problem (For readers familiar with solid geometry.)

In proving equation (10) we took it for granted that the point P would lie on the line CD, i.e., that the plane APB would be perpendicular to the surface of the water. Explain this.

Page 108: Calculus an Introductory Approach

CHAPTER 5

THE FUNDAMENTAL THEOREM

5.0. The fundamental theorem of calculus shows the close relation between differentiation and integration, enabling us to extend inte­gration far beyond the few cases handled in Chapter 3. It also enables us to prove the existence of the integral for a wide class of functions; the extension to an even wider class of functions is given in Appendix A.

5.1 Derivatives and Integrals. The number of functions f(x) whose integrals were evaluated in Chapter 3 is much smaller than the number of functions whose derivatives were evaluated in Chapter 4. Furthermore, in Chapter 4 we had many more general theorems to enable us to work from the differentiation of one or two functions to a wide class of related functions. This deficiency in our theory of inte­gration will now be remedied in part by a broadening of the class of functions we are able to integrate. To do this, we establish a version of the fundamental theorem of calculus, which (roughly stated) says that integrations and differentiation are reverse processes.

To give an example, in § 3.3 it was established that

fb xn dx = bn+l - an+l

a n+l n+l

in case n = 0, l, 2, and the case n = 3 was given as a problem. Inasmuch as the derivative of xn+lj(n+ l) is xn this suggests the general pro­position that

( l) J>(x) dx ~ F(b)-F(a),

where F(x) is a function whose derivative is f(x). Another example of this formula (l) is given by Theorem 4 of§ 3.6,

s: sin x dx ~ (-cos b)- (-cos a),

lOO

Page 109: Calculus an Introductory Approach

SEC. 2 THE FUNDAMENTAL THEOREM 101 --------------------------------

which is (I) withf(x) replaced by sin x, and F(x) replaced by -cos x. Equation (I), with appropriate qualifications on the functions f(x)

and F(x), is the fundamental theorem of calculus. But first we discuss a preliminary result.

5.2 Upper and Lower Sums. We now return to the definition of an integral as it was given in equation (4) of§ 3.2:

(2)

where Xo = a, Xn = b,

~~ is any value satisfying x;-1 ;;; ~~ ;;; Xj, and the largest of x1-xo, x2-x1, ... , Xn-Xn-1 tends to zero as n tends to infinity. This notation and the accompanying conditions will be assumed throughout the rest of the chapter.

In this section we assume that f(x) is monotonic on the interval a ;;; x ;;; b (see § 4.6 for the definition of a monotonic function). Next we define the sums Sn and Sn by the equations

(3) Sn = (x1-xo)f(xo)+(x,-x1)f(x1)+(xa-x2)f(x,)+ ...

+ (xn- Xn-1)f (Xn-1),

(4) Sn = (x1-xo)f(x1)+(x,-x1)f(x,)+(xa-x,)f(xa)+ ...

+ (xn -Xn-1)f(xn)·

THEOREM I. If f(x) is monotonic increasing, then

n Sn ;;; '2, (Xj-XH)f(~j) ;;; Sn

j~1

If f(x) is monotonic decreasing, the inequalities are reversed.

PROOF. Consider the monotonic increasing case; the other case is analogous. We observe that

f(Xj-1) ;;; f(~j) ;;; f(Xj),

(Xj-Xj-1)f(xH) ;;; (Xj-Xj-1}f(~j) ;;; (Xj-Xj-1)f(xj),

and the theorem follows when we sum these inequalities from j = I to j = n.

Page 110: Calculus an Introductory Approach

102 CALCULUS: AN INTRODUCTORY APPROACH CHAP. 5

THEOREl>I 2. If f(x) is monotonic, then

lim (Sn -sn) = 0. n~ro

PROOF. Again we confine our attention to the monotonic increasing case, since the other case is analogous. We see that

(5) n

Sn -Sn = L (x;-x;-1)[f(xJ)-f(x;-1)] ;;:; 0. ;~1

Let tn denote the maximum of x;-X;-1 for all values j = 1, 2, 3, ... , n. Then we can write

n n (6) Sn -Sn ~ .L tn[f(x;)-f(x;-r)] = tnL [j(x;)-f(XJ-I)J.

j=l J-1

The last sum involves much cancellation because it is

[j(xr)- f (xo)] + [f(x2)-f(xr)] + [f(xa)- f(x2)J.+ ...

+ [f(xn)- f(Xn-1)] = - f(xo) + f(xn) = f(b)- f(a).

Hence (5) and (6) imply that

0 ;;; Sn -Sn ~ tn[f(b)- f(a)].

But f(b)- f(a) is a constant, and tn -> 0 as n -> oo, so we can apply Theorem 9 of§ 2.3, with an replaced by tn[f(b)- f(a)] and bn replaced by Sn-Sn. Thus Sn-Sn-+ 0 as n -> oo, and the proof is complete.

THEOREM 3. Let F(x) be a differentiable function whose derivative is f(x) on the interval a ~ x ~ b. Let f(x) be monotonic on this interval. Then with Sn and Sn defined as in (3) and (4),

(7) limSn = limsn = F(b) -F(a). n-)c:o n-)oo

PROOF. Again it suffices to prove this in case f(x) is monotonic increasing. We apply the mean value theorem of§ 4.7 to F(x), with a and b replaced by x0 and x1. We conclude that there is a value 81 between Xo and xr such that

F(xr) -F(xo) = (xr -xo)f(Or).

l ! I

~

Page 111: Calculus an Introductory Approach

SEC. 3 THE FUNDAMENTAL THEOREiil 103

Similarly there are real numbers 82 between x1 and x2, 83 between x2 and x3, ... , Bn between Xn-1 and Xn, such that

F(x2) -F(x,) ~ (x, -x,)f(82)

F(xa)-F(x2) ~ (x.-x,)J(Bs)

Adding all these equations we get

F(xn)-F(xo) ~ (x,-xo)f(B,)+(x,-x,)f(B2)+(xa-x2)f(Ba)+ ...

+ (xn- Xn-l)f (Bn)·

The sum on the right lies between Sn and Sn, by Theorem 1 above. Also Xn = b ind xo = a, so we have

Sn ~ F(b) -F(a) ~ Sn.

To this we apply Theorem 14 of§ 2.4, with c replaced by F(b)-F(a), an replaced by Sn, and bn replaced by Sn. In view of Theorem 2, we obtain equations (7).

Problems

1. Computesn,SnandF(b)-F(a)incasej(x) ~ 2x,F(x) ~ x 2,a ~ xo ~ 0, x1 = .2, xz = .4, xa = .6, x4 = .8, b = xs = l.

2. Compute Bn, Sn, and F(b)-F(a) in case j(x) ~ 2x, F(x) ~ x', a~ 0, b ~ 1, n ~ 10, Xj ~ j/10 for j ~ 0, 1, 2, ... , 10.

3. Compute· Bn, Sn, and F(b)-F(a) in case j(x) ~ 2x, F(x) ~ x', a~ 0, b ~ 1, n ~ 20, Xj ~ j/20 for j ~ 0, 1, 2, ... , 20.

4. The same as Question 1, but with j(x) ~ 3x', F(x) ~ x•. 5. The same as Question 2, but with j(x) ~ 3x', F(x) ~ x•. 6. The same as Question 3, but with f(x) ~ 3x2, F(x) ~ x•.

5.3 The Fundamental Theorem of Calculus. Although we do not establish the fundamental theorem for the widest possible class of functions, nevertheless the results we obtain are adequate for all functions within our purview.

THEOREM 4. Let F(x) be a Junction with a derivative f(x) on the interval a ~ x ~ b. If J(x) is monotonic on this interval, then the integral (2) exists and its value is

J:f(x) dx ~ F(b) -F(a).

Page 112: Calculus an Introductory Approach

104 CALCULUS: AN INTRODUCTORY APPROACH CHAP. 5

PRooF. Again we treat only the case where f(x) is monotonic increas­ing. Write an for the sum in equation (2), that is

(8)

so that

n an= 2,(xj-Xj-!)f(gj),

j~!

fb f(x) dx = lim an a n-)co

if this limit exists. By Theorem 1 we have

Sn ;;; an ~ Sn.

To these inequalities we apply Theorem 17 of § 2.4. In view of Theorem 3 above we conclude that

lim an = limSn = !imsn = F(b) -F(ct). n->co

Thus Iim an exists, and so by (8) the integral exists.

a

FIG. 5.1 Piecewise monotonic function.

Next we generalize Theorem 4 to functions f(x) that are piecewise monotonic, which means that the graph of f(x) consists of a finite number of pieces each of which is monotonic.

DEFINITION. A function f(x) is said to be piecewise monotonic on an interve<l a; ;:;; x ;:;; b if the interval can be separated into a finite number of pieces.

a ~ X ~ Cl, C1 ~ X ;;::; C2, ... , Cj-1 ~ X ~ Cj, Cj ~ X ~ b,

such that f(x) is monotonic on each piece.

THEOREM 5. Let F(x) be a function with" derivative f(x) on the interval a ;:;; x ;:;; b. If f(x) is piecewise monotonic on this interval, then the integral (2) exists and is equal to

J:f(x) dx = F(b)-F(a).

Page 113: Calculus an Introductory Approach

SEC. 3 THE FUNDAMENTAL THEOREM 105

PROOF. This follows from Theorem 4 and formula (5) of § 3.2. First, by repeated application of this formula we get (dropping the f(x) dx in the integrals for convenience)

To each integral in the last sum we apply Theorem 4, and hence

J:f(x) dx = [F(c,) -F(a)] + [F(c2) -F(c,)] + [F(c3) -F(cz)]

+ ... + [F(c;) -F(c;-I)] + [F(b) -F(ct)J.

All the terms on the right side cancel except two, namely -F(a) and F(b ). Thus the theorem is proved.

REMARK ON NOTATION. The notation

b x=b [F(x)]a or [F(x)Jx~a

is often used to denote F(b) -F(a).

EXAMPLE l. Evaluate the integral.

J7x3 dx.

5

Solution. Here we have f(x) = x3 and so we can take F(x) to be x4J4; hence we get

J'x3 dx = [x4]' = 2401- 625 = 444. 5 4 5 4 4

ExAMPLE 2. Evaluate the integral

Jn/2

0 cosx dx.

Solution. In this case f(x) = cos x and we can take F(x) to be sin x. Hence we have

Jn/2

0 cosxdx = [

• ]''" • 1T • SlllX =Sill-- Slll0 = l-0 = 1. 0 2

Page 114: Calculus an Introductory Approach

106 CALCULUS: AN INTRODUCTORY APPROACH CHAP. 5

Problems

l. Prove Theorem l of § 3.3 by the methods of the present section. 2. Prove Theorem 4 of § 3.6 by the methods of the present section. 3. Evaluate the following integrals.

(a) r4x'dx (h) J"l' 0

cosx dx

(b) J:x3 dx (i) J"/6 0

cosx dx

(c) J:rx"-x)dx ( j) J: sinx dx

(d) rdx (k) J" sinx dx 1 x2 •/2

J'x'+ l r dx (e) --dx (I) 1 x' ol +x2

(f) r2dx 2 x3 (m)

rS/3 _.'!"'____ o l+x2

r· 51 2+x' (g) 0

cosxdx (n) --dx o l+x2

4. Prove that, for n #- -1, a;?; 0, b ~ 0,

Jb bn+1-an+l xndx = .

a n+l

5.4 The Indefinite Integral. In applying Theorem 5, we have seen that the first question to be faced is to find a function F(x) whose derivative is the integrandf(x). For example, in evaluating the integral

s:6x5 dx

we look for a function F(x) whose derivative is 6x5. The funct.ion x• will serve as F(x), but so will x6 + 3, x•- 5, and in general x6 + c, where c is any constant. Are there any other functions that will serve as F(x)1 The answer is no, as the next two theorems show.

THEOREM 6. Let H(x) be a differentiable function whose derivative H'(x) = 0 on an interval et. :<; x :<; {3. Then H(x) = c on the interval, where c is some constant.

Page 115: Calculus an Introductory Approach

SEC. 4 THE FUNDAMENTAL THEOREM 107 ·---

Intuitively this seems plausible, because H'(x) ~ 0 implies that the slope of the curve y ~ H(x) is zero, so that the curve has everywhere a horizontal tangent line. This suggests that y ~ H(x) is a straight line parallel to the x-axis, i.e., I:l(x) ~ c.

PROOF. By the mean value theorem of§ 4.7, withj(x) and F(x) now replaced by I:l'(x) and H(x), we see that given any two values x ~ a and x ~ b in the interval a ;;; x ;;; {3, then for some value k between a and b

H(b)-H(a) ~ (b -a)H'(k) ~ 0, or I:l(b) ~ H(a).

Thus H(x) has the same value for all numbers x in the interval a ;;; x ;;; f3, so that H(x) ~ c.

THEOREM 7. Let F(x) and G(x) be differentiable functions with equal derivatives on an interval a ;;; x ;;; b. Then F(x) ~ G(x) + c on the interval, where c is some constant.

PROOF. Write H(x) for F(x)- G(x), so that by Theorem 4 of § 4.3

H'(x) = F'(x)- G'(x) = 0.

Hence we can apply Theorem 6 to get

H(x) = c, F(x)- G(x) ~ c, F(x) ~ G(x) +c.

Having proved Theorem 7, we see that it justifies our remark just before Theorem 6 tbat x6 + c is the most general function whose derivative is 6x5. We call x 6 +c the antiderivative or the indefinite integral of 6x5, and we write

J 6x5dx ~ x6+c.

More generally, if the derivative of F(x) is f(x), we write

J f (x) dx = F(x) +c,

and this is called the indefinite integral or the antiderivative of j(x).

ExAMPLE. Evaluate the indefinite integral

J(x4+x2) dx.

Page 116: Calculus an Introductory Approach

i I

108 CALCULUS: AN INTRODUCTORY APPROACH CHAP. 5

Solution. Since we know that

d -(x5) = 5x4 and dx

d -(x3) dx

3x2,

we see that

I I I (x4+x2)dx = -x5+ -x3+c.

5 3

Problems

Evaluate the following indefinite integrals.

I xdx I dx I. 7. I +x2

I 3x2 dx I dx 2. 8. 2+2x2

3. I (4x-6x2) dx 9. I si:hxdx

4. rx 10. I cosxdx x2

5. rx ll. I sin2xdx xs

6. rx I2. I cos3xdx x4

I3. Prove that the most general function satisfying dyfdx = x is y = tx2 + c, where c is an arbitrary constant. Suggestion: Use Theorem 7 of § 5.4.

14. Find the most general solution of the equations:

(a) dy - = x3 dx

dy (b) - = 2 cosx.

dx

*I5. Let F(x) and G(x) be functions having the same derivative, say J(x), on the interval a ~ x ~ b. Let J(x) be piecewise monotonic on this interval. Give two proofs that F(b)-F(a) = G(b)-G(a), one by using Theorem 5 of § 5.3, the other by using Theorem 7 of § 5.4.

5.5 Inequalities for integrals. The following results are needed for subsequent work with integrals.

Page 117: Calculus an Introductory Approach

SEC. 5 THE FUNDAMENTAL THEOREM 109

THEOI<EM 8. If two integrable functions F(x) and G(x) satisfy the inequality F(x) :;;; G(x) for all values of x on an interval a :;;; x :;;; b, then I: F(x) dx :;;; I: G(x) dx.

From a geometric standpoint this theorem has a simple interpreta­tion. As shown in Figure 5.2 it asserts that the are~ between y ~ G(x) and the x-axis is at least as large as the area between F(x) and the x-axis, both taken from x ~ a to x ~ b.

y-axis _ ~y=G(x)

~:

~y=F(x) lx-a I I - x=b I 1 1 x-axis

FIG. 5.2

PROOF. Define the function f(x) as the difference F(x)- G(x), so that f(x) :;;; 0 for all x in the interval a :;;; x :;;; b. Then by equation (8) of § 5.3, we see that if we write

then

n

an ~ 2 (XJ-XH)f(gj), j~l

Ib f(x) dx ~ lim an. n-)oo

a

Furthermore since x;- Xf-1 ;;; 0 and f(g!) :;;; 0 we see that an :;;; 0, aud so by Theorem 15 of § 2.4

I:f(x) dx :;;; 0.

Butf(x) ~ F(x)- G(x), and we apply Theorem 3 of§ 3.5 to get

I: [F(x)- G(x)] dx ~ I:F(x)dx- J: G(x) dx:;;; 0,

Page 118: Calculus an Introductory Approach

110 CALCULUS: AN INTRODUCTORY APPROACH CHAP. 5

and

s:F(x) dx ;;; s: G(x) dx.

THEOREM 9. Let f (x) be an integrable function on the interval a ;;; x ;;; b. Furthermme, let m and M be constants such that m ;;; j(x) ;<; .M over the interval. Then

m(b-a) ;;; s:j(x) dx ;<; M(b-a).

This result has a simple geometric interpretation, illustrated in Figure 5.3, that the area bounded by the curve y = j(x) and the x-axis, from x = a to x = b, lies between the areas of the two rectangles ABDC and ABFE.

y-axis

x -axis

y=M E,------------.F

A y=O B

Fw. 5.3

y= f(x)

PROOF. First we apply Theorem 8 with F(x) replaced by m, and G(x) replaced by j(x) to get

J:mdx;;; J>(x)dx, i.e. m(b-a);;; J>(x)dx.

Next we apply Theorem 8 with F(x) replaced by j(x) and G(x) replaced by M to get

J>(x) dx;;; J: M dx, i.e. I:j(x) dx ;<; M(b-a).

Problems

1. Prove that 2(b3-a3) ;;; 3(b'-a') for any numbers a and b satisfying b ;;; a ;;; 1. Suggestion: Apply Theorem 8 with F(x) = x, G(x) = x'.

2. Prove that 3(b4-a') ;;; 4(b3-a3) for any numbers a and b satisfying b ;;; a ;;; 1.

Page 119: Calculus an Introductory Approach

SEC. 5 THE FUNDAMENTAL THEOREM Ill

3. Prove that b' -a' ;,; 3(b2 -a2) for any numbers a and b satisfying b ;,; a ;,; l.

4. Compute b- a and the best possible values of m and M as in Theorem 9 for the following integrals.

rdx fn/3 (a) (e) sinxdx

1 X n/6 r dx r3 (b) (f) tanxdx al+x n/6

(c) rdx 2 x2

(g) r~ 0

cosxdx

(d) L x'dx (h) fn/2

0

(sinx+cosx) dx

Page 120: Calculus an Introductory Approach

CHAPTER 6

THE TRIGONOMETRIC FUNCTIONS

6.0. The purpose of this chapter is to obtain infinite series expansions for the functions sin x and cos x. To avoid a protracted discussion of the general theory of infinite series, we get at these expansions by a procedure involving inequalities. By our approach we gain brevity, but this is achieved by confining attention to a limited class of functions. As an application of the results, the computation of trigonometric tables is discussed. '

6.1 Two Important Limits. As a preliminary, we first want to establish that for any positive integer n and any positive number {3,

( 1)

This is a consequence of the binomial expansion, according to which

n(n-1) n(n-1)(n-2) (1 + {3)n = 1 +nf3 + f32 + f33+ ... + {3n.

2 6

Since every term here is positive, the inequality (1) follows if we simply discard all terms except nf3.

THEOREM. l. Ler r be any real number strictly between - 1 and + 1, that is - 1 < r < l. Then

limrn = 0. n->ro

PROOF. Consider first the case where r is positive. For any positive number <, we mnst prove that rn < < for all but a finite number of integers n. Since r < 1, there is a positive number f3 such that r = 1/(1 + {3). In fact, by solving this equation we see that

1-r /3=-,

r

112

Page 121: Calculus an Introductory Approach

Slic. l THE TRIGONOMETRIC FUNCTIONS 113

and this is clearly positive. Then from (l) we see that

l l (2) rn = (l +,B)n < n,B'

by using Theorem 6 of§ 2.1. Now for any given E the relation (1/n,B) < E

holds for all but a finite number of positive integers n; in fact it holds for all integers n exceeding l/(,8<). So by (2) we see that rn < E holds for all but a finite number of positive integers n.

Passing over the case r = 0, suppose now that r < 0. The same proof as above holds with r replaced by lrl, so that lr•l < E holds for all but a finite number of positive integers.

The next theorem contains the notation n!, read "n factorial", which denotes for any positive integer n the product of all positive integers from l to n, thus

n! = n(n-l)(n-2) ... 3 · 2 · l.

THEOREM 2. For any constant c,

en lim- = 0.

n->oo n!

PROOF. As in Theorem l we prove this for c > 0, and the result then follows in case c < 0 by the definition of limit of a sequence in § 2.3. Define le as the least integer satisfying le > 2c, so that le is fixed. Writing k +m for n we see that

ck+m c c c c

n! (k+m)! = k!. k+l.lc+2. k+3 . .... lc+m'

Now c"fk! is a constant, say O,and each of the other factorscj(lc+l), cf(k+2), ... , cj(k+m) is less than}. Hence we have

en (1)m - < 0- . n! 2

As n tends to infinity, so does m, and so we see that

limGf = o,

by applying Theorem l of the present section, Theorem ll of § 2.4, and Theorem 9 of § 2.3.

Page 122: Calculus an Introductory Approach

114 CALCULUS: AN INTRODUCTORY APPROACH CHAP. 6

Finally we prove a result about subsequences: a subsequence of a sequence a1, az, as, ... is any partial set of the terms, as for example a4, as, a12, a16, ....

THEOREM 3. If an infinite sequence {an} tends to a limit a, so does any infinite subsequence.

PROOF. Let <be any given positive number. By definition, all except a finite number of the inequalities

-E<an-a<E, n=l,2,3, ... ,

hold. It follows a fortiori that these inequalities are satisfied by all but a finite number of terms of the subsequence.

Problems

1. Prove inequality (l) by using Theorem 8 of § 5.5 with F(x) ~ l, G(x) ~ (l +x)n-1, a ~ 0, b ~ (3.

2. Find the following limits: 2n 5n6n+l

(a) lim--- (d) lim---n-•00 (n+l)! n~oo n!

n!+3n (b) lim--- (e) lim 8(4/5)n

n->oo n! n-)oo

n! n! (c) lim--- (f) lim--

n-•ro(n+l)! n-•ro (2n)!

6.2 Infinite Series Expansions for sin x and cos x. We begin with the inequality cos x ;;; l, which holds for all values of x. By applying Theorem 8 of § 5.5, with F(x) and G(x) replaced by cos x and l, we conclude that

(3) J: cos x dx ;;; J: l dx

for any positive number t. Since the derivative of sin x is cos x, and since sin 0 ~ 0, the integral on the left side of (3) equals sin t, and so (3) states that sin t ;;; t. This holds for any positive value of t, so we can apply Theorem 8 of § 5.5 with F(x) and G(x) replaced by sin x and x; thus

J: sinxdx;;; J:xdx

Page 123: Calculus an Introductory Approach

SEC. 2 .':fH_E_• _TRIGONOMETRIC FUNCTIONS: ____ __:1=15

for any positive number t. Since the derivative of cos x is -sin x, the evaluation of these integrals gives

t' 1-cost:::; ~ or - 2

t' 1-- ~cost. 2 -

Consequently we can apply Theorem 8 of § 5.5 with F(x) and G(x) replaced by 1- x'/2 and cos x, to get

f}- :}x ;2

Integrating, we find that

t3 t--~sint

3!-

for any positive number t. The repetition of this process leads to a series of inequalities which are valid for any positive number x:

cosx ~ 1 sinx ;;:; X

x' x3 cosx;;; 1-- sin x f; X--

2! 3!

x' x4 x3 x5 (4) cosx ~ 1--+- sinx ~ X--+-

2! 4! 3! 5!

x2 x4 x6 x3 x5 x? cosx ~ 1--+--- sinx ~ X--+---

2! 4! 6! 3! 5! 7!

etc.

Let us look first at the inequalities for cos x. We define a1, a 2, aa, ... by the equations

and in general

x' x4 1--+-

2! 4!'

x2 x4 x6 xs 1--+---+-

2! 4! 6! 8!'

x2 x4 x6 xS xlO x12 1--+---+---+-

2! 4! 6! 8! 10! 12!'

x2 x4 x6 x4n-6 x4n-4 1--+---+ ... - + .

2! 4! 6! (4n-6)! (4n-4)!

Page 124: Calculus an Introductory Approach

116 CALCULUS: AN INTRODUCTORY APPROACH CHAP. 6

Similarly we define br, b,, ba, ... by the equations

br = x'

1--2!'

b, = x2 x4 xB

1--+---2! 4! 6!'

x2 x4 x6 xB xlO ba = 1--+---+---

2! 4! 6! 8! 10!'

x2 x4 x6 xB xlO xl2 x14 1--+---+---+---

2! 4! 6! 8! 10! 12! 14!'

and in general x2 x4 x6 x4n-4 x4n-2

1--+---+ ... + ----. 2! 4! 6! (4n-4)! (4n-2)!

With these definitions we see that the inequalities on cos x can be written in the following form:

[ cosx ~ a1 cosx ~ bl

(5) cosx ~ a2 cos x 'f; bz

cosx ~ aa cosx ~ bs

COSX ~ a4 cosx ~ b4

etc. Let us think of x as a fixed positive number. Then we can apply

Theorem 14 of § 2.4, because x4n-2 x4n-2

an-bn = , lim(an-bn) = lim = 0 (4n-2)! n-•oo (4n-2)!

by Theorems 2 and 3 of § 6.1. The conclusion is that an -+cos x and bn ->cos x, so that cos x is expressible as an infinite series

x2 x4 x6 xB ( 6) cosx = 1--+---+-- ....

2! 4! 6! 8!

A similar argument can be applied to those inequalities ( 4) involving sin x, to give the result

x3 x5 x7 x9 (7) sinx = x-- +---+--

3! 5! 7! 9!

Equations (6) and (7) have been established only for positive values of x. However, it is easy to see that they must be valid for all values of x, positive, negative, or zero. Let us begin with (6). If x = 0, equation

Page 125: Calculus an Introductory Approach

SEC. 2 THE TRIGONOMETRIC FUNCTIONS 117

(6) states that cos 0 ~ l, which is correct. For negative values of x, we recall from trigonometry that cos(- x) ~ cos x, and we note that the right side of equation ( 6) has this same property that it is unchanged when x is replaced by - x, because

( -x)2 ~ x2, ( -x)4 ~ x•, ( -x)B ~ xB, etc.

Hence equation (6) holds for every real number x. As for equation (7), it becomes the well-known result sin 0 ~ 0 when

xis replaced by zero. For negative x, we know that sin(- x) ~ -sin x, and the right side of equation (7) also changes sign when x is replaced by - x, because

( -x)" ~ -x", ( -x)5 ~ -x5, ( -x)' ~ -x', etc.

Equations (6) and (7) are thus seen to be valid for all real numbers x, where of course x must be taken in radian measure. The reason for this is that all our formulas and theorems about the differentiation and integration of sin x and cos x were based on radian measure. However, for an angle measured in degrees, equations (6) and (7) can be applied by first converting the angle into radian measure.

ExAMPLE. Evaluate cos ·} with accuracy to five decimal places. Solution. Using equation (6) we see that

l l (1)2 l (1)4 cos-~ 1--- +-- - ... ;;: .98007,

5 2 5 4! 5

where the symbol ;;: stands for "is approximately equal to". The approximate value .98007 of cos!; has been obtained here by

using only the first three terms of the series. How can we be certain that . 98007 is accurate to five decimal places 1 Might it not be that the rest of the terms in the infinite series have influence in determining the first five decimal places! The answer to this question is no, as the follow­ing analysis will demonstrate.

When an approximation a is used in place of a number rx, we shall say that the error is la- rxj. For example, if the number is 22.9 and the approximation is 23, the error is .l. Similarly, if the number is 23.1 and the approximation is 23, the error is .l. Thus for our purposes we are making no distinction between approximations larger than and those smaller than the number.

THEOREM 4. If the first k terms of the series (6) are used to compute an wpproximation to cos x, then the error is at most the absolute value of the le+ 1-st term. An analogous result holds for equation (7).

Page 126: Calculus an Introductory Approach

ll8 CALCULUS: AN INTRODUCTORY APPROACH CHAP. 6

Before proving this in general, let ns begin with the example :;hove. In view of the inequalities (5), we can conclude that

1-~(~)\ __1:_(~)4

- __1:_(~)6

;o; cos~ ;o; 1- ~(~)\ __1:_(~)4

2 5 4! 5 6! 5 5 2 5 4! 5 ,

or, in abbreviated notation, bz ;;;; cos i ;£ a2. Thus cos llies be­tween b2 and a2, two numbers whose difference is

(8) 1 (1)6 - - ;;;: .00000009. 6! 5

Now if a2 is used as an approximation to cost, the error is az- cos}, and this is at most a 2 - b2. That is, the error is at most the number (8) above. Thus we are certain that the approximation .98007 to cost, obtained by using the first five decimal places of az, is accurate to five decimal places.

The proof of Theorem 4 in general is virtually the same. It is based on the simple idea that if cos x lies between two approximations a and A, the error in using either approximation is at m~st lA -4 We note that in view of (4), cos x lies between the sum of the first k terms of (6), namely,

x2 x4 x6 (9) 1--+---+ ...

2! 4! 6!

x2k-2 + ( -J)k-l-:-:----c-:­

(2k-2)!'

and the sum of the first k + 1 terms, namely,

x2 x4 x6 x2k-2 x2k 1--+---+ ... +(-1)>-1 +(-1)'~-.

2! 4! 6! (2k- 2)! (2k)!

Hence the error in using (9) as an approximation to cos xis at most the absolute value of the difference between these sums, namely

1(-l)k(::!l = l(::J This is the absolute value of the le+ 1-st term of the series (6), and so Theorem 4 is proved.

Problems

l. Compute sin i to 5 decimal places of accuracy. 2. Compute sin 2° to 5 decimal places of accuracy. Suggestion: The angle

2° becomes 7T/90 when converted to radians. Use 7T = 3.14159. 3. Compute sin 87° to 5 decimal places of accuracy. Suggestion: sin 87°

= cos 3°.

-~

Page 127: Calculus an Introductory Approach

SEC. 3 THE TRIGONOMETRIC FUNCTIONS 119

4. Compute cos 86° to 5 decimal places of accuracy. 5. Presuming that it is correct to differentiate an infinite series term by

term (which it is in this case) differentiate equation (7) to get (6). 6. Verify that when (6) is differentiated term by term the result is (7)

with all signs changed. 7. Replace x by 2x in every term in (7) to get an infinite series expansion

for sin 2x. Then verify as far as powers up to x7 that this is the same as 2 sin x cos x by multiplying equations (6) and (7).

6.3 Remarks on Infinite Series. The series expansions of cos x and sin x in (6) and (7) are special cases of what are called Taylor series. These are series which give an alternative form of a function f(x),

(10) f(x) = ko+k,x+k,x2 +lc3x3 + ... ,

in the same sense that 1 + 2x + x' is another form of the function (1 +x)2• Only a restricted class offunctionsf(x) can be expanded into a series of the form (10). We do not discuss here the restrictions on f(x), nor do we deal with the question of how the coefficients Ieo, le,, k,, ... can be determined from f(x).

We shall get the Taylor series expansions for several functions, but as in the cases of cos x and sin x, we shall use special methods to obtain the series. From the manner in which we get the series, there will be no need to discuss fully what is called "convergence" and "divergence" of series. But we shall make a few remarks about this important topic.

To begin with a couple of examples, the series

.3 + .03 + .003 + .0003 + ... is convergent. It converges to the limit }. On the other hand the series

1 + 1 + 1 + 1 + 1 + 1 + ... is not convergent; it is divergent. In general, the sum of an infinite series of terms

(ll)

is said to be convergent if the sequence of "partial sums" 81, 8 2, 83, 84, .••

tends to a limit, wbere the partial sums are defined by the equations

81 = a1, 82 = a1 +a2, 83 = a1 +a2+a3, n

... ,sn = L a1. j=l

Each of the terms a1, az, a3, ... may be positive, negative, or zero. If the partial sums s1, s,, ss, ... tend to a limit s, then the series (ll) is said to converge to s.

Page 128: Calculus an Introductory Approach

120 CALCULUS: AN INTRODUCTORI' APPROACH CHAP. 6

In § 6.2 we used the theory of limits as developed in Chapter 2 in establishing equations (6) and (7). The approach that we took assures us that the series in (6) and (7) are convergent for every value of x, positive, negative or zero.

Page 129: Calculus an Introductory Approach

CHAPTER 7

THE LOGARITHMIC AND EXPONENTIAL FUNCTIONS

7.0. What was done in the preceding chapter for trigonometric functions is now done for logarithmic and exponential functions, specifically for y = log x and y = e". The number e is a fundamental mathematical constant whose nature and properties will be revealed in the subsequent discussion.

It might have been noted that in our treatment of differentiation in Chapter 4, the derivatives of various trigonometric functions were obtained, but no derivatives of exponential or logarithmic functions. Such derivatives are more difficult, and in order to get at them, we shall start by defining the logarithmic function in terms of an integral. But because the reader is already familiar with log x defined in another way, and because it is not apparent that these definitions amount to the same thing, we shall call our function L(x) at first. Later in the discussion it will become clear that L(x) is the logarithmic function, and then we shall change its label from L(x) to log x.

From a strict logical standpoint, the work of Appendix A should precede the present chapter, and the reader can take them in that order if he wishes. We have placed the existence proof in an appendix in order to break the theoretical development into smaller pieces, and to postpone the more difficult part of the theory.

7.1 A Function Defined. The equation

fb x2dx = ~- ~ 1 3 3

is quite familiar; it is a special case of Theorem l of § 3.3. If we regard b as a variable quantity we see that the integral is a function of b,

121

Page 130: Calculus an Introductory Approach

122 CALCULUS: AN INTRODUCTORY APPROACH CHAP. 7

namely (b3- I )/3. To change from b to the more standard variable x we can write

(I) fx x3 1

1v2dv =a- 3.

This integral, then, is just another way of writing (x3 -1)/3. Geometri­cally, the integral signifies the area under the curve y = v2, taken from v = I to v = x as shown in Figure 7 .I. The difference between this and previous areas is that we have no definite boundary on the right, since x can be any value we want to specify. It is understandable, then, that the integral (I) is a function of x, but not a function of v because v is merely a dummy variable of integration.

Now analogous to (I) we define a function of x, denoted by L(x), by the equation

(2) fx dv f" L(x) = - = v-1 dv, 1 V 1

for any x > 0.

This can be conceived of geometrically* as the area between the curve y ljv and the v axis, taken from v = I to v = x. The restriction x > 0 is imposed in equation (2), and the graph in Figure 7.2 suggests

v-axis v -axis

FIG. 7.1 Fw. 7.2

* The theorems on the existence of the integral Jb f(x) dx depended on our know­ledge of a function F(x) whose derivative is f(x). In °the present case the variable of integration is v, andf(v) = v-1 • But we have no knowledge of the existence of a function F(v) whose derivative, with respect to v, is v-t. Thus while we can conceive of some meaning for the integral {2) from a geometric picture, we are not certain of the existence of this integral because a careful development of calculus should not be based on geometric intuition. To avoid an unduly complicated analysis at this point, we postpone the proof of the existence of such integrals as (2) until Appendix A.

Page 131: Calculus an Introductory Approach

SEc. I THE LOGARITHMIC AND EXPONENTIAL FUNCTIONS 123

why. As v tends to 0 through positive values, the y values increase beyond all bound, so that y has no value when v = 0. Thus the limits of integration must not include v = 0, nor cross v = 0. Figure 7.2 illustrates the case x > I, for which the functiou L(x) has positive values. We note that

(3) fl dv

L(l)= -=0, 1 V

and L(x) is negative for any x in the range 0 < x < I. These conclu­sions follow from the general definition of an integral in § 3.2.

THEOREM I. The derivative of L(x) is Ijx, i.e., L'(x) = ljx.

PROOF. By the definition of derivative

, . L(x+!lx)-L(x) . I (JxMxdv J"dv) L (x) = hm = hm - - - - .

L\X->0 D._x L\X->0 _6.x l V l V

From equation (5) of§ 3.2 we see that the integrals can be combined, thus

[ ] JXMX dv] L'(x) = lim - - .

L\X--o>O ~X X V

First let us consider the limit as !lx tends to zero through positive values. In this case we see that the integrand has maximum value ljx on the range of integration, and minimum value Ij(x+ !lx). Hence we can apply Theorem 9 of § 5.5. with a and b replaced by x and x+!lx, and we can take m = Ij(x+!lx) and M = Ijx. Hence we get, since b -a is replaced by !lx,

(4) ~ :$ fx+L\x dv ~ Llx. X+Llx - x V - X

Dividing through by !lx we get

( 5) __ 1_ < _1_ JX+L\X dv < ~ x+!lx = !lx x v = x'

As !lx tends to zero we see that the center term is caught between ljx and a fraction which is tending to Ijx. It follows that L'(x) = Ijx.

But what if !lx tends to zero through negative values? In this case we apply Theorem 9 of § 5.5 to the integral

fx dv with m = :,

x+.6.x V X

I M = ---, b-a = -!lx.

x+!lx

Page 132: Calculus an Introductory Approach

124 CALCULUS: AN INTRODUCTORY APPROACH CHAP. 7

Thus we get, in place of (4),

-t!.x ~-;:;; f

x dv - <

x+6.x V =

-t!.x

X

We multiply by the positive quantity -1/t!.x, and interchange the upper and lower limits on the integral; cf. Theorem 2 of § 2.1 and equa­tion (6) of § 3.2. This gives

~ :::; _1_ fx+6.x dv :::::; l ' x - t!.x x v - x+t!.x

which is (5) with the inequalities reversed. As t!.x tends to zero it fol­lows again that L'(x) ~ 1/x.

From Theorem 1 we can easily get the fundamental property of the function L(x), as follows.

THEOREM 2. For any positive numbers a and x,

L(ax) ~ L(a)+L(x).

The reader will notice that this resembles one of the basic properties of logarithms, namely log ab ~ log a+ log b. This resemblance does not entitle us, however, to say that the function L(x) we are studying must be the logarithmic function. This identification will come later in the analysis.

PROOF. We will regard a as a constant, and we consider the deriva­tive of L(ax) with respect to x. We write u for ax and apply Theorem 11 of § 4.5 to y ~ L(ax), thus

y = L(u), u = ax, dy 1 du - ~ L'(u) ~ -, dx ~ a, du u

d dy dy du 1 l 1 -{L(ax)} ~ - ~ - ·- ~-·a ~-·a~-. dx dx du dx u ax x

Thus we see that L(x) and L(ax) have the same derivative, namely 1Jx, and so by Theorem 7 of § 5.4 we conclude that L(x) and L(ax) differ by a constant,

(6) L(ax) ~ L(x) +c.

To determine the value of the constant, we set x ~ 1 and use equation (3) to get

L(a) ~ L(1)+c ~ O+c ~c.

Thus c ~ L(a), and substituting this in (6), we have proved Theorem 2.

Page 133: Calculus an Introductory Approach

SEc. 2 THE LOGARITHMIC AND EXPONENTIAL FUNCTIONS 125

Problems

l. Express f~ (dx(x) in terms of the L function. 2. Prove that L(a') = 2L(a), L(a') = 3L(a) and L(a4) = 4L(a) for any

. positive number a. 3. Prove that L(}) = -L(2). Suggestion: Substitute a = i and x = 2

in Theorem 2. 4. Prove that L(/r) = -L(3). 5. Prove that L(!) = -L(!). 6. Prove that L(6)+L(lO)+L(l5) = 2L(30). 7. Prove that

8. Combine the sum

into a single iritegral. 9. Corn bine the sum

into a single integral.

J3 dx f' dx f! dx -+ -+ -!X !X !X

7.2 Properties of L(x ). If we replace a and x by any positive numbers b1 and b, in Theorem 2 we get

If we replace a and x by b1b2 and bs, we see that

L(b1bzb,) = L(b1b2) + L(bs) = L(b1) + L(b,) + L(bs).

Next we can replace a and x by b1b2bs and b4 to get

L(b1b2b3b4) = L(b1b2bs) + L(b4) = L(b1) + L(b2) + L(b3 ) +L(b4).

Continuing in this fashion we see that for any positive numbers b1> b2, ba, ... , bn,

(7) L(b1b2bs ... bn) = L(b1) + L(b,) + L(bs) + ... + L(bn)·

If the numbers b1, b,, ... , bn are all equal, say equal to b, we get

(8) L(bn) = L(b)+L(b)+L(b)+ ... +L(b) = nL(b).

Page 134: Calculus an Introductory Approach

126 CALCULUS: AN INTRODUCTORY APPROACH CHAP. 7

For later use it will also be convenient to know that

(9) L(2) < l, L(3) > 1.

To prove the first of these we apply Theorem 9 of § 5.5 to

J2 dv J312 dv f2 dv

L(2) = - = - + - = It +I. (say). 1 V 1 V 3/2 V

With the integrand lfv, we can take M = l in !1 and M = % in I. and so we have

I. ;;; ~ (2- ~) = ~ 3 2 3'

l l 5 L(2) ,; - + - = - < 1.

- 2 3 6

To prove the second part of (9) we do much the same thing, except that we apply the other part of Theorem 9 of § 5.5. It is necessary to break the integral into more parts; thus we take

J3 dv J514 J312 J714 J2 J512 J3

L(3) = - = + + + . + + . 1 V 1 5/4 3/2 7/4 2 5/2

In the application of Theorem 9 of § 5.5 to these six integrals, we can take m to be{', t, t, }, §, and i in the respective cases. Hence we can write

+ ~ ( ~ - z) + ~ (3- ~), 5 2 3 2

l l l l l l L(3) ;, - +- +- +- +- + -.

5 6 7 8 5 6

It follows by a .straightforward calculation that L(3) > 1. From the definition (2) of L(x), we can see that since the integrand

lfv is positive, the value of L(x) increases as x increases, and decreases as x decreases. In other words, L(x) is a monotonic increasing function. As x increases indefinitely, that is to say, as x tends to infinity, what happens to L(x) 1 The answer is that L(x) also increases indefinitely as the following argument shows.

From equation (8) with b = 3, and from (9), we conclude that

L(3n) = nL(3) > n.

Page 135: Calculus an Introductory Approach

SEc. 2 THE LOGARITHMIC AND EXPONENTIAL FUNCTIONS 127

Hence we can make L(3n) as large as we please by taking n sufficiently large. In other words, L(3n) increases indefinitely as n increases in­definitely, and so L(x) tends to infinity with x.

The function L(x) is defined for x > 0. What happens to L(x) as x tends to zero? To answer this, we replace a and x in Theorem 2 by 3n and 3-n, and so by equation (3) we get

L(3n · 3-n) = L(3n)+L(3-n), L(3n · 3-n) = L(1) = 0,

L(3n)+L(3-n) = 0, L(3-n) = -nL(3).

Thus as n increases indefinitely, L(3-n) tends to infinity, L(3-n) tends to minus infinity. But as n tends to infinity, (3-n) tends to zero. Thus we can write*

lim L(3-n) = - oo, n~oo

lim L(x) = - oo. x->0

decreases indefinitely: as n

y-axis

With this information about L(x) we can make an approximate sketch of its graph. The function has a derivative, and so by Theorem 1 of§ 4.1 it is continuous. The graph of the function is shown in Figure 7.3.

FIG. 7.3 Sketch of y = L(x) or y = log x.

Problems

l. ~rove that L(3) ;<; t. Suggestion: Separate J~ into the parts s: and J2 •

2. Prove that L(4) ;'i 1 + !+i. 3. Prove that

4. Prove that

1 L(n) ;<; 1+!+}+ ... +--.

n-1

1 l L(n);;; }+}+ ... +- +-.

n-1 n

* In writing lim L(x) = - ro, our notation is somewhat at variance with our lan-x->o

gunge about limits. If we ask the question, "Does the function L(x) have a limit as X

tends to zero?", the answer is "No". The reason is that in saying that a sequence or a function has a limit we mean a finite limit.

Page 136: Calculus an Introductory Approach

128 CALCULUS: AN INTRODUCTORY APPROACH CHAP. 7

5. Prove that L(i) > -l. 6. Prove that L(!;) < -l. 7. Prove that L(%) ;;; l. 8. Prove that L(cjb) = L(c)-L(b) for positive numbers c and b. 9. Prove that Lh/c) = tL(c) for c > 0. 10. Prove that L(f/c) = !;L(c) for c > 0. ll. Prove that L(\)'c) = (1/n)L(c) for any positive integer n, and c > 0.

7.3 The Exponential Function. Since the function y = L(x) is monotonic increasing and differentiable, withL'(x) # 0 on the interval 0 < x < oo, we can apply Theorem 12 of § 4.6, and conclude that the inverse function is also differentiable. Denoting the inverse function by x = E(y), we see by the same theorem that

L'(x) · E'(y) = 1, E'(y) = x,

since L'(x) = 1jx. Now x = E(y) so we see that E'(y) = E(y). Writing this with the usual variable x, we conclude that

d (10) dx E(x) = E'(x) = E(x).,

The graph of the equation y = E(x) can be obtained at once from the graph of y = L(x) in Figure 7.3 by interchanging the roles of x and y,

Fm. 7.4 Sketch of y = E(x) or y = ez.

and hence

as shown in Figure 7.4. Note that the domain of the function y = E(x) is - oo < x < oo, and that E(x) > 0 for every value of x.

The fundamental property L(ax) = L(a)+L(x) of the function L(x) entails a similar basic result for the inverse function. Let us write b for L(a), y for L(x), and w for L(ax), thus

b = L(a), y = L(x),

w = L(ax), w = b+y.

In terms of the inverse function we see that

a = E(b), x = E(y),

ax = E(w), ax = E(b + y),

(ll) E(b) ·E(y) = E(b+y).

Page 137: Calculus an Introductory Approach

SEc. 3 THE LOGARITHMIC AND EXPONENTIAL FUNCTIONS I29

This basic property of the E function is now used to analyze the function. If we set y ~ 0 we get E(b) · E(O) ~ E(b), and since E(b) is positive whatever the value of b, we conclude that

(I2) E(O) ~ I.

If we set b ~ I and y ~ I in (ll) we get E(I) · E(I) ~ E(2) or E(2) ~ {E(I)}2• Then if we set b ~ 2 and y ~ I we get

E(3) ~ E(2) · E(I), whence E(3) ~ {E(I)}".

Similarly, if we set b ~ 3 and y ~ I we get E(4) ~ {E(IJ)4, and a continuation of this process leads to

(I3) E(n) ~ {E(I)}"

for any positive integer n. For any positive rational number nfr let us form the product with r factors,

E(njr)E(nfr)E(njr) ... E(njr) ~ {E(njr))'.

Applying (ll) repeatedly we see that this product equals E(n), and so from (I3) we conclude that

(I4) {E(n/r))' ~ E(n) ~ {E(IJ}n, E(njr) ~ {E(I))nlr.

Next, if we replace b by njr and y by -njr in (ll) we get, by use of (I2) and (I4),

(I5) E(njr)E( -njr) ~ I, E( -nfr) ~ {E(I))-nlr.

It may be noted that E(I) plays a central role in these equations, and so to simplify the notation we replace it by e. Thus (I3), (I4), and (I5) can be written

E(n) ~ en, E(njr) ~ enlr, E( -njr) ~ e-nlr.

Hence we can conclude that

(I6) E(x) ~ ex

for all rational values of x. What about the irrational values of x1 From our approach we know

that E(x) is defined and has meaning for irrational x, but how is e" defined 1 Let us fix our ideas on one specific irrational number a, and let us say that

is a sequence of rational numbers having (1, as a limit. For example if

" were v2 we could take the sequence to be

I, I.4, l.4I, I.4I4, l.4I42, ....

Page 138: Calculus an Introductory Approach

130 CALCULUS: AN INTRODUCTORY APPROACH CHAP. 7

The natural thing to do is to define e" as the limit of the sequence

(17)

But does this sequence have a limit! Our approach assures us that it does. For the sequence (17) can be written, because of (16), as

(18) E(a,),E(a,),E(as),E(a4), ....

Furthermore, since the function E(x) is differen,...iable, it is continuous by Theorem I of§ 4.2, and consequently the limit of (18) exists and is equal toE( a). Hence the sequence (17) has the same limit, and we have proved that equation ( 16) holds for all values of x, rational and irra­tional.

This means that E(x) is an exponential function ex, where e is some constant. We will determine the precise value of e in the next section. For the moment we recall that e is E(1), and so

E(l) ~ e, L(e) ~ I.

Now we proved that L(2) < 1 and L(3) > 1 in equation (9) of § 7.2, and since L(x) is a monotonic increasing funqtion, it follows that e lies between 2 and 3.

Furthermore since E(x) is an exponential function, the inverse function must be a logarithmic function. Specifically, the inverse of the function

y ~ E(x) or y ~ ex

can be written in either form.

x ~ L(y) or x ~ log,y.

Logarithms to base e are called natural logarithms, and throughout this book these logarithms are written without any base being shown. Thus log y means Iog,y. Logarithms to base 10 are called common logarithms, and we write such logarithms always with the base showing, e.g., log1oY·

In view of all this we can rewrite Theorem I with log x in place of L(x):

THEOREM 3. Jj y ~ log X, then dyjdx ~ 1jx.

Also, the basic property in Theorem 2 can now be written in the form

(19) logax ~ Ioga+logx

for positive numbers a and x, and similarly (11) can be written in the form (20) eb · eY = eb+v.

Page 139: Calculus an Introductory Approach

SEc. 4 THE LOGARITHMIC AND EXPONENTIAL FUNCTIONS I3I

Finally we state equation (10) as a theorem, with e" in place of E(x).

THEOREM 4. Jf y = ex, then dyjdx = ex.

Problems

I. Write E(l), E(2) and E(3) in terms of e. 2. Evaluate log e, log e' and log e3. 3. Find the derivatives of

(a) y = log2x

(b) y = log4x

4. Differentiate

(c) y = log(x+l)

(d) y = log(2x+I)

X y =log--.

x+I

Suggestion: log ajb =log a-log b. 5. Differentiate y = log- x3. 6. Find the derivatives of

(a) y = ex+l

(b) y = e2x

(c) y = xex

(d) y = e-x

7. For what positive value or values of xis y = (1/x)+log x a maximum or minimum1

*8. Prove that the areas bounded by the following two configurations are equal: (a) the curve y = Ijx from (}, 2) to (2, }) and the straight-line segments from (2, }) to (2, 0), from (2, 0) to (}, 0), and from (!1, 0) to (!, 2); (b) the curve y = ljx from (I, I) to (4, tJ and the straight-line segments from (4, !J to (4, 0), from (4, 0) to (I, 0), and from (I, 0) to (I, I).

9. Prove that the tangent line to the curve y = log x at the point (I, 0) crosses the x-axis at an angle of 45°.

10. At what point on the curve y = ex does the tangent line have slope I! I I. Differentiate y =log cos x. Then prove that

J: tan x dx = log cos a -log cos b

for a and b strictly between -rr/2 and rr/2.

7.4 The Series Expansion for ex. By a device very similar to the one used in § 6.2 to get expansions for sin x and cos x, we now derive an infinite series expansion for ex. First we want to differentiate e-x. We write

y = eu, u = -x,

Page 140: Calculus an Introductory Approach

132 CALCULUS: AN INTRODUCTORY APPROACH CHAP. 7

and apply Theorem 11 of § 4.5. From Theorem 4 of the preceding section we get

dy - = eu du '

du dy d -=-I, - = (eu)(-1) =-e-x, -(e-x)= -e-x. dx dx dx

Next we use the fundamental theorem of calculus (from § 5.3) to write

(21) I: e-xdx = [ -e-x]g = -e-b+eO = 1-e-b.

For positive x we know that ex > 1, so that by Theorem 6 of § 2.1,

1 e-x=-< I.

ex

Thus for x f; 0 we see that e-x ;;; 1. We now apply Theorem 8 of § 5.5 with F(x) replaced by e-x, G(x) replaced by 1, a = 0 and b any positive number. We get, by use of (21),

I: e-x dx ;;; I: 1 dx, I- e-b ;;; b, 1- b ;;; e-b.

This holds for any positive value of b, so 1- x ;;; e-x holds for any positive x; it also holds for x = 0. Again we apply Theorem 8 of § 5.5, this time with F(x) replaced by 1-x and G(x) replaced by e-x. We get

Ib (1-x) dx ;;;Ibe-Xdx, b- bZ;;; 1-e-•, e-b;;; 1-b+ b

2.

0 0 2 2

This holds for any b > 0, so e-x ;;; 1-x+x2j2 holds for any positive x; it also holds for x = 0. Once again we apply Theorem 8 of § 5.5, with F(x) replaced by e-x and G(x) replaced by 1-x+x2j2. We get

I: e-x dx;;; I: (1-x+x2j2) dx,

b2 b3 b b2 b3 1-e-b,; b-- +- =---+-

- 2 6 1! 2! 3!'

b b2 b3 1--+---,; e-b.

1! 2! 3! -

Again we switch from b to x and apply Theorem 8 of § 5.5. Iteration

Page 141: Calculus an Introductory Approach

SEc. 4 THE LOGARITHMIC AND EXPONENTIAL FUNCTIONS 133

of this process leads to the following chain of inequalities, valid for any non-negative x,

x x2

1--+~ l! 2!

xx2 x3x4 1--+---+-

1! 2! 3! 4!

etc.

X 1-­

l!

x x2 x3 1--+--­

l! 2! 3!

xx2x3x4x5 1--+---+--­

l! 2! 3! 4! 5!

Thus for any non-negative x we see that e-x is caught between succes­sive partial sums of the series

x x2 x3 x4 x5 x6 x7 (22) 1--+---+---+---+ ....

1! 2! 3! 4! 5! 6! 7!

Denoting the partial sums by sr, sz, sa, 84, ••. we have

X X x2 x x2 x3 s1 = I, sz = 1- ~, sa = 1-- + -,

1! l! 2! 1--+--- ....

1! 2! 3!,

Thus no member of the sequence

(23)

is less than e-x, and on the other hand, no member of the sequence

(24)

is greater than e-x. Let us regard x as a fixed non-negative number for a moment, so

that e-x is fixed. We apply Theorem 14 of § 2.4, with e-x playing the role of c, the sequence (23) in place of the sequence {an}, and the sequence (24) in place of the sequence {bn}· In order to apply Theorem 14 of § 2.4, we must also know that lim(an- bn) = 0, in the notation of that theorem. In the present context the difference an- bn is S2n-1- s,n, and we see that

x2n-1

B2n-1 -Szn = . (2n-1)!

As n tends to infinity this tends to zero, by Theorems 2 and 3 of § 6.1. Thus all the hypotheses of Theorem 14 of § 2.4 are satisfied and we conclude that both sequences (23) and (24) have limit e-x. But the

Page 142: Calculus an Introductory Approach

134 CALCULUS: AN INTRODUCTORY APPROACH CHAP. 7

sequences (23) and (24) are the partial sums of the sequence (22), and so we conclude that

x x2 x3 x4 x5 x6 x? (25) ex= 1--+---+---+---+ ....

1! 2! 3! 4! 5! 6! 7!

This expansion of e-x into a series has been established only for non-negative values of x, that is, for all x ;;; 0. If we replace x by -x in (25) we get the more common form

xx2x3x4x5 (26) ex = 1 + - + - + - + - + - + ....

1! 2! 3! 4! 5!

Since (25) was proved for x ;;; 0, we note that (26) has been established for x ;:;; 0, because x was switched to -x. Formula (26) is a very basic result which holds in fact for all values of x. To prove this, we must establish that (26) is correct for positive values of x, and this we now do.

We begin afresh with the function j(x) defined by

j(x) = 1+xe'-ex.

Here c is any fixed positive number, and the domain of the function is restricted to 0;:;; x;:;; c. Note that j(O) = 0 and that j'(x) = e'-ex We apply Theorem 15 of § 4.7. Since ex is monotonic increasing, we see that f '(x) > 0 for 0 < x < c. Hence j(x) is monotonic increasing for 0 ;:;; x ;:;; c, and it follows that

f(x) ;;; 0, 1 +xe'-ex ;;; 0, 1 ;:;; ex ;:;; 1 +xe'.

We apply Theorem 8 of § 5.5 to get

J: 1 dx ;:;; J: ex dx ;:;; J: (1 +xe') dx

for any number b satisfying 0 ;:;; b ;:;; c. Evaluating these integrals we have

Replacing b by x we see that

xz

b2 o;:b+-e'. - 2

S x+ -ec - 2 for 0 ~ X ~ C,

Page 143: Calculus an Introductory Approach

SEc, 4 THE LOGARITHMIC AND EXPONENTIAL FUNCTIONS 135

and we apply Theorem 8 of § 5.5 again to get

I: x dx ~I: (ex_ 1) dx ~I: (x+ :2 e') dx,

b2 b2 b3

- :o; eb- 1 - b :o; - + - e' for 0 :o b :o c. 2- -2 6 --

Replacing b by x and applying Theorem 8 of § 5.5 again, we obtain

IbxZdX ~ Ib (eX-1-x)dX ~Ib (Xz + XS ec)ax, o2 o o 2 6

b" b2 b3 b4 -:oe•-1-b--:0-+-e' for O~b~c. 6- 2-6 24

Repetition of this process leads to

bn b2 b3 bn-1 bn bn+l -~eb-1-b----- ... ----:::;;-+ ec, n! - 2! 3! (n-1)! - n! (n+ 1)!

valid for any b satisfying 0 ~ b ~ c. By Theorem 2 of§ 6.1 we conclude that

bn lim- = 0, n-HJ) n! lim - + e' = lim- +e'lim---(

bn bn+l ) bn bn+l

n-•oo n! (n+ 1)! n! (n+ 1)!

= O+e' · 0 = 0.

Consequently, as n tends to infinity it follows from Theorem 17 of § 2.4 that

b2 b3 e•-l-b----- ... =0.

2! 3!

Replacing b by x we obtain (26) for any x in the domain 0 ~ x ~ c. But c is any positive number, and so (26) has been established for all positive values of x. Since (26) had already been proved for x ~ 0, we have now completely proved this formula for all values of x.

We are now in a position to compute the value of e, by setting x = 1 in (26). This gives

1 1 l l 1 e = 1 + - + - + - + - + - + ...

l! 2! 3! 4! 5! ,

e = 2.71828 ....

Page 144: Calculus an Introductory Approach

136 CALCULUS: AN INTRODUCTORY APPROACH CHAP. 7

Problems

l. Check the arithmetic in the above computation of e to five decimal places.

2. At the opening of this section we found the derivative of y = e-x. Verify the result by writing the function in the form y ~ ljex, and using Theorem 6 of § 4.3.

3. Presuming that it is correct to differentiate an infinite series term by term (as it is under certain conditions) differentiate (26) term by term and verify Theorem 4 of § 7 .3.

4. Compute the value of ye to three decimal places by substituting the appropriate value of x in (26). Verify your answer by taking the square root of 2.71828.

*5. What function j(x) has the properties f '(x) ~ j(x) and j(O) ~ 41 6. Presuming that it is correct to multiply two infinite series term by

term, multiply equations (25) and (26), computing all powers up to x5 in the product.

7.5 The Number e. We now develop a well-known expression fore as a limit, namely formula (27) below. We start with Theorem 3 of§ 7.3, which can be written in the form

. log(x + 6.x) -log x I hm ~ -.

t:.x->0 b..x x

Taking x ~ I, and using log l ~ 0, we get

lim [-1-log(l + 6.x)] ~ l.

t..x->0 Llx

We let 6.x tend to zero through the positive values l, !, !, !, ... so that the above limit becomes the limit of a sequence. Writing lfn for 6.x, we see that Llx --* 0 implies n -+ eo, and so we have

lim [n log(!+~)] ~ l. n->oo n

To the left member of this equation we apply equation (8) of § 7.2 to write n log(!+ lfn) in the form log(!+ lfn)n, and thus we get

lim log( I+ lfn)n ~ l. n~oo

For convenience at this point, let us write an for log(!+ lfn)n, so that we have a sequence a1, a 2, a3, a4, ... with limit 1.

Now as we noted in our discussion following (18) in§ 7.3, the function E(x) or e" is continuous because it is differentiable. Consequently, by

Page 145: Calculus an Introductory Approach

SEc. 5 THE LOGARITHMIC AND EXPONENTIAL FUNCTIONS !37

the definition of a continuous function given in § 2.9, we conclude that the sequence ·

has limit e', or e. But since E(x) and L(x) are inverse functions, we see that

ea" = E(an) = E(log(l + 1/n)n) = E(L(l + 1/n)n) = (! + 1/n)n,

and

(27) e = lim(l + 1/n)n. n~ro

Some calculus books establish the properties of e, ex, and log x by using equation (27) as a definition of e. In our approach, (27) is a property of e, and we now use this equation to show an interesting interpretation of this basic mathematical constant. If one dollar could be invested at !00 per cent rate of interest for I year, it would accumu­late to I+ I dollars at the end of the year. If the rate were lOO per cent compounded semi-annually, the amount accumulated would he l + l dollars at the end of six months, and so

(I +tl +W +ll = (I+})(I+:ll = (I +:!)2

dollars at the end of the year. If the interest rate were l 00 per cent compounded three times during the year, the effect would he to multiply by a factor of I+! at the end of each interest period, l +! because we need a factor I to give the original amount and a factor } to account for the interest. Hence the accumulated amount at the end of the year would be

(! + })3

dollars. Similarly, quarterly additions of compound interest would yield, by repeated multiplication by I+!, a total of

(I+ i)4

dollars at the end of the year. In general, if the interest rate is lOO per cent compounded n times a yea.r, one dollar would increase to

(! + 1/n)n

dollars at the end of the year. But by (27) the limit of this expression as n tends to infinity is e, and so e can be interpreted as the accumulated amount at the end of one year when one dollar earns interest at the rate of lOO per cent compounded continuously. Since ordinary com­pound interest represents growth in spurts, we see that e is related to

Page 146: Calculus an Introductory Approach

138 CALCULUS: AN INTRODUCTORY APPROACH CHAP. 8

a steady, continuous growth. And so is ex; for example if instead of using an interest rate of lOO per cent compounded continuously we had used a rate of (IOOx) per cent, then the continuous compounding of interest would have increased the principal of one dollar to an amount of ex dollars. In other words, we analyzed the special case with x = I. Because of such properties as these, the graphs of the function y = ex and of similar exponential equations are called the growth curves.

Problems

I. Compute(!+~)',(!+·})" and(!+;})' to two decimal places. 2. Evaluate the limits

( I r ( I ( (a) lim I+- (c) lim 1+-n__,ro 2n U-JOO n2

( I r+l ( I r (b) lim 1+-- (d) lim I+-n->oo 2n+l n-HxJ 3n

Suggestion: Use Theorem 3 of § 6.1. *3. Evaluate the limits

( Ir' 1 1 )3n

(a) lim I+- (d) lim (I+-n->oo n n->oo n'

( lr lim (I+_!_ r (b) lim I+- (e) n->co n n-"co 3n

(c) lim (I+_!_ r n->co 2n (f) ( I r+l lim I+-

n->oo 2n

Suggestion: Use Theorem 12 of § 2.4.

7.6 The Series Expansion for the Logarithmic Function. Although the reader is probably familiar with the concept of a geo­metric progression and its sum, nevertheless we develop the basic idea here because it is quite essential to our method. Consider the sum

S = I +r+r'+r3+ ... +rn-2+rn-r,

where n is some positive integer. To find an algebraic formula for this sum, we multiply the above equation by r to get

rS = r+r'+r"+r4+ ... +rn-l+rn.

Page 147: Calculus an Introductory Approach

SEc. 6 THE LOGARITHMIC AND EXPONENTIAL FUNCTIONS 139

Subtracting these two equations we get

S -rS ~ 1-rn, or S(l-r) ~ 1-rn,

since most of the terms on the right sides cancel. Dividing by 1-r we get S ~ (1-rn)j(l- r), and so

l-rn I rn (28) I +r+r2+r3+ ... +rn-2+rn-l ~ -- ~ -----.

1-r 1-r 1-r

This equation is not valid in case r ~ l, because in this case l - r ~ 0 and the division is impossible. But (28) holds for any number r except r ~ l, and for any natural number n. So also the following variation on (28) holds for x # l,

l xn (29) -- ~ l+x+x'+x3+ ... +xn-2+xn-1+ --.

1-x l-x

We plan to integrate this equation. To begin, we want to find the integral of l/(1-x). Theorem 3 of§ 7.3 suggests that we should look at the derivative of y ~ log(l-x). Using this theorem, and also Theorem !I of § 4.5, we can write

(30)

y ~ logu, u = 1-x,

dy dy du - = -·- = dx du dx

dy I - =-, du u

du

dx

I -1 -- = --,

u 1-x

d -1 -{log(l-x)} ~ -. dx 1-x

-I,

With this background, we integrate (29) to get

(31) f

b dx J" J" Jb Jb -- ~ I dx+ xdx+ x2dx+ x3 dx o 1-x o o o o

+ ... + Jb xn-2dx+Jb xn-ldx+Jb ~dx. o o ol-x

In view of (30) we see that

fb dx ~ [ -log(l-x)]g ~ -log(l-b)+logl -log(l-b), o l-x

and so (31) can be written as

b2 b3 b4 bn-1 bn fb xn (32) -log(l-b) ~ b+-+-+-+ ... +--+-+ --dx.

2 3 4 n-1 n · o 1-x

Page 148: Calculus an Introductory Approach

140 CALCULUS: AN INTRODUCTORY APPROACH CHAP. 7

This last integral will not be evaluated, but we will establish that it tends to zero as n tends to infinity, provided that b is suitably restricted. For convenience we write

(33) Jb xn

In~ --dx. o 1-x

Now the fact is that In -> 0 as n ~, oo for all values of b satisfying - 1 :;;; b < 1; but we shall prove this only for the values satisfying 0 :;;; b < 1, since this is all that we shall need for our purposes. We apply Theorem 9 of § 5.5. With b satisfying 0 ;'; b < 1 we notice that the values of x involved in the integral (33) satisfy

O~x~b, xn~bn, 1-x~l-b,

1 1 xn bn -- :<;; --, -- :<;; --. 1-x- 1-b 1-x- 1-b

Hence we can take the M of Theorem 9 of§ 5.5 to be bnf(l-b), and so we conclude that

Jb xn bn bn+l

In~ --dx ;'; --(b-0) ~ --. o 1-x 1-b 1-b

Also we observe that In ;;; 0, since the integrand is non-negative for the values of b under consideration. Thus we can write

(34) bn+l

0 :0; In :0; --. - - 1-b

Now lim bn ~ 0 as n tends to infinity, by Theorem 1 of§ 6.1, and so

~H b b lim -- ~ lim bn · lim -- ~ 0 · -- ~ 0. n->oo 1-b n-)co n->oo 1-b 1-b

Hence lim In ~ 0 by Theorem 9 of § 2.3. It follows that if we let n tend to infinity in (32) we get

b2 b3 b• b5 -log(1-b) ~ b+-+-+-+-+ ....

2 3 4 5

This holds for b restricted to 0 ;'; b < 1, and so we have the infinite series for log(1-b),

b2 b3 b4 b5 (35) log(1-b) ~-b--------- ....

2 3 4 5

Page 149: Calculus an Introductory Approach

SEc. 7 THE LOGARITHMIC AND EXPONENTIAL FUNCTIONS 141

A more thorough analysis of equations (32) and (33) can be given to show that (35) is valid for values of b satisfying -l ;:; b < l, but we will not go into that.

Problems

l. Compute log .9 to 4 decimal places. Suggestion: Set b = .1 in (35). 2. Compute log .99 to 4 decimal places. 3. What does (35) give us if we set b = 0?

7.7 The Computation of Logarithms. Equation (35) can be used to compute logarithms. For example if we want to find the value of log 2, we take b = } in (35), and use the fact that log·} = -log 2 to get

log~= -log2= -~-~(~)2

-~(~)3

-~(~)4

- •••• 2 2 22 32 42

The value of log 2 to 4 decimal places is .6931. In the same way we could take b = § in (35) to obtain the approximate decimal value of log 3. But it would be better to take b = ·}, because the smaller the value of b, the more rapidly does the series (35) converge, i.e., the more rapidly does it tend to its limit. With b = -} we would get log §, and this could be subtracted from log 2 to give log 3:

log2-log! = log3.

But we drop this approach, and now show another way of getting at log 3 by using series that converge even more rapidly. First we use b = i and b = -,'6 in (35), thus

4 Jog-=

5 - ~- ~ (~)2 -~ (~)3- ~ (~)4

5 2 5 3 5 4 5 - - .2231,

log1~ = -110 -~c~r -H1~r- H1~r - ... ;; -.1054'

where the symbol ;; stands for "is approximately equal to". Then we can write

Jog5 = 2log2-log4/5;; 2(.6931)+.2231 = 1.6093,

Jog10 = Jog2+log5;; .6931+1.6094 = 2.3024.

At this stage we have picked up an error in the last decimal place, which is to be expected because the errors in the approximations are

Page 150: Calculus an Introductory Approach

142 CALCULUS: AN INTRODUCTORY APPROACH CHAP. 7

piling up as we combine our results. The value of log 10 to 4 decimal places is 2.3026, and that of log 5 is 1.6094. Finally we have

log9 ~ logl0+log(9/IO)"" 2.2036~.1054 ~ 2.1972,

log 3 ~ t log 9 "" 1.0986.

All these are natural logarithms, that is, logarithms to base e. To compute logarithms to base 10, common logarithms as they are called, we use the identity

(36) log10a = logea · log10e, for a > 0.

To prove this, let us write Cf., fJ, and y for these three logarithms, that is

(f. ~ logwa, fJ ~ log,a, y ~ logwe.

Then we can establish (36) by observing that

a ~ 10", a ~ eP, e ~ IOY, a ~ (IOY)P ~ IOYP, 10" ~ IOYP,

(f.~ yfJ. In order to use (36) we need the value of logwe, and we can get this by setting a~ 10: ·

logwiO ~ I ~ log,IO ·logwe.

We have already established that log,IO "" 2.3026, and so we find that

I I logwe ~ -- "" _ .4343.

log,IO 2.3026 -

Substituting this in (36) we get

logwa "" .4343 log,a.

To compute logw2, for example, we use the previously computed log 2 "" .6931 to get

log102 "" (.4343)(.6931) "" .3010.

Problems

I. Compute the following to 4 decimal places.

(a) logw3 (c) log106

(b) logw5 (d) log10!.6

2. Compute the following to 4 decimal places.

(a) log7

(b) log,o7

(c) log 11

(d) log,oll

Page 151: Calculus an Introductory Approach

CHAPTER 8

FURTHER APPLICATIONS

8.0. This chapter, containing some additional applications of calculus, differs from the earlier chapters in that there is no sequential development of the material. The sections can be read independently, and therefore in any order, except for these restrictions: §§ 8.1 and 8.2 constitute a unit, an item of theory followed by an application; the mixing problem in § 8.5 leads to essentially the same equation as in § 8.4, and so one step of the argument is carried over intact.

8.1 The Series Expansion for the Arc Tangent Function. We can get an infinite series for arc tan x by using a scheme quite similar to that for log(1-b) in§ 7.6. We replace r by -x2 in equation (28) of § 7. 6 to get, after switching terms around,

( 1)

Since (28) was an identity, valid for all values of r except r = 1, we see that this equation (1) holds for all real values of x without exception. This is so because - x2 = I has no solution in real numbers; and we have no concern with complex numbers in our analysis.

In order to integrate equation (1), we first use Theorem 13 of § 4.6 to get

Jb dx -- = [arctanx]g = arctanb-arctanO = arctanb.

o 1+x2

143

Page 152: Calculus an Introductory Approach

144 CALCULUS: AN INTRODUCTORY APPROACH CHAP. 8

Then if we integrate equation (I) term by term, with limits 0 to b on the integration, we get

b3 b5 b' b9 (2) arctanb ~ b--+---+-- ...

3 5 7 9

b2n-3 b2n-l Jb x2n + ( -l)n-2~~ + ( -l)n-1~~ + ( -l)n ~~dx.

2n-3 2n-l o I+x2

We show that the last term tends to zero as n tends to infinity if b satisfies 0 ;;; b ;;; I. Since x2 ;;; 0 for every value of x, it follows that

1 x2n l+x2 ~I, ~~ :-s; 1, ~~ S x2n.

- l+x2 - I+x2 -

We apply Theorem 8 of § 5.5 with F(x) and G(x) replaced by xnj(l + x') and xn respectively, to get

Jb x2n Jb [ x2n+l ] b b2n+l --dx S x 2n dx = ~~ ~ ~~.

o l+x2 - o 2n+I q 2n+l (3)

Since we are restricting b to the interval 0 ;;; b ;;; I we see that

b2n+1 I lim-~- ~ limb2n+l·lim~~ ~ 0. n-•oo 2n+ I 2n+ I

0 ~ b2n+l ~ I,

It follows from (3) that the limit of the integral is also zero as n ->- w, by the usual application of Theorem 9 of § 2.3. Thus equation (2) becomes, as n -:.. oo

(4) b3 b5 b' b9 bll

arctanb ~ b--+---+---+ .... 3 5 7 9 ll

What is the range of validity of equation (4)1 In developing the equation we restricted b to the interval 0 ;;; b ;;; 1. However, since arc tan( -b)~ -arc tan b, as illustrated in Figure 4.IO of§ 4.6, and since all the terms on the right side of ( 4) also change sign when b is replaced by -b, we see that (4) is valid for bin the interval -I ;;; b ;;; I.

8.2 The Computation of rr. Equation (4) of the preceding section enables us to compute the value of rr. If we take b ~ I, and recall that arc tan I ~ rr/4, we get

4

I I I I I I--+---+---+ ....

3 5 7 9 ll

Page 153: Calculus an Introductory Approach

Sli:c. 2 FURTHER APPLICATIONS 145

This equation, elegant though it is, does not lend itself to computation because the series converges very slowly, so that a considerable cal­culation is necessary for accuracy to, say, four or five decimal places.

A better equation for computational purposes can be had by taking b = v3j3 in (4). From elementary geometry we know that

" ..;3 tan-=-,

6 3 ..;3 " arc tan-=-, 3 6

and so

; = ~3 _ ~ (~3)"+ ~ (~~r _ ~ ( ~3) \ .... Multiplying by 6 and simplifying we have

(5) 7T=2y31--+-----+-----+ .... -[ 1 1 1 1 1 ] 3.3 5.32 7.33 9.34 11.35

Yet another equation for " can be arrived at by using the trigono­metric identity

(6) tan o: +tan f3

tan(o: + /3) = . 1-tano: tanf3

If we take o: = arc tan l and f3 = arc tan 1 we get

1+1 7T tan(o:+/3) = = l, and so o:+/3 = -.

1-l· t 4

This leads to the following relation, by use of equation (4),

7T 1 1 (7) -=arc tan- +arctan-

4 2 3

[1 l 1 1 ] 2 - 3.23 + 5.25 - 7 .2' + ...

The value of" can be computed to as many decimal places as we please by use of equation (5) or equation (7).

Problems L Use (5) to compute the value of 7T to four decimal places. 2. Use (7) to compute the value of n to four decimal places. 3. Compute arc tan .l to four decimal places.

Page 154: Calculus an Introductory Approach

146 CALCULUS: AN INTRODUCTORY APPROACH CHAP. 8

4. Compute arc tan 10 to four decimal places. Suggestion: Equation (4) cannot be used with b ~ 10. Prove that arc tan 10 +arc tan .I ~ rr/2.

5. Compute arc tan 2 to four decimal places. 6. Verify that the error in using 22/7 as an approximation to rr is slightly

larger than .0012.

8.3 The Velocity of Escape. Consider the following problem. At what initial velocity need an object be shot upwards from the surface of the earth in order to escape, so that it does not fall back to the surface of the earth 1 To attack this question, let us begin by stating what the conditions of the problem are. We shall disregard the re­sistance caused by the atmosphere at the surfa9e of the earth. We shall presume that the initial thrust applied to the object is not directed vertically, but in the appropriate direction to counteract the earth's rotation. Thus the object is presumed to move in a straight line point­ing directly away from the earth, so that if this line were continued backwards it would pass through the center of the earth. Furthermore the problem is idealized in the sense that the entire universe consists only of the earth and the object. Thus we are dealing with a simple case of a two body problem; the complications that arise with more than two bodies are considerable. Finally we shall presume that after the initial thrust there is no further force applied to the object except the gravitational pull of the earth.

In solving the problem we shall use the inverse square law of gravi­tation. The precise meaning of this will be formulated at the point in the discussion where it is used.

0~·------~f~-+~~--x x+Lix

V :FIG. 8.1

t+ L'.t v+ L'. v

Consider the object at a time t (in seconds) after the initial thrust, so that it left the surface of the earth at timet ~ 0. Suppose that at time t the object has reached a point P, as in Figure 8.1, at a dis­

tance x miles from the center of the earth. Since we are measuring dis­tance from the center of the earth, we note that x ~ 4000 when t ~ 0, taking the radius of the earth as 4000 miles. Next consider the object as it passes a point Q a little farther away from the earth. Suppose that the elapsed time is 'now t + L'.t and that Q is at a distance x + L'.x from the center of the earth. Since the object is moving away from the earth, the changes L'.x in the x coordinate and L'.t in the t coordinate are positive quantities.

Now in moving from P to Q the object has traversed a distance L'.x in time L'.t, and so its average velocity in this motion is L'.xjL'.t. We are

Page 155: Calculus an Introductory Approach

SEC. 3 FURTHER APPLICATIONS 147

using here the basic idea that velocity is distance divided by time. If we let llt tend to zero we get the velocity v of the object as it passes the point P, thus

(8) llx

v ~ lim-. M~O flf

This equation is the definition of velocity at a point, and it should be noted that the limit concept is central to this definition.

Now if v denotes the velocity of the object as it passes the point P, it is natural to let v + !lv denote the velocity of the object as it passes Q. Because the flight of the object is being retarded by the gravitational pull of the earth, that is to say, the object is slowing down as it moves from P to Q, it follows that !lv is negative. Just as velocity is the ratio of distance to time, so acceleration is the ratio of velocity to time. Thus the average acceleration of the object as it moves from P to Q is llvjllt, and so the acceleration a at the point Pis

(9)

Since .1.v is negative, so is a.

llv a~ lim-.

at~o llt

We are now in a position to say what is meant by our basic assump­tion of the inverse square law: the force exerted on the object by the gravitational pull of the earth is inversely proportional to the square of the distance from the center of the earth to the object. Further, since acceleration is directly proportional to the force exerted (this is Newton's second law of motion, which the reader may have met in its mathematical form, F ~ ma), we conclude that the acceleration is inversely proportional to the square of the distance,

(10) k

a=--. x2

The symbol k is the constant of proportionality. Why do we write a minus sign in equation (10)? For this reason, that a is negative in our problem, and so if we want a positive constant of proportionality k, we must include a minus sign. If we had omitted the minus sign in (10), then k would be negative, which would have been less convenient although not incorrect.

The constant of proportionality k can be determined from standard experimental data. At the surface of the earth the acceleration due to gravity has the value - .00609, the units of distance and time being

Page 156: Calculus an Introductory Approach

148 CALCULUS: AN INTRODUCTORY APPROACH CHAP. 8

miles and seconds. (This is more commonly given as -32.16, in units of feet and seconds.) Taking the radius of the earth as 4000 miles, we get from (10)

k (11) - .00609 = - (4000)2' k = 9.74 X 104.

Next if we write (8) in the reciprocal form 1/v = lim(Ll.tf~x), and if we multiply this by (9), we get

a ~V ~t ~V ~~ ~V dv - = lim - · lim- = lim -- · - = lim- = --, V ~t ~X ~t ~X ~X dx

where these limits are taken as ~t, and so also ~x, tend to zero. Also we replace a by its value from (10) to get

dv a k dv (12) - - 2v- = - 2kx-2,

dx - ; - - vx2' dx

where we have multiplied by 2v for a special reason. Defining V as the square of the velocity, V = v2, we see by the chain rule of § 4.5 that

dV dV dv --=--·-= dx dv dx

dv 2v-,

dx

so that (12) can be written as

(13) dV -- = -2kx-2 . dx

This says that there is a certain function V whose derivative is - 2kx-2. But 2kx-1 has this same derivative, by Theorem 7 of § 4.3. When two functions, V and 2kx-1, have the same derivative, what can we con­dude! By Theorem 7 of § 5.4 we can conclude that these functions differ by a constant,

(14)

2k V= 2kx-l+c, v2 =- +c,

X

1.95 X 105 v2 = -----+c.

X

By interpreting equation (14) we now solve the problem posed at the beginning of this section. Equation (14) gives the velocity v of the object as a function of the distance x from the center of the earth,

Page 157: Calculus an Introductory Approach

SEC. 3 FURTHER APPLICATIONS 149

but in terms of a constant c which has not been determined. Let us try a numerical example to see how c can be evaluated. Suppose that an object is hurled into space with an initial velocity of 10 miles per second, just to take a number at random. Then in equation {14) we have v = 10 when x = 4000; thus

l.95x 105 100 = ----+c = 48.7+c, c = 51.3.

4000

Thus c is determined when we specify an initial velocity for the object. Notice that in this case of an initial velocity of 10 miles per second, equation (14) can be rewritten as

1.95 X 105 v2 = +51.3.

X

From this equation we see that as x increases v decreases, as is to be expected. But v never becomes zero no matter how large we take x to be, because of the presence of the positive constant 51.3. As x gets indefinitely large, v2 tends to 51.3.

On the other hand let us see what happens in equation (14) if we presume an initial velocity of 9 miles per second for the object. In this case {14) gives us

81 = 1.95 X IQ5

4000 +c = 48.7 +c, c = 32.3,

and so for initial velocity 9 equation (14) becomes

1.95 X 1Q5 v2 = ----+32.3.

X

As x gets indefinitely large, v2 tends to 32.3. Thus we conclude that while 9 miles per second is enough of an

initial velocity to get and keep the object away from the earth, it is more than enough because the object will have a velocity of at least V 32.3 no matter how large x is. Hence the smallest possible initial velocity is the one that yields c = 0 in {14), and so we write

1.95 X IQ5 v2=----

X

Another way of establishing that c = 0 for the smallest possible initial velocity that permits escape is to observe that the velocity v

Page 158: Calculus an Introductory Approach

150 CALCULUS: AN INTRODUCTORY APPROACH CHAP. 8

should tend to zero as x tends to infinity, and so from (14) we get

[ 1.95 X [05 ]

lim v2 ~ lim ----- +c ~ 0, X-700 X-'OC! X

[ 1.95 X J05 ]

lim x +lim[c] ~ O+c ~ 0.

In any event we have (14) with c ~ 0, and we determine the initial velocity by setting x ~ 4000 to get

1.95 X J05 v2 = V ~ 6.98.

4000

'l'hus we have established that the minimum velocity of escape is 6.98 miles per second.

Problems

*1. Let g denote the acceleration of gravity at the surface of any planet. Assuming the planet to be a sphere of radius r, prove that the velocity of escape equals vzgr.

*2. Given the following table of approximate values,

EARTH MooN MARS

RADIUS IN

l\ULES

4000 llOO 2100

MEAN DENSITY

(WATER~ 1) 5.5 3.3 4.0

prove that gzfg, ~ .165 and g3 jg, ~ .38, where g,, g,, g3 are the accelera­tions of gravity on the Earth, the moon, and Mars. Suggestion: Newton's law of gravitation implies that g is directly proportional to the mass of the planet, and inversely proportional to the square of the radius.

*3. Use all preceding data to show that the velocities of escape from the moon and Mars are approximately 1.5 and 3.1 miles per second.

8.4 Radioactive Decay. We now give another example leading to a differential equation, which is the name for any equation involving derivatives, like equation (12) of the preceding section. Any radioactive material disintegrates by giving off radiation, and there remains a substance of lower atomic weight. If we make the reasonable assump­tion that the amount of radiation is proportional to the amount of radioactive material present at any time, it follows that the rate of decrease of the radioactive substance is proportional to the amount of radioactive substance remaining at any given time.

Page 159: Calculus an Introductory Approach

SEC. 4 FURTHER APPLICATIONS 151

To translate this observation into mathematical symbols, let u denote the amount of radioactive substance present in a sample of material at any time t, so that u, being dependent on t, is a function oft. As time passes, u decreases, and since dujdt is the rate at which u changes, this derivative is negative. But the rate at which u changes is proportional to u, say - ku, where a minus sign is included since the rate of change is negative. Thus k is some positive coustant of propor­tionality, and we have the differential equation

du 1 du (15) - = -/,;u or - ·- = -k.

dt u dt

In the absence of a systematic treatment of differential equations, again as with equation (12) we find a solution by introducing a new variable. If we define y = log u, then by Theorem 3 of § 7.3 we have

dy 1

Now by Theorem ll of§ 4.5, with tin place of x, we see that

dy dy du 1 du dt = du . dt = ; . dt ·

Combining this with (15) we obtain dyjdt = - k, and this is a dif­ferential equation whose solution is readily available, namely

y = -kt+c.

By including the constant c in this equation, we know that we have the most general representation of y as a function of t, by Theorem 7 of § 5.4. Now y is the same as log u, so the solution of (15) is

(16) logu = -kt+c.

This equation is equivalent to

(17) u = e-kt+c,

and this gives u as a function of t. The constants c and k can be determined for any given sample of

radioactive material by experimental data. For suppose we measure the amount of radioactive material at an initial time t = 0 and again at some later time t1, finding that the amount is u0 at time t = 0 and u1 at t = t1• Now if k and care properly chosen, (16) gives the amount u of radioactive material present at time t for all values oft. At t = 0, equation (16) asserts that log u = c. Since the amount u is uo at

Page 160: Calculus an Introductory Approach

I["

11

'I' I I lie

tli

, I .I

!52 CALCULUS: AN INTRODUCTORY APPROACH CHAP. 8

t = 0 we have log uo = c, which tells us the value of c. Similarly at t = t1 we find from equation (16) that log u1 = -kt1 +c. These results enable us to determine k as well as c, in the ±allowing way:

c = loguo, logu, = - kt, + c,

logu1 = - ktr + logu0,

k = loguo -logu, log(uofu,)

t, t,

Thus we have expressed c and k in terms of uo, ur, and tr. In numerical problems, it is not always necessary to evaluate k

itself. For example, suppose we know that u = 10 when t = 0, and u = 9 when t = 50, and we want to find the value of u when t = 150. Then (17) gives

10 = ec, 9 = ec-50k, u = ec-150k.

The first two equations imply that e-50k = 19

0 and so we get

u = e'. e-150k = 10. (,',)" = 7.29.

Problems

I. By substituting uo for u when t = 0 in (17), prove that (17) can be re\vritten as u = uoe-kt.

2. Verify that if equation (17) is differentiated, the result is (15). 3. A capsule contained 20 milligrams of a certain radioactive substance

60 years ago. Disintegration has reduced the amount to 19 milligrams today. How much will remain 240 years hence?

8.5 A Mixing Problem. To keep matters from getting too com­plicated, we shall examine a very specific mixing problem. Consider a tank containing 1200 gallons of brine, at a concentration of I pound of salt per gallon. Suppose that pure water is poured steadily into the tank at a rate of 20 gallons per minute; that the mixture is kept at a uniform concentration throughout the tank by vigorous stirring; and that the mixture flows steadily out of the tank also at a rate of 20 gallons per minute, so that the total volume under consideration is a constant 1200 gallons. The question is, what is the amount of salt in the tank at any time t1 We may presume that t is measured in minutes, and that t = 0 at the time that the pure water started pouring into the tank.

Let there be x pounds of salt in the mixture in the tank at any time t, so that x = 1200 when t = 0. Then from time t to a later time t + 1\.t the amount of salt in the tank decreases from x to x + Ll.x, where 1\.t is positive but 1\.x is negative. At time t the amount of salt per gallon

Page 161: Calculus an Introductory Approach

SEC. 5 FURTHER APPLICATIONS 153

of the mixture is x/1200, and at timet+ [',t the amount is (x + /',x)/1200. Over a time interval /',t, exactly 20(/',t) gallons of water enter the tank, and at the same time 20(/',t) gallons of the mixture leave the tank. This outflow of 20(/',t) gallons contains between

X -- · 20(/',t) and 1200

x + /',x -- . 20(/',t)

1200

pounds of salt. The second of these is the smaller because /',xis negative. But the change in the amonnt of salt in the tank from time t to time t + /',t is /',x; in other words the amount of salt leaving the tank in this time interval is - /',x pounds. Hence we conclude that

x x+/',x -- · 20(/',t) > -/',x > -- · 20(/',t). 1200 1200

Dividing by [',t, we get

x /',x x + /',x > -->---

60 /',t 60

Taking limits as [',t and /',x tend to zero, we see that

X dx dx X 1 dx

60

1

60

Notice that this amounts to the differential equation (15) of the preceding section, with x in place of u, and ,', in place of k. Conse­quently we can go at once to the solution (17) of (15), which gives, when adapted to the present situation,

x = ec-t/60 = ec . e-t/60.

Using the information that x = 1200 when t = 0, we see that 1200 = ec, and so the above equation becomes

x = 1200 e-t/60.

This equation represents the amount of salt x as a function of the time t, valid for t ;;; 0.

Problems

The reader will find tables of natural logarithms and the exponential function of considerable use for the following problems.

l. With reference to the problem outlined above, how much salt remains in the tank after the pure water has been pouring in for l hour'? for 2 hours?

Page 162: Calculus an Introductory Approach

154 CALCULUS: AN INTRODUCTORY APPROACH CHAP. 8

2. At what time will the concentration of salt in the tank be ! pound per gallon 1 i pound per gallon?

3. Draw a rough graph of the equation x = 1200e-tl6o from, say, t = 0 to t = 180. Suggestion: The units of length on the axes should be carefully chosen; the t-axis should be horizontal, the x-axis vertical.

8.6 The Focusing Property of the Parabola. The parabola y = x' was used to illustrate some basic problems in calculus in earlier chapters. This particular equation can be regarded as a special case of 4py = x 2, with p = l;. If we choose another value for p, say p = 1, we get another parabola of a different size, 4y = x2. Thus the equation 4py = x', or y = x'J(4p), is a general formula which includes all possible parabolas asp is given different positive values. The following analysis can be simplified if p is assigned the value 1 throughout; our reason for not taking p = 1 is that we want to establish the focusing property for all parabolas, not just for the parabola 4y = x'.

y -axis

L 0 R (u,-p) K

FIG. 8.2 Graph of x2 = 4JY1J.

In Figure 8.2 an arbitrary point P with coordinates (u, v) is selected on the parabola x' = 4py. The coordinates u and v are thus related by the equation u2 = 4pv, and this relation will be used several times. The fixed point F with coordinates (0, p) is called the focus of the para­bola. The line LK is drawn parallel to the x-axis, at a distance p below. Hence the equation of the line LK is y = - p, since the y coordinate of every point on this line is - p. What we want to establish first is the following property of the parabola: that every point on the curve is equally distant from F and from the line LK. In Figure 8.2 a perpendicular PR is drawn from the point P to the line LK, inter­secting the line LK at R. Thus the property to be proved can be written in the form of an equation, PF = PR.

Page 163: Calculus an Introductory Approach

S'Ec. 6 FURTHER APPLICATIONS 155

The line LK, with equation y = - p, is called the directrix of the parabola. The property PF = PR states that every point P on the parabola is equally distant from a fixed point, the focus, and a fixed line, the directrix. Now whereas we prove this property PF = PR as a consequence of the equation x2 = 4py (or u2 = 4pv), the parabola is often approached the other way around in books on geometry. Tbat is to say, the parabola can be defined as the set of points (in a plane) equally distant from a fixed point and a fixed line (in the plane). Then the relation x2 = 4py can be proved as a property, provided the axes are located properly. The logic is correct no matter which is the definition and which the proved property. For our purposes it is most convenient to define the parabola in terms of its equation, and then derive the property PF = PR as a consequence.

Proof that PF = PR. Again with reference to Figure 8.2, let the line PR intersect the x-axis at the point S, whose coordinates are clearly (u, 0). Finally, let a perpendicular FT be drawn from the focus F to the line PR; the coordinates ofT are seen to be (u, p). We note the relations

PS = v, TS = p, PT = v-p, PR = v+p, FT = OS = u.

Using Pythagoras' theorem applied to the triangle PFT we get

PF' = PT'+FT' = (v-p)'+u' = vL2pv+p'+4pv = (v+p)2.

It follows that PF2 = PR2, and so PF = PR = v+p. We now turn to a second geometric property of the parabola, namely

that the tangent line to the parabola at the point P makes equal angles with PF and PQ, where PQ is the line through P parallel to the y- A axis, as illustrated in Figure 8.3. Our plan is to prove this geometric result, and then interpret its meaning as "the focusing property of the parabola".

Let the tangent line to the parabola at the point P intersect the y-axis at B, so that as shown S (u, 0) in Figure 8.3, APB is the tangent line. Then the geometric property under discussion is that LFPB C = LQPA, and this we now prove.

As a preliminary, we establish Fw. 8.3

Page 164: Calculus an Introductory Approach

!56 CALCULUS: AN INTRODUCTORY APPROACH CHAP. 8

that FP = FB. We differentiate the equation

I

to get

y = -x2 4p

dy I X -- = -(2x) = -. dx 4p 2p

Hence the slope of the line PB is uj(2p), since u is the x coordinate of P. But the slope of PB is also PCjBC, and consequently

u PC PS+SC -=-=~---=

2p BC BC

PS+ OB

os v+OB

u

Multiplying by u we get

u2 - = v+OB, 2p

,.z 4pv OB = - - v = - - v = 2v- v = v.

2p 2p

Hence the length of FB is expressible as

FB = FO+OB = p+v.

But FP = p + v also, so FP = FB. Thus the triangle FPB is isosceles, and LFPB = LFBP. But LFBP = LQPA, and so LFPB = L QPA, which is what we wanted to prove.

Hence if we conceive of a ray of light emanating from F along the

FIG. 8.4

line FP, it will be reflected at P along the line PQ. Let a mirror be forll)ed by revolving the para­bola about the y-axis. Then the light rays emanating from F will be reflected by the mirror along paths parallel to the y-axis, as in Figure 8.4. A spotlight results from putting a light source at F. Or, if we think of the light moving

in the other direction, the parabolic mirror will focus the light at the point F. Thus in a reflecting telescope the eye or the camera is located at point F.

Problems

1. Figure 8.2 has been drawn with the point P "above" the point T, i.e., with v > p. Prove that PF = PR in case v < p; also in case v = p. Hraw diagrams in both cases.

Page 165: Calculus an Introductory Approach

SEC. 7 FURTHER APPLI:.::CccAc:T:.::I.:cO~N:c:S:_ _______ 157

2. What are the coordinates of the focus of the parabola y = 2x21 of the parabola 6y = x'? of the parabola 6x = y''!

8.7 Newton's Method for Solving Equations. Consider the equation

(18) x3-3x-12 = 0.

This equation can be solved by exact methods, that is by methods yielding solutions in terms of radicals in the same sense that the for­mula

-b ± yb2 -4ac X=

2a

is an exact solution of the equation ax' + bx + c = 0. If we are looking for a solution of (18) which is accurate only to a few decimal places, it is usually quicker to solve the problem by the use of an approximation method, such as the one we are about to describe, than to use the exact solution. Furthermore, while there is an exact solution of cubic equa· tions like (18), and more generally there is a solution in terms of radicals for equations of degree 4 or less, it is known that it is impossible to solve the general equation of degree 5 or more by means of radicals. Hence to solve an equation such as x5-2x4+2x3-3x2+l = 0, we would use an approximation method.

Returning to equation (18), let us denote x3-3x-l2 by f(x), and we can readily compute the following values:

f(-2) = -14, f(-1) = -10, f(O) = -12, j(1) = -14,

f(2) = - 10, f(3) = 6.

The pair of functional values f(2) = -10 and f(3) = 6 suggests that as x varies continuously from x = 2 to x = 3, f(x) varies continuously from - 10 to 6, and so there is some intermediate value of x where f(x) = 0. This is so, and although it is perhaps intuitively obvious, it is not easy to give a rigorous mathematical proof. We give no proof here, but simply take it as axiomatic that if a continuous function f(x) has functional values f(a) and f(b) with opposite signs, then there is at least one value of x between a and b for which f(x) = 0. Thus equation (18) has a root between x = 2 and x = 3, and we can take either of these as a first approximation to the root. The approxima­tion x = 3 is preferable, because 6 is closer to 0 than - lO is.

In general, suppose we have a first approximation x = a to a root of an equation f(x) = 0. The idea is that we will use a tangent line to

Page 166: Calculus an Introductory Approach

158 CALCULUS: AN INTRODUCTORY APPROACH CHAP. 8

the curve y = j(x) at the point where x = a, as illustrated in Figure 8.5, to get a better approximation to the root. In Figure 8.5, P is the point where the graph of y = f(x) crosses the x-axis, so the x coordinate of the unknown point P is the root of f(x) = 0 that we are trying to

y-axis approximate. Beginning with the approximation x = a, we draw a tangent line to the curve at the point A with coordinates (a, f(a)). This is the line AB, and the point B on the x-axis has an x coordinate

_.j_ _ _l__..l~c......----'x~--'ax=is which is in general a better approx­K

Fm. 8.5

imation than x = a. Repeating the process, we use a tangent line

y~ f(x) at C to get to the point D, whose x coordinate gives a still better approximation.

This geometrical process leads to a simple formula by the following analysis. The slope of the line AB equals the value of the derivative f'(x) at x = a, namely f'(a). But if we use (b, 0) as the coordinates of B, the slope of the line AB is also seen to be, from equation (I) of § 1.1,

0-_.fc...:.(a...:._) - , and so

b-a 0-f(a) = j'(a). b-a

Using simple algebra, we get

( 19)

-f(a) = (b-a)J'(a), b-a =

f(a) b =a--­

f'(a)'

Thus from a first approximation x = a we get a second approximation x = b as in (19), and this in turn gives a third approximation x = c, where

j(b) c = b - --, etc.

j'(b)

Returning to equation (18) as an application of the use of (19), 've have

j(x) = x3-3x-12, j'(x) = 3x2-3, a= 3,

f(a) 6 11 f(a) = 6, j'(a) = 24, b =a--= 3-- = -.

J'(a) 24 4

Page 167: Calculus an Introductory Approach

SEC. 7 FURTHER APPLICATIONS

Turning to decimals at the next stage, we get

f(b) c = b-~- =

f'(b)

to two decimal places.

.55 2.75- ~-"'

19.7 -2.72,

159

Had we started with the approximation a = 2, we would have ob­tained

f(2) = -10, !'(2) = 9, b = f(a)

a-~-=

f'(a) 3.1.

We appear to be losing ground, but 3.1 is actually a better approxima­tion than 2 is. A repetition of the process would lead from 3.1 back toward the root.

FIG. 8.6

The use of (19) will not always give a succession of values a, b, c, ... which converge to the actual root. For example, in Figure 8.6 there is an example of a function such that the approximation a leads to b, but iteration of the process leads back to a again. The mathematical theory giving conditions under which success of the procedure is assured is beyond the scope of this book. We shall confine our attention to cases where the successive approximations do lead to the root with increasing accuracy. To minimize the amount of calculation required, we shall require accuracy to two decimal places only. When two suc­cessive approximations are identical to two decimal places, we shall regard this as assurance of accuracy. For example, in the case of the equation (18), we obtain 2.72 as an approximation to the root after two applications of (19). Another iteration of the process leads from 2.72 to the same number, 2.72, to two decimal places of accuracy. Hence we conclude that 2. 72 is a root of x3- 3x- 12 = 0, accurate to two deci­mal places.

Newton's method is not restricted to the solution of polynomial equations, but can also be applied to such equations as sin x = xj2

Page 168: Calculus an Introductory Approach

160 CALCULUS: AN INTRODUCTORY APPROACH CHAP. 8 ~~~~~~-

and ex = x + 7. No new idea is involved in such applications, so we shall confine attention to polynomials, for ease of computation.

Problems I. By drawing a rough sketch of the graph of y ~ x'-3x-l2, establish

that equation (18) has only one real root. Suggestion: The derivative of j(x) ~ "'"-3x-12 isj'(x) ~ 3x'-3. and so by the theory of§ 4.7 the curve has a local maximum point at ( -1, -10) and a local minllmun point at (1, -14). Furthermore, as x increases, j(x) is increasing for x < -1 and x > I. but j(x) is decreasing for -I < x < I.

2. Establish that x3 + x- 5 = 0 has only one real root, and that this root lies between x = I and x = 2. Find the root to two decimal places of accuracy.

3. Find all real roots of x3+x2-7 = 0. 4. Establish that the equation x'+ 12x2+36x-4 ~ 0 has only one real

root, and that this root lies between x = 0 and x = I. Find this root to two decimal places of accuracy.

5. Find all real roots of x3 + 3x2 + x- 6 = 0 to two decimal places of accuracy.

6. Find all real roots of x3-3x2-x+4 = 0 to hvo decimal places. 7. Find the value of v3 to two decimal places by solving the equation

x'- 3 ~ 0 by Newton's method. 8. Evaluate -{7'36 to two decimal places. *9. (a) Let k and a be positive numbers such that a is an approximation

to Vk. Prove that Newton's method applied to the equation x2-k ~ 0 gives ~(a+kja) as the next approximation.

(b) Prove that the approximation i(a-+kfa) is larger than Vk, whether or not a is larger than vk. Suggestion: The inequality }(a+kja) > Vk is equivalent to a+kfa > 2Vk, a-2Vk+kfa > 0. (Va- Vkja) 2 > 0.

(c) Suppose that the approximation a is too large, i.e., a > Vk. Prove· that ~(a+kfa) is a better approximo.tion in the sense that it is closer to Vk; that is, prove that

i(a+kfa)-yk < a-yk.

(d) Suppose that the approximation a is too small, i.e., a < Vk, but large enough so that 3a > vk. Prove that }(a+kfa) is a better approxima­tion; that is, prove that

}(a+kfa)- yk < yk-a.

Suggestion: This inequality is equivalent to a+kfa-2yk < 2yk-2a, 3a-4vk+kja < 0, and

- -

(3a- yk)(a- yk) -------- < 0.

a

*10. Let a be an approximation to .v'k. Prove that Newton's method applied to x3-k = 0 gives ~(2.a+lcja2) as the next approximation.

Page 169: Calculus an Introductory Approach

APPENDIX A

ON THE EXISTENCE OF THE DEFINITE INTEGRAL

In § 5.3 we established the existence of the definite integral

(I) J>(x) dx,

but only in the presence of a function F(x) whose derivative F'(x) is equal to f(x). In the case of such an integral as

J>!x3+ldx

we know of no function F(x) whose derivative is V x' +I. Nevertheless this integral exists, as we shall now show. The general result that we establish is this: the integral(!), !M defined in equation (2) of § 5.2, or equation ( 4) of § 3.2, exists for any monotonic function f(x).

The existence proof that we are about to give is necessary in order to complete the argument of Chapter 7, where we defined the function L(x) by the integral

Jxdv

L(x) = -. 1 V

We now establish that L(x) really exists, and so we will have a proper foundation for the logarithmic and exponential functions.

The proof is given for a monotonic increasing function f(x). For a monotonic decreasing function the proof is analogous with all inequalities reversed.

In the course of the proof we shall use a fundamental proposition about real numbers, which we now explain. It will be labeled an "axiom", although in any thorough study of the real number system

161

Page 170: Calculus an Introductory Approach

162 CALCULUS: AN INTRODUCTORY APPROACH APPENDIX A

it can be given as ·a theorem, Le., it can be proved. We say that a sequence a1, az, as, a4, ... is non-decreasing in case a1 ~ a2, a2 ~ a3, a3 ~ a4 , etc. The sequence is non-increasing in case all the reverse inequalities hold. A sequence is bounded above if there is a constant c such that a 1 < c, a 2 < c, a3 < c, etc. Similarly a sequence is bounded below if there is a constant k such that k < a,, k < a2, k < a3, etc.

For example consider the sequences

(2) I, 2, 3, 4, 5, 6, 7, 8, ... ,

(3) 1234567 2'3' 4'5'6'7'8' ... '

(4) 1234567 2' 22' 23' 24' 25' 26' 27' ....

The sequence (2) is non-decreasing, but is not bounded above, al­though it is bounded below. (The sequence (2) is actually increasing, but "non-decreasing" is a more useful concept in this Appendix.) The sequence (3) is non-decreasing and is bounded above and below. The sequence ( 4) is non-increasing and is bounded above and below.

AxroM ON BoUNDED SEQUENCES. A non-decreasing sequence that is bounded above has a limit. A non-increasing sequence that is bounded below has a limit.

This axiom has nothing to say about the sequence (2) above, but it applies to the sequences (3) and (4) and assures us that these sequences have limits. The limits are 1 and 0 respectively, but the axiom only assures us of the existence of a limit, not that we can find the actual limit in all cases.

We shall use this axiom later in the proof, but first we prove a result about limits.

THEOREM l. If the terms of the sequences {an), {bn) and {en) satisfy the inequalities bn ;<; an ;<; Cn for all positive integers n, and if an -+ k and Cn- bn -+ 0 as n -> oo, then bn -+ k and Cn -+ k.

PROOF. Noting that

0 ~ an-bn ~ Cn-bn

we see that an- bn -+ 0 by Theorem 9 of § 2.3. Then by Corollary 13 of § 2.4 we conclude that bn -> k. Also Cn- bn -> 0 implies that bn- Cn -+ 0, and so the same Corollary 13 implies that Cn -+ k.

Page 171: Calculus an Introductory Approach

APPENDIX A EXISTENCE OF THE DEFINITE INTEGRAL 163

It will be convenient to use some simple notation from the theory of sets. By a set C we mean a collection of elements; all the sets we use will have real numbers for elements. If C and C' are two sets we will write C c C' in case every element of C is also an element of C'. For example if C comprises the elements I, 3, 5, 7 and C' comprises 1, 3, 5, 7, 9, ll, then we can write Cc C'.

The notation C u C' will denote the union of the sets C and C', that is, the set of all elements which are in C or in C'. For example if C comprises the elements I, 3, 7 and C' comprises 3, 5, 9, 13, then the set C u C' comprises the elements I, 3, 5, 7, 9, 13. In this example it is not true that C c C'.

Now the sets that we discuss come from the definition of an integral in § 5.2, namely

(5)

where a = Xo ~ x1 ~ X2 ~ ... ~ Xn-1 ~ Xn = b, and Xj-1 ~ f1 ~ Xj.

We write On for the set of real numbers a, x1, x2 , ... , Xn-l, b. Every set that we discuss will contain the numbers a and b. Furthermore, every set will contain only a finite number of elements, and no set will con­tain any number less than a or any number greater than b.

The upper and lower sums, Sn and Sn, of equations (3) and (4) of § 5.2; are now assigned new notation:

S(Cn) = (x,-xo)f(x,)+(x2-XI)f(x2)+ ... +(Xn-Xn-I)f(xn),

s(Cn) = (x1- xo)f(xo) + (x2- x,)j(x,) + ... + (xn- Xn-I)f(xn-l)·

Hence Theorem 2 of § 5.2 can be reformulated thus: If j(x) is a mono­tonic function. then

(6) lim {S(Cn) -s(Cn)} = 0. n~ro

In general if C is any set of numbers including a and b say

(7) G : a, ZJ, z2, za, ... , Zr-1, b,

where

a = Zo ~ Zl ~ Z2 ~ za ~ ... ~ Zr-1 ~ Zr = b,

then we define s(C) and S(C) by the equations

(8) s(C) = (z,-Zo)f(Zo)+(z2-ZI)f(z,)+ ... +(zr-Zr-I)f(zr-1),

(9) S(C) = (z,-zo)f(z,)+(z2-ZI)f(z2)+ ... +(z,.-zr-I)f(z,).

Page 172: Calculus an Introductory Approach

164 CALCULUS: AN INTRODUCTORY APPROACH APPENDIX A

THEOREM 2. If f(x) is monotonic increasing over a ;'; x ;'; b, and if C c 0', then

s(C) ;<; s(C') and S(C) ?; S(C').

PROOF. It suffices to establish these results in case the set 0' con­tains exactly one more real number than does C. For then we can work from C to C' in general by bringing in one new element at a time, step by step from C to C'.

So let us suppose that C is as given in (7), and let us write 0' as the same set with the addition of one new element u,

0' : a, z1, zz, za, ... , Zf, u, Zf+l, .•. Zr-b b,

with ZJ ;'; u ;'; ZJ+l· Then s(C') is the same as s(C) except that the term (ZJ+l -ZJ)f(z;) is replaced by

(u- ZJ)f(z;) + (ZJ+r-u)f(u).

Thus to prove that s(C) ;'; s(C') we must establish that

(10) (ZJ+l- ZJ)f(zJ) ;'; (u- zJlf(zJ) +(ZJ+r-u)f(u).

Bringing the terms involvingf(zj) together on the left, we see that (10) amounts to

( 11)

Now ZJ+r -u is non-negative, and f(zJ) ;'; f(u) since f is a monotonic increasing function, and so (11) and (10) are established.

As to S(C) ?; S(C'), the proof is much the same. In place of (10) we want to prove

(ZJ+l-ZJ)f(ZJ+l)?; (U-ZJ)f(u)+(ZJ+l-u)f(ZJ+l),

which reduces to

(u- ZJ)f(zJ+r) ?; (u- ZJ)f(u).

This can be established in the same way as was (11).

There is a simple geometric in­terpretation of (10), shown in Figure A.l. The functional values f(z1),j(u), andf(zJ+l) are represen­ted by the lengths AD, BG, and CK. Inequality (10) simply asserts that the area of the rectangle ACFD is not more than the sum of the areas of the rectangles ABED and BCHG.

A B u

FIG. A.l

c

Page 173: Calculus an Introductory Approach

APPENDIX A EXISTENCE OF THE DEFINITE INTEGRAL 165

Next consider once again the definition (5) of the integral. For each positive integer n, the set On consists of n + l points a, xr, x2, ... , Xn-b b. Define On* as the union of all sets 0 1, 0 2, ... , On, thus

(12) On* = Or U 02 U 03 U ... U On.

In other words, On* consists of all numbers in all the sets Or, 02, ... , On. It follows at once that

0 1* c 0 2* c 0 3* c C4* c ... c On* c ... ,

and so from Theorem 2

(13) s(Or*) ;;; s(02*) ;;; s(03*) ;;; s(04*) ;;; ....

In order to apply the Axiom on Bounded Sequences, we want to show that the sequence (13) is bounded. Every term in (13) is of the form (8), and since j(x) ;;; j(b) for all x satisfying a ;;; x ;;; b, we see that

(zr- zo)f(zo) + ( z2- zr)f ( zr) + ... + (zr- Zr-r)f (zr-r)

;;; (zr- zo)f(b) + ( Zz- zr)f(b) + ... + (zr- Zr-r)f(b)

= j(b)[(zr- zo) + (z2- zr) + ... + (z,- z,_,)J

= f(b)[zn- zo] = f(b)[b-a].

Now f(b) and b-a are fixed, and so the terms of (13) are bounded above. Hence s(On*) tends to a limit, say k, as n tends to infinity.

Next we observe that

s(On) ;;; s(On*) ;;; S(On*) ;;; S(On)·

The first and last of these follow from (12) and Theorem 2. The middle inequality can be seen at once from (8) and (9), since j(x) is monotonic increasing. Applying equation (6) and Theorem 1 with bn, an, and Cn replaced by s(On), s(On*) and S(On), we conclude that as n -> oo,

s(On) -+ k, S(On) -> k.

Reverting to our earlier notation, we can rewrite these in the forms Sn -> k, Sn -> k. Then by Theorem 1 of § 5.2 we have Sn ;;; a,. ;;; Sn, where an is defined by

n an = 2 (xj-Xj-l)j(gl)·

j~l

Applying Theorem 17 of § 2.4 we conclude that an -> k. Thus we have established the existence of the integral because it is, by equation (5), the limit of "n·

Page 174: Calculus an Introductory Approach

ANSWERS TO SOME PROBLEMS

Section 1.1

2. (a) t (b) -% (c) (b-2)/(a-1) (d) 0

3. 7/2 6. 15 8. -3

11. (-4, -ll) 14. (b-d)f(a-c) 15. dfc

l. 12 3. 2

Section 1.2

Section 1.3

Section 1.4 These are incorrect: 2, 6, 11, 13

Section 1.5

l. 500,500 3. 297,925 7. }n3- -§-n2+ifn

10. n(n +I )(n + 2) 12. tn2(n +I )2

l. ·~

3. -i 5. ~

Section 1.6

Section 2.1 8. T ~ 42 and T ~ 82

!66

Section 2.2

These will do:

3. lOO < X < 200; 200 > x > lOO; lx-1501 < 50; 50 > lx-1501

5. These will do: 70 ;;; T ;;; 74; 74 ;;; T ;;; 70; IT-721 ;;; 2; 172-TI ;;; 2

7. These will do : lx-yl ;;; •; ly-xl ;;; •; •;;; lx-yl; •;;; ly-xl; - e ~ x-y ~ e; - e ~ y-x ~ e; E ~ x-y f;; -e; e ~ y-x f; -e.

8. (a) -6<x<6 (c) -7 < x-3 < 7

or -4 < x < lO

9. (a) lxl < 9 (c) lx-221 < 8

Section 2.3 l. (a) 0 (b) 0 (c) 1 3. (i) -a (ii) l+a

Section 2.4

l. (a) 2 (b) 4 (c) 0 3. (i) 3a (ii) b- 2a

Page 175: Calculus an Introductory Approach

ANSWERS ANSWERS TO SOME PROBLEMS 167 ____ __::_::_:

Section 2.5

1. (a) 0 < x < oo (b) 0 ;;; X < 00

2. (a) Domain-2 ;;; x;;; 2, range 0 ;;; f(x) ;;; 2

(b) Domain -v2;;; x;;; v2,

range 0 ;;; f(x) ;;; 2 (d) Domain

-CO < X < eo, range 0 < f(x) < oo

3. /(1) = 3,/(2) = 6,/(4) = 24

Section 2.6

2. (a) 0 (e) 1 (g) v3/2

Section 2.8

1. 1 3. 1 and 3 5. 0 9. (a) -1 (d) t (e) -2

Section 3.1

7. 1

Section 3.3

1. t' lf-'\ 42 3. lo

Section 3.4

1. (a) 'J- (c) ~l (e) - \~

Section 3.5

1. (a) ~.~ (c) l,± 2. (a) 1 (c) H

Section 3.6

1. (a) (2- v3)/2 (c) 2 3. (a) 2 (c) (2- v2)/2

Section 3.7

2. 1"(2~''- 3r2k +k3) 3. !prr2h

Section 4.1

1. (a) dyjdx = 2x +I (c) dyjdx = 6x2

3. dyfdx = 4x3

5. (a) (0, 0) and(},,',) (b) (0, 0) and (3, 27)

7. Yes,atthepoint(-4, -64)

Section 4.2

dy . sin(x+6.x)-sinx 3. - = hm -'-----'--

dx Ax~o 6.x

Section 4.3

1. (a) dy - = 14x dx

(c) dy - = 14x-4 dx

(e) dy - = -3x-4

dx

(g) dy -= 4 dx

dy 3 5. (a) -=--

dx (3 + x)2

dy 6x-x4 (c)

dx (x3+ 3)2

dy 3-x (e)

dx 2yx(x+ 3)2

Section 4.4

1. (a) y' = -cosecx cotx

(c) y' = -cosec2x

(e) y' = x cosx-sinx

x" (g) y' = 2 sinx cosx

(i) y' = -2sinxcosx

Page 176: Calculus an Introductory Approach

168 CALCULUS: AN INTRODUCTORY APPROACH ANSWERS

2. (a) sec2x (c) -sinv

3. (a) y' ~ sinx+x cosx

x cosx-sinx (c) y' ~

Section 4.5

1. (a) y' ~ - 7 sin 7x

1 (c) y' ~ 2v'x+ 1

(e) y' ~ 18x2(2+x")'

(g) y' ~ -2 cosx sinx

X ( i) y' ~ ---,-------,--,­

(a2 + x2)1

1 5.~==

2v'x+ 3

Section 4.6 3. None

2 4. (a) y'- -­

- 1 +4x2

2 arc tanx (c) y' ~

1 +x2

Section 4.7

1. (i) c ~ 1 (iii) c ~ t (v) c ~ -~

Section 4.8

1. (a) y is a minimum if x ~ 2 (c) y is a maximum if

x = -3, a minimum if x~2

(e) None (g) y is a maximum if

x = - 2, a minimum if X~ 0

(i) y is a minimum if X ~ -v'9j2

2. (a) x ~ k1

(c) x = -}(k1 +k2+ks) 3. pj4 by p/4 5. cj2 and c/2 7. (a) x = 1.l-, y = l 3Q.

(b) X ~ l['-, y ~ j}

Section 4.9

1. r ~ 5j-v":;;, h ~ 1oj-v';; 5. 6 inches by 8 inches

7. r ~ V2/3, h ~ v' 4/3

9. v11

Section 5.2 1. Sn ~ .8,' Sn ~ 1.2,

F(b)-F(a) ~ 1 3. Sn ~ .95, Sn ~ 1.05,

F(b)-F(a) ~ 1 5. Sn = .855, Sn ~ 1.155,

F(b)-F(a) ~ 1

Section 5.3

3. (a) 80 (c) 2 (e) CL3"

(g) 1 (i) t (k) 1 (m) "'/6

Section 5.4

1. !x2 +c

3. 2x2-2x3+c

-1 5. -+c

2x2

7. arctanx+c.

9. -cosx+c

11. -I cos2x+c

Page 177: Calculus an Introductory Approach

ANSWERS ANSWERS TO SOME PROBLEMS

Section 5.5 4. (a) b-a ~ I, m~"},

M~ I (c) b-a~2,m~,~, M~ l:

(e) b-a ~ 7Tj6, m~ lt M~ VS/2

(g) b-a ~ 1Tj2, m~ 0, M~ I

Section 6.1

2. (a) 0 (b) I (e) 0

Section 6.2 1. .I9867 3 .. 99863

Section 7.1

1. L(5)-L(2), or L(5/2) 9. s: (dxfx)

Section 7.3 1. e, e2, e3. 3. (a) y' ~ 1fx

(c) y' ~ 1/(x+ I) 5. y' ~ 3/x 7. y is a minimum if x ~ I

Section 7.4 5. f(x) ~ 4ex

Section 7.5

2. (a) e (c) e 3. (a) e (c) v'e (e) ~e

Section 7.6 1. -.I054

Section 7.7 1. (a) .4771 (c) .7782 2. (a) 1.9459 (c) 2.3979

Section 8.2

3. .0997 5. 1.1071

Section 8.4 3. I5.5 milligrams

Section 8.5 1. 441.5 lb., 162.4 lb.

Section 8.6

2. (0, !), (0, f), (f, 0)

3. 1.63 5. 1.09 7. 1.73

Section 8.7

I69

Page 178: Calculus an Introductory Approach
Page 179: Calculus an Introductory Approach

INDEX

absolute value, 21 acceleration, 147 angle measure, 35 antiderivative, 107 approximation notation1 117 arc tan x, 88, 89, 143 area, 12, 15, 16

under parabola, 12 axiom on bounded sequences, 162

bounded sequences, 162

calculation, see computation chain rule for differentiation, 82 common logarithms, 130

relation to natural logarithms, 142 computation, of logarithms, 141

of,., 144 of trigonometric functions, 117

continuous functions, 46, 72 cosine (cos x), 80, 116 cylinder, minimum surface area of, 4,

96

derivative, 68 ff. chain rule for, 82 definition of, 71 notation for, 72 of a sum, 75 of a product, 75 of a quotient, 76 of arc tan x, 89 of cos x, 80 of em, 131 of log x, 130 of sin x, 79 'of tan x, 81 of X11

, 76, 87 differentiable, 71 differential equation, 151

171

differentiation, 68; see also derivative domain of a function, 34

e, base of natural logarithms, 129 as a limit, 137 value of, 135

ex, derivative of, 131 series expansion of, 131, 134

equations, solution of, 157 error, 117 escape velocity, 146 exponential function, 120 ff.

factorial notation, 113 focus of a parabola, 154 function, 3.3

continuous, 46, 72 differentiable, 71 domain of, 34 exponential, 128 ff. inverse, 84 limit of, 36, 39 logarithmic, 121, 130 monotonic, 86 piecewise monotonic, 104 range of, 34 trigonometric, 79, 88, 116

functions, with same derivative, 107 with zero derivative, 106

fundamental theorem, 103, 104

graph, of y = x2, 5 of cos B, 41 of f(x) ~ [x[, 72 of y ~ x, 75 of y ~ c, 75 of y ~ VX,78 of y ~ x', 85 of y = tan x, 88 of y = arc tan x, 89

Page 180: Calculus an Introductory Approach

~17~2~--------------------~INDEX

graph (cont.) of 11 ~ 1/v, 122 of y = log x, 127 of y = e"', 128

gravitation, 146

indefinite integral, 107 inequalities, 18

for integrals, 108 inflection point, 70 integer, 7 integral, 52

evaluation of, 56-64, 105-108 existence of, 103, 104, 161 indefinite, 107 inequalities, 108 interpretation as an area, 58 of sin x, 62 of tan x, 64, 131

inverse function, 84 inverse square law, 147 inverse trignonometric function, 88

law of gravitation, 146 limit, 6, 25

fore, 137 of a function, 36, 39 of a sequence, 25 of a sum, 48 of (sin e)le, 36-39 of cn!n!, 113 of rn, 112 of (1 + 1/n)n, 137 theorems on, 27, 29, 42, 112

logarithm, 121 ff. common, 130 computation of, 141 natural, 130 properties of, 125 series expansion of, 140

maxima and minima, 92 applications, 96-99

mean value theorem, 90 minimum principle, 98 minimum surface area, 4, 96 mixing problem, 152 monotonic function, 86

natural logarithm, 130, 142 Newton's method, 157

notation, for a sum, 7 for approximate values, 117 for evaluation of integrals, 104, 105 for factorial, 113 for logarithms, 130 for sets, 163 see also symbols

ordinate, 50

parabola, focusing property, 154 7T, computation of, 144 piecewise monotonic function, 104

radian measure, 35 radioactive decay, 150 range of a function, 34 real number, 17 refraction of light, 97

sequence, 24 bounded, 162 limit of, 25

series, convergence of, 119 partial sums of, 119

series expansion, 114 for arc tan x, 144 for cos x, 116 for ex, 134 for log(!- x), 140 for sin x, 116

sine (sin x), 79, 116 slope, 1, 2

of a muve, 2, 5, 69 Snell's law, 97 sphere, volume of, 65 subsequences, 114 snm of squares, 10 symbols, J f(x)dx, 51, 104, 105

dy!dx,69 y', 72 f'(x),72 11!' 113 "',117 see also notation

tan· x, 81 trigonometric functions, 79, 88, 114, 143

computation of, 117

velocity of escape, 146 volume of a sphere, 65

Page 181: Calculus an Introductory Approach

I

'