CALCULUS F. M. AND C. AN ELEMENTARY TREATISE ON CALCULUS A TEXT BOOK FOR COLLEGES AND TECHNICAL SCHOOLS WILLIAM S. FRANKLIN, BARRY MacNUTT AND ROLLIN L. CHARLES OT LEHIGH XJNIVBHSITT SOUTH BETHLEHEM, PA. PUBLISHED BY THE AUTHORS 1913 All rights reserved o(\ o 03 ^7 Copyright, 1913 By Franklin, MacNutt and Charles Set up and electrotyped. Published April, 1913 PRESS or THE NEW ERA PRINTING COMPANY LANCAbTER. PA. PREFACE. 1. We believe that the most important thing in the teaching of calculus is to lead the student to a clear understanding of principles. Therefore our chief endeavor has been to develop the subject as simply and as directly as possible. 2. We fully appreciate the importance of extensive practice in the handling of algebraic transformations. Therefore we have given an adequate collection of formal problems in differentiation and integration. 3. We are convinced, however, that it is a pedagogical mistake, in a text book on calculus for young men, to break the thread of the textual discussion by unnecessary algebraic developments and by large and frequent groups of purely formal problems. We believe indeed that the proverbial unintelligibility of calculus is to a very great extent a psychological consequence of this almost universal and really hideous feature. Therefore we have arranged the greater portion of our formal problems in an appendix. 4. Nearly every elementary science text in existence carries a false suggestion of completeness and finality, and there are two things which young men should understand in connection with their study of the mathematical sciences. The first is that such study is exacting beyond all compromise, involving as it does a degree of constraint which it is beyond the power of any teacher greatly to mitigate. And the second is that the completest science stands abashed before the infinitely complicated and fluid array of phenomena of the material world, except only in the assurance which method gives. We hope that this elementary treatise on calculus may prove to be sufficiently definite and exacting to be useful; next to this there is nothing we coul3Vish to have more in evidence from beginning to end than its incompleteness. 273389 VI PREFACE. In order to emphasize the incomplete character of this text we have introduced references to more complete treatises throughout the text and we have given in Appendix C a carefully selected list of treatises on mathematics and on mathematical physics. Teachers who use this text should, we believe, direct the students* attention to this appendix. In our brief reference to the infinitesimal method in article 20 we do not wish to be thought of as taking sides in the old contro- versy as between the " method of limits " and the " method of infinitesimals" in calculus. Article 20 is unquestionably falla- cious as it stands ; and the same may be said of the discussion of divergence and curl in articles 126 and 129. Indeed articles 20, 126 and 129 may be characterized as mere plausibilities, and the harder one tries to understand them the more vague and unintel- ligible they become. The fact remains, however, that the infini- tesimal method contributes very greatly to directness and sim- plicity of speech in the discussion of physical problems, and the idea of infinitesimals is therefore used throughout this text. Any one who is disturbed by the element of easy plausibility that is involved in the s raight-forward use of the infinitesimal method in the discussion of physical problems should heed the advice given by D'Alembert to a young student, "Go ahead, young man, go ahead! Conviction will come to you later." "The absolute requisites for the study of this work are: a knowledge of elementary algebra to the binomial theorem (accord- ing to the usual arrangement), plane and solid geometry, trigono- metry, and the most simple parts of the usual apphcations of algebra to geometry." This was said by De Morgan in the preface to his great treatise on Differential and Integral Calculus (London, 1842), in com- parison with which this book is a primer. Franklin, MacNutt and Charles. South Bethlehem, Pa., March 22, 1913. CALCULUS. In the study of phenomena which depend upon conditions which vary in time, that is, upon conditions which vary from instant to instant, it is necessary to direct the attention to what is taking place at an instant; or, in other words, to direct the attention to what takes place during a very short interval of time; or, borrowing a phrase from the photographer, to make a snap-shot, as it were, of the varying conditions. In the study of phenomena which depend upon conditions which vary from point to point in space, the attention must be directed to what takes place in a very small region. This paying attention to what takes place during a very short interval of time or in a very small region of space does not refer to observation but to thinking, it is a mathematical method and it is called calculus. Thus the density of a body is its mass divided by its volume. This definition, and indeed the idea of density, appHes only to a homogeneous substance. To apply the idea of density to a non-homogeneous substance one must think of a very small portion of the substance. The same thing is true of every measurable property of a substance. Thus the idea of elastic strain as a measurable effect is easily established when every part of a body is similarly strained as in the case of a stretched rod, but to apply this precise idea of strain to a bent beam or to a twisted rod one must think of a small portion of the beam or rod. Two distinct methods are involved in the directing of the attention to what takes place during very short intervals of time or in very small regions of space, as follows: (a) The meth gdj:iljii^[srmtial cakiihisv - A phenomenon may be prescribed as a pure assumption, and the successive instantaneous viii CALCULUS. aspects may then be derived from this prescription. Thus we may prescribe uniform motion of a particle in a circular path, and then determine the acceleration of the particle at each instant. Or, a condition in space may be prescribed as a pure assumption, and the minute aspects may then be derived from this prescrip- tion. Thus we may prescribe the distribution of temperature along a rod, and then determine the temperature gradient at each point from this prescription. (6) The method of integral calculus. It frequently happens that we know and can easily formulate the action which takes place at a given instant or in a small region of space. The problem then is to build up an idea of the result of this action throughout a finite interval of time or throughout a finite region of space. Thus the acceleration of a body may be known at each instant, and from this knowledge we may find the velocity gained and the distance traveled in a finite interval of time. Or, consider a disk rotating at a given speed. It is easy to establish a formula for the energy of a very small particle of this rotating disk, and it is then possible to derive an expression for the energy of the entire disk. **% TABLE OF CONTENTS. PAGE Preface and Introduction v-x Chapter I. A General Survey of Differential and Integral Calculus 1-37 Chapter II. Formulas for Differentiation and Inte- gration 38r 73 Chapter III. Integration 74- 86 Chapter IV. Partial Differentiation and Integration 87-108 Chapter V. Miscellaneous Applications of Calculus 109-152 Chapter VI. Expansions in Series 153-167 Chapter VII. Some Ordinary Differential Equations 168-189 Chapter VIII. The Partial Differential Equation of Wave Motion 190-209 Chapter IX. Vector Analysis 210-253 Appendix A. Problems 1-20 Appendix B. Table of Integrals 21- 28 Appendix C. Some Important Books on Mathematical Theory 29- 38 ix " In view of the acknowledged difficulty of calculus \the student must be willing to stop in his course luntil he can form exact notions and acquire precise ndeas.^' Augustus De Morgan. CHAPTER I. GENERAL SURVEY OF DIFFERENTIAL AND INTEGRAL CALCULUS. 1. Constant quantities and variable quantities. — Elementa ry algebra deals only wil^h quantities which do not change in value (constant quantities). For example let x be the quantity which subtracted from 12 leaves a remainder equal to Jx. That is, 12 — a; = ^x, whence x = S. The quantity x has the same value throughout this discussion and it is therefore called a constant quantity. Calculus is a branch of algebra and it deals not only with constant quantities but also and especially with quantities which change in value {variable quantities). Thus one of the simplest problems of calculus is to consider how rapidly x^ grows in value when x is imagined to grow at a definite rate. Variable quantities fall intQjthree^ fajxl^jd i s t i nct .classES,. naiaely : (a) Quantities which vary in time, (h) quantities which vary in space, and (c) quantities which are arbitrarily assumed to vary. An example of the first class is the varying temperature of the air at a given place as the seasons come and go. An example of the second class is the varying temperature of the air from place to place at a given time. Variables of the first and second classes are sometimes called natural variables to distinguish them from variables of the third class, examples of which are considered in Arts. 10 and 11. 2. Value at a given instant. Value at a given point. — Let y represent the amount of water in a pail into which a small stream of water is flowing. Evidently y is sl changing quantity. If the flow of water were to be shut off at any given instant there would be a definite quantity of water left in the pail, and, therefore, there is a definite amount of water in the pail at each instant even while 2 1 2 /'V r -' :.•:».• ^ . ■ CALCULUS. the inflow continues. Any changing quantity y has a definite value at each instant* A quantity which varies in space has a definite value at each point.* Thus everyone understands that there is a definite tem- perature at Philadelphia although it may be different from the temperature at New York or Washington. An iron rod with one end in the fire has a definite temperature at each point. 3. Increments and decrements. — It is frequently desirable to consider an increase (or a decrease) in the value of a variable quantity. Such an increase is called an increment, A decrease is called a decrement. For example, let y he & variable quantity. An increment of y is usually represented by the symbol Aj/. When Ay is negative it is a decrement. This symbol Ay does NO T mean A multiplied by y, it is a single symbol, and it may be read increment-of-y or delta-y. The latter is preferable because of its brevity. The Greek letter A used as a prefix always means increment of. 4. Rate of change of a quantity which varies in time. — Consider a pail into which water is flowing in a small stream. Let y be the amount of water in the pail. Then y is evidently a changing quantity. Consider the amount of water that flows into the pail during a short interval of time At; this amount of water is evi- dently the increment of y during the short intervaj.-of timCj and Ay it is to be represented by the symbol Ay. The quotient -j At is called the average rate of change of y duHng_ihe -interval — ^."f^ If the inflowing stream of water is constant, say, always 2 cubic inches of water per second, then if At is chosen smaller and smaller, the value of Ay will be proportionately smaller and smaller, and Ay the value of the quotient -^ will be exactly 2 cubic inches per second whatever the duration or length of the time-interval A^. * This may not be true of a discontinuous variable. t Thus if y is the amount of money a man has saved, then, if he saves $20 more during 30 days, we have Ay = $20 and At = 30 days, and the aver- age rate of saving during the 30 days will be 66 1 cents per day. DIFFERENTIAL AND INTEGRAL CALCULUS. 3 If the inflowing stream is variable, then the value of the quotient ~ will not be the same for different values of Ai, but the amount of water which flows into the pail during a very short interval of time will be very nearly proportional to the interval. For example, a certain amount of water W flows into the pail during a particular one-thousandth of a second. Imagine this particular one-thou- sandth of a second to be divided in halves; then only a very Httle more than ^W will flow into the pail during one of the half- thousandths of a second and a very httle less than |TF will flow into the pail during the other half-thousandth of a second. // a shorter and shorter interval of time he taken the amount of inflow of water will be more and more nearly in EXACT proportion to the duration of the interval. This means that the quotient -Tj- approaches a definite limiting value as At and Ay both ap- All proach zero; and this limiting value of -rf is called the rate of change of y at a certain instant or the instantaneous rate of change of y. All dii The limiting value of -^ is always represented by -^ which means the rate of change of y at a given instant. Nearly everyone falls into the idea that such an expression as 10 feet per second means 10 feet of actual movement in an actual second of time, but a body moving at a velocity of 10 feet per second might not continue to move for a whole second or its velocity might change before a whole second has clasped. Three cubic inches per second is the same rate of inflow of water into a pail as 2,070 cubic yards per year, but to specify the rate of inflow in cubic yards per year does not mean that a whole cubic yard of water flows into the pail nor does it mean that the inflow continues for a year. A man does not need to work for a whole month to earn money at the rate of $60.00 per month, nor for a whole day to earn money at the rate of $2.00 per day. A falling body has 4 CALCULUS. a velocity of 19,130,000 miles per century after it has been falling for one second, but to specify its velocity in miles per century does not mean that it falls as far as a mile nor that it continues to fall for a century ! The units of length and time which appear in the specification of a velocity are completely swallowed up, as it were, in the idea of velocity; and the same thing is true of the specifica- tion of any rate. 5. Continuous variables and discontinuous variables. — A quan- tity which changes by sudden jumps is called a discontinuous variable. For example the amount of money one has is a dis- continuous variable, because a debt is made on the instant that one decides to accept a purchase; that is to say, money is spent in lumps, and a lump of money is spent during an indefinitely short time. The amount of money spent during an interval of time does NOT become more and more nearly proportional to the interval as the interval grows shorter and shorter, and consequently it is meaningless to speak of the rate of spending money at a given instant. If y is a discontinuous variable and if the time-interval A^ happens to include a jump in the value of y, then the value of Ay remains finite as At approaches zero and the quotient -r^ approaches in- finity. The rate of change of a discontinuous variable at a given instant is unthinkable. The amount of water in a pail, as considered in Art. 4, is an example of what is called a continuous variable. If y is a con- tinuous variable and if Ay is the increment of y during the time- interval At, then Ay becomes more and more nearly proportional to At as A^ approaches zero; that is to say, the quotient -J a,pproaches a definite limiting value as At approaches zero, and this limiting value of -~ is called the instantaneous rate of change zir of y. In the study of calculus we deal almost exclusively with con- tinuous variables. DIFFERENTIAL AND INTEGRAL CALCULUS. 5 6. Example showing determination of instantaneous rate of change. — Let us assume that the amount of water in the pail in the discussion of Art. 4 is proportional to the square of the time t which has elapsed since a chosen instant. Then we may write y = kt' (1) where /: is a constant. A moment later t has increased and of course y has increased also. Let us represent the new value of t by t + At and the new value of y by y -\- Ay. Then, since equation (1) is assumed to be true for all values of t and y, we have y + Ay = k{t + Aty or y + Ay = kt^ + 2kt'At + h{AtY (2) Subtracting equation (1) from equation (2) member by member, we have: Ay = 2kt'At -f- k{AtY (3) Whence, dividing both members by Aty we have: ^ ^2kt + k'At (4) Now it is evident that k-At approaches zero as At approaches Av zero, and therefore the limiting value of -ry (as At approaches dv Av zero) is 2kt. Hence, writing -~ for tY.e limiting value of -r^, we have : | = 2« (5) That is, the rate of change of y is at each instant equal to 2kt. Observation or thinking; which? — The discussion, in Art. 4, of the rate of increase of the amount of water in a pail presents a serious difl&culty. How is one to know the increment of water which occurs during an indefinitely short interval of time? One 6 CALCULUS. certainly cannot measure it. Observation and measurement are entirely usele!er inch. 6. (o) Find the average temperature gradient along the iron rod of problem 4 between x = and x = 20 inches. (6) Find the average temperature gradient between x = 10 inches and X = 11 inches. Ans. (a) 12 d^;rees per inch, (6) 12.6 degrees per inch. 9. Arbitrary variations. — One of the most important things in calculus is to consider how fast such an expression as x* or x* or sin X changes when x is arbitrarily assumed to grow steadily in value. Such -variations may. be called arbitrary variations. The ver>' great importance of arbitrary- variations will be understood when the student reaches Arts. 22 and 23. 10. A purely algebraic example of an arbitrary variation. — Having given the equation y = aa^ (1) let it be required to find the limiting value of the quotient — when the arbitrary increment of x, namely Ax, approaches zero. Writing y -{- Ay for y and writing x + Ax for x in equation (1) we have y + Ay = a(x + Ax)* (2) or y + Ay = ax» 4- 2axAx + a(Ax)* (3) whence, subtracting equation (1) from equation (3) member by 10 CALCULUS. memoer, we have Ai/ = 2aa:-Aa; + a(Ax)« (4) and dividing both members by Ax we have ^ = 2ax + a-Ax (6) Ax But a 'Ax approaches zero when Ax approaches zero, and there- fore it is evident that the limiting value of -r^ is 2ax. The Ax . Ay dv limitmg value of -r^ is always represented by ~. Therefore we i^c dx have: g = ^ (6) 11. Two more purely algebraic examples of arbitrary variations. — (a) Having given the equation y = as? (1) let it be required to find the limiting value of -^ when the arbi- trary increment of x approaches zero. Writing y -\- Ay for y and writing x + Ax for a; in equation (1) we have y-^Ay = a{x + AxY (2) or y-\- Ay ^ a3? -\- Sax^-Ax + 3ax(Ax)2 + a{^xY (3) whence, subtracting equation (1) from equation (3) member by member, we have Ay = 3ax2.Ax + 3ax(Ax)2 + a(Ax)« (4) or ^ = 3ax2 + 3ax-Ax + a{Axy (5) Now 3ax-Ax and a{Axy both approach zero when Ax approaches Ay eero, and therefore it is evident that the limiting value of ~ is DIFFERENTIAL AND INTEGRAL CALCULUS. 11 Sax^; that is, # _ dx (6) Having the equation 3ax2 (6) y = l (7) let it be required to find the limiting value of ~ when the arbi- trary increment of x approaches zero. Writing y -\- Ay for y and writing x + Ax for x in equation (7) we have y + ^y-^^ («) Subtracting equation (7) from equation (8) member by member, we have A2/ = -^ - ? (9) ^ x + Ax X Reducing the two fractions to a common denominator we have: whence Ay (11) Ax x^ -\- x-Ax but x^ -\- X'Ax approaches x'^ as its limit when Ax approaches zero, and therefore the limiting value of ■— is ^ ; that is : dx X' ^^^^ dy The derivative — is sometimes called the rate of change of y with respect to x. 12 CALCULUS. PROBLEMS. The following problems can be solved by the methods employed in Arts. 10 and 11. In each case £uid the rate of change of y with respect to x. l.y^ax, g = a. 2. 2/ = ax», ^ = 5ax^, 3. y = ms? + w, -p = 27nx. 4. 2/ = oar^ + 6a; + c, ^ = 2ax + 6. 5. ^ = ax^ + 6x2, ^ = 3ac' + 2&a;. - a di/ 2a _ g dy _ — a(2hx + 1) **^~6a;'» + x' 5i n 10.2/ = =^, * V '^ 6 — x' da; 11. y ^ 2X^-5 ^ 12. 1/ X dy ^ (x-l)2' d!x {x-iy (6X2 + 3.)2 9 (x + sr ab (6 - x)2- 9 (x + 2)2- x + 1 12. Functions. — When a spring is subjected to a stretching force Ff a certain elongation e is produced as shown in Fig. 2. For DIFFERENTIAL AND INTEGRAL CALCULUS. 13 every value of e there is a definite corresponding value of F. When two quantities are associated in this way either one is said to be a un stretched spring Fig. 2. The elongation of the spring is a function of the stretching force. function of the other. Thus the elongation e produced by a given force F in Fig. 2 is a function of F; or the force F re- quired to produce a given elongation e is a function of e. The 1 1 K -^ i j| w 1 1 1 1 area^lw Fig. 3. The area of a square is function of length of side. Fig. 4. The area of a rectangle is a function of length and width. meaning of the word function is further illustrated by Figs. 3, 4, 5, 6, 7 and 8. 13. Dependent and independent variables. — If I in Fig. 3 is assumed to be arbitrarily variable it is called an independent variable, and the area of the square is called a dependent variable because its value is determined or fixed when the value of I is given. The quantity of air pumped into the cylinder in Fig. 7 may be as much as one pleases, the temperature of the cylinder 14 . CALCULUS. and air may be increased or decreased at will, and the volume of the cylinder may be changed by moving the piston. Therefore tr- thermometer ^ "l-cloaed • '. : vessel •'•' containing ': air . • - Fig. 5. The pressure of the air is a function of the amount of air pumped into the tire. Fig. 6. rhe pressure of the air in the vessel is a function of the amount of au- pumped into the vessel and of the temperature. the three quantities, (a) quantity of air, (6) temperature of the air, and (c) volume of space occupied by the air are called independent variables, and the pressure of the air is called a dependent variable, thermometer pressure gauge piston cylinder containing \ air SiLnp Fig. 7. The pressure of the air in the cylinder is a function of (a) the amount of air pumped into the cylinder, (6) the temperature, and (c) the volume of the space occupied by the air. DIFFERENTIAL AND INTEGRAL CALCULUS. 15 given curve because its value is fixed or determined when the quantities (a), (6) and (c) are given. From this discussion it is evident that a dependent variable may be a function of one, two, three or more independent variables. 14. Natural functions and artificial functions. — We have hereto- fore taken the point of view that two quantities must be connected physically if either is to be a function of the other. Such a function may be called a natural func- tion. It is, however, of very great importance to consider func- tional relations which grow out of Or are expressed by algebraic equations. Such functions may be called artificial functions. Some idea of the importance of artificial functions may be obtained by looking back at Articles 4 and 6, and 7 and 8. Natural functions can be in many cases expressed algebraically^; indeed no function can be handled in calculus unless the function is expressed algebraically. 15. Tabulation of a function. — ^When one quantity is a function of another it is often convenient to express the functional relation by means of a table giving pairs of corresponding values of the two quantities. An ordinary table of logarithms is such a table, a table of sines or cosines is such a table, the mortality table* used in life insurance is such a table. When a functional relation is determined by experiment the * If one were to consider all men in the United States who are now forty years of age and keep a future record of their deaths one would find that their average age at death would be, say, sixty-five years. ^This is expressed by saying that the expectancy of fife at forty years is twenty-five years, because on the average a man forty years old may expect to live twenty-five years. The expectancy of life is, of course, less and less with increasing age, and there is a definite expectancy of life corresponding to each age. I That is, expectancy of life is a function of age. 7 Fig. 8. The ordinate of a point on a curve is a function of the abscissa of the point. 16 CALCULUS. Fig. 9. observed quantities are always arranged in tabular form. For example Fig. 9 represents an am- meter arranged to measure the electric current flowing through a coil of wire, and a spring scale ar- ranged to measure the force with which an iron plunger is pulled into the coil; and the accompanying table shows a series of observed values of current in amperes and the corresponding pulls in pounds. OBSERVED RELATION BETWEEN CURRENT AND PULL IN FIG. 9. Current in Pull on Plunger Amperes. in Pounds. 1.0 3.9 2.0 23.6 S.0 40.0 10 48.1 a. 16. Graphical representation of a fimction. — The relation be- tween the pull of the spring and the reading of the am- meter in Fig. 9, as shown in the table of Art. 15, may be represented graphically by a curve of which the ab- scissas represent ammeter readings and of which the ordinates represent corre- sponding readings of the spring scale. Such a curve is shown in Fig. 10. In the same way an algebraic func- tion may be represented graphically, represents the function y = x^. 50 40 30 20 10 ^ ^ /^ / / / / ^ [/ ' 345 amperes Fig. 10. Thus the curve in Fig. 11 DIFFERENTIAL AND INTEGRAL CALCULUS. 17 17. Derivative of a function. let A?/ be the increment of y due to an arbitrary incre- ment of X. Then the lim- iting value of the quotient -^ as Ax approaches zero is called the derivative of y mth respect to x or the rate of change of y with respect to X* When one wishes to speak in general terms of any func- tion of X and its derivative, the function may be repre- sented by f{x) or by <^(x) and the derivative of the -Let 2/ be a function of x and I X'OxiiB Fig. 11. function with respect to x may be represented hy f'{x) or by '{x). Meaning of a derivative. — Let i/ be a function of x and let x dy be assumed to grow steadily at a definite rate, then y grows -r- times as fast as x at each instant. For example, let y = x^; then dif ~ = 2x, according to Art. 10, and the following relations exist: y increases 2x{= 20) times as fast as x value X — 10, y increases 2a; (= 22) times as fast as x value X = 11, y increases 2a: (= 24) times as fast as x value X — 12, ^ etc., etc., while ; while ; while ; As another example, let y = -; then -^ X dx is passing through the is passing through the is passing through the etc. 1_ x^' according to Art. 11, and the following relations exist: * Sometimes also called the differential coefficient of y with respect to x. 18 CALCULUS. y decreases ^ ( ~ 4 ) as fast as x increases when x is passing through the value 2, y decreases -j ( = qJ as fast as x increases when x is passing through the value 3, etc., etc., etc. --> A negative value of ^ shows that y decreases as x increases. ax The derivative of a function of x is itself a function of x. — If y = dx^. then ^ = 2ax, according to Art. 10. In this case ax dti it is evident that -p is a function of x because it is equal to 2ax and it has a definite value for every value of x. If /(x) represents any function of x, then its derivative }'{x) is in general a function of X also. 18. Graphical representation of derivative. Slope of a curve at a point. — Consider the curve CCy Fig. 12. This curve represents Fig. 12. a definite function, that is, the ordinate ?/ is a definite function of the abscissa x, the point p being anywhere on the curve. If x is increased by the amount Ax, then y will be increased by the DIFFERENTIAL AND INTEGRAL CALCULUS. 19 amount A^ as shown, and the ratio -f- is equal to the tangent of the angle a. Draw the hne tt touching the curve at p as shown. As Aa; approaches zero the angle a becomes more and more nearly equal to the angle 6, in fact 6 is the limiting value of a, and, of course, the limiting value of -r^ is ~. Therefore, since tan a = — ^ , we must have tan ^ = -r^- That is, the function y dti being represented by the ordinates of a curve, the derivative -r- is represented by the steepness or grade of the curve at each point. 19. Derivative notation and differential notation. — Consider the function y = ax^ (1) of which the derivative is: I = ^ (2) dv Heretofore we have looked upon ^- as a single symbol the mean- ing of which is explained in Art. 17. Convenience of notation sometimes makes it desirable to write equation (2) thus: dy = 2ax'dx (3) This equation expresses the limiting relation between the incre- ments A^ and Ax, that is to say, the relation which is approached as Ay and Ax both approach zero; dy is called the differential of y (that is to say, the differential of ax^, because y here stands for ax^), and dx is called the differential of x. These two differentials may be thought of as indefinitely small increments of y and x respectively.* * When y = ax^ we have Ay = 2ax • Ax + ^(Ax)^ according to Art. 10. That is to say, the increment of y is not equal to 2ax times the increment of x unless both increments are indefinitely smaU. 20 CALCULUS. Or, if one wishes to think of dy and dx as having a physical dt meaning they may be thought of as abbreviated expressions for dx and -rr respectively, that is, dy may be thought of as the rate of change of y and dx may be thought of as the rate of change of x, PROBLEMS. 1. A pail is 10 inches in diameter, so that the volume of water in the pail is y = —^ Xj where x is the depth of the water in the pail. Find how fast water must be poured into the pail to cause the water level to rise at a velocity of 4 inches per second. Ans. IOOtt cubic inches per second. */t2. Water flows at a constant rate of 30 cubic inches per second into a metal cone of which the dimensions are as shown in Fig. p2. Find the velocity at which the water level rises (a) when x = b inches, and (6) when a; = 15 inches. Ans. (a) 2.715 inches per second, (6) 0.302 inch per second. €m t'c-- ^ 3. The sides of a square are growing at the rate of 5 inches per second. Find the rate of growth of the area of ^ the square (a) when the sides of the square are 10 inches and (6) when the sides of the square are 20 inches. Ans. (a) 100 square inches per second, (6) 200 square inches per second. 4. The radius of a circle is growing at the rate of 5 inches per second. Find the rate of growth of the area of the circle (a) when the radius is 10 inches and (h) when the radius is 20 inches. Ans. (a) lOOir square inches per second, (h) 2007r square inches per second. 5. The edges of a cube are growing at the rate of 5 inches per second. Find the rate of growth of the volume of the cube when the edges are 10 inches long. Ans. 1,500 cubic inches per second. Fig. p2. DIFFERENTIAL i^fe INTEGRAL CALCULUS. 21 m 9 of iR] - 6. Air is blown into a soapi«3ble at the rate of 10 cubic inches per second. Find the rate of Slrease of the radius of the bubble when the radius has the values (^ 1 inch and (6) 4 inches. Ans. (a) 0.795 inch per second, (6) 0.050 inch per second. -7. A man 6 feet tall walks at a speed of 4 miles per hour under a lamp which is 10 feetfrgmthe ground. Find how fast ^€-tip • ef- the man's shadow wa^eS-mien the horizontal distance from the lamp to the man is 20 feet. Ans. 6 miles per hour. 8. The two sides of a right angled triangle are 40 inches and x^ and the hypothenuse is y. The side z is growing at the rate of 5 inches per second. How fast is y growing when x = 30 inches? Ans. 3 inches per second. 9. The two sides of a right angled triangle are x and y, and the hypothenuse is 50 inches. The side x is growing at the rate of 4 inches per second. How fast is y growing (a) when x = 30 inches, and (6) when x = 40 inches. Ans. (a) 3 inches per second. (6) 5.33 inches per second. 10. The observed temperatures of a vessel of cooling water after 1 minute, after 2 minutes, and so forth are as follows: Elapsed Time in Minutes, ObserTed Temperaturea, t T. 92.0 1 85.3 2 79.5 3 74.5 5 67.0 7 60.5 10 53.5 15 45.0 20 39.5 Plot these values of t and T, using the best grade of squared paper, draw a smooth curve through the plotted points, draw tangents to the curve at the points corresponding to (a) T = 80°,, (h) T = 65°, and (c) T = 50°, and determine the rate of decrease of the temperature at each point by measuring the intercepts; of these tangents on the axes of reference. 22 CALCULUS. 11. Find for what values of x the function ?/ = Gx^ — 12x is an increasing function and for what values of x it is a decreasing function. Ans. 2/ is an increasing function when x is greater than 1, and a decreasing function when x is less than 1. 12. Plot the values of x and y^ where y = x^, using the best grade of squared paper, draw a smooth curve through the plotted points, draw tangents to the curve at the points corresponding to X = 1, x = 2 and x = 3, and determine the corresponding values of -p by measuring the intercepts of the tangents on the dii axes of reference. Compare the values of -— thus found with dxi the values as calculated from the formula -— = Sx^. ox 13. Plot the values of x and y where y = logio x. Take the values of logio x from an ordinary table of logarithms and use the following values of x: 2, 3, 4, 5, 6, 7, 8, 9 and 10. Draw a smooth curve through the plotted points, draw tangents to the curve at the points corresponding to x = 4, x = 6 and x = 8, and deter- dv mine the corresponding values of -^ by measuring the intercepts of these tangents on the axes of reference. Compare the values of dti -^ thus found with dx dy ^ logioe ^ 0.4343 ^ dx X X ' 20. Determination of limiting relation between Ay and Ax by Aiy consideration of infinitesimals. — The value of -~ is found in Ax Arts. 10 and 11 as the sum of a finite qvxintity and a quantity which approaches zero as Ax approaches zero. Thus when y = as? we found that -r^ = Zax^ + 3ax-Ax + a(Ax)2, in which Sax^ is a ZaX finite quantity and 3ax-Ax + a(Ax)^ is a quantity which ap- proaches zero as Ax approaches zero. In such a case it is very dti -^ thus found with the values as calculated by the formula DIFFERENTIAL AND INTEGRAL CALCULUS. 23 A?/ easy to see what the limiting value of -~- must be. In many cases, >oweve% it is desirable to find the limiting rela- tion between ^y and Ax when Ax approaches zero without deriving an expression for ~ . For example let y = a^. Then from Art. 11 we have: - - - -^ ^ _/■' ^ , Ay = Sax^'Ax + Zax{Axy + a{Axy (1) Both members of this equation approach zero when Ax approaches zero, and it would therefore seem rather difficult to determine the hmiting relation between Ay and Ax (when Ax approaches zero) from this equation as it stands. But when Ax is made as small as you please, thejM|(Aa;)2 is infinitely smaller than Ax, and {AxY is infi]||||fily sn^^r than (Ax)^. For example let Ax be a millionth of^Mpait^i^K (Ax)^ is a million-millionth of a unit, and (Ax)^ is^3fcpioB;^Blion-millionth of aunit. Therefore the terms 3ax • (Ax)^and ' ^J\x)^ become more and more nearly negligible in comparison with Sax^ • Ax as Ax grows smaller and smaller in equation (1), The limiting relation between Ay and Ax may be found by writing dx and dy for Ax and Ay to indicate that we have proceeded to the limit, and by dropping every term which contains the square or any higher power of Ax (or the square or any higher power of Ay). This gives dy = Sax^-c^x (2) In this equation dy and dx are as small as you please and they are called infinitesimals; but {dyY and {dxY being infinitely smaller than dy and dx are called infinitesimals of the second order J {dyY and {dxY being infinitely smaller than {dyY and {dxY are called infinitesimals of the third order, and so on. In any differential expression the lowest order infinitesimals, only, are significant; all terms containing higher order infinitesimals as factors may he dropped. 24 CALCULUS. Example 1. — Let it be required to differentiate the expression 2/2 = ax» (3) that is to say, let it be required to determine the relation between Ay and Ax when Ax is as small as you please Writing {y -{• Ay) for y and writing (x + Ax) for x, we have y^-\-2yAy-i- {Ay)^ = ax^ + Sax^-Ax + 3ax(Ax)2 + a(Ax)» (4) whence, subtracting equation (3) from equation (4) member by member, we have 2y'Ay -\- (Ay)^ = 3ax2-Ax + 3ax(Ax)2 -f- a{Ax)^ (5) from which the desired limiting relation is found by dropping second and third order infinitesimals, so that we have: 2y-dy = dax^-dx (6) The value of y from equation (3) may be substituted in equation (6), giving: 2^[a^-dy = Zax^'dx (7) which can be simplified by cancellation, giving: dy = j-^ax-dx (8) This dropping of higher order infinitesimals from differential expres- sions does not lead to merely approximate results, because in every case it is the limiting rela- tion between Ay and Ax which is to be determined, that is, the relation when Ay and Ax are both as small as you please. Equation (8), for example, is rigorously true. Example 2. Differential of the arc of a parabola. — The length s of the heavy portion of the curve cc, 'X^' DIFFERENTIAL AND INTEGRAL CALCULUS. 25) Fig. 13, from x = b tox = x is a function of x, and from the in- finitesimal triangle whose sides are dx, dy and ds we have ds = ^l{dxy + {dyy or *= >r+ ity-^ (9) (10) For example suppose the curve cc, Fig. 13, to be a parabola whose equation is: y = kx^ (11) then dy _ dx = 2kx which, substituted in equation 10, gives: ds = ^ll-\-4:k^x^'dx (12) (13) 21. Functions which have the same derivative. — The curve A, Fig. 14, defines i/' as a function of a;, and the curve B defines y y-axiB Fig. 14. as a function of x. The steepness or grade of curve A is every- where the same as the steepness or grade of curve B, that is, the derivative -j- is equal to the derivative — for each value of a;, 26 CALCULUS, or expressed as an equation we have dx dx (1) From the figure it is evident that the difference 1/ — y is everywhere equal to the constant quantity C. Consequently when equation (1) is true we have y' — y = a. constant (2) That is to say, two functions whose derivatives are equal have a constant difference. Examples. — ^When two men save money at the same rate there is a constant difference between the amounts they have saved; or, if they start even, their savings are equal. When two trains travel at the same speed they remain at a constant distance apart. The two functions ax^ and ax^ + h have the same derivative with respect to x, h being a constant. 22. Example showing the use of calculus. Work required to stretch a spring. — Let W be the work required to stretch a spring from condition A to condition B, Fig. 2. It is evident that W is a function of e, and it is desired to find an algebraic expression for this function; that is, it is desired to find an equation expressing W in terms of e. To do this the first step is to find an expression for dW the derivative -t~ by considering the amount of work ATF that "^ "'- /^ must be done to produce a slight additional elongation Ae. The force F in Fig. 2 is, according to Hooke's law, proportional to e; that is, F = ke (1) where A; is a constant. Imagine the force F to be increased by the amount AF so that the elongation would be increased by the amount Ae. Then the work done would be greater than F-Ae and less than {F -\- AF)-Ae because the average value of the force which acts DIFFERENTIAL AND INTEGRAL CALCULUS. 27 on the spring while the added elongation Ae is being produced is greater than F and less than F + AF. Therefore AW is greater than F • Ae (2) and AW is less than (F -\- AF) - Ae (3) whence, dividing by Ae, we have : aw -^ is greater than F and less than F + AF (4) Ae AW Therefore —— approaches F as its limit as AF (and also Ae) approaches zero. That is, dW or, using the value of F from equation (1), we have f=^ ,,, Now in Art. 10 it is shown that ax^ is a function whose derivative is 2ax. Therefore: ae2 is a function whose derivative is 2ae (7) Therefore, substituting k = 2a (or a = \k) in (7) we have: \ke^ is a function whose derivative is ke (8) The work W is also a function whose derivative is ke, according to equation (6). Therefore, according to Art. 21, we must have W = W + C (9) where C is a constant. To determine the value of the constant C we must know the value of W for some particular value of e, and of course we know that TT = when e = 0, that is, no work is required when the spring is not stretched at all. Therefore, 28 CALCULUS. substituting this pair of values of W and e in equation (9), we have = C (10) Substituting this value of C in equation (9) we have W = ike'' (11) which is the desired expression for the work W. 23. Another example showing the use of calculus. The area under a parabola. — The curve cc. Fig. 15, represents a parabola of which the equation is: y = px^ (1) and it is required to find an expression for the area A. Imagine the value of x in Fig. 15 to increase by the amount Ax^ X X — Fig 15. Fig. 16. then the increment of A is the area of the narrow strip in Fig. 16. The width of this strip is Ax and its height is y (= px^). There- fore whence A A = px^'Ax AA (2) (3) Now the width Ax in Fig. 16 must be infinitely small in order that the height of the strip may be considered to be equal to DIFFERENTIAL AND INTEGRAL CALCULUS. 29 y {= px^) . That is, we have already proceeded to the limit in writing down equation (2), and it is evident that equation (3) gives the AA limiting value of — . Therefore we have In Art. 11 it was shown that: ac^ is a function whose derivative is Soi* (5) Therefore substituting ^ for a we have : Ipa^ is a function whose derivative is px'^ (6) The area A is also a function whose derivative with respect to X is px^ according to equation (4), and therefore, according to Art. 21, we must have: A = ip:x^+C (7) where C is a constant. To determine the value of the constant C we must know the value of A for some particular value of x. An inspection of Fig. 15 shows that A = when x = L There- fore, substituting this pair of values of A and x in equation (7) we have: = ipP + C (8) so that C = - ipP (9) Substituting this value of C in equation (7) we have: A = ipa^ - \pP (10) which is the desired expression for the area A in Fig. 15. It is to be noted that a^ is a variable in the above discussion because we have arbitrarily made it vary in the derivation of equation (4), whereas p and I are both constant. 30 CALCULUS. PROBLEMS. In each of the first six following problems, find the rate of change of y with respect to x. 1. y^ = 3x, 2. y^ = ax^ — 3x, 3. y^ = 2ax^ + 3x, 4. y^ = a^ - x\ dx 2 >x* d^ _ 3(ax^ - 1) d^ ~X^ax^~^Zx ' dy _ 4:ax + 3 dx ~i (2aa;2 + 3x)« 5 .2 = 1 ^=-,J. ^* ^ x' cx^'dx -\- {cx^ + l)(2ax -\-h)'dx 33. Differentiation of the quotient of two functions. — Let u and V be any two functions of Xy and let y = - (1) Then ^^^vdu-u-dv (2) Proof. — From equation (1) we have . . u-\- Au .^. whence, subtracting equation (1) from equation (3) member by member, we have . u -\- Au u ,.. Ay = — — (4) ^ V + Av V Reducing ; — -— and - to a common denominator, equation V -\- Av V * ^ (4) becomes: V'Au — u-Av ,.v v^ + v-Av But as Au and Av approach zero the denominator approaches v^ as its limit. Therefore equation (5) reduces to equation (2). 42 CALCULUS. Example. — Let it be required to differentiate the function: Let u = as? -^rhx and let v = cx^ -\- d. Substitute these values of u and v, and the values of du{= 2ax-dx -\-h'dx) and d/o{ — Zcx^ ' dx) in equation (2) and we get the desired result. 34. Differentiation of a function of a function. — Let y he a Junction of x and let z he a function of y. Then y changes -^ dz times as fast as x, and z changes j- times as fast as y. There- dz dv fore it is evident that z changes t" X -t^ times as fast as x. That is, ^ = - X ^ (I) dx dy ^ dx ^^l dxi dz The symbols -^ and -7- are so cumbersome that the proposition upon which this equation is based is not easy to understand when it is stated as above. Reduced to its simplest terms the proposi- tion is as follows: If z changes 5 times as fast as y and if y changes 7 times as fast as x, then z changes 5X7 times as fast as x\ Example. — Let 2 = a(x2 -h a; -f- l)^ To differentiate this expression one can write y for a;^ + a; -}- 1. Then z = ay^ and j- = Zay"". But ^ = 2x-\-l. Therefore, dz using equation (1), we have t~ = Sa{x^ + x -\- l)2(2x + 1). For problems see group 1 in the Appendix. FORMULAS FOR DIFFERENTIATION AND INTEGRATION. 43 DIFFERENTIATION OF LOGARITHM. 35. Convergent series. — Before the derivative of logarithm of x can be found it is helpful to consider the two series of fractions: a h c d e . . . n 1st series 1 2 1 4 1 8 1 16 1 32 . . . 1 2» 2d series 1 II 1 1 1 II 1 1 1 where |n stands for the product of all positive integers from 1 to n inclusive, factorial n, as it is sometimes called. The first series may be brought before the reader most distinctly by means of the following schedule; Fraction of original sum which is spent each day. Half of a sum of money is spent the first day ^ Half of the remainder is spent the second day i Half of the remainder is spent the third day i Half of the remainder is spent the fourth day . . . tV etc., etc., etc. If one were to follow this schedule indefinitely the original sum of money would never be entirely spent, because there is always an unspent remainder; but the remainder would grow smaller and smaller, and the amount spent would approach, as nearly as you please, to the entire original sum. This is equivalent to saying that i + J + i + T^ + etc. can never be equal to unity, but it can be made to approach unity, as nearly as you please, by adding together a larger and larger number of the successive fractions of the series. Such a series of fractions is called a convergent series, and the limit of J + i + i + etc., as a greater and greater number of the successive fractions are included, is called the sum of the series. The sum of the series ^ + i + i + xV + etc. is unity. A series need not be convergent merely because the successive fractions of the series grow smaller and smaller. Thus | + i + i+i + i+etc. grows 44 CALCULUS. a h c d 1st series i hoi a iof6 ioic 2d series J J of a iof6 ioic larger and larger without limit as a greater and greater number of the successive fractions are included. Such a series of fractions is called a divergent aeries. To show that the series ,7:, , o > rj » ,-^) etc. is a convergent series |2 |3 14 [5 it is sufficient to arrange the two series, as in the following table, showing how each fraction is related to the preceding fraction in each series: etc. etc. Each fraction of the second series is equal to or smaller than the corresponding fraction of the first series. Therefore, if the first series is convergent, the second series must also be convergent. The particular series which is used in finding the derivative of a logarithm is ^=^+i+L2+ii+ii+ii+^*''- ^'^ and this series is convergent because it is identical to the second series in the above table with the exception of the first two terms. The sum of this series, correct to seven decimal places, is: e = 2.7182818 (2) 36. The limit of ( 1 + - j as z approaches infinity. — This limit is important in finding the derivative of a logarithm. Let us assume that z is always a positive integer. In this case ( 1 -f - j be expanded by the binomial theorem,* giving: (■+»'-+,l'©+,i*-'>©' can + ^z{z-l)(z-2)(^iy+ etc. (3) * See foot-note to Art. 30. Substitute o = 1, 6 = - and n = z, and we z have equation (3). FORMULAS FOR DIFFERENTIATION AND INTEGRATION. 45 But as 2 approaches infinity (1 j, (1 J, etc., approach unity, and therefore the limiting value of ( 1 H — ) , as 2 approaches infinity, is found by writing unity for each of the expressions (l ), (l ), etc., in equation (4), and when this is done the second member of equation (4) reduces to 1 + il "^ 1 9 "^ I q "^ fl "^ etc. Therefore the limiting value of ( 1 + - ) as z approaches infinity is 2.7182818, or e, as explained in the previous article. To prove that the limit o/ f 1 + - ] is e when z approaches infinity, hiU when z is not always a positive integer. Let s be any positive value whatever of z, and let w be a positive integer such that z lies between m and m + 1. Then ( 1 + - ) lies between [ 1 -\ — ) and { 1 + — r— r ) . Now as \ zj \ mj V ' m + 1/ (I \m / 1 \m+l 1 -\ j and {l-\ ^— r 1 both approach e as a Umit; and therefore f 1 + - J , which is always between ( 1 -^ — j (1 \m+i 1 4 ^— r j , also approaches e as a limit. To show that ( 1 H — ) approaches e as a limit when z approaches negative infinity. Let z = — r. Then fl-1 — j =fl j where z is negative and r is positive. Now but (l+^^J approaches unity as r approaches infinity and ( H — ^ ] 46 CALCULUS. approaches e as r approaches infinity. Therefore every member of equation (5) approaches e as r approaches infinity. Consequently f 1 H — ] ap- proaches e when z approaches negative infinity. PROBLEM. One dollar placed at compound interest for 10 years at 6 per cent, amounts to (1 + 0.06)^^ dollars, when accrued interest is added to the principal once a year. If accrued interest is added to the principle n times per year, one dollar after 10 years would / 0.06 \ ^^ amount to ( 1 + — — I . Find what one dollar would amount ('+"-?) to after 10 years at 6 per cent, compound interest (a) when accrued interest is added to principle once per year, that is, when n = 1, (6) when n = 2, and (c) when accrued interest bears interest without any delay whatever, that is, when n = oo. Ans. (a) $1,791, (6) $1,803, (c) $1,822. Note.- ( 1 + M§ )^"= [(l +\yj where . = ^ and a = 0.6. 37. Differentiation of the logarithm of x. — Let y be the number of times that a given number a must be multiplied by itself to give X. Then x = a^, and y is called the logarithm of x to the base a. The base of the common system of logarithms is 10, and the base of the Napierian system of logarithms is 2.7182818 or e, Napierian logarithms are used throughout this treatise except where it is otherwise expressly stated.* Let y = log X (1) Then dy = | (2) Proof. — ^Writing y + Ay for y and writing x + Ax for x in * Napierian log x = logio X logxo e FORMULAS FOR DIFFERENTIATION AND INTEGRATION. 47 equation (1) we have: y + Ay = log (x + Ax) (3) whence, subtracting equation (1) from equation (3) member by- member, we have: Ay = log (x + Ax) — log X (4) or Ay _ log(a; + Ax) - log x , . (X + Ax\ J or to log ll-j-~y, and ^log (l+"^) is equal to logfl+-^j^. Therefore equation (5) reduces to £ = '-( l+f)^ (6) and this may be written A^ Ax = l,og(i+^)£ (7) X or, writing z for -— , we have: Ax :=ilog(l + y' (8) But, as Ax approaches zero ^ {='ir-) approaches infinity, and the limiting value of ( 1 H — 1 as z approaches infinity is e, as explained in Art. 36. Therefore as z approaches infinity equation (8) becomes: But the Napierian logarithm of e is unity because e is equal to eK 48 CALCULUS. Therefore equation (9) reduces to ^= - d =— dx X X Differentiation of loga x. — Let y = loga X (10) This means that x = a« (11) Therefore, taking Napierian logarithm of each member of this equation, we have log X = 2/ log a or where * is, of course, a constant. Therefore, according to loga equation (1) and (2) we have dy = ,-i-.^ = l5i^l:^ (13) •' log^ X ^ X ^ 38. Differentiation of the exponential function. — Let y = Ae^- (1) where A and k are constants and e = 2.7182818. Then: g = kAe^- = ky (2) Proof. — Taking logarithms of both members of equation (1) we have log y = \ogA + kx (3) But the differential* of log y is — according to Art. 37. There- * It is very important that the student keep in mind what a differential is. The differential of log y is the infinitesimal increment of log y due to an infinitesimal increment, dy, of y. FORMULAS FOR DIFFERENTIATION AND INTEGRATION. 49 fore, by differentiating, both members of equation (3) we have: ^ = k-dx (5) y dv whence equation (3) is obtained by solving for ^; and the value of y{= Ae^") may be substituted for y in the resulting equation. Differentiation of a**. — Let y = a*' (6) where a and k are constants. Taking logarithm of each member of (6), we have log y = kx log a (7) Differentiating we have '^^kloga-dx (8) dv Therefore, solving for -~ and substituting the value of y from equation (6), we have: dv ^ = k log a-a** = k log a-y (9) PROBLEM. Given the logarithm of 2 to base 10 find the approximate value of logarithm of 2.01. Ans. 0.3032. Note. — Consider logio x. The derivative of logio x multiplied by a small increment of x gives the approximate value of the increment of logio x. 39. General proof of equation (2) of Art. 30. — Let y = ax"* (1) where n has any value, integral or fractional, positive or negative. Taking logarithms of both members of (1) we have: log y = log o + n log a: (2) Differentiating, we have ^^ = n^ (3) y X 5 50 CALCULUS. log o being a constant. Substitute ox" for y in this equation and solve for dy, and we have dy = ruix'*~^'dx (4) For problems see group 2 in the Appendix. DIFFERENTIATION OF TRIGONOMETRIC FUNCTIONS. 40. Limit of sin 2r sin p Fig. 18. chord ob is equal to 2r sin 4>. as approaches zero. — It is a fundamental principle in geometry that if a polygon of n sides be inscribed in a circle,* the sum of the sides of the polygon approaches the length of the circular arc as a limit when n approaches in- finity. The idea that a circu- lar arc has a definite length de- pends upon this principle. Consider Fig. 18. The arc ah is equal to 2r0, arc ah 2r0 But approaches imity as its limit when approaches arc zero; and therefore sm (f> must also approach unity as its limit when (f) approaches zero. 41. Differentiation of sin x. — Let y = sin X (1) dy = cos x-dx (2) Proof. — Writing y -\- Ay for y and writing x + Ax for x, we Then * Or any continuous curve. FORMULAS FOR DIFFERENTIATION AND INTEGRATION. 51 have : y -{• Ay = sin(x + Ax) (3) Whence, subtracting equation (1) from equation (3) member by- member and solving for -^ , we have Ax Ay _ sm(x + Ax) — sin x Ax ~ Ax (4) Using the trigonometric formula: sin a — sin 6 = 2 sin ^{a — h) cos ^{a + 6) (5) equation (3) may be reduced to: Ay _ sin jAx •cos(x + J Ax) (6) Ax |Ax but as Ax approaches zero cos(x + J Arc) approaches cos x as its limit, and -^~ — approaches unity as its limit. Therefore, from equation (6) we have : -^ = cosx or dy = cosx-dx 42. Differentiation of cos x. — Let y = cos X (1) Then dy = — sinx-dx (2) Proof. — Proceeding as in the previous article we get Ay _ cos(a; + Ax) — cos x Ax Ax (3) and using the trigonometric formula: cos a — cos h — — 2 sin i(a — 6)sin i(a + h) (4) we obtain Ay sinjArc . / , i. n /kx 52 CALCULUS, and from this we obtain -^ = — emx or dy — — Qmx'dx Far problems see group 3 in the Appendix. EXAMPLES SHOWING USE OF FORMULAS FOR DIFFERENTIATION. 43. Differentiation of tan x. — In order to differentiate tan x sin X the familiar formula tan x = may be used as follows: cos a; , sinx ,^. y = tan x = (1) cos a; Proceeding according to Art. 33, let w = sin x so that du = cos x • dx, and let v = cos x so that dv = — sin x • dx. Then using equation (2) of Art. 33, we have: J cos^ X'dx + sin^ x-dx dx /„^ dy = ^ = — ^ (2) COS^ X COS^ X Any trigonometric function can he differentiated hy expressing the function in terms of sine and cosine and using the formulas of Arts, 41 and 4^. As a further example, consider i = I sin (at (3) in which I and to are constants and t is clasped time reckoned from any chosen instant, and let it be required to find the rate of di change of i, namely, -jz. Now cot is a function of t and sin is the angle whose tangent is ^- * dt^ Show from equations (1) to (6) of Art. 47 that the resultant acceleration (or total acceleration) of the ball B in Fig. 20 is equal to -j , and show that it is parallel to the radius A at each instant, v being the resultant velocity \/ ( ;^ ) + ( 7^ ) • 10. If the distance traveled by a body is s = ae*' + 6e~**, show that the acceleration is a^s. If t is expressed in seconds and s in feet, (a) in what units is a expressed, and (6) in what units is ah expressed? Ans. (a) reciprocal seconds, (6) feet per second per second. 48. Harmonic motion. Example showing the use of a second derivative. — A body m is suspended by a spring and the body stands in equilibrium in the position shown in Fig. 21. If the body is displaced x feet downwards (or upwards) an unbalanced force F will act upon the body, and this unbalanced force will be proportional to x. Therefore we may write F = -kx (1) FORMULAS FOR DIFFERENTIATION AND INTEGRATION- 59 The negative sign is chosen because F is upwards when x is downwards, and F is downwards when x is upwards. If the body is pulled upwards or downwards and released, it will vibrate up and down, and x will dx vary with the time t. Then -rr is the velocity dt d^x of the body at each instant, and -^ (the rate of dx\ change of -77 ) is the acceleration of the body at each instant. But the unbalanced force which acts upon a body is equal* to mass X acceleration. Therefore F = m > -p, or, substituting the value of F from equation (1), we have: spring df m (2) To determine the motion of the body in Fig. 21 it is necessary to find a; as a function of t which will satisfy the law of growth as expressed by equa- tion (2), and the problem is solved if we find a k function of t whose second derivative is equal to multiplied by the function itself. This function would be at once recognized if one were familiar with the derivatives of all kinds of functions. Thus we have another example showing how important it is to be familiar with the derivatives of functions. Compare Arts. 22 and 23. Consider the function a; = a sin (coi + 6) (3) where a, w and are undetermined constants. Differentiatingf * When mass is expressed in pounds, acceleration in feet per second per second, and force in poundals; or when C.G.S. units are used throughout, t Let z = Oil ■\- B and use the formula for the differentiation of a function of a function as explained in Art. 34. 60 CALCULUS, equation (3) with respect to t we have: ^ = coa cos (o)t -f 0) and differentiating again, we have: '«! cPx ^ = - co^a sin (ut + 6) (6) or, substituting x for a sin {cot -\- 0), we have: i Fig. 26. of this circle is ■a — ~x- Fig. 27. {x - ay + (2/ - hf (12) where a and h are the coordinates of the center of the circle and r is the radius of the circle. Differentiating equation (12), we have 2{x - a) ' dx + 2{y -h) ' dy = or {x-a) + {y- h)u = (13) du where u is written for -~. Differentiating equation (13), re- membering that y and u are both functions of x, we have: dx -\- {y — h) • du -{- u * dy = Q 68 CALCULUS. Dividing through by dx and remembering that w • ^ = w^ we have: 1 + (2/ - 6) • g + w^ = du (Pv But ^ is the second derivative ~. Therefore, representing ^ by V, we have: l-\-{y-h)v-\-u' = or 2/ - 6 = -— (14) and, substituting this value of {y — h) in equation (13), we have: V Substituting the values of (x — a) and {y — h) from equations (15) and (14) in equation (12), we get: , = a±j^' (16) Now u and v and x and y are all known at a given point p of the curve cc of Fig. 26, and these four quantities u, v, x and y have the same values for the osculating circle where it touches the curve cc at the point p. Therefore a and h can be found from equations (14) and (15), and r can be determined from equation (16) so that the osculating circle at the point p is com- pletely determined. PROBLEMS. 1. Find the radius of curvature at the point p of the parabola shown in Fig. pi. Ans. 40.5 inches. FORMULAS FOR DIFFERENTIATION AND INTEGRATION. 69 2. Find the radius of curvature of the ellipse shown in Fig. p2, (a) at the point p, and (6) at the point p'. Fig. pi. Ans. (a) 2.67 inches, (b) 18 inches. dy . Fig. p2. Note. — The value of -^ is infinite where a curve is parallel to the i/-axis and generally the value of ^ is also infinite at such a point. Therefore the expression for radius of curvature at such a point of a curve usually becomes indeterminate. It is evident, however, that x and y may be interchanged throughout the entire discussion of Art 49. This interchange gives an ex- dx d^x pression for radius of curvature in terms of t- and ^ where x is thought of as a function of y] and this expression can be used to determine the radius of curvature at a point on the curve where the curve is parallel to the y-axis. Fig. p3. 18 inches Fig. p5. 3. A stone arch is to be made as shown in Fig. p3, the two dotted circles having the same curvature as the ellipse at the junction points pp. Find the radius of the circles. Ans. 14.07 feet. 70 CALCULUS. center of curve 4. Find the radius of curvature of the curve y = sinx at the points (a) a; = and (h) a: = ^. Ans. (a) infinity, (6) 2.6. 5. Find the radius of curvature at the point p of the sine curve shown in Fig. p5. Ans. 5.47 inches. 6. A straight portion of railway track merges gradually into a circular curve as indi- cated by the dotted line in Fig. p6. The dotted portion is called a transition curve. The curvature of the track (which is sensibly equal to ^~ RM. 100 feet Fig. pQ. because dy . , ~ IS very nearly zero) increases uniformly from zero at x = to t^ftf: at a: = 100. Find the equation of the dotted transition curve. Ans. 900,000 y = a^. Note. — If the curvature v 2 increases uniformly its rate of change, ^-j, is a constant. 50. Geometric differentiation. The acceleration of a particle which travels in a circular orbit. — There are cases in which the instantaneous rate of change of a given quantity may be found when the quantity is specified in geometric terms. An important example is the following: A particle travels round and round a circular path or orbit at velocity v, and it is required to find the deceleration ^ of the particle. At a certain instant the particle is at P, Fig. 28, and its velocity FORMULAS FOR DIFFERENTIATION AND INTEGRATION. 71 at this instant is represented in direction and magnitude by the Hne vi. During a short interval of time the particle will travel Fig. 28. Fig. 29. the distance v.dt, the particle will then be at Q, and its velocity will be as represented by the line v^ which has the same length as Vi. From any fixed point 0' draw two lines parallel and equal to vi and V2 respectively, as shown in Fig. 29. Then the line dv represents the velocity which must be added* to Vi to give i;2, that is the line dv represents the increment of the velocity of the particle during the interval dt. Now when dt is chosen smaller and smaller, the angle <(> ap- proaches zero, and the angles at P' and Q' approach 90° so that dv becomes more and more nearly parallel to the radius OP. Also when dt is very small the arc PQ ( = v-dt) is sensibly a straight line and the two triangles OPQ and O'P'Q' are similar. Therefore from these similar triangles we have : r : v.dt : :v:dv (1) In this proportion v is written instead of Vi or V2, both of which * We are here concerned with what is called vector addition or geometric addition. The most familiar example of this kind of addition is the addition of two forces by the principle of the "parallelogram of forces." 72 CALCULUS. have the same numerical value. Solving this proportion we have; dv dt (2) We have already proceeded to the limit in considering the arc PQ a straight line. That is, the acceleration of a particle moving in a circular orbit is equal to the square of the velocity of the particle divided by the radius of the orbit, and the acceleration is at each instant towards the center of the circle because the infin- itesimal increment of velocity dv in Fig. 29 is in the direction of PO in Fig. 28. 51. Geometric differentiation. The inward force per unit length of a barrel hoop. — Another important case of geometric differentiation is the following: Figure 30 represents a hoop around a tank. It is required to find the value of -r where dF is the force with which a short length ds of the hoop pushes inwards on the tank. The radius r of the tank and the tension T of the hoop are given. The tension of the hoop is the force with which a portion of the hoop pulls on an adjoining portion. Thus the two forces tank Fig. 30. Fig. 31. FORMULAS FOR DIFFERENTIATION AND INTEGRATION. 73 Ti and T2, Fig. 30, which pull on the portion ds of the hoop in Fig. 30, are each equal to the tension T, and the resultant of the two forces Ti and T2 is the total inward force dF exerted on the portion; and this total inward force acting on ds is the force with which ds pushes against the tank. The resultant of Ti and T2 is shown in Fig. 31, and the triangle aVh' in Fig. 31 is similar to the triangle aOb in Fig. 30. There- fore we have r :ds::T :dF so that ? = ^ (1) ds r We have already proceeded to the limit in considering aOh to be a triangle, that is, in considering ds to be a straight line, which is only true when ds is infinitely small. The practical meaning of equation (1) can be best understood dP by thinking of -p as the really important quantity under con- sideration, not by thinking of the meaning of F* * Indeed one gets into unfamiliar vector theory when one tries to consider the meaning of F. CHAPTER III. INTEGRATION. 52. Integration as an arithmetical argument. — ^Any argument which leads to a proposition concerning total change when a rate of change is given may properly be called integration. Thus the following arguments are the simplest existing examples of inte- gration: (a) A man earns money at the constant rate of two dollars per day. Therefore the man earns two dollars each day, and in 100 days he would earn 200 dollars. Or, to get the amount of money earned in 100 days multiply two dollars per day by 100 days.* (h) One man A earns money twice as fast as another man B. Therefore in any interval of time whatever A earns twice as much as B. 53. Integration as a mechanical process. Integrating machines. — The simplest and most familiar integrating machine is the ordinary cyclometer. The wheel of a bicycle turns at a speed which is proportional to the velocity of travel of the bicycle. Therefore the number of revolutions of the wheel is proportional to the distance traveled, and the cyclometer is a revolution counter arranged to iesA one additional unit for each mile traveled. The ordinary "electricity meter "f which is installed in con- nection with electric lamps is an integrating machine. The * If one can get 200 dollars by multiplying two dollars per day by 100 days, one might well ask how to perform such a profitable operation? The answer is: Work hard for 100 days! That is what this particular multiplication means. Nearly every mathematical process rightly understood relates to physical reality in a manner as definite and as exacting as this particular case of multiplication. t Properly called a watt-hour meter. 74 1-lfer- INTEGRATION. 75 spindle of the meter turns at a speed which is proportional to the rate of delivery of energy (in watts) to a customer. Therefore the number of revolutions of the spindle is proportional to the total amount of energy delivered. The clock-work of the watt-hour meter is a revolution counter which is arranged to read one addi- tional unit for each watt-hour of energy delivered. 54. The planimeter.* — The planimeter is an integrating machine for measuring area, and it depends upon the following kinematical principle : Consider a line AB, Fig. 32, moving in any manner whatever in the plane of the paper. The motion of the line at each instant may be thought of as a motion of translation combined with a motion of rotation about an arbitrary point p.f Let V be the component at right angles to AB of the velocity of translation, as shown in Fig. 32, and let to be the angular velocity of AB about the point p. It is desired to find the rate at which the line AB sweeps over area, area swept over by the line being considered as positive when the line sweeps over it from right to left to an observer looking from A towards B. During a short interval of time, A^, the motion of translation carries the line from AB to A'B' as shown in Fig. 34, and the * The planimeter which is here described is the Amsler planimeter. The student wishing to become familiar with various kinds of integrating machines should consult Les Integraphes Abdank-Abakanowicz, Paris, Gauth- ier-Villars. A discussion of integrating machines is also given in Encyclopddie der Mathe- matischen Wissenschaften, Vol. II. This great work is now appearing in a re- vised form in a French translation. See also Report on Planimeter s, British Association Annual Report, 1894, pages 496-523. See also Mechanical integrators, Shaw, Proceedings of the Institution of Civil Engineers, London, 1885, pages 75-143. The earhest type of integrating machine for use in harmonic analysis is that of Lord Kelvin. This machine is described in Franklin's Electric Waves, pages 240-242, The Macmillan Co., New York, 1909. t See Franklin and MacNutt's Mechanics and Heat, Arts. 74-77, pages 174 and 175, The Macmillan Co., N. Y., 1910. 76 CALCULUS. area swept over by the line is I X v-At. Dividing this area by At gives the rate at which the Une sweeps over area because of its motion of translation. / Sf center of / < I Fig. 32. During a short interval of time, A<, the motion of rotation carries the Une from AB to A"B" ^ as shown in Fig. 35, and the Fig. 35. \t ■ area swept over consists of two parts as shown. The positive area \(l \2 in Fig. 35 is equal to 2(2 + '^) X w-A<, the negative area in Fig. 35 is equal to \%-) X oo'At, and the total area swept over is equal to IDo) • At. Dividing this area by At gives the rate at which the line sweeps over area because of its motion of rotation. INTEGRATION. 77 Therefore the total rate at which the line sweeps over area /dA\ . is equal to Iv + IDoo, or, using -77 for w, we have Let AB be a metal bar carrying a wheel mounted as shown in Fig. 36 (which is a side view), and let the rim of this wheel roll on the paper as the bar AB moves, as shown in Figs. 36 and 37. y^middle of bar W^u„^ ao "\|— 4— U paper j.,,roUinff I * \ '^y/ wheel side view paper Fig. 36. Fig. 37. Let ^ be the total angle turned by the wheel in a given time; then ^- is the angular velocity of the rolling wheel, and it is equal to -, where r is the radius of the wheel, as shown in Fig. 37. There- fore we may substitute r -^ for v in equation (1) and we have: in which Ir -^ is the rate of sweeping over area because of the do translatory motion of AB, and ID -r. is the rate of sweeping over area because of the rotatory motion of AB. Now if the rate of sweeping area due to the translatory motion of AB is Ir times the rate of turning of the wheel, then the total area swept over because of the translatory motion of AB is Ir times the total angle rp turned by the wheel. Similarly, if the rate of sweeping area due to the rotatory motion 78 CALCULUS. of AB is ID times the rate of turning of AB, then the total area swept over because of the turning of AB is ID times the total angle 6 turned by AB. Therefore the total area swept over by AB is Irrj/ + IDd.* That is, A = H-\- IDS (3) If the bar AB comes back to its initial position without turning completely round, then = and equation (3) becomes A =^lri (4) That is, the total area swept over by AB is proportional to the angle ^ turned by the rolling wheel, and the value of Ir may be so chosen that one square inch corresponds to each revolution of the wheel. In this case square inches of area are indicated by whole revolutions of the wheel, and the circumference of the wheel may be divided so as to indicate fractions of a square inch. In the actual planimeter one end of the bar AB is constrained to move back and forth along a circlef CC, Fig. 38, while the other end A travels once round the closed curve CC which bounds the area to be measured. Then any area outside of CC which is swept over by AB at all is swept over as many times to the left as to the right, but every portion of the shaded area is swept over once more to the left than to the right. Therefore, * How much more easily miderstood is this argument than to say integrating equation {2) we have equation {3) ! and yet this brief statement means exactly what is given in the above argument. t Along a straight line in some forms of planimeter. Fig. 38. .m^ INTEGRATION. 79 in view of the agreement as to algebraic signs, the total area A (= lr\J/) swept over by the line AB is equal to the shaded area. 55. Integration by steps. — In the laying out of a new electric railway the engineer usually makes a study of the probable schedule speed of the car in order to be able to estimate the traffic income and to be able to choose suitable driving motors. In the i€ « 8 1^' — / "*^ r^ /" ^ / / ■^ / / / 9' V / 1/ ^^ 1 24 4 •«: 10 20 30 40 50 60 70 80 J f P y v /\ - Kr ^ ? ' ^45 g / § / Vj <. / B ^•'" / s / g / s / «7r / s / i / % /^ / / _ 5, JO 15 .20 25 30 ebieUsi, bandredtba tot curve A, tea-tbouaandUi9 for curv» B- Fig. pi. from a neutral condition to B = 16,000 lines per square centi- meter. Ans. 326 X 10^ ergs or 2.406 foot-pounds. Note, — ^The work done in magnetizing iron is: '^ = If^ dB 82 CALCULUS. where W is expressed in ergs and V is the volume of the iron in cubic cenf i- meters. There are 30.5 X 453.6 X 980 ergs in one foot-pound. TABLE OF B AND H FOR WROUGHT IRON. H B 10 12,400 20 14,330 30 15,100 40 15,550 50 15,950 60 16,280 70 16,500 3. The force F which acts on a body is given as a function of the time in the accompanying table. Force being expressed in poundals and time in seconds. The mass of the body is 1,000 pounds. Find the velocity produced from ^ = to i = 40 seconds. Ans. 27.1 feet per second. dv F F Note. — The acceleration -j-. = — or dv = —'dl or the velocity v is equal dt m m J -i to — times the integral of F-dt; velocity being expressed in feet per second. TABLE OF F AND L t in Second!. F in Poundali. 500 10 625 20 700 30 745 40 785 56. Algebraic integration. — Integration, in the sense in which this word is used throughout this text, is the recognition of a given differential expression as the differential of a known function, and the rules for differentiation which are given in Chapter II are rules for integration also. Thus the differential of log x + C dx dx is — , and therefore, by definition, the integral of — is loga;+C, X X INTEGRATION. 83 where C is any constant. The rules given in Chapter II cover differentiation completely, or nearly so, but they do not cover integration completely. Indeed complete rules cannot be given for integration.* In the recognition of a given differential expression as the differential of a known function, two things are helpful, namely, (a) A table of standard forms or functions with their differentials (such a table is usually called a table of integrals), and (6) some degree of familiarity with the transformations which may be used to reduce a given differential expression to a standard form or to a combination of standard forms. A table of integrals is given in Appendix B. 57. Transformation of differential expressions. — When a differ- ential equation has been found by differentiating a given function, the integral of the differential equation is of course known. Many of the forms in the table of integrals have been established in this way as problems in differentiation in Chapter II. When a given differential equation has not been found by differentiating a known function it is in many casesf possible to invent a scheme for transforming the given differential expression so as to reduce it to one or more of the standard forms in the table of integrals. The most important^ of these transformations are embodied in the following rules, and a few of the lesser important transforma- tions are illustrated by problems. As a very simple example consider the differential equation * Integration by series is a universal and systematic method of integration. dv Any given algebraic expression for -^ can be expanded by Maclaurin s theorem, each term of the series so obtained can be integrated according to form 1 in the table of integrals, and the resulting series, when it is convergent is the integral of the given expression for -p . t Not in every case, by any means; many integrals cannot be expressed in terms of the elementary functions which are used in the table of integrals. X Transformations depending upon the use of imaginaries are also very important. See Art* 93 in Chap. VI. 84 CALCULUS. which was found in the discussion of the arc of a parabola in Art. 20, namely: ds = VT+lfc2^2 . dx (1) This expression is easily reduced to: ds=2ky]^,-hx''dx (2) and according to rule I, below, we have: h^li' + *= • dx = 2kf^± +x^.dx (3) Therefore, writing a^ for — , we get an expression which is identical to form 57 in the table of integrals. Rule I. — Any constant factor may be removed from under the integral sign. Thus: J a ' du = ajdu where a is a constant and u is any differential expression what- ever. Rule n. — The integral of the sum of two differential expressions is equal to the sum of the integrals of the respective expressions. Thus: J(du + dv) = f du -\- jdv where du and dv are any differential expressions whatever. Rule in. — When a differential expression is of the form w" • du its integral is, of course, given by form 1 or by form 2 of the table of integrals, whatever the function u may be. For example consider the differential equation: dy — sin^ x cos x * dx (4) Let mix = u then sin^ x = u^ and co^x • dx = duj and equation (4) becomes: dy = u^ • du ."ii^ii- INTEGRATION. 85 whence y^iu' + C or, substituting sin a; for u we have. y = i sin^ x -j- C (5) For problems see group 7 in the appendix. Rule IV. Integration by parts. — This rule is expressed thus: fu'dv = uv —fv • du (6) This formula is sometimes useful when a given differential ex- pression can be resolved into two factors, namely, (a) any function u, and (6) a familiar differential expression dv. This rule is based on Art. 32 where it is shown that: d(uv) = u • dv -\- V ' du which is to say: uv =fu • dv +fv ' du so that, by transposing we get equation (6). For example consider the differential equation: dy = X log X ' dx (7) In this case x - dx(= dv) is a familiar differential expression, that is, it is the differential of ix^{= v). Therefore, using equa- tion (6), we have: J u • dv = u ' v —f V ' du (6) f{\ogx)(x'dx) = {\ogx)(ix')-fiix')(j) (8) But (ia;2) f — j is Jx • dx, and its integral is ix^ + C. Therefore equation (8) becomes: fx log X ' dx = ix^ log X - ix^ - C (9) For problems see group 8 in the appendix. 86 CALCULUS. Special methods. Some differential expressions require more or less elaborate transformations or substitutions to get them into forms whose integrals can be recognized, or easily obtained by the simple transformations of rules III and IV above. Among these special methods are the following: (a) Integration of rational fractions by resolution into partial fractions. For 'problems illustrating this method see Group 9 in the Appendix. (6) Integra tion of expressions involving Va^ —x ^, V a;^ + a* or V x^ — a^. Such expressions can usually be reduced to recog- nizable ^ormsJ)y making the following substitutions. In Va^ — x^ substitute a sin d for x. In V^^^J^ substitute a tan d for x. In Vx^ -f a^ substitute a sec $ for x. Problems illustrating these trigonometric substitutions are given in Group 10 of the Appendix. (c) Miscellaneous Substitutions. A few problems involving mis- cellaneous substitutions are given in Group 11 in the Appendix. A good discussion of integration transformations is given on pages 32-70 of W. E. Byerly's Integral Calculus ^ Ginn & Co., Boston, 1881 J second e dition^ 1^ CHAPTER IV. PARTIAL DIFFERENTIATION AND INTEGRATION. 58. Differentiation of a function of two variables. — Consider a rectangle of length x and breadth y as shown in Fig. 40. The area of the rectangle is: A = xy (1) that is, the area is a function of the two variables x and y. Writing A -\- AA for A, X + Ax for X, and y -\- Ay for y, we have: A + AA = (x + Ax){y + Ay) or A-\-AA='xy-\-X'Ay-\-y'Ax-{-Ax'Ay (2) whence, subtracting equation (1) from equation (2) member by- member, we have AA ==^ X ' Ay ■\- y ' Ax ■\- Ax ^ Ay (3) But when Ax and Ay are infinitely small we may drop the second order infinitesimal Ax • Ay, and we get: dA = X • dy ■{• y ' dx (4) Now X ' dy is the infinitesimal incre- ment of A when y alone increases, and y ' dx \& the infinitesimal incre- ment of A when x alone increases. Therefore, according to equation (4), the infinitesimal increment of A when x and y both increase is equal to the sum of the two infinitesimal increments: — (a) that which is due to the increase of x alone and (6) that which is due to the increase of y alone. 87 88 CALCULUS. This proposition is a special case of a general theorem which is expressed by the equation: dz where — is the derivative of z with respect to x on the assump- dz tion that y is constanty and — is the derivative of z with respect dz Uyyonthe assumption that x is constant; that is -^ * dx is the ox infinitesimal increment of z due to an increment of x alone, IT ' dy is the infinitesimal increment of z due to an increment dy of y alone, and dz I =-r~ ' dx + — - dy ) is the infinitesimal increment of z when x and y both increase. It is customary to use the symbol d instead of d in partial differentiation, but to do so is apt to be misleading because it gives the impression that there is an unexplained and mysterious difference between ordinary and partial differentiation. It is indeed allowable to use d instead of d because one always knows when one is dealing with a function of more than one independent variable. 59. Successive partial differentiation. — Let y he a. function of a single independent variable x. Then the rate of change of y with respect to x is called the derivative of y with respect to x. But this derivative is itself a function of x and its rate of change with respect to x is called the second derivative of y with respect to x; and so on as explained in Art. 46. Let 2 be a function of two independent variables x and y. Then the two first derivatives r- and t- are themselves functions ax dy of x and y, and each of these first derivatives has its derivative with respect to x and its derivative with respect to y. For PARTIAL DIFFERENTIATION AND INTEGRATION. 89 example, if 2 = T^y^y then — = 3xy(=a), and — =3a;y(=/3); and each of these first derivatives is a function of x and y. The •\ »\ derivative t~ (= ") ^^s two derivatives, namely, t- ( = 6x2/^) ox ox and ^(= OxV); and the derivative — (= /S) has two deriva- tives, namely, ^ (= 9xV) and 7- (= 6x^1/), dz d^Z The derivative of — with respect to x is indicated as r^. 62 d^Z The derivative of — with respect to y is indicated as —. uX oy ' ox dz d^z The derivative of — with respect to x is indicated as -r r-. dy ^ dx ' dy The derivative of — with respect to y is indicated as r-r. In the above example, namely, when z = oi^y^j the two second d^z d^z derivatives, -r-^- and r-r-, are each equal to 9xV. That is dydx dxdy ^ ^ these two second derivatives are identical. Indeed the two deriva- tives T-^ and ^-T- are always identical when z is a function of X and y. This proposition is very important in mathematical physics. 60. Partial differential equation.* — Let 2 be a function of x and y concerning which it is known that 1= ax^ (1) and let it be required to find an expression for z. Now y is dz assumed to be a constant when the derivative r- is found from a OX * A very interesting example of a partial differential equation is discussed in Art. 90. 90 CALCULUS. given function, and therefore y is to be looked upon as a constant in equation (1) when one tries to think of the function from which equation (1) is derived, that is, ay in equation (1) is to be con- sidered as a constant, so that we may write ay = h (2) and consider what function of x has a derivative equal to hx^. Evidently the desired function of x is [^bx^ + C]. But in this argument y is assumed to be constant, so that the constant of integration C can be any function whatever of y. Therefore, writing f(y) for C and using ay for h, we have Hayx^ -{- f{y)] as the most general algebraic expression which gives ax^y when differentiated with respect to x, and this is therefore the desired expression for z. That is: z = iai^y+f{y) (3) where f(y) represents any function of y whatever. Of course this function f{y) may include a term which does not contain y, as in ay^ + hy + c, or indeed it may be a simple constant which does not contain y at all. Equation (1) expresses the law of growth of z with respect to X, and it is called a partial differential equation because z is a function of more than one independent variable. Example of an ordinary differ- ential equation. If then dz dx ax^ (4) (5) z==iax^-{-C where C is a constant which may have any value whatever. If we place x = inequation Example of a partial differ- ential eqvxition. If then 52 dx = ax"^ (6) z^\a:>?-\-f{y) (7) where f{y) is any function whatever of y which does not depend on x. PARTIAL DIFFERENTIATION AND INTEGRATION. 91 (5) we have z = C. Therefore the constant C is determined if we know the value of z when a; = 0. The constant C is called the \ constant of integration. If we place x = in equa- tion (7) we have z = f{y). Therefore the function f{y) is determined if we know 2 as a function of y when a; = 0. The function f{y) is called the function of integration. The above discussion refers to very simple partial differential equations. As another example let us set up the partial differ- ential equation which characterizes a given function. Let z be any function whatever of s where s stands for x + ay. Then ds dx = V , ds and -y- = a. dy Also, according to Art. 34, we have: dz dx dz ds -r = -r • -7- and -r = -i- dx Therefore, using the values, dz ds dx = 1 and d2 dx and dz _ dz dy ds ds dy dz dz dy dy = a. we have ds and consequently: dz dy = a dz dx (8) This is, of course, a partial differential equation, and it is satisfied by any function whatever of {x -\- ay). For example it is satis- fied by z = X -{- ay, by z = {x -\- ayy, by sin {x + ^y), by log {x + ay), by e^^^', etc. 6L Differentiation of an implicit function. — Any equation between x and y defines 2/ as a function of x^. If the equa- tion is solved for y, we have y as an explicit function of x; otherwise the equation defines y as an implicit function of x. * In this discussion the symbol d is used instead of d for partial differentia- tion. t Or X as a function of y. 92 CALCULUS. Thus 6X2 ^ IQ^yS + 2/4 + 5 == (1) defines y as an implicit function of x. It is convenient to repre- sent the entire left-hand member of equation (1) by the letter /. Imagine for a moment that / need not be equal to zero as required by equation (1), then ^ • dx would be the increment of / due to an increment of x, and ^ • dy would be the increment of / due to an increment of y. But according to equation (1) / is always equal to zero. Therefore the increments dx and dy must be so related to each other as to make the sum of the two increments of / equal to zero. That is we must have: — ) (2) (3) when y is an implicit function of a; as defined by the equation / = 0, where / is any algebraic expression involving x and y. In the above expression the two differentials dx and d-^ are called increments for the sake of brevity and clearness. A differ- ential is of course not an increment because it is infinitely small. PROBLEMS. dv Using the method of Art. 61, find -^ in the following expressions. 1 ?!_L.^'_i =n dy^_^ *• a^'^b^ ^' dx a'y ' df dx dx + dy 'dy or, solving for dy dx' we have dy dx d^ " ^1 dy This formula gives the value of dy dx dy dx dy x^ - ay ' ax - t' 6a: + 52^ dx dy y\\hx + 22/) x^ + y^x^ - y^ dx dy _ dx x{y + Vx2 - 2/2) y - X ' y + x' PARTIAL DIFFERENTIATION AND INTEGRATION. 93 2. a:^ + 2/' - ^axy -3 = 0, 3. Qx^ + lOxy^ + 2/' + 5 = 0, 4. log (x^ - 2/2) - 2 sin-i - = 0, X 5. 2 tan-i - - log {x^ + y^) = 0, y 62. Slope of a hill. — Imagine a hill built upon the plane which contains the x and y axes of reference, and let z be the height of the hill above the point on the plane whose coordinates are X and y. Then an equation expressing z as a function of x and 2/ is the equation of the surface of the hill. Thus z = Vr2 - a:2 - 2/2 (1) is the equation of a hemispherical hill of radius r. Consider the derivatives -r- and r- ; these derivatives are both dx dy we have functions of x and y. Thus if 2 = Vr^ — x^ - dz X -f, dx Vr2 - x^ - y^ and dz y dy Vr2-a:2-2/2* Let a:', 2/' and 2' be the coordinates of a particular point on the surface of the hill as shown in Fig. 41. If the values x' and 2/' are substituted for x and y in the general expressions for — ox dz and — we get the values of these derivatives at the point p, and we will represent these values by ( t" ) and ( t- ) • Then ( — 1 94 CALCULUS. is the slope of the tangent Hne qq and it is equal to tan a where a is the angle shown in Figs. 41a and 416; and ( 7- I is the slope \2-axi8 >^ x^axht y-axis Fig. 41a. A ~S^^ ^ ^S y ^r X / X-axis y '«\ / X' /f*' // >' /v-axi» Fig. 416. of the tangent line rr and it is equal to tan /3 where /3 is the angle shown in Figs. 41a and 416. In Fig. 41a — and — are both negative; that is, z decreases as X increases, and z decreases as y increases. In Fig. 416 ■r- and -r- are both positive. dx dy 63. Equation of the tangent plane. — The derivation of the equation of the tangent plane which touches a surface at a point is somewhat similar to the derivation of the equation of a tangent line which touches a curve at a point. Therefore it is worth while to give the derivation of the equation of a tangent line before considering the equation of a tangent plane. Let a;' and z' be the coordinates of the point p where the tangent line touches the plane curve cc, as shown in Fig. 42a. PARTIAL DIFFERENTIATION AND INTEGRATION. 95 Starting at the point p in Fig. 42a, which is at a height z' above the base Hne, travel along the tangent line tt covering the hori- Fig. 42&. zontal distance {x — x') and the rise will be {x — x') tan a or dz (x — x') -J-. We thus reach any given point B in the tangent line of which the coordinates are x and z, and the value of z is s = 2' + (^ - a;') S (1) which is the desired equation of the tangent line tt. Let x', y' and z' be the coordinates of the point p where the tan- gent plane touches the given surface (see Fig. 426) . Starting at the point p, which is at a distance z' above the base plane, travel along the tangent line rr (which lies in the tangent plane) covering the horizontal dis- t^ tance (y — y') towards the reader in Fig. 426, and the rise will be (2/-2/')tan/3=(2/-2/')(|)^ 96 CALCULUS. Then continue along the Hne q'q' parallel to the tangent line qq (the line q'q' lies in the tangent plane) covering the horizontal distance (x — x') to the right in Fig. 426, and the rise will be {x - x') tan a = (x - ^')\r£) We thus reach any given point B in the tangent plane of which the coordinates are x, y and z, and the value of z is: 2 = 0' + which is the desired equation of the tangent plane. Of course the derivatives in this equation refer to the equation of the hill. 64. Component slopes and resultant slope. — Let X be the slope of a hill at the point p, Fig. 41, in a direction parallel to the the rise* per unit-horizontal-distance- parallel-to-the-x-axis, and it is equal to Also let Y be the slope of the hill at p in a direction parallel to theiz-axis. Then Y = (^) . Let X x-axis. That IS, X is x-axia Wp and Y be represented to a chosen scale by the lines X and Y in Fig. 43. Then the diagonal R in Fig, 43 represents what may be called the result- ant or actual slope of the hill at p. That is, the line R points directly up hill, and the length of R represents (to the chosen scale) the rise of the hill per unit-of-horizontal-distance-in-the-direction-of-/2. This is a funda- mental and important theorem in the mathematical theory of electricity and magnetism, and the proof of the theorem is as follows: * If the slope is negative, X is the drop per unit horizontal distance, etc. y-axis Fig 43. The X and y axes are shown as they would appear if seen from above in rigs.41a and 416. PARTIAL DIFFERENTIATION AND INTEGRATION. 97 Let the line AB in Fig. 44 be the line of intersection of a horizontal plane through p with the plane which is tangent to the hill at p. Then the line AB is evidently at right angles to y-axia Fig. 44. R. Let the line ah he the line in the tangent plane at every point of which the tangent plane is unit distance above the horizontal plane through p. Then, by definition: and R = X = Y = pr 1 ps J_ pq (1) (2) (3) where pr, ps and pq are the distances shown in Fig. 44. But 1 , — vr 1 ps pr _ cos 6 R-cosO , and pq — sin R ' sin 0' Therefore and X = RcosB Y = Rsind Z2 + F2 = R^ (4) (5) (6) 98 CALCULUS. These equations establish the proposition that R is the diagonal of a rectangle of which X and Y are the sides as shown in Fig. 44. 65. Example of gradient in two dimensions. — The observed pull in Fig. 10 is a function of one independent variable, the current as indicated by the ammeter, and in such a case it is helpful to plot a curve of which the abscissas represent observed currents and of which the corresponding ordinates represent observed values of pull. This curve is shown in Fig. 11. Frequently one is concerned with a quantity which depends upon two independent variables. For example one might be concerned with the distri- bution of temperature over a flat metal plate. In such a case it is helpful to think of the temperature at each point p as represented by the height at p of a hill. Then the component gradi- ents of temperature are rep- resented everywhere by the component slopes of the hill, and the resultant gra- dient of temperature is rep- resented everywhere by the resultant slope of the hill, as indicated by the sides and diagonal of the rectangle in Fig. 45. 66. Example of gradient in three dimensions. — One might be concerned with the distribution of temperature throughout a solid body. Of course, the temperature would have a definite valiie at each point in a body, or, as expressed in mathematical language, the temperature T at each point would be a definite function of the coordinates x, y and z of the point, and the derivatives ^— , -r- x-axi8 metal plate '< ^ i x" Y U^axis H Fig 45. and -^r- would also be functions of x, y and z. oz dx^ dy That is, the PARTIAL DIFFERENTIATION AND INTEGRATION. 99 three derivatives would have definite values at each point in the body, and the three derivatives, --, -^- and -r-, are the com- ponent gradients of the temperature at the point in directions parallel to the respective axes of reference. Representing the resultant temperature gradient at a point by R and the three components of R by X, Y and Z, we have R^ = X'+Y^ + Z2 (1) 2 = S (4) These equations constitute an extension to three dimensions of the theorem of Art. 64. In Fig. 45 the direction at right angles to the surface of the plate is available for purposes of geometrical representation, and one can think of an actual hill being built upon the plate so that the temperature of the plate at each point is represented by the height of the hill at that point, and of course this " hill" becomes a "valley" dipping below the plate where the temperature of the plate is below zero. The temperature at each point of a solid body, however, cannot be represented geometrically as a height because space is filled in every direction by the body itself. That is, every direction in space is used for representing the independent variables x, y and z, and no direction is left for the representation of T. It is convenient, however, even in this case, to speak of the "temperature hill,^' the "height" of the hill at each point of the body being the temperature itself, high or low as the case may be. 67. Partial integrations.* — In many cases the derivative of a * What is here called partial integration is usually called multiple integra- tion, double, triple, quadruple, etc., as the case may be. 100 CALCULUS. function of one independent variable can be set up or established by the use of elementary principles of physics and arithmetic, and the function itself can be found by one integration, as exempli- fied in Arts. 22 and 23. l But when the function to be established is a function, say, of two independent variables {x and y for example), then it is usually necessary to perform an integration with respect to x in setting up the derivative of the desired function with respect to 2/, or to perform an integration with respect to y in setting up the derivative of the desired function with respect to x. Such integrations may be called partial integrations because they are related to partial differentiation in the same way that ordinary integration (with respect to one independent variable) is related to ordinary dijfferentiation (with respect to one independent variable). 68. Volume of a hill. Example of partial integrations. — For the sake of simplicity let us consider a particular case, namely, a hemispherical hill with its center at the origin of coordinates. Then the equation of the surface of the hill is = Vr2 - x2 - r (1) Thus Fig. 41a represents one quarter of a hemispherical hill, and it is required to find its volume. Iz-axU I ^ff-axis Fig. 46a. Fig. 466. i PARTIAL DIFFERENTIATION AND INTEGRATION. 101 Let V be the volume of the portion ahcABC of the hill, as shown in Fig. 46a. Then y is a function of y (see figure), and the increment of v due to an infinitesimal increment of y is the volume of the thin slab in Fig. 466. Therefore we have: dv = A ' dy (2) where A is the area of the face of the slab as shown in Fig. 466. Therefore we must find an expression for A before we have a known expression for -7-. Now ape in Fig. 46a is a plane curve, and equation (1) is the equation of this curve (ordinate z, abscissa x) if y is constant. Let a be the shaded area in Fig. 46a; then the strip with double shading is da and it is equal to z • dx. Therefore using the value of z from equation (1) we have: da = Vr2 - ^2 - 2/2 . dx ■ (3) and the total area A of the face abcp is the integral of this expression from a; = to x = ha = Vr^ — y^* Therefore: Vr2 - x^- ?/2 . dx (4) In performing this integration y is a. constant. This integral gives an expression for A, and the value of A so found can be substituted in equation (2). Then equation (2) may be integrated between the limits y= to y = r to give the desired expression for the volume V of the hill. That is v=r *Jy=0 A ' dy (5) The above discussion applies to a particular case, namely, to a quarter of a hemispherical hill; but the volume of any hill what- * This value of ha is the value of x (for the given constant value of y) when 2 = 0, as found from equation (1). 102 CALCULUS. ever can be found in the same way. Find by a first integration, the area ^4 of a plane section of the hill parallel to the xz plane and at a distance y therefrom, and then integrate A • dy to get the required volume; or find by a first integration the area A' of a plane section of the hill parallel to the yz plane and at a distance x therefrom, and then integrate A' * dx to get the required volume. The prismoid formula. — A prismoid is a solid with parallel top and bottom faces with sides generated by straight lines. Thus a cylinder is a prismoid, any wedge even if its base is a polygon or curve is a prismoid, Fig. 47a represents a prismoid. The volume of a prismoid is given by the formula T + ^M + B 6 Xh (6) where T is the area of the top face, B is the area of the bottom face, M is the area of the middle section (which is parallel to top and bottom faces) and h is the altitude of the prismoid. Equation (6) also gives the volume of any figure bounded by a second degree surface ss in Fig. 476 between parallel top and Fig. 47a. Fig. 476. bottom faces, and equation (6) can be used to give a very close approximation to the volume of any solid between parallel ends with sides smoothly curved from end to end. PARTIAL DIFFERENTIATION AND INTEGRATION. 103 PROBLEMS. 1. Integrate equations (4) and (5), above, and find the volume of one quarter of a hemispherical hill as shown in Fig. 41a. Verify the result from the formula: volume of a sphere equals ^tt times its diameter cubed. 2. Find the volume of the entire wedge (20 inches long) in Fig. p2, the wedge being 8 inches wide in a direction perpendicular to the plane of the paper. Ans. 800 cubic inches. Note. — The volume of a wedge is, of course, easily found by using the prin- ciples of elementary geometry. It is intended, however, that this problem be solved by integration as follows: Let v be the volume of the shaded portion of the wedge as shown in Fig, p2. Then t; is a function of x, and the increment of V due to an increment of a: is the volume of the thin slab shown in Fig. ^2. The area of the face of the slab is — X 10 inches X 8 inches = 4a: square inches. Therefore the volume of the slab is 4a;-dx cubic inches. That is, dv = Ax'dx ^ x-axU Fig. 2)2. and the desired volume is found by integrating this expression from ic = to aj = 20 inches. 3. Find the volume of the shaded portion of the wedge in 30 inches Fig. p3. jj i0_inche8 Fig. p4. 104 CALCULUS. Fig. p3, the wedge being 6 inches wide in a direction perpendicular to the plane of the paper. Ans. 92 L6 cubic inches. 4. Find the volume of the frustum of a cone which is shown in Fig. p4:. Ans. 229.2 cubic inches. 5. Find the volume of the paraboloid of revolution which is shown in Fig. p5. Ans. 1060.7 cubic inches. Note. — ^The equation of the paraboloid is y = px', where the axis of revolution is the ^/-axis of reference; and the value of p is determined from the condition that x = 7}4 inches when y = 12 inches. ' 7 inches f 5 inches Fig. p5. Fig. p6. 6. Find the volume of the shaded portion of the paraboloid in Fig. p6. Ans. 699.8 cubic inches. 7. Find the volume of the spherical segment which is shown in Fig. p7. Ans. 276.79 cubic inches. 8. Find the volume of the segment of a sphere which is shown in Fig. pS. Ans. 119.71 cubic inches. ' 9. Find the volume between the xy plane and the plane z = a -{-hx + cy and inside of the cylinder {x — ey-\-{y—fY = g^ when a = 10 inches, 6 = 3, c = 2, e = 8 inches, / = 9 inches and gr = 5 inches. Ans. 4085.7 cubic inches. PARTIAL DIFFERENTIATION AND INTEGRATION. 105 Note. — Let A-dy be the expression which must be integrated with respect to y to give the desired volume. Then A is obtained by integrating z • dx between Umits which are the two values of x found from (x — e)^ + (?/ — ff = g^ for the given value of y, and then A-dy is integrated between the limits y ^ f -\- g and y = f — g. I Is Fig. p7. finches Pinches^ ^ '^ Fig. p8. I I II li 1^. -^^ 10. Find the total volume which is common to the two cylinders: 1/2 _|_ 2-2 _ 25 and x^ + y^ = 25, everything being expressed in inches. Ans. 666.7 cubic inches. Note. — One should be able to determine the proper limits for each integra- tion in this problem if one understands problem 9. 11. The inside dimensions of a barrel are 20 inches in diameter at each end, 23 inches in diameter at the middle and 31 inches long. What is the capacity of the barrel in gallons? Ans. 51.3 gallons. 69. Area of the surface of a hill. Another example of partial integration. — It is desired to find the area of the portion caCA of the surface of the hill shown in Fig. 48 when the equation of the surface of the hill is given; this equation, of course, expresses the height z of the hill at a point as a function of x and y as explained in Art. 68. Let S be the area of caCA. Then the 106 CALCULUS. normal to dt increment of S due to an infinitesimal increment of y {= Bb in the figure) is the area dS of the narrow strip ca. To find the area of this narrow strip let s be the area of the portion from c to ds in the figure, then s is a function of x, and the increment of s due to an infinitesimal in- crement of X is the small ele- ment of area ds which caps the prism dx . dy. This small ele- ment of area is sensibly coinci- dent with the tangent plane at ds, and therefore the normal to Fig. 48. ds makes an angle 7 with the z-axis whose tangent is equal to ^/(i^(i)' which is the resultant slope of the hill at ds, according to Art. 64. Now the area of the base of the prism, dx of the area ds, and consequently we have ; dx ' dy — cos 7 • ds dx • dy dy, is the projection or ds = But if we find: cos 7 (1) <-^'4(MFW)' COS 7 = ^Rl)■-(sJ so that equation (1) becomes: (2) '^ = ^' + {'iy+{fJ-'^-'y (2) .Jjl^ PARTIAL DIFFERENTIATION AND INTEGRATION. 107 10 inches Now the expression under the radical is a function of x and y which may be found by differentiating the equation of the surface of the hill (which is given); and the total area of the strip ca (which is equal to dS) is the integral of (3) from x = to x = ha* In this integration y and dy are constants. The value of dS so found is the required infinitesimal increment of the area caCA due to an infinitesimal increment of y, and the value of the area caCA is found by integrating this expression for dS between the limits y = and y = y (meaning any value of y). If the entire area of the hill in Fig. 48 is to be found, the expression for dS must be integrated from 2/ = to y = the intercept of the hill on the ^/-axis.f PROBLEMS. 1 . The hill shown in Fig. 48 is a quarter of a hemisphere. Find the area of its surface and verify your result from the formula: area of a sphere = 47r times its radius squared. 2. Find the area of the curved surface of the cone shown in Fig. p2, Ans. 135.8 square inches. Note. — Let a be the area of the i jp ly-i shaded portion of th cone in Fig. p2. " *" Then a is a function of x; and the in- Yig. v2. crement of a due to an increment of x is the area of the narrow strip shown in the figure. The width of this strip (parallel to the slant height of the cone) is ^^^ "^ ^^ X dx, and the length of the strip (cu-cumference of the cone) is ~.XS X tt inches. Therefore we get: da = OM^irX'dx * The value of 6a is found from the given equation of the surface of the hill. It is the value of x for the given constant value of y when 2 = 0. t This intercept is found from the equation of the surface of the hill by plac- ing X and z both equal to zero and solving for y. 108 CALCULUS. and the required area is found by integrating this expression from a; = to X = 10 inches. 3. Find the area of the curved surface of the frustum of the cone in Fig. p3. Ans. 144.38 square inches. 4. Find the area of the part of the surface of the cylinder .^^' ^>.. j^ JUf.Jnche8 ^ Ilg.p3. ^1 4 inches Pinches, m 1^ I Fig. p5. a;2 -f ^2 _ 100 which is inside of the cylinder x^ -\- y^ = 100, the radii of the cylinders being measured in inches. Ans. 800 square inches. 5. Find the area of the convex surface of the segment of a sphere which is shown in Fig. p5. Ans. 989.5 square inches. 6. Find the area of a zone on the earth's surface between 40° and 80° north latitude, taking the earth as a sphere 4,000 miles in radius. Ans. 34,500,000 square miles. CHAPTER V. MISCELLANEOUS APPLICATIONS. 70. Influence of errors of observation upon a result. When one measures a thing with great care repeatedly the successive measurements never agree with each other, and it is easy to determine what is called the probable error* of the measurement by considering the discrepancies in a set of observed values. Thus the length of a bar might be found by repeated measure- ment to be 103.625 centimeters with a probable error of, say, db 0.0036 centimeter; the mass of a body, as determined by re- peated weighing on a balance, might be 67.2382 grams with a probable error of, say, ± 0.0006 gram. In many cases, however, it is not the directly measured quantities that are important but a result which is calculated therefrom; and it is often desirable to calculate the probable error in such a result when the probable errors in the individual measurements are known. For the sake of simplicity we will discuss a special case, as follows: A steel ball is weighed on a balance and its mass is found to be M grams with a probable error of ± m grams, and the diameter of the ball is measured and found to be L centimeters with a probable error of ± I centimeters. The density of the steel as calculated from the measured values of M and L is: ^ = & (^) and it is required to find the probable error in D. The probable *See Franklin, Crawford and MacNutt's Praclicol Physics, Vol. I, pages 1-9 for a very simple discussion of errors of observation. For more complete discussion see Merriman's Least Squares, John Wiley & Sons, New York. See also Palmer's Theory of Measurements, McGraw-Hill Co., New York, 1912. 109 110 CALCULUS. errors m and I are usually very small so that it is sufficiently accurate to calculate ADm (the probable error in D due to =t m) and ADi (the probable error in D due to ± I) by the formulas: AI>« = l^-m (2) and AD, = gj (3) Let AD be the probable error in the result due to m and I together. Then AD is not equal to the sum AD;;^ + AD^, but according to the theory of probability we have: AD = V(ADJ2 + (ADO^ (4) PROBLEMS. 1 . The scale of an alternating current ammeter reads in degrees, and the value of the current is i = k^, where fc is a constant and is the deflection in degrees due to a current of i amperes. The error of reading is probably ± i^ degree. What is the probable error in the result, t, in fractions of an ampere (a) when the deflection is in the neighborhood of 200° and (6) when the deflection is in the neighborhood of 20°? Take the value of k to be such that 10 amperes gives 100° deflection. Ans: (a) =t 0.0089 ampere, (6) ± 0.028 ampere. 2. What is the percentage error (probable) in the value of i under conditions (a) and (6) in problem 1? Ans. (a) =t 0.06 per cent., (6) ± 0.6 per cent. Note. The percentage error is — X 100. 3. The diameter of a steel sphere is repeatedly measured and found to be 0.8762 inch with a probable error of ± 0.0016 inch. Calculate the volume of the sphere and calculate the probable error in the calculated volume. Ans. 0.3525 cubic inch =b 0.0019 cubic inch. MISCELLANEOUS APPLICATIONS. Ill Note. It is interesting to note that the percentage error in the volume i3 three times as great as the percentage error in the measured diameter, that is 0.0019 . 0.3525 is three times as great as 0.0016 0.8762 ' 4. Show that the percentage error of x^ is twice as great as the percentage error of x, when a; is an observed quantity and x^ is a calculated result. 5. By measurement the length of a rectangle is 25 inches with a probable error of =^ 0.02 inch, and the width of the rec- tangle is 10 inches with a probable error of ± 0.015 inch. What is the probable error in the calculated area of 250 square inches? Ans. ± 0.421 square inch. 6. The diameter of a steel ball is 2.542 centimeters ± 0.001 centimeter, and the mass is 116.25 grams ± 0.02 gram. What is the density and what is the probable error in the density? Ans. 7.781 grams per cubic centimeter =t 0.0168 gram per cubic centimeter. 71. Maximum and minimum values. It is evident that a growing thing reaches its maximum size when it stops growing; but a thing may stop growing for a while and then go on growing Fig. 49. . . dv . ... A maximum; -^ is positive when x is less than a, and negative when x is greater than a. Fig. 50. Not a maximum; -^ is positive when X is less than a, greater than a. also when x is again. Also it is evident that a decreasing thing reaches its minimum size when it stops decreasing; but a thing may stop decreasing for a while and then go on decreasing again. Thus 112 CALCULUS. ■^ is equal to zero when x = a in all four of the following figures, 49, 50, 51 and 52, but 6 is a maximum value of y in Fig. 49 only, and 6 is a minimum value of y in Fig. 51 only. y-ojdB I x-axi8 . — a ->l Fig. 51. dti A minimum; -^ is negative when x is less tlffen a, and positive when x is greater than a. y-€ixi8 Fig. 52. dv Not a minimum; -^ is negative when X is less than a, also when X is greater than a. To determine the values of x for which a given function of x is a maximum or a minimum, differentiate the function, place its derivative "equal- to zero,-and 'solve for x. Then for each value a of X so found, determine the algebraic sign of the derivative for a value of X slightly less than a and also for a value of x slightly greater than a. Interpret the results by referring to Figs. 49 to 52* PROBLEMS. In the first five problems, following, find for what values of x is 2/ a maximum or minimum and compute the maximum or minimum value in^each case. • 1. 2/ = a:'' — 6x2 + 9x -f 3. ^^s. x = 1) y = 7, a maxi- mum: X = 3; 2/ = 3, a minimum. ^ x^ + 4x + 44 2. 2/ = Ans. a^ = 8; 2/ = I, a mimmum: x2 + 8x + 40 X = — 6; 2/ = 2, a maximum. * A slightly easier method is to consider the values of the second derivative of the given function but the above method is sufficient. MISCELLANEOUS APPLICATIONS. 113 3. y = (x — 4)3 {x + ly. Ans. x = — 1; y = Oj a maxi- mum : X = 1; y = — 108, a minimum. 4. y = 9e2=^ + 25e~2^. Ans. a: = log Vf; 2/ = 30, a minimum 5. 1/ = 3 sin a; + 4 cos a;. Ans. a; = tan~^f in the first quad- rant; y = 5, a maximum: x = tan-^J in the third quadrant; y = — 6, a minimum. 6. Divide the number 20 into two parts such that the product of one part by the square of the other part shall be a maximum. Ans. 13.33 and 6.67. 7. What number added to one half the square of its reciprocal gives the least possible sum. Ans. 1. 8. The cost of an electrical power transmission line is approxi- mately proportional to the weight of copper used because the cost of poles and the cost of erection can be expressed with a fair degree of accuracy as an addition of so many cents per pound on the cost of copper. Therefore the annual money cost represented by interest on investment and depreciation, being reckoned as so many per cent, of first cost, is proportional to the weight of copper used, or it is equal to kw where A; is a constant and w is the weight of copper used. The amount of power lost in a transmission line is halved if the amount of copper is doubled. Therefore the annual money loss due to loss of energy in the line is universely proportional to the weight of copper, or it is equal to — where m is a con- stant. Therefore the total annual money loss is kw H . Find the w value of w (expressed in terms of k and w) which will make this total annual money loss a minimum. Ans. w = ^llH Note. This matter is discussed in electrical engineering books under the head of Kelvin's law. See Franklin' s Electric Light ing^ Art. 26, pages 66-70. The Macmillan Co., New York, 1912.' 8. It is desired to construct of sheet metal a cylindrical gallon 114 CALCULUS. measure without a cover. Find what its depth must be in order that the amount of sheet metal used shall be a minimum. Ans. y = jj?^ (= 4,19) inches. Note. The amount of sheet metal used will be a minimum when the surface area S of the vessel is a minimum; but S = 2irxy + ttx^, where x is the radius of the base of the cylinder and y is its depth. Since the measure is to contain 1 gallon (=231 cubic inches), irx^y =231; and, substituting for z its value in terms of y, we have: ^ = 2>/23i^ + ??i The value of y which makes this expreasion a minimum is to be obtained. 10. Determine the altitude of the cylinder of greatest convex surface that can be inscribed in a sphere whose radius is 10 inches. Ans. 14.142 inches. 11. Determine the altitude of the cylinder of greatest volume that can be inscribed in a sphere whose radius is 10 inches. Ans. 11.547 inches. 12. Determine the volume of the smallest cone which can be circumscribed about a sphere whose radius is 10 inches. Ans. — ^— (= 838.1) cubic inches. o 13. A man who can row at a speed of 4 miles per hour and run at a speed of 6 miles per hour wishes to reach the gi \ point p from a boat at h %^->.^ ^\ as shown in Fig. pl3 in the ^1 \ \^^ least possible time. Find the T^ch — ^ ^ V. ' beach distance ap that the man t;,. '"'-t* must run on the beach. Ans. rig. pl3. 1.056 miles. 14. The strength of a rectangular beam is proportional to its breadth and to the square of its depth. Find the breadth and depth of the strongest beam which can be cut from a cylindrical 'N MISCELLANEOUS APPLICATIONS. 115 log whose radius is 18 inches. Ans. Breadth, 20.8 inches; depth, 29.4 inches. 15. For what angle of deflection does the old-fashioned tangent galvanometer give the greatest percentage accuracy in the value of the current? Ans. 45°. Note. The equation of the tangent galvanometer is: i = k tan 6 where i is the current in amperes, 6 is the observed angle of deflection, and fc is a constant. Let dd represent the probable error of the observed value of 0, then the probable error in the value of i is: ,. k.de di = COS2 d and the percentage error in the value of i is: _di _ dd ^ ~ i "" sin COS 9 in which dd being the probable error of 6 is to be treated as a constant. A general discussion of this important matter of precision as dependent upon the choice of magnitudes of the quantities to be observed is given in Chapter XII of Palmer's Theory of MeasurementSy McGraw-Hill Co., New York, 1912. 72. The problem of the bent beam. The formulas used in connection with a bent beam are based on Hookers law, and the F sectional area a p p sectional area 8 p original length L m! original length L y increase of length I --^ decrease of length l-^ Fig. 53. Fig. 54. derivation of two of the simpler formulas furnishes a good example of the use of calculus. Hookers law. Figure 53 represents a rod subjected to a tp stretching force F. Experiment shows that — is proportional s 116 CALCULUS. to y (Hookers law). Therefore we may write L or a L F = E-^ (1) in which ^ is a proportionality factor and it is called the stretch modulus of the substance of which the rod is made. Equation (1) also gives the compressing force F required to produce a slight shortening of the rod as represented in Fig. 54. When a beam is bent the filaments on one side of the beam are lengthened, and the filaments on the other side are shortened. If we wish to determine the degree of lengthening or shortening it is necessary to consider a very short portion of the beam. But a very short portion of a bent beam has the same curvature as the osculating circle as shown in Fig. 55, and the easiest way ^^ mediaa Une very short portion beam Fig. 55. Fig. 66. to think of the very short portion of the bent beam is to think of it as a portion of a long beam which is actually bent into the arc of a circle. Thus Fig. 56 represents a beam bent into the arc of a circle. A certain filament in the bent beam is unchanged in length. MISCELLANEOUS APPLICATIONS. 117 pulU This filament is located in what is called the median line (or plane) of the beam. The median plane of a beam of rectangular section is the middle plane of the beam. Let us consider the forces which act across a section qq of a bent beam as shown in Fig. 57. Considering these forces as acting on A5, they are a set of pulls on A and a set of pushes on B, and together they con- stitute a turning force or torque, T, about an axis perpen- dicular to the plane of the paper. It is the object of this discussion to derive an expression for T in terms of 6, D, R and Ej where h is the breadth of the beam, D is the depth of the beam, R is the radius of curvature of the median line of the beam, and E is the stretch modulus of the material of which the beam is made. Let us consider the portion of the beam which lies between the radii RR in Fig. 56. This portion of the beam is shown to I I median line jIL. — .TP . \x length MR+x)^ length s=Re length =(R-jc;d Fig. 58. a larger scale in Fig. 58. The original length of every filament of this portion of the beam is Rd. Consider the upper filament 118 CALCULUS. Jf of the beam. The stretched length of this filament is {R + x)dj because // is a circular arc of radius (R + x) and it subtends the angle 6. Therefore the filament }} has been increased in length by the amount xB by the bending. Similarly it may be shown that the lower stretched _ 5 /'Stretch eompresaed- •ectional viewjof Jtean Fig. 59. filament /'/' in Fig. 58 has been shortened by the amount xd by the bending. Figure 59 is a sectional view of the beam with its depth arbi- trarily increased by added layers of material on top and bottom, and we wish to find the incre- ment of T due to this assumed increase of depth of the beam. The stretching force in the top layer can be found from equation (1) by substituting xd for Z, h . dx for s and RO for L, which gives F = -r^'X'dx this stretching force is a pull at A in Fig. 57, and its torque action about the axis 00 in Fig. 59 is counter-clock wise as seen in Fig. 57, and it is equal to Fx = ^'xHx The compressing force in the bottom layer is a push at B in Fig. 57, and its torque action about the axis 00 is also counter- clock wise as seen in Fig. 57, and it is equal to Fx = Eh R x^'dx Therefore the total torque action due to the two added layers is: dT = ?|^-x2.daj (2) MISCELLANEOUS APPLICATIONS. 119 Whence, by integrating, we have: 2 Ehx^ T = 3 R + a constant (3) But it is evident that T = when x = 0, because a beam of zero depth would of course have no stiffness at aU. Therefore the constant of integration in equation (3) is zero, and equation (3) becomes: (4) It is somewhat simpler to express T in terms of the depth D of the beam {= 2x). Therefore substituting x = ^ inequation (4), we have: 12 B (5) This equation applies to a beam bent in any way whatever, T being the torque action across a particular section of the beam, and R the radius of curvature of the median line at that point. Thus Fig. 60 shows a beam with one end fixed in a wall and carrying a weight W at the other end. In this case the bending torque at p is equal to W(a — x)j and the torque action across the section of the beam at p is equal and opposite to W{a — x) Therefore, ignoring algebraic signs, we have Fig. 60. T = W(a - x) (6) 120 CALCULUS. Substituting this value for T in equation (5) and solving for R we have: _ 1 EhD^ ^~T2W(a-x) ^^^ or R = -^ (8) a — X where c = -i.^^ (9) Let it be required to find the equation of the curve formed by the median line of the beam in Fig. 60. According to Art. 49 the radius of curvature of any curve at a point is: n.hM If the bending of the beam in Fig. 60 is very slight the value of -p will be everywhere very small so that, as a first approxima- tion, we may neglect (~f) in the numerator of equation (10) where it is added to the much larger constant quantity 1. Therefore from equations (8) and (10) we have i = ^ (Py a — X dx" or d^y _ a x dx^ c c whence, by integrating,* we have (11) (12) dy ax x^ , . . ;p = — + Si constant uX C JiiC (13) * By recognizing the function of which is the derivative. MISCELLANEOUS APPLICATIONS. 121 But ~ is actually* equal to zero when a; = in Fig. 60, that the constant of integration in equation (13) is zero, or: so dy dx ax c 2c Integrating this expression, we have: y = ax2 2c ~ 6c + a constant (14) (15) But y = when x = in Fig. 60, so that the constant of integration in equation (15) is zero, and equation (15) becomes: y ax" 2c 6c (16) in which c the constant defined by equation (9). 73. Average value of a function. Let y he sl function of x as represented by the curve cc in Fig. 61. The average value of y between x = a and X = h is the shaded area in Fig. 61 divided by the base {b — a). This gives the height of a rectangle of base {h — a) the area of the rectangle being equal to the shaded area in the figure. From this definition of average value we have: {average value of y between x = a and x Fig. 61. =i]^^a£y- dx (1) PROBLEMS. 1. The velocity of a falling body is v = gt. Find the average value of V from i = to t = 10. Ans. 5g. * Not approximately. 122 CALCULUS. 2. The velocity of a falling body, starting with an initial velocity w, is v = u -\- gt. Find the average value of v from t = 5 to < = 10. Ans. u + 7.5 g. 3. Find the average value of y — sinx from a; = to a; = tt. The answer is shown in Fig. p3. y=8inx Fig. p3. 4. Find the average value of e = E sin (at during a whole cycle from cot = to w^ = 27r. Ans. Zero. 5. Find the average value oi y = sin^ x during a half cycle (x = to X = tt) and during a whole cycle (x = to x = 27r). The answer is shown in Fig. p5. Fig. p5. 6. Find the average value oi y = sin a; sin {x — 0) during a whole cycle {x — to x — 2t). Ans. J^ cos d. MISCELLANEOUS APPLICATIONS. 123 CENTER OF GRAVITY. 74. Resultant of a set of parallel forces. The single force R in Fig. 62 is equivalent to the three parallel forces A, B and C combined, and it is called the ^ ^ resultant of A, B and C. k X >j The value of R is equal to it__a— ^ ! j A -{- B -\- C, and the point of i^ j j^. ji apphcation of R is determined ^ M |b 1^1^ by the condition that the } torque action of R about any Ijij arbitrarily chosen axis X (perpendicular to the plane of p- g2. the paper) must be equal to the combined torque action of the given forces A, 5, and C about that axis. That is: R^ A + B + C (1) (2) A + B + C ^^^ Definition of the center of gravity of a body. Every portion of a body is pulled downwards by gravity, and the forces which thus act upon the various parts of a body are together equivalent to a single force (their resultant) the point of application of which is called the center of gravity of the body. Thus the short arrows /// in Fig. 63 represent the forces with which gravity pulls on a body, these forces are together equivalent to the single force F, and the point of application of F is the center of gravity, C, of the body. The force with which gravity pulls on a body is proportional to the mass, m, of the body. Therefore, if we use the pull of gravity on a one-pound body as our unit of force, the pull and RX = Aa + Bh + Cc so that ■*, V _Aa + Bh-{-Cc 124 CALCULUS. of gravity on any body is numerically equal to the mass of the body in pounds. Thus a 10-pound body would be pulled by 10 units of force. This unit of force is called a pound. It is evident therefore that the word pound has two meanings; it signifies a unit of mass when we speak of so many pounds of sugar or coal, and it signifies a unit of force when we speak of so many pounds of force. Consider a small particle of the body in Fig. 63 of which the mass is dm pounds and of which the abscissa is x as shown. The pull of gravity on the particle is dF = dm, the torque action of this force about the axis (perpendicular to the plane of the paper) is x-dm, the combined torque action about of all the forces fff may be expressed as the integral J x-dm, and this total torque action is equal to the torque action about of the resultant F. Therefore we have: Fig. 63. FX = Jx'dm (4) Now the force F is equal to the total mass M of the body as above explained. Therefore from equation (4) we have: X = fx'dm M (5) Similarly, the y and z coordinates of the center of gravity are given by the equations: F = fy-dm M (6) MISCELLANEOUS APPLICATIONS 125 and Z = f Z'dm M (7) Equations (4) and (5) have exactly the same significance as equations (2) and (3); and when the value of fx-dm is found for the whole body, then the distance X of the center of gravity from the origin can be calculated if the total mass of the body is known. 75. Center of gravity of a bent rod. It is desired to find the coordinates X and Y of the center of gravity of a rod which is represented by the heavy black portion of the parabola which is, shown in Fig. 64 and whose equation is y = kx^. Let h be y-dxis 1 1 I.. V i \ \ \ / r ( 1 •- 1 '^x. --■»< x-axifi t-"!-,- t 1 1 / / 1 \ \ J / i /l>i N. ,X \ Ijc-ojcis ^ JC -^ Fig. 64. Fig. 65. the mass of the rod per unit of length (pounds per foot). Then }i'd^ = hA{dxY-\- {dyf is the mass in pounds of the portion dfs of the rod, as may be understood with the help of Fig. 65. But from the equation of the parabola , y — fcx^, we have dy — 2kx'dx. 126 CALCULUS. Therefore h ^{dxY + {dyY is equal to ^ Vl + ^k'^x'^ • dx and this expression is to be used instead of dm in equation (5) of the previous article so that: £6 Now the mass of the rod in Fig. 64 is the length of the rod multiplied by ^, and the length is given by the integral of Vl + Wx^ • dx from x = a to x = h. That is /'»x=6 M = h \ Vl + Wx'^'dx (2) When M is determined by this equation its value can be used in equation (1) to give the value X. In a similar manner equation (6) of Art. 74 gives, for the bent rod shown in Fig. 64: hh £=6 xHl -^Wx^'dx a Y — M (3) from which Y can be calculated when M is known Jsee equation ^ (2)]. 76. Center of gravity of a solid cone. Consider the solid cone shown in Fig. 66. To determine the location of the center of gravity of this cone equations (5), (6) and (7) of Art. 74 can be used as in the case of a bent rod, but it is worth while to give a slightly different argument as follows: Let T be the torque action of the pull of gravity on the cone, T being reckoned about the axis (perpendicular to the plane of the paper). Then T is evidently a function of the length x of the cone {x is shown in Fig. 67), and the increment of T due to an increment of x is: dT = X'dF (1) J^ MISCELLANEOUS APPLICATIONS. 127 where dF is the force with which gravity pulls on the thin slab of material in Fig. 66, the radius of the slab being y and its thickness dx. The volume of this slab is iry^ • dx and its mass is Fig. 66. X'OxU Fig. 67. irDy^'dx (= dm) where D pounds per cubic foot. But of the half-angle of the cone, so that equation (1) becomes is the density of the material in y = kx, where k is the tangent Therefore dm = dF = irkWx^'dx, whence dT = TrkWx^'dx T = — -i — x^-\- a. constant (2) (3) but it is evident that T is zero when x = because when x = there is no cone at all. Therefore the constant of integration in equation (3) is equal to zero, and equation (3) becomes T=^-m.^ (4) Now the volume of the cone in Fig. 66 is a function of x whose increment dV is: dV = Try^'dx (5) 128 CALCULUS, or, using y = A;a; as before, we have dV = wk^x^dx (6) so that V = iirk^x^ (7) the constant of integration being zero because V = when X = 0. The mass, Af, of the cone is equal to D7 pounds, and the total pull of gravity on the cone is therefore F = DV pounds. The torque action of this force about the axis is FX = DVX = DX times JttAjV and this torque action is equal to T. Therefore, substituting DX times iirk^x^ for T in equation (4) and solving for X, we have: X = ix (8) 77. Average distance of the particles of a body from a plane. A new view of center of gravity. Imagine a body to be made up of minute particles of equal mass. Then the number of particles in a piece of the body would be proportional to the mass of the piece, that is, equal to A^ X mass of piece, where A^ is a constant. Let dm be the mass of a small piece of a body; then N'dm is the number of particles in dm. If x is the abscissa of dm, then x X N'dm is the sum of the abscissas of all the particles in dm, and the integral J Nx • dm or Nj x • dm (extended so as to include the entire body under discussion) is the sum of the abscissas of all the particles of the body. Further- more NM is the total number of particles in the body, M being the mass of the body. Therefore, dividing Nfx-dm by NM we have the sum of the abscissas of all the particles divided by the number of particles, which is the average abscissa of the entire body. Representing this average abscissa by X we have: NjX'dm fx-dm ^ ^ NM ^ ~~M~~ MISCELLANEOUS APPLICATIONS. 129 Therefore, comparing this with equation (5) of Art. 74 we see that the abscissa of the center of gravity of a body is the average abscissa of all the particles of the body, and similar statements can be made with reference to Y and Z as given by equations (6) and (7) of Art. 74. From the point of view of averages the center of gravity of a body is usually called the center of mass of the body. PROBLEMS. 1. Four downward forces of 100 pounds, 125 pounds, 200 pounds and 50 pounds act upon a bar at distances of 10 inches, 16 inches, 18 inches and 24 inches from the end of the bar, respectively. Find the value of the resultant of the four forces and the distance from the end of the bar to the point of applica- tion of the resultant. Ans. 475 pounds, 16.4 inches. 2. Find the center of gravity of a straight bar 15 inches long and of uniform sectional area on the assumption that the density of the material is given by the equation D = kx^; where D is the density of the material at a point, x is the distance of the point from the end A of the rod, and fc is a constant. Ans. 12 inches from end A. B-\ i ^-. O 1f-axi8 circulitr rod \ Fig. pSa. Fig. p36. 3. Find the location of the center of gravity of a rod bent into the arc of a circle as shown in Fig. p3a. Ans. 8.27 inches from the center of the circle in the line AB, 10 130 CALCULUS. Note, Figure p36 shows how this problem may be formulated. The length dx of the portion ds of the rod is where 6 is the angle between ds and ^ cos » * the X-axis and it is equal to the complement of the angle between r and the X-axis. That is cos ^ = -. The total added mass corresponding to dx is the mass of the two short portions ds. Therefore where h is the mass of the rod per unit length. With this start it is easy to carry the problem through to a conclusion. 4. Find the distance from the 16 inch base to the center of gravity of the entire wedge (30 inches long) which is shown in Fig. p4. Ans. 10 inches. 5. Find the distance from the 16 inch base to the center of gravity of the frustum of a wedge (12 inches long) which is shown in Fig. p4. Ans. 5.5 inches. 6. Find the distance from the apex to* the center of gravity of the entire cone (10 inches long) shown in Fig. pQ. Ans. 7.5 inches. 7. Find the distance from the apex to the center of gravity of tlie frustum of a cone (5 inches long) which is shown in Fig. p6. Ans. 8.04 inches. 8. Find the position of the center of gravity of a solid hemi- ^^- MISCELLANEOUS APPLICATIONS. 131 sphere whose radius is 16 inches. Ans. 6 inches from the center of the sphere on the axis of symmetry. Note. This problem can be formulated in a manner very similar to the formulation of the problem of the cone which is discussed in Art. 76; and Fig. p8 will suggest the method of formulation. 9. Figure p9 represents a segment of a solid sphere. Locate its center of gravity. Ans. 14.32 inches from the center of the sphere on the axis of symmetry of the segment. y'^^^^y^segment of sphere (r-x) x-axis \ T ■*—8lab of I thickness dx / / / / Fig. p8. I 11 Fig. p9. 10. Locate the center of gravity of a thin hemispherical shell whose radius is 16 inches. Ans. 8 inches from the center of the sphere on the axis of symmetry of the shell. iVote. Let Fig. p36 represent a thin spherical shell. Then 2-Ky.ds is the area of the portion of the shell corresponding to dx. The mass dm of this portion of the shell is 2-Ky .ds X a, where a is the mass of the shell in pounds per square foot. With this start it is easy to carry the problem through to a conclusion. 1 1 . Locate the center of gravity of the solid paraboloid which is shown in Fig. pll. Ans. 8 inches from the vertex of the paraboloid on the axis of the paraboloid. 12. Locate the center of gravity of the frustum of a solid 132 CALCULUS. paraboloid which is shown in Fig. pl2. Ans. 9.91 inches from the vertex of the paraboloid on the axis. 13. Locate the center of gravity of the thin paraboloidal shell / L7inch±8_ \ 5 inches Fig. pU. Fig. pl2. which is shown in Fig. pll. Ans. 6.73 inches from the vertex of the paraboloid on the axis. Note. In the solution of this problem it is necessary to integrate an ex- pression of the form J a: Va + 6aJ • dx. This expression can be put into a form whose integral is easily recognized by substituting * z^ := a -\- bx 80 that X = dx = z" - c h 2zdz 14. Figure pl4 represents a flat metal plate cut with a parabolic edge. Locate the center of gravity of the whole plate. Ans. 5.4 inches from the vertex of the parabola on the axis. 15. Locate the center of gravity of the portion AB of the plate shown in Fig. pl4. Ans. X = 6.66 inches, 7 = 0. 16. Locate the center of gravity of the portion A of the plate Bhown in Fig. pl4. Ans. X — 6.66 inches, Y = 3.42 inches. Note. Consider a small element of the plate as represented by the shaded square in Fig. pl6. The area of this element is dx.dy and its mass is dm = k'dx'dy, where & is a confitant. Therefore the integrals fx-dm and ■au.^: MISCELLANEOUS APPLICATIONS. 133 J y-dm become kjjx'dx-dy and kjfy-dxdy, respectively. The double integral signs indicate that two integrations are necessary in each case, one with respect to x and the other with respect to y. The limits of these integra- tions are most easily assigned when the integration with respect to y is made first. In this case the integration with respect to y is between the Umits y = and y = Vax (where y^ = ax is the equation of the parabola). In this integration x and dx are treated as constants. Then the integration with respect to a; is made between the limits x = 4 inches and x = 9 inches. 17. Locate the center of gravity of a fiat plate which is in the shape of a sector of a circle as shown in Fig. pl7. Ans. In the line AB at a distance 6.36 inches from A. y-axi8 hemiapfiere thin slab x^axia Fig. pl7. Fig. pl8. 134 CALCULUS. 18. Suppose that the density of the material of which a sphere is made is kr where A; is a constant and r is any distance from the center of the sphere. Locate the center of gravity of half of the sphere, its radius being 2 feet. Ans. X = 0.8 foot. Note. Consider a ring-shaped element of material lying in the thin slab in Fig. pl8; the radius of the ring being y and the thickness and breadth of the ring being dx and dy respectively. The volume of this ring is ^irydx-dy and its mass is equal to kr X 2iry'dx'dy, where r = Vx* + y^ as shown in the figure. Therefore the pull of g ravity (parallel to the y-axis, say) on the ring-shaped element is 2irky Vx^ + y^ -dx-dy and the torque action of this pul about the axis (perpendicular to the plane of the paper) is: dT = 2Trkxy\x^ + y^' dx-dy With this start the problem can be formulated without difficulty. The total mass of the hemisphere can be found by integrating, between proper limits, the expression: dM = 2irky^x^ -{- y^ 'dx-dy 19. What is the average distance of the points on a semi-circle from the diameter which bounds the semi-circle? Ans. 0.64 of the radius. Note. Let N be the nimiber of points per unit length of a line. Then N-ds \a the number of points in the line element ds, and iVTrr is the number of points on the semi-circle. 20. What is the average distance of all the points in a semi- circular area from the diameter which bounds the semi-circle? Ans. 0.42 of the radius. Note. Let N represent the number of points in one unit of area. Then Ttr^ N-dA is the number of points in an element of area dA and A^* -7^- is the number of points in the semi-circle. CENTER OF PRESSURE. 78. Force exerted on a water gate. Figure 68 shows a water gate covering an opening in a dam. The short arrows fff repre- sent the forces with which the water pushes against the gate. These forces are together equivalent to the single force F, their .^ MISCELLANEOUS APPLICATIONS. 135 water level r rrrr dam Fig. 68. resultant, the point of application of which is called the center of pressure on the gate. The force F is equal to the sum of all the forces ///, and the torque action of F about an arbitrarily chosen axis (see Fig. 69) is equal to the sum of the torque actions about of the forces ///. Consider a horizontal strip of the gate of which the width is dx as shown in Fig. 69, and of which the length is w. The area of this strip is w-dx and the force dF ex- erted on the strip is:* dF = Dwx-dx (1) Therefore I X'dx (2) x=a Fig. 69. * The pressure in pounds per square foot at a place x feet beneath the surface of water is p = Dx, where D is the density of water in pounds per cubic foot (= 62K). Furthermore, the force exerted on the strip is found by multiplying the area of the strip in square feet by the pressure in pounds per square foot. 136 CALCULUS. which gives F = iDw{¥ - a^) (3) The torque action of the force dF about the axis is x-dF or Dwx^'dXf and the total torque action about of all the forces /// is '~\^'dx (4) T = DwT *Jx=a , which gives T = \DwQy' - a^) (5) This torque action must be equal to XF, Therefore using the value of F from equation (3) we have \Dw{h'' - a2)X = \Dw{W ~ a^) (6) whence 79. Force exerted on a curved surface by a stationary fluid under pressure. The short arrows /// in Fig. 70 represent the forces with which the water pushes on the curved surface of a dam. All the forces /// are together equivalent to a single force F (see Fig. 73) which is called their resultant. Let Fx and Fy be the x and y components of F as shown in Fig. 73. Then Fx is the sum of the a:-components of the forces fff in Fig. 70, and Fy is the sum of the ^/-components of the forces ///. Let X and Y be the coordinates of the point of appHcationof the resultant force F (see Fig. 73). Then we have the following four conditions which determine Fxy Fy, X and Y: (a) Fx is the sum of the aj-components of the forces fff. (h) Fy is the sum of the ^/-components of the forces fff. (c) The torque action of Fx about the arbitrarily chosen axis is FxY, as may be seen from Fig. 73, ahd this torque action is equal to the sum of the torque actions about of the a;-components of the forces ///. MISCELLANEOUS APPLICATIONS. 137 (d) The torque action of Fy about is FyX, and this torque action is equal to sum of the torque actions about of the ^/-components of the forces ///. In order to formulate these four conditions it is necessary to derive expressions for the x and y components of the force dF which is exerted by a fluid on an element of a curved surface. Consider the surface element AB, Fig. 71, the width of the element being ds and its length I (perpendicular to the plane of the paper). The area of AB is l.ds square feet, and the force in pounds exerted on AB is dF = pl.dSf where p is the pressure ,umter level Fig. 70. of the fluid in pounds per square foot. The x and y components of dF are, respectively: and dF:, = sin0-dF dFy = cos (f)'dF dx Therefore, using pl-ds for dF, using -p for cos 0, and using -r for sin 6, we have ds and dF. ==pl-dy (1) dFy = phdx (2) To determine the total force exerted on the face of the dam in 138 CALCULUS. Fig. 70, that is, to determine the two components Fx and Fy of the total force and the coordinates X and Y of its point of appHcation as shown in Fig. 73, consider the element of area hds as shown in Fig. 72, where I is the length of the dam per- pendicular to the plane of the paper. The pressure of the water at this element is Dy pounds per square foot as explained in the footnote to Art. 78. Therefore, using Dy for p in equations (1) and (2), we have for the x and y components of the force dF in Fig. 72: dFx = Dlydy (3) and dFy = Dlydx (4) From equation (3) we find the x-component of the total force exerted on the dam by integrating between the limits y = Q to y = hj which gives: F, = iDlh^ (5) The equation of the parabola in Fig. 72 is ^ 2 y = p-^^ Therefore, using this value of y in equation (4) and integrating MISCELLANEOUS APPLICATIONS. 139 between the limits x = to x = h, we find the 2/-component of the total force exerted on the dam, namely: Fy = \Dlhh (6) To find the distance X in Fig. 73 it is necessary to find the total torque action about of all the forces dFy and place the result equal to FyX. The torque action of dFy (the i/-com- ponent of dF in Fig. 73) about is a; . dFyj which by equation (4) becomes Dlxy-dx. But so that the torque action of dFy about is DlroOc^-dx, and the torque action of all the forces dFy about is Dlrh I x^-dx O^Jx=0 which is equal to \I)lhh^. Therefore we have XFy = \DlhW (7) or, using the value of Fy from (6), we have: X = 16 (8) The torque action of dFx (the x-component of dF in Fig. 73) about is y-dFx, which by equation (3) becomes Dly^-dy; and the total torque action of all the forces dFx is Jf*V=n which is equal to iDl¥. Therefore we have: YFx = iDW (9) or, using the value of Fx from (5), we have Y = ih (10) 140 CALCULUS. y-axis PROBLEMS. 1. Find the total force F acting on the gate in Fig. 68 and find the distance X, when a = 4 feet, 6 = 8 feet and ly = 4 feet. Ans. 6000 pounds, 6.22 feet. 2. A dam has a circular hole through it 6 feet in diameter. The center of the hole is 7 feet beneath the surface of the water. The hole is covered with a gate. Find the total force with _ which the water pushes on the gate 4---(vr/77r^ 1 ^^^ fiiid the distance from the surface I V///////1 "V r I of the water to point of application of this force. Ans. 12,370 pounds, 7.32 feet. X-axis Note. Consider the element of area which is » represented by the small shaded square m Fig. p2. ^^' ^ • The pressure at this area is kx where A; is a constant. Therefore the force acting on the element of area is kx-dx'dy, and the total force on the gate is jj kx'dx'dy. To specify the limits of this integration let y = f(x) be the equation of the circle in Fig. p2 (the gate). Then if we integrate first with respect to y the limits are y = — fix) to y = +/(x). We may then integrate with respect to x between the Umits X = h — r to x = 6 + r. The torque action of the force kx-dx'dy about the axis (perpendicular to the plane of the paper) is kx^-dx-dy, and the total torque action of the force acting on the entire gate is given by the integral J J kx^'dx-dy. The limits are the same as stated above. 3. Find the abscissa X and the ordinate Y of the center of pressure of the shaded half of the circular water gate in Fig. p2, where 6 = 7 feet and r = 6 feet. Ans. X = 7.32 feet, Y = 1.27 feet. MOMENT OF INERTIA. 80. Kinetic energy of a rotating body. Definition of moment of inertia. Consider a wheel which is rotating n revolutions MISCELLANEOUS APPLICATIONS. 141 per second. Its speed in radians per second is CO = 2irn (1) because there are 2% radians in one revolution. The speed of a rotating body in radians per second is called the angular velocity or the spin velocity of the body. Consider a particle of the wheel at a distance r from the axis of rotation. As the wheel rotates this particle travels in a circle of which the circumference is 27rr and it travels n times round this circle per second so that the velocity v of the particle is V = 2Trnr (2) or using w for 27rn according to equation (1), we have: v = ojr (3) If the spin velocity of the wheel is doubled it is evident from this equation that the velocity of every particle in the wheel will be doubled so that the kinetic energy of every particle in the wheel will be quadrupled. Therefore the total kinetic energy, W, of a rotating wheel is quadrupled if the spin velocity co of the wheel is doubled; that is, for a given wheel, W is pro- portional to £0^, and for the given wheel there is a definite constant by which co^ may be multiplied to give W. That is: W = iKo)'' (4) where (^K) is the proportionality factor for the given wheel. The constant K is called the moment of inertia of the wheel,* and it depends upon the size, shape, and mass of the wheel. * The kinetic energy of a particle is equal to ^mv^, where m is the mass of the particle in pounds, v is its velocity in feet per second, and kinetic energy is expressed not in foot-pounds, but in foot-poundals. Throughout this dis- cussion of moment of inertia distance is expressed in feet, velocity in feet per second, mass in pounds, force in poundals, torque in poundal-feet, work or energy in foot-poundals, and moment of inertia in pound-feet.^ 142 CALCULUS. 81. General integral expression for moment of inertia. Imag- ine a small particle of mass dm to be added to the spinning body at a distance r from the axis of spin. Then the velocity V of the added particle is cor, and the kinetic energy of the added particle is ^dm X coV which is, of course, the infinitesimal increment of the kinetic energy of the spinning body due to the added particle. Therefore dW = ioy^rHm and by integration* we have W = Wfr^-dm (5) Comparing this equation with equation (4) it is evident that K^fr^'dm (6) 82. Average value of the square of the distances of all the particles of a body from an axis. Definition of radius of gyration. Using the ideas of Art. 77, N-dm is the number of particles in a small piece dm of a body. If r is the distance of the small piece dm from a chosen axis, then Nr^-dm is the sum of the squares of the distances of all the particles in dm from the axis, and the integral jNr^.dm or Nfr^-dm extended so as to include an entire body is the sum of the squares of the distances of all the particles of the body from the axis. But NM is the number of particles in the body. Therefore N f r^ ' dm fr^ • dm — ^ = ^ (I) is the average value of the squares of the distances of all the particles of a body from the axis. This average may be repre- * This integration can only be indicated. Before the actual value of the integral can be found r and dm must be expressed in terms of one or more independent variables, and the limits of the integration must be such as to include the entire spinning body. MISCELLANEOUS APPLICATIONS. 143 sented by p^ so that f r^ • dm M or p2M = fr^-dm (2) (3) but J 7-2 • dm is the moment of inertia K of the body with respect to the chosen axis, according to equation (6) of Art. 81. There- fore we have: K = pm (4) The distance p, which is the square-root-of-thc-aVerage-value- of-the-squares-of-the-distances-of-all-the-particles-of-a-body-from- a-chosen-axis, is called the radius of gyration of the body with respect to the chosen axis. 83. Moment of inertia of a circular saw about its axis of spin. The moment of inertia of a circular saw could be easily derived from equation (6) of Art. 81, but it is instructive to give the complete argument again as follows: Let W be the kinetic energy of a circular disk r feet in radius rotating at a constant spin velocity of co radians per second, as shown in Fig. 74, and n— -.^- added , material Fig. 74. let it be required to find the infinitesimal increment of W due to an arbitrary infinitesimal increment of r. Let t be the thick- ness of the disk and let D pounds per cubic foot be the density of the material of which the disk is made. Imagine the radius 144 CALCULUS. of the spinning disk to be increased by the addition of material as indicated in Fig. 74. The volume of the added material is 2irr X t X dr, the mass of the added material is 2-Ktr'dr X D pounds, the velocity of the added material is or, the kinetic energy of the added material is equal to one half the product of its mass times the square of its velocity, and the kinetic energy of the added material is the desired infinitesimal increment dW. Therefore : dW = rDUa'^r^'dr (1) W and r being the only variables. Therefore, by integration we get: W = JttD^coV^ + a constant (2) But W must evidently be zero when r is zero. Therefore the constant of integration must be equal to zero, so that equation (2) becomes W = iirDtJ'r' (3) Now irrHD is equal to the mass m of the disk in pounds so that equation (3) becomes: W = iw^mr^ (4) But according to Art. 80 the kinetic energy of any rotating body can be expressed as ^Ku^, where K is the moment of inertia of the body. Therefore we have ^v W = i«2mr2 = ^Kc^ (5) from which we have: K = imr^ (6) That is, the moment of inertia of a circular saw about its axis of spin is equal to one half the mass of the saw in pounds multi- plied by the square of the radius of the saw. Remark. The thickness i in the above discussion may be anything whatever. Therefore equation (6) expresses the mo- ment of inertia of a circular cylinder of any length rotating about its axis of figure. MISCELLANEOUS APPLICATIONS. 145 84. Moment of inertia of a rectangular bar. To determine the moment of inertia of the rectangular bar shown in Fig. 75 the general equation (6) of Art. 81 will be used. Figure 76 represents a top view of the bar, being the axis of rotation. Consider the ele- ment of the barwhich is rep- resented by the small black square in Fig. 76. This ele- ment is understood to ex- tend entirely through the bar parallel to the axis paper in Fig. 76). Fig. 75. (at right angles to the plane of the Therefore the volume of the element is t'dx'dy and its mass is tD-dx' dy. The distance of the element from the axis is r = Vx^ + y^, so that r^ = x^ + y^ Therefore the general equation (6) of Art. 81 becomes: K = tnffix^ + 2/2) 'dx'dy (1) in which the double integral sign is used because two integrations, one with respect to x and one with respect to y are necessary, as explained in the note to problem 16 on page 132. If we integrate with respect to y (treating x and dx as con- stants) between the limits L y = to t/ = + - we Fig. 76. get the moment of in- ertia with respect to the axis of the thin slab be- tween the dotted lines in Fig. 76. We may then inte- l . . I grate with respect to x between the limits x = —-^ to x 11 = + 2 146 CALCULUS. and we get the desired expression for K, namely: K = ^\V + ^) (2) But Dtwl is the mass m of the bar in pounds, so that equation (2) becomes: X=^(P + «,»)« (3) Remark. The same final result is obtained if we integrate first with respect to x (treating y and dy as constants), and then with respect to y. What is the meaning of the result of the first integration in this case? 85. Torque required to increase the spin- velocity of a body. The velocity of a particle in a rotating wheel is t; = rw, according to equation (3) of Art. 80. Therefore when « increases, v increases r times as fast as w. That is: dv du) ,^v This -r. is the acceleration of the particle in the direction at right angles to r*, and dv J / d(a J \ is the sidewise force (at right angles to r) which must act on the particle to produce the sidewise acceleration. The torque action of this sidewise force about the axis of rotation is: dT = r^'dmXr dt or dT = ~ -r'-dm (2) * We are not here concerned with the radial acceleration of the particle which is discussed in Art. 50. MISCELLANEOUS APPLICATIONS. 147 so that the total torque action required to increase the spin velocity of the body at the rate -t: is : ^''^r'^-dm (3) T = — C dtj The integral is of course understood to be extended over the whole body and it is equal to the moment of inertia of the body according to equation (6) of Art. 81. Therefore equation (3) may be written r = Kf (4) That is, the spin acceleration -^ of a body multiplied by the moment of inertia of the body is equal to the torque which must be exerted on the body to produce the spin acceleration. 86. Moments of inertia about par- ^dsf allel axes. Let K be the moment ^y"^ of inertia of a body about an axis ^^^ A (perpendicular to the plane of the ^^ / 1 paper in Fig. 77) which passes through *i-.^^2 ^-. the center of gravity of the body, pjg 77^ and let K' be the moment of inertia of the body about an axis P (perpendicular to the plane of the paper) which is at a distance a from 0. Then K' = Z + aW (1) where M is the total mass of the body in grams or pounds as the case may be. Proof. From the triangle in Fig. 77 we have: gf2 = (j2 _j_ j,2 _j_ 2ar cos B or 52 = (j2 4. y.2 _|_ 2ax (2) where x is the abscissa of the element of material dm referred 148 CALCULUS. to the center of gravity as an origin. Now according to equation (6) of Art. 81 we have K' = fq'-dm (3) Therefore, using the value of q^ from equation (2), we have: K' = fa^-dm + fr^-dm + f2ax-dm K' = a'fdm + Jr^-dm + 2afx'dm (4) But J dm = the total mass M of the body, jr^-dm is the moment of inertia K referred to the axis 0, and jx-dm is zero according to equation (5) of Art. 74, because the center of gravity in Fig. 77 is taken as the origin. Therefore equation (4) becomes: K' = am + K (1) 87. Moment of section of a beam. Consider equation (2) of Art. 72, namely: dT = ^^x^'dx (1) The product 2b -dx is the area of the shaded strips in Fig. 59, and X is the distance of this area from the axis 00, Let us use dA for 2b 'dx, then equation (1) becomes: dT = ^x^'dA (2) 80 that ^ "* :^'dA (3) -f/' The integral fx* • dA is called the moment of section of the beam; and according to equation (3) the torque T which bends any beam is equal to -h- » where E is the stretch modulus of the K material of the beam, R is the radius of curvature of the median line of the beam, and S[ =fx* • dA) is the moment of section of the beam. MISCELLANEOUS APPLICATIONS. 149 The integral fx^-dA is similar in form to the integral fr^-dm in equation (6) of Art. 81, and because of this similarity engineers call J x^ ' dA the "moment of inertia " of the section of the beam. The term wmnent of section is however the correct term. PROBLEMS. 1. Find the moment of inertia of a circular saw six feet in diameter and of which the mass is 125 pounds. Ans. 562.5 Ib.-ft.^ 2. Find the amount of energy in foot-pounds stored in the saw of problem 1 at a speed of 600 revolutions per minute. Ans. 34482 foot-pounds. Note. In the formula W = }4Ku^, K is expressed in pound-feet-squared, w is expressed in radians per second, and W is expressed in foot-poundals (not in foot-pounds). There are 32.2 foot-poundals in one foot-pound. 3. Find the moment of inertia of a cylinder 2 feet in diameter and 3 feet long with respect to its axis, the density of the material being 420 pounds per cubic foot. Ans. 1980 Ib.-ft.^ 4. Find the moment of inertia of a hollow-cylinder, the external dimensions being the same as in problem 3, inside diameter being 1 foot, the axis of revolution being the axis of the cylinder, and the density of the material being 420 pounds per cubic foot. Ans. 1114 lb.-ft.2 5. Find the moment of inertia of the rectangular bar in Figs. 75 and 76. The bar is 5 feet long, 1 foot wide and 0.5 foot thick and it has a mass of 1000 pounds. Ans. 2167 lb.-ft.2 6. Find the moment of inertia of the bar in problem 5 when the axis of rotation coincides with the edge t in Fig. 75. Ans. 8667 Ib.-ft.^ 7. Find the moment of inertia of a very thin circular disk when If-axis ,^axi8pf^ revolution x-axi8 Fig. p7. 150 CALCULUS. the axis of rotation is a diameter of the disk; the mass of the disk being 120 pounds and its diameter being 6 feet. Ans. 270 Ib.-ft.^ Note. Let the circle in Fig. p7 represent the disk. Take the narrow vertical black strip as dm. Then dm = k X2y-dx where A; is pounds per square foot of disk area. 8. Find the moment of inertia of the thin circular disk with respect to the axis as shown in Fig. p8, the density of the front view Fig. p8. / 7?2 \ material being D. Ans. K = tR^D l-j-^x^j-dx. Note. The mass of the narrow strip in Fig. p8 is 2D a//? _ y^'dx'dy = dm and the distance of the strip from the axis is Va:» + 2/2 = r Of course x and dx are constants in this problem. 30 inches Fig. p9. hoop 3 feet in diameter Fig. plO. MISCELLANEOUS APPLICATIONS. 151 9. Find the moment of inertia of the soUd cylinder shown in Fig. p9; the density of the material being 0.28 pound per cubic inch. Ans. 1833 Ib.-ft.^ Note. The method of formulating this problem is suggested by problem 8. 10. A slender circular ring or hoop has a mass of 10 pounds and it is 3 feet in diameter, (a) Find its moment of inertia about the axis C (perpendicular to the plane of the paper in Fig. plO), and (6) find its moment of inertia about the axis ab. Ans. (a) 22.5 Ib.-ft.^ (6) 11.25 lb.-ft.2 11. Find the moment of in- ertia of a solid steel sphere 10 inches in diameter, the axis of revolution being a diameter of the sphere. The density of steel is 0.28 pound per cubic inch. Ans. 22.4 Ib.-ft.^ cuds of revolution solid sphere cylindrical shell Note. Let the circle m Fig. pll represent the sphere. Take for dm a thin cylindrical shell of radius r and thickness dr. Fig. pll. 12. Find the moment of inertia of a steel governor ball, 4 inches in diameter, with respect to the axis of revolution of the governor, the center of the ball being 6 inches from the axis. Ans. 2.45 Ib.-ft.^ TABLE OF MOMENTS OF INERTIA. Axis through center of mass in each case, m = mass of body in grams or pounds. K = moment of inertia. I and having uniform section, 1. Thin straight bar of length axis at right angles to bar. K ■■ 2. Rectangular parallelopiped, axis parallel to one edge, and b being lengths of other edges. K = j^ia^ + b'^)m a 152 CALCULUS. 3. Cylinder or disk of radius r, referred to axis of cylinder. K = ir^m. Referred to axis at right angles to axis of cylinder --(5+0 m where I is the length of the cylinder. 4. Hollow cylinder oirsidu R and r, referred to axis of cylinder. K = ^{R^ + r^)m. 5. Sphere of radius r referred to a diameter. K — \r^m. Note. If the axis of rotation does not pass through the center of mass, use equation (1) of Art. 86. CHAPTER VI. EXPANSIONS IN SERIES. USE OF COMPLEX QUANTITY. 88. Maclaurin's theorem. — Let y be any function whatever of X which is finite and continuous and of which all of the derivatives with respect to x are finite and continuous. Then: y = A4-Bx+^Cx2 + -ilDx«+|^Ex^+... (1) where A, B, C, D, etc., are constants as follows: A is the value of y when X = B is the value of y^ dx when X = C is the value of -r^ when X = D is the value of , , when X = E is the value of ^ when X = etc., etc. Equation (1) expresses what is known as Maclaurin^s theorem. Proof. — Let the curve cc in Fig. 78 represent the given function. Then the value of y which is to be expressed by equation (1) is the ordinate of the point p. To establish equa- tion (1) we will make a series of approximations and consider the limit towards which this series of approximations trends, as follows: First approximation. — To get a first approximation let us dv dv assume that -^ is equal to the constant B (the value of -^ 153 154 CALCULUS. when X = 0) everywhere between the points p and q in Fig. 78. That is by assumption we have: dx ^ B (2) Integrating this differential equation we have: y = Bx -\- a. constant but y = A when x = 0, so that the constant of integration is Fig. 78. evidently equal to A. Therefore as a first approximation we have: y = A + Bx (3)* dv It is interesting to note that to assume -^ = B everywhere is the same thing as to take the inclined dotted line in Fig. 78 as an approximation to the curve cc; and the ordinate of this inclined straight line is A + Bx. * Figure 79 shows a curve pq for which dy dx B everywhere, but the ordinate of p is not equal to A -{■ Bx because of the discontinuity or jump at d. The function y and all of its derivatives must be finite and continuous everywhere between p and q, as stated at the beginning. Any discontinuity outside of the region between p and q does not vitiate equation (1) for the region between p and q. EXPANSIONS IN SERIES. 155 Second approximation. — To get a second approximation let us assume that -j-^ is equal to the constant C (the value of -T^ when a; = 0) everywhere between p and q in Fig. 78. That is, by assumption, we have: d^y Integrating once we have: d.^ = ^ w ^ = Cx + a constant dx dii But -f — B when a; = 0, so that the constant of integration is equal to 5, giving: % = B + Cx (5) Integrating again we have : y = Bx + -x Cx^ + a constant But y = A when x = 0, so that the constant of integration is equal to A, giving as a second approximation: y = A + Bx + ^Cx2 (6) Third approximation. — To get a third approximation let us assume that -r^ is equal to the constant D (the value of -~ when X = 0) everywhere between p and q in Fig. 78. That is, by assumption, we have: By three successive integrations (the constant of each Integra- 156 CALCULUS. tion being determined as above) we get, as a third approximation: y = A + Bx4-|cx2 + iDx3 (8) nth approximation. — To get an nth approximation let us assume that 7-^ is equal to the constant N (the value of t^ when a; = 0) everywhere between p and q in Fig. 78. That is, by assumption, we have: g = ^ (9) By n successive integrations this differential equation gives, as the nth approximation: y= A + Bx +^Cx2 + ||Dx3 + . . . + j^Nx" (10) To show that the nth approximation approaches the true value of y as a limit as n approaches infinity. In the first place it is evident that equation (10) gives equation (1) when n is in- creased more and more, but it remains to be shown, that y in equation (10) approaches the correct value of ?/ as a limit as n is increased. d^ti The nth derivative -t\ is assumed to be everywhere finite and it must have therefore a definite largest value V and a definite smallest value v between p and q. If the largest value V is used instead of N in equation (10) we get too large a value, a, for y. If the smallest value v is used instead of N in equa- tion (10) we get too small a value, b, for y.* That is: a = A + Bx+~Cx^+"'n'Vx'' (11) * This statement happens to be plausible and therefore the reader is apt to accept it as true without actually perceiving its truth, which is indeed evident if one takes the trouble to think about it. EXPANSIONS IN SERIES. 157 and h = A -i- Rv 4- ^^ , 2 ' 1^ A + Bx+lcx" + . . . ^. t;x" (12) and the true value of y lies between a and b. But, subtracting equation (12) from equation (11) member by member we get: and this difference approaches zero as n approaches infinity, because when n is increased by 1 the numerator is multiplied by the finite quantity x, whereas the denominator is multipUed by the quantity n + 1, which becomes as large as you please. It is evident therefore that the value of a as given by equation (11) becomes more and more nearly equal to the value of 6 as given by equation (12). But the true value of y Hes between a and h. Therefore equations (11) and (12) both approach the true value oi" y as a limit as n is increased; and equations (11) and (12) reduce to equation (1) when n is indefinitely great. Taylor's series. — Heretofore any algebraic expression contain- ing X has been spoken of as a function of x. Thus ex -\- e is a function of x, c and e being costants. Also such expressions as e^a;+/i)^ sin(x + h), tan (a; -f h) are functions of x. If, however, a: is a constant and h a variable we would think of these expres- sions as functions of h. Let y be any function whatever of {y + h), let a; be a constant and h a variable, and let it be understood that y and all of its derivatives are finite and continuous. Then y may be expanded by Maclaurin's theorem, giving: y [any function of (x + h)] -=A+Bh-\- ~Ch^ + ^D¥ + • • • (14) where A, B, C, D, etc., are constants, as follows: A is the value of y when h = 158 CALCULUS. dv B is the value of -4 when ^ = an C is the value of -r^ when h = etc., etc. Equation (14) is sometimes called Taylor^s series, but it is identical to Maclaurin's series and to give it a separate name is to create a false distinction. 89. Examples, (a) Expansion of e*. — The successive deriva- tives of e* are as follows: y = e' which is equal to 1 when x = dxi -^ = e^ which is equal to 1 when a; = j\ = e* which is equal to 1 when a; = etc., etc. Therefore A = B = C = D = etc. = 1, and equation (1) of Art. 88 gives: e' = 1 + X -^-^x^ + j^ + ^x' -^ etc. (1) (6) Expansion of sin x. — The successive derivatives of sin x are as follows: 2/ = sin a; which is equal to when re = 3^ = cos a; which is equal to 1 when a; = ax -t5 = — sin a; which is equal to when a; = -M = — cos a; which is equal to — 1 when a; = t4 = sin a; which is equal to when x = ax* and so on in endless repetition. EXPANSIONS IN SERIES. 159 Therefore A = 0, 5 = 1, C = 0, D = ~ 1, ^ = 0, etc., and equation (1) of Art. 88 gives: />t3 /v»5 /v»7 /v»v Binx = x-j- + j--j- + -^ etc. (2) (c) Expansion of cos a;. — The successive derivatives of cos x are as follows: y = cos X which is equal to 1 when x = -^ = — sin X which is equal to when x = — I = — cos X which is equal to — 1 when x = -^ = sin ic which is equal to when a; = -T-^ = cos X which is equal to 1 when x = and so on in endless repetition. Therefore A = 1, 5 = 0, C = - 1, D = 0, ^ = 1, etc., and equation (1) of Art. 88 gives: /ji»2 /v«4 /J.6 /»8 COS X = 1 -^ + ^ - y + -|g- etc. (3) 90. Maclaurin*s theorem applied to a function of two inde- pendent variables. — If u is a function of x and y which is itself finite and continuous, and if all of its derivatives (partial derivatives) are finite and continuous, then: u = A + B.X + Byy + i(C..x2 + 2C.yXy + Cyyy') + ||-(i>xXXa^ + SD,,yX^y + 3D,yyXy^ + Dyyyf) (1) etc. etc. etc. 160 CALCULUS. Where A is the value of u when x and y are both zero. Bx is the value of -j- when x and y are both zero. d/li By is the value of -r- when x and y are both zero. Cxx is the value of ^J when a: and ?/ are both zero. Cxy is the value of , , or -r-^ when x and 2/ are both zero. Cyy is the value of -r-y when a; and y are both zero. etc. etc. etc. The proof of Maclaurin's theorem as applied to a function of two or more variables is essentially identical to the proof of the theorem as applied to a function of one variable. To enable one to appreciate the modified character of argument and especially as an interesting example of partial integration, let us consider what we may call the second approximation, that is the expres- sion we get for u on the assumption that the respective second derivatives are everywhere constant and equal to Cxx, Cxy and Cyy (their respective values when a; = and j/ = 0) as follows: dhi ^ Cxy (3) dhi dxdy |S = C. (4, Integrating equation (2) with respect to x and equation (3) with respect to y we have T- = CxxX + any function of y (5) and T- = CxyV + any function of x (6) Now -T- is equal to Bx when x and y are both equal to zero. Therefore the constant term in "any function of y" in (5) is equal to Bx, and the constant term in "any function of a;" in (6) is equal to Bx. Therefore, using the EXPANSIONS IN SERIES. 161 expression "vanishing function of ?/" for a function of y which is equal to zero when y = 0, equations (5) and (6) become: -3- = CxxX + (a vanishing function of y) + Bx (7) and ;t- '= Cxyy -\- {Q' vanishing function of x) -\- Bx (8) But these two expressions must be identical, consequently the "vanishing function of ?/" in (7) must be dyy and the "vanishing function of x^' in (8) must be CxxX. Therefore both of these equations reduce to ^ = C.x + C.,7/ + jB. (9) In a similar manner we may integrate equation (4) with respect to y and equation (3) with respect to x, and get the equation: ^=Cyyy+CxyX-\-By (10) Now equations (9) and (10) may be integrated, giving: u = \CxxX^ + Cxyxy + BxX + any function of y (11) and w = ICyyy^ + CxyXy + Byy + any function of x (12) But u = A when x and y are both equal to zero, therefore " any function of ?/" in (11) must be ("a vanishing function of ^") + -4, and likewise, "any function of x" in (12) must be ("a vanishing function of re") -{-A. Further- more, equations (11) and (12) must be identical so that the "vanishing func- tion of 1/" in (11) must be }4Cyyy^ + Byy, and the "vanishing function of x" in (12) must be }4CxxX^ + BxX. Therefore equations (11) and (12) both reduce to: u = A + BxX + Byy + liCxxX^ + 2Cxyxy + Cyyy^) (13) which is the approximate value of u as derived from equations (2), (3) and (4). PROBLEMS. 1. Expand log(l + x) by Maclaurin^s theorem. Ans. />»2 ^3 /v»4 /v*5 log(l+a;) = x-2+3--4 + 5 12 162 CALCULUS. Note. — The function log x and all of its derivatives become infinite for a; — 0, and therefore log x cannot be expanded in powers of x. 2 . Using the answer to problem 1 make an attempt to calculate the logarithm of by placing x = — 1 and adding together a number of terms of the series. Note. — The series obtained by Maclaurin's theorem for log (1 + x) cannot be used for values of x which lead up to or beyond a point where log (1 + x) or any of its derivatives become discontinuous or infinite. 3. Expand cos(a; + h) in a series of ascending powers of h, Ans. cos (a; + ^) = cos a; — /i sin x — i^cos x + "to" sin a; + • • • 4. Expand y = tan x by Maclaurin's theorem. Ans. , , x^ , 2x^ , tana: = a;+3-+-j5-+ ••• Note. — ^When x = - the given function and its derivatives become infinite. Therefore the series found in answer to this problem does not give the value of tan X for values of x equal to or greater than ^. 91. Demoivre*s Theorem. — An important algebraic identity, which was discovered by Demoivre, is expressed by the equation: e'* = cos a; + j sin X (1) where e is the Napierian base and j = V— 1.* This relation is known as Demoivre^s theorem, and it is very useful in certain transformations for purposes of integration, and it is also useful in the theory of alternating currents. To establish equation (1) write jx for x in equation (1) of Art. 89, remembering that f = — 1, f — — jj j^ = + 1, etc., and we have: -I x^ . x* x^ . . , . / a^ x^ ^' I i. \ /o\ * It is customary in the th eory o f alternating currents to use i for electric current, and j is used for V — 1* EXPANSIONS IN SERIES. 163 But the real terms in this series give a series identical to equation (3) of Art. 89, and the imaginary terms give a series identical to equation (2) of Art. 89. Therefore from equation (2) we get g/^ = cos X + j sin X. Definition of complex quantity. Geometric representation of complex quantity. — Any expression like a-\-h V — 1, which is part real and part imaginary is called a complex quantity. Thus the right-hand member of equation (1) is a complex quantity; and of course e^'* is a complex quantity because equation (1) shows that e^'* is part real (cos x) and part imaginary (j sin x). The accepted method of rep- resenting a complex quantity geo- metrically is shown in Fig. 80. ' pjg go. The vector E is the complex quan- tity, its a:-component is thought of as a real quantity a, its ^/-component is thought of as the imaginary quantity jh, and the vector is the sum of its two components. That is E = a+jh* * This algebraic expression of a vector as a complex quantity is one aspect of the important use of complex quantity in the theory of alternating currents. See Frank lin and Esty's Elements of Electrical Engineering^ Vol. II, Chapter V; The Macmillan Co., New York, 1908. Another aspect of the use of complex quantity in the theory of alternating currents is exhibited in Chapter VII of this treatise where the fundamental differential equations of alternating currents are integrated with the help of transformations involving the use of complex quantity. The practical use of complex quantity in alternating-current theory is due chiefly to C. P. Steinmetz; but the use of complex quantity in the solution of linear differential equations as explained in Chapter VII contains everything that is now known of the use of complex quantity in alternating-current theory in a manner which is self-evident, and the use of complex quantity as exemplified in Chapter VII is much older than electrical engineering. 164 CALCULUS. 92. Euler*s expressions for sin x and cos x. — Many differential expressions can be reduced to simple recognizable forms for purposes of integration with the help of Demoivre's theorem, and it is in some cases convenient to modify equation (1) of Art. 91 so as to express sin x and cos x in terms of exponentials Such expressions are due to Euler and they are; sm X = — ^ — (1) and cos X = » (2) The use of these equations is exemplified below. They are derived as follows: From Demoivre's theorem we have: g;* = cos a; + 3 sin x (3) Write — a; for a; in this expression, remembering that cos(— x) = cos a; and that sin(— x) = — sin x and we have: g-yx = cos X — j sin x (4) Subtracting equation (3) from equation (4) member from member we get equation (1), and adding equations (3) and (4) member to member we get equation (2). 93. Example showing use of Euler's equations. — Consider the function : z = sin mx cos nx (1) To find the average value of this function between x = and a; = 27r it is necessary to integrate z-dx between the Hmits x = and a; = 27r as explained in Art. 73. Now the function given in equation (1) is a function whose average value is of fundamental importance in connection with Fourier's theorem (see Chapter VIII) and therefore it is important to be able to EXPANSIONS IN SERIES. 165 reduce the differential expression sin mx cos nx -dx to a combina- tion of fundamental forms which can be found in Class A of the table of integrals in appendix B. This reduction may be easily made with the help of Euler's equations as follows: It is required to integrate the differential equation: dy = sin mx cos nx • dx (2) Writing mx for x in Euler's expression for sin x we get an expression for sin mx, and writing nx for x in Euler's expression for cos X we get an expression for cos nx. Substituting these expressions for sin mx and cos nx in equation (2), we get: dy = ^. e^"("'+">^-rfa; + ^.e''^"'-''^'-"dx - }^ e-i(»»+")=^ 'dx-^. e-J («-">^ • dx (3) Each term in the second member of this equation is of the form ae^^-dx of which the integral (ignoring constant of integration) is r • &^. Thus, in case of the first term a = —. and 5 = 2{^ + n) so that the integral of the first term is 1 4(m + n) e' {m-\-n)x Proceeding in a similar manner with each term and arranging the result systematically we get: ^ " 4(m + n) ^^ ^ ^ 4(m — n) Fg— y(m-n)x _ g;(m— n)il /^\ And this expression can be easily reduced to form 24 in the table of integrals by using Euler's equations. 166 CALCULUS. PROBLEMS. 1. Reduce dy = sm^x-dx to a combination of standard forms as given in the table of integrals. Ans. dy ^ h-. (e~'"'* - e^' + Z&' - Se"'') *dx 2. Reduce dy = coB>^X'dx to a combination of standard forms. Ans. dy = -hie^"^ + e-''^ + ^&^ + ^e-^^ -\-^)'dx 94. Hjrperbolic sines and cosines.* — In the solution of the differential equation of the alternating-current transmission linef the expressions (e* — e~^) and {e' + e"') occur, and trans- mission line calculations are facilitated by the use of tables giving the values of (e* — e~^) and (e* + e~^) for various values of x, and of course the discussion of such calculations is simplified by having names for {e" — e~*) and {e' + e~'). Indeed the expressions ^ and ^ are related to the equilateral hyperbola in the same way that ^. ( = sin x) and ^ ( = COS x) are related to the circle. Therefore these expressions are called the hyperbolic sine of x (sinh a;) and the * Tables giving values of a, h, c and d in the expressions cosh(aj + jy) = a -^-jb and sinh(x -^jy) = c -\-jd for various values of x and y are published in a supplement to the General Electric Review for May, 1910. t The simplest discussion of this subject is that which is given in Franklin's Electric Waves, pages 141-153, The Macmillan Co., New York, 1909. This discussion is essentially complete, although no mention is made of hyperbolic sines and cosines. EXPANSIONS IN SERIES. 167 hyperbolic cosine of x (cosh x) respectively. That is: and sinh X = ^ (1) cosha; = ^^i^ (2) PROBLEMS. 1 . Differentiate y = sinh x» 2. Differentiate y = cosh x, 3 . Integrate dy = sinh x • dx. 4. Integrate dy = coshx-dx. For answers see forms 29 and 30 of table of integrals in Appendix B. CHAPTER VII. SOME ORDINARY DIFFERENTIAL EQUATIONS. 95. Degree and order of a differential equation. — The simple equation az -{- h = is said to be linear with respect to z because it contains no power of z higher than the first power. A differential equation of the form: is called a first degree or linear differential equation because it contains no products of y, ~ and -—, and no powers of 2/, ^, etc., higher than the first. The coefficients A, B and C in a linear differential equation may contain the independent vari- able X, but we shall confine our attention in this chapter chiefly to linear differential equations with constant coefficients. If the first derivative,* only, appears in a differential equation, the differential equation is said to be of the first order. If the second derivative occurs (with or without the first derivative) the differential equation is said to be of the second order; and so on. 96. Ordinary and partial differential equations. — A differential equation expressing the law of growth of a function of one inde- pendent variable is called an ordinary differential equation. Many examples of ordinary differential equations are given in Chapters I and V. See Art. 24 in particular. * The terms degree and order apply to partial differential equations as well as to ordinary differential equations. A function of two variables, however, has two first derivatives, three second derivatives, and so on as explained in Art. 59. Therefore it is somewhat misleading to speak of the first derivative, or the second derivative in explaining what is meant by a differential equation of the first or second order. 168 i ORDINARY DIFFERENTIAL EQUATIONS. 169 A differential equation which expresses the law of growth of a function of two or more independent variables is called a partial differential equation. Some examples of partial differential equa- tions are given in Chapter V. See Arts. 60 and 90 in particular. This chapter is devoted to the discussion of a few important ordinary differential equations, and several important partial differential equations are discussed in Chapters VIII and IX. 97. Pure and mixed differential equations.* — A pure differ- ential equation contains but one derivative (the first or any- higher derivative) and does not contain the dependent variable y. Thus dy = Qx- dx, -^ = sin x, ^\ = log x are pure differential equations. A mixed differential equation contains more than one derivative, or it contains one or more derivatives and the dependent variable y. Thus are mixed differential equations. Solution of pure differential equations. — A pure differential equation can be integrated by looking up the appropriate form in the table of integrals. This is exemplified by every problem in integration heretofore given in this treatise. In the case of a pure differential equation of the second or third order, successive integrations are of course necessary. For example, consider the third order pure differential equation: Let q represent ~, then equation (1) becomes: 1 = -' (2) *This and the following articles refer primarily to ordinary differential equations. 170 CALCULUS. This equation can be integrated, giving; 9 = 3 = 4<^ + C (3) Let p represent -p, then equation (3) becomes: %-ia^ + C (4) This equation can be integrated, giving: P = ^ = i\ax' + CX + D (5) This equation can be integrated, giving y = -hax^ -f ^Cx^ + Dx + E (6) Solution of mixed differential equations. Separation of variables. — The linear differential equation (1) of Art. 95 is of course a mixed equation, and the general solution of this equation with constant coefficients is given in a subsequent article. We will here consider one or two examples of a simple transformation, called the separation of variables, which can sometimes be used to bring a mixed differential equation into a form which can be integrated by looking up appropriate forms in the table of integrals. As a first example consider: This equation reduces to: 1 = ^ (^) ^ = x'^'dx (8) y and this equation can be integrated with the help of forms 1 and 2 of the table of integrals, giving log2/ = |x3 + C (9) ORDINARY DIFFERENTIAL EQUATIONS. 171 As a second example consider dij Let p represent -^ and this equation becomes g = x«p (12) and the integral of this equation, according to equation (9), is: \ogv = \:x? -\- C or P = I = ei^'^" (13) This is a pure differential equation and its integral can be found by looking up the appropriate forms in the table of integrals. PROBLEMS. Solve the following differential equations: 1. x3 ^ = 2. Ans. 2/ = log a; + Cx" ■\- Dx -^ E, 2. ^ = xe', Ans. y = {x - 2)e' + Cx + D. 3. ^ = 27 sin^x. Ans. y = 21 cos x- cos^ x -\-Cx^ +Dx -{-E, 4. 1 - 2/ + (1 + a;) ^ = 0. Ans. log \-^ = C. ax 1 ~~ 2/ 5. 1 - 21/ = 3^. Ans. C(l - 2y) = e"?. 6. sin 2/-cix + a; cos i/-(f?/ = 0. Ans. xsiny = C. 7. sin a; cos y-dx — cos x sin y - dy = 0. Ans. cos y = C cos a;. 8. fa + a;i/ )rfa; + (a; — xy) dy = 0. Ans. logxy = C — x + y. 9. Vl — 2/^ • (ia; + V 1 — a;^ • (^2/ = 0. Ans. sin""^ x-\- sin"^ y=C. ^°-S + .-| = "- Ans.Clogx = . + D. 172 CALCULUS. 11. ^ + (^f^y + 1=0. Ans. y = log cos {x - C) + D. 12.(l+.^)g + .| + ax = 0. Ans. y = D — ax -\- C log [x -\- Vl + x^]. 98. The general solution and particular solutions of a differ- ential equation. — The general solution of a differential equation is an expression for y (the dependent variable) in terms of x (the independent variable) which includes every possible* func- tion which satisfies the differential equation. Thus equation (6) of Art. 97 is the general solution of equation (1) of that article. Concerning the three constants of integration it is evident that E is the value of y when a; = 0. Also it is evident from equa- tion (5) of Art. 97 that D is the value of -^ when x = 0, and it is evident from equation (3) that C is the value of — g when X = 0. The general solution of a differential equation of the nth. order contains n undetermined constants of integration. When one or more of the constants of integration in the general solution of a differential equation have particular values assigned to them we have what is called a particular solution of the differential equation. For example, let C = 2, let D = and let E = in equation (6) of Art. 97, then the equation becomes ax^ y = j^ -\- x^ which is a particular solution of equation (1) of Art. 97. 99. Discussion of the first order linear differential equation with constant coefficients. — Consider the differential equation: y + Ag = o (1) * This statement is subject to some qualification because of what is called the singular solution. See Johnson's Ordinary and Partial Differential Equa- tions, page 43, John Wiley & Sons, New York, 1890. ORDINARY DIFFERENTIAL EQUATIONS. 173 The most obvious method for solving this differential equation is to "separate the variables" as explained in Art. 97. Thus equation (1) is easily reduced to: A^=-dx (2) y ^ ^ and this equation can be integrated by looking up appropriate forms in the table of integrals, giving: Alogy = -x-VC (3) It is desirable, however, to solve equation (1) by the following method because the method applies to a linear differential equa- tion of any order when the coefficients are constants. Let y = Ce*^ (4) where C and k are constants, and e is the Napierian base. Then: Substituting the values of y and ~- from (4) and (5) in equation (1), we have: Ce^^ + AkCe^' = (6) whence by cancellation we have: 1 + AA; = or t = — A k = -1 (7) Therefore if h in equation (4) has the value — -j then equation (4) satisfies equation (1); that is, equation (4) is a solution of (1) if k = — -T-', indeed equation (4) is the general solution of 174 CALCULUS. (1) because it contains one undetermined constant C, the constant of integration. 100. The principle of superposition. — A principle of extremely wide application in physics is the so-called principle of super-posi- tion. From the physical point window No, 1 window No. 2 of view a general statement of this principle is scarcely possi- ble and therefore the following examples must suffice: (a) A person at A (Fig. 81) can see window No. 1 and another per- son at B can see window No. 2 Fig. 8L at the same time. This means that two beams of light a and h can travel through the same region at the same time and not get tangled up together as it were, each beam behaving as if it were traveling through the region alone. (6) Two sounds can travel through the same body of air simultaneously, each sound behaving as if it were traveling through the body of air alone, (c) Two systems of water waves can travel over the same part of a pond simultaneously, each system behaving as if the other were not present, {d) Two messages* can travel over a tele- graph wire at the same time and not get mixed up. (e) Two forces F and G exerted simultaneously upon an elastic structure produce an effect which is the sum of the effects which would be produced by the forces separately, provided the sum of the forces does not exceed the elastic limit of the structure; therefore each force may be thought of as producing the same effect that it would produce if it were acting alone All of the effects in physics which are superposable — and this * Indeed any number of messages can travel over a telegraph wire in either direction or in both directions simultaneously. The only limiting feature in multiplex telegraphy, when line-loss is negligible, is in the design of the sending and receiving apparatus; and the same is true in wireless teleg- raphy. In each of the above examples the word two means iwo or more. ORDINARY DIFFERENTIAL EQUATIONS. 175 includes the greater part of the effects in mechanics, heat, electricity and magnetism, light and sound, and a great many- effects in chemistry — are expressible in terms of linear differential equations with constant coefficients, and the principle of super- position may be thought of as a property of such a differential equation as follows . If y is a function of x which satisfies a linear differential equation with constant coefficients, and if z is another function of x which satisfies the same differential equation, then (y + z) is a function of x which satisfies the equation* Proof. — Let the given linear differential equation be: , A du . „ d?u . f. /^v If a function y satisfies this equation, then: If another function z satisfies equation (1), then: Now ^^y + ^'> = ^ + ^ and ' ^^^ + ^> _ ^ . ^ There- ^""^ dx dx^ dx *""* dx^ dx' + dx»- "^"^^ fore, adding equations (2) and (3), we get: (, + .)+A'?(^ + B^J-i)+...=0 (4) But this is the same form as equation (1) which shows that the function {y + z) satisfies (1). * This proposition is true for a partial linear differential equation also. Indeed most of the superposable effects in physics are expressible in terms of partial linear differential equations with constant coefl&cients. Examples are given in Chapters VIII and IX. 176 CALCULUS. 101. Discussion of the second order linear differential equa- tion with constant coefficients. — Consider the differential equa- tion: It is not possible to "separate the variables" in this differential equation and therefore the second method of Art. 99 must be used. Therefore let: y = Ce** (2) then | = iCe*. (3) and Substituting these values of y, -^ and -r-^ in equation (1) and cancelling the common factor Ce**, we have : 1 4- A/b + M2 = f\ ^ (5) whence Therefore using a for one of these values of k and using j8 for the other value of A;, we get two solutions of (1), namely: w = Ce'^ (7) and z = De^' (8) But according to Art. 100 the sum (w + z) is also a solution. Therefore using y for (w -{- z), we have as a solution of (1): 2/ = C6- + i)e^- iW^-v-ii--^ (9) and this is the general solution of (1) because n contains the two undetermined constants C and D. ORDINARY DIFFERENTIAL EQUATIONS. 177 102. The starting of a boat. — At a certain instant (t = 0) a constant force E begins to act on a boat, and it is desired to find an expression for the increasing velocity i of the boat. At very low speeds the backward drag or friction of the water on a boat is proportional to the velocity of the boat. Let us assume that this proportional relation is exact, then the frictional drag of the water on a boat is equal to Ri, where i is the velocity of the boat and R is sl constant which depends on the shape and size of the boat. Therefore the net accelerating force acting on the boat is E — Ri. This force is equal* to the product of di the mass L of the boat and the acceleration -r;. Therefore we at have: L%^E-Ri (1) To get this equation into the standard form of a linear differential equation, let y = E-Ri (2) Then ^ = _ p^ dt ^dt and equation (1) becomes: R dt ^ or ^ + 1-1 = (^) The general solution of this equation is given in Art. 99, but it is worth while to work it out anew, as follows: Let y = Ce^« (4) * If force is expressed in poundals (or dynes), mass in pounds (or grams), and acceleration in feet per second per second (or centimeters per second per second). 13 178 CALCULUS, then dt ^^ = A;Ce« (5) dv Substituting these values of y and -^ in equation (3) and cancelling out the factor Ce*', we have: 80 that " L Therefore equation (4) becomes: y = Ce"^"' or, substituting E — Ri for yj we have: (6) E- Ri = Ce ^ (7) Now i — Q when t = 0. Therefore placing this pair of values in equation (7) we have: E ^C (8) so that the constant of integration is determined, and equation (7) becomes: E - Ri ^ Ee ^ or R r" |-|e-^- (9) This equation expresses the value at each instant of the increasing velocity of the boat as a function of the elapsed time t. 103. The stopping of a boat. — A boat is moving at velocity I when the propelling force ceases to act, and it is required to find an expression for the decreasing velocity, i, of the boat as it ORDINARY DIFFERENTIAL EQUATIONS. 179 gradually comes to rest. In this case the only force acting on the boat is the retarding force Ri, and in this case the accelera- di tion T7 is negative so that: or ^ + §•1 = The general solution of this differential equation is: (1) (2) (3) But i = / when t = 0. Therefore C = /, and equation (3) becomes: i = le (4) Remark. — The notation used in this discussion of the starting and stopping of a boat is the standard notation used in electrical theory. This notation is used here and in the following articles 35 30 25 «20 ■ -5 p- . ^ -^ ' ^ >^ / . / grc R= wing '3 oh current ms 1 R / L = 0.04 110 hem polts V J_ 3 4 5 6 7 8 9 hundredths of a second Fig. 82. 180 CALCULUS. because the mechanical problems discussed in this and the following articles are strictly analogous to certain electrical problems which constitute the foundation of the theory of alternating currents. Thus the curve in Fig. 82 is a graphical representation of equation (9) of Art. 102; the ordinates of this curve represent the growing values of the current i in a circuit \- 35 K J^ ^ •^b \ dei R = :ayin -3 oh 0.04 367 g current img ^20 \ \ henry amperes \ \ lO \ \. S , • 2 3 mndr edth 5 ( 8 of < 3 I sec 1 ond 3 < ? Fig. 83. of resistance R and inductance L after a battery of electro- motive force E is connected to the circuit. The curve in Fig. 83 is a graphical representation of equation (4) of Art. 103; the ordinates of this curve represent the decaying values of the current i when an initial current of / amperes is left to die away in a circuit of resistance R and inductance L. PROBLEMS. Solve the following differential equations. Ans. y = Ce-' + De^'. -3-2-^^ = 0- ^- da? 4y. Ans. y = Ce^' + De-^', "- -^ ^^ ORDINARY DIFFERENTIAL EQUATIONS. 181 ^-4^ + 3)/ = 0. Ans. y = Ce' + D^'. differentiating again we have : S=«5^-l= '"•(•) ») Similarly we have: S4M=--'(^) ao) 192 CALCULUS, and Therefore, comparing equations (9) and (11), we have and it may be shown, as above, that equation (3) leads to this same differential equation. A partial differential equation may be recognized as one which contains more than one independent variable. Therefore it is unnecessary to use the symbol d instead of the symbol d. Here- after the symbol d will be used exclusively. General solution of eqvxition {12). — Equations (2) and (3) both satisfy equation (12), and therefore equations (2) and (3) are both solutions of (12). Therefore, according to the principle of super- position as explained in Art. 100, the sum of F{x — vt) and fix + vt) or y = F(x-vt)+f{x + vt) (13) is a solution of equation (12). Indeed this is the general solution of (12) because it contains two undetermined fimctions.* PROBLEMS. l.Findthevaluesof |,3,Jandgwhen y = A sin (x — vt); also when y = A sin (x -\- vt). Show that both of these functions satisfy equation (12) of Art. 107. * The general solution of an ordinary differential equation of the second order contains two undetermined constants, and the general solution of a partial differential equation of the second order contains two undetermined functions. DIFFERENTIAL EQUATION OF WAVE MOTION. 193 2. Show that A sin {x — yQ + A sin (x + vi) = 2A sin x cos vi^ and plot the curve y = 2A sin x cos vi for the following values of vtj namely, 0, ^, 2' ^» 'T^ -4 » -2^ -^ and ^tt. iVote. — The curves thus plotted show the successive configurations of a string which is vibrating in what is called a simple mode. See Franklin and MacNutt, Light and Sound, page 243, The Macmillan Co., New York, 1909 108. Equation of motion of a stretched string. — When a stretched string is in equilibrium it is of course straight. Let us choose this equilibrium position of the string as the a:-axis of reference as shown in Fig. 90, and let us set up the differential equation which expresses the mode of motion of the string while the string is vibrating or while a bend is traveling along the string as a wave. We will assume that each part of the string moves only up and down (at right angles to the a;-axis in Fig. 90) and we will assume that the string is perfectly flexible.* Under these conditions the a;-component of the tension of the string is always and everywhere equal to a constant T. Let the curve ccc, Fig. 90 be the configuration of the string at a given instant t, that is, ccc is what a photographer would call a snapshot of the moving string. The shape of the curve ccc * This means that the only thing that keeps the string straight is its tension. 14 194 CALCULUS. defines y as a. function of x, and the steepness of the curve at a dv point is the value of -^ at that point. Consider the very short portion ab of the string. The length of this portion of the string when the string lies along the rc-axis (in equilibrium) is dx, and the mass of the portion is m - dx pounds or grams as the case may be, where m is the mass per unit length of the string. An enlarged view of the very short portion of the string is shown in Fig. 91. The adjacent parts of the string pull on the portion ab in the direction of the string at a and at 6, and the forces R and R' are thus exerted on the portion ah. The a;-component of R is the force T towards the left, and the a;-component of R' is an equal force T towards the right. Therefore the downward force D on the end a of the portion ah \s D = T tan B, and the upward force U acting on the end h is U = T tan 6'. Therefore the net upward force exerted on ah is dF = T tan 6' - T tan (1) dv But tan 6 is equal to the value of -^ at a, and tan 0' is equal dv to the value of -p at h. Therefore the difference, tan 0'— tan 0, dv is the increase of -^ from a to h, and this increase is equal to cPv d^v -t4 • dx. This is evident when we consider that the value of -j^ dv means the rate of increase of -f- with respect to x. Therefore, d^v substituting j^ • dx for tan 6' — tan d in equation (1), we have: dF =T^^'dx (2) Now according to Newton's laws of motion the net upward force dF acting on the portion ah of the string is equal to the ^<^- DIFFERENTIAL EQUATION OF WAVE MOTION. 195 mass m ' dx oi the portion multiplied by the upward acceleration -z^ of the portion. The in equation (2), we have: -z^ of the portion. Therefore, substituting m -^ ' dx for dF d^ d^ ^ dt' ^ dx' or dp- m dx^ ^^ This differential equation is here derived for the case of wave motion on a string, but identically the same form of equation applies to sound waves in air, to electric waves, to waves of light (which indeed are electric waves), and to certain kinds of water waves.* Equation (3) is therefore very important. Its general solution as found in Art. 107 is: y = F{x-vt)+f(x + vt) (4) where _ V = a/- (5) The term F(x — vt) represents a wave (a bend of any shape) traveling to the right at velocity v^ and , the term f{x + vt) represents a wave (a bend of any shape) traveling to the left at velocity v. 109. Idea of wave motion established without the help of equation (3) of Art. 108. — It is evident from Art. 107 that travelj purely and simply, is about the only thing that is estabUshed by the solution of equation (3) of Art. 108. Therefore one might expect to obtain equations (4) and (5) without the help of equation (3) by introducing the idea of travel at the beginning. With this end in view let us consider a bend of any shape and let us imagine that this bend is traveling along the string to the right at any * We are here considering only the simple case in which there are no appreci- able energy losses as the wave travels along. 196 CALCULUS. velocity v. One could make a bend of definite shape travel along a stretched string by threading the string through a bent tube and moving the bent tube along, and this state of affairs would be entirely unchanged if we imagine the tube to he stationary and the stretched string to he drawn through it as indicated in Fig. 92. Fig. 92. It is assumed that the string slides through the tube without friction. At any point p the string pushes against the side d of the tube because of its tension, and the string pushes outwards against the side c of the tube because of centrifugal action. Let r be the radius of curvature of the tube at p. Then the particles of the string near p may be considered to travel along a circular path of radius r. Therefore, according to Art. 50, the radial* acceleration of a particle of the string at p is - . Consider r an infinitely short piece of the string whose length is ds and whose mass is m • ds. Then m ' ds X - is the radial force r which must act upon the short piece of string to produce the specified acceleration — . If the string has no tension, then all of this radial force is exerted by the side c of the tube. If the string is under tension T, then the tension produces a T radial force equal to - • ds on the portion ds of the string, according to Art. 51; and if the string is not moving this force is exerted against the side d of the tube. * In the direction of r and towards the center of the osculating circle. DIFFERENTIAL EQUATION OF WAVE MOTION. 197 If the string is moving at a velocity which satisfies the equation: or V' T m - = - r r (1) then the radial force due to the tension of the string is just sufficient to produce the necessary radial acceleration, and no force at all is exerted on the string by the guiding tube. Under these con- ditions the guiding tube might be removed and the bend would stand in a fixed position on the traveling string; or if the string were standing still the bend would travel along the string at velocity v. In this discussion the bend can be of any shape and it can travel at velocity v in either direction. 110. The vibration of a plucked string.* — A string is pulled to one side as shown in Fig. 93 and released. The string is thus Fig. 93. * The vibration of a string which is struck with a hammer as in a piano is easy to formulate. The vibration of a string which is set vibrating by a bow as in a violin is somewhat more difficult to formulate. A good discussion of this subject is given in Byerly's Fourier's Series and Spherical Harmonics, pages 30-145, Ginn and Co., Boston, 1893. A discussion of the motion of piano strings and of violin strings is given in Appendices V and VI of Helmholtz's Tonempfindungen. English trans- lation by Alexander J. Ellis is entitled Sensations of tone. Published by Longmans, Green and Co., 1885. 198 CALCULUS. set vibrating. The vibrating string has at each instant a definite shape, that is, it forms a definite curve. It is desired to find an equation expressing y in terms of x and elapsed time t, which equation will be at each instant the equation of the curve formed by the moving string. There are two conditions which must be satisfied by the ex- pression for ?/, namely: (a) y must be zero at all times when x = and when x = tt, and (6) at the instant t = the ex- pression for y must be the equation of the curve formed by the plucked string at the moment of release. Also, of course, the expression for y must satisfy the differential equation (3) of Art. 108, namely, (Py^T^ dt^ rndx" ^^ Now it can be shown (see problem 2 on page 193) that the following expressions satisfy equation (1) y = A sin nx sin nvt (2) 2/ = A sin nx cos nvt (3) y = A cos nx sin nvt (4) y = A cos nx cos nvt (5) where A and n have any values whatever. But only (2) and (3) give ?/ = when a; = 0, and n must he an integer to give 2/ = when x = w. But (2) cannot be used because it gives 2/ = everywhere when ^ = 0. Therefore our problem is to take 2/ = A sin nx cos nvt (3) which satisfies condition (a), and add a large number of such solu- tions together (using different values of A and n in each) to get an expression for y which satisfies condition (6), above. The possibility of doing this was discovered by Fourier* and is embodied in what is called Fourier's theorem. * Fourier's original discussion is very simple and interesting. See Fourier's Theory of Heal, translated by Alexander Freeman, Cambridge, 1878. DIFFERENTIAL EQUATION OF WAVE MOTION. 199 111. Fourier's theorem. — ^Let y be any given function of x. Consider a certain portion AB of the curve which represents y as shown in Fig. 94, and let the distance AB be 27r (this is AB equivalent to choosing -^r— as our unit of length). Then, ac- cording to Fourier we may express the function y within the region AB as follows: y = Ao -{- Ai sin X + Az sin 2x + ^3 sin 3x + + Bi cos X + B2 cos 2x + B3 cos Sx ::::) (1) where the coefficients Ao, Ai, A2, • • • and Bi, B2, B3, • • • are constants whose values are : 1 /»a;=2rr Ao = ^ J ydx (2) -j r*x=2ir An=' ~ \ y minx ' dx (3) '^ Jx=0 -t nx=Z7r — I y cosnx ' dx (4) '^ Jx=Q =277 B. I '" lx=Q A complete proof of Fourier's theorem would involve a proof that the right-hand member of equation (1) is a convergent \^^' series.* This, however, will here be taken for granted. The proof then reduces to a derivation of equations (2), (3) and (4). * This matter is discussed in Chapter III of Byerly's Fourier's Series and Spherical Harmonics, Ginn and Co., Boston, 1893. 200 CALCULUS. In this proof a consideration of average values is most important, and the student should therefore look over Art. 73 again. Another matter of import- ance is to understand clearly 'V^A^uin fix ^^ ^YiQ meaning of the equation y = sin nx (and y = cos nx) when n is an integer. Now y — sin nx is the equation of a sine curve. If n = 1 this sine curve makes a posi- tive "arch" between x = and a; = TT, and a negative "arch'* between x = t and x = 27r. In general the sine curve y = sin nx makes a positive arch between x = and x = -, and a negative arch between x = - n' ^ n and X = — as shown in Fig. 95. Therefore the sine curve n ^ y = sinnx makes n pairs of positive and negative arches between X = and x = 2t, The derivation of equations (2), (3) and (4) depends upon the following propositions, n and m being integers. Proposition I. — The average value of sin nx (or of cos nx) between X — and x = ^tt is zero. This is evident when we consider that sin nx (or cos nx) passes through exactly similar sets of positive and negative values which exactly offset each other between a; = and x = 27r. Proposition n. — The average value of sin^ nx (or of cos^ nx) between x = and x = 2x is equal to \. This may be shown -I /*Z=ilt by finding the value of ^ I m^nx-dx (or of the correspond- ing expression for cos^ nx) as explained in Art. 73. Proposition m. — The average value of sin nx sin mx between X = and x = ^ir is zero when n and m are different integers. DIFFERENTIAL EQUATION OF WAVE MOTION. 201 This may be shown by finding the value of the expression 'oZ I sin nx sin mx • dXy which may be easily done with the help of form 24 in the table of integrals. Proposition IV. — The average value of sin nx cos mx between X = and x = 2t is zero whether the integers n and m he the same or not. This may be shown with the help of form 25 in the table of integrals. Proposition V. — The average value of cos nx cos mx from x = to X = 2t is zero when m and n are different integers. This can be shown with the help of form 26 in the table of integrals. Derivation of equation (2). — Consider the average value of each member of equation (1) between x — and x = 27r. The average value of the first member is, by definition, equal to — I y-dx, and the average value of the second member is Ao according to proposition / above. Therefore ^ I y'dx = Ao which is equation (2). Derivation of equation (3). — Multiply both members of equa- tion (1) by sin nx and we have: y sin nx = Ao sin nx + Ai sin nx sin x -{- • • • -{-An sin^ nx + • • • + -Bi sin na; cos X + • • • . Consider the average value of each member of this equation between a; = and x = 27r. The average value of the first member is ;r- ( y sin nx-dx. The average value of every term ^T^ Jx=0 of the second member is zero according to the above propositions except the term An sin^ nx, and the average value of this term 202 CALCULUS. is }An according to proposition IL Therefore ^1 y sinnx ' dx = iAn which is equation (3) Derivation of equation (4). — ^Multiply both members of equa- tion (1) by cos nx, consider the average value of each member of the resulting equation between a; = and x = 27r, and equation (4) is obtained. Simplification of equations (3) and (4) due to symmetry of given curve. — Let ccc, Fig. 96a, be the curve which is to be expressed by equation (1). It is permissible to take the base of ccc as tt. Fig. 96a. Under these conditions there are two important methods for choosing the second half of the complete curve, namely, the portion between tt and 27r, as follows: Curves having sine terms only. — The second half of the complete curve may be chosen like ccc turned upside down and turned end for end, as shown by the dotted curve c'c' in Fig. 96a. In this case y = — y' as shown in Fig. 96a. Now sin nx has equal and opposite values at p and at p', therefore y sin nx • dx has the same value and sign at p and at p', and therefore ry %\Ti. nx • dx = 2 \ y ^\ sin nx • dx DIFFERENTIAL EQUATION OF WAVE MOTION. 203 Consequently equation (3) becomes: An = - I y sinnx ' dx (5) TT Jo Furthermore cos nx has the same value and the same sign at p and at p', therefore y cos nx • dx has the same value but has /»27r opposite signs at p and at p\ and therefore I y cos nx • dx is zero. Consequently all of the cosine terms drop out of equation (1). Furthermore it is evident from the symmetry of the com- plete curve cccc^c' in Fig. 96a that 1 y-dx = so that Jo Ao = 0. Therefore the complete curve in Fig. 96a is given by: y = Ai sin X + Az sin 2x + A3 sin 3x + • • • (6) Curves having cosine terms only. — The second half of the complete curve may be chosen like ccc turned end for end but not turned upside down as shown by the dotted curve c'c'c' in Fig. 966. y'tixia K x-axis ^ 0' 1 — X -ii ! K (x) ^ — . — ~-;r-^ ^ Fig. 966. Under these conditions it may be shown, by an argument some- what similar to the above, that equations (1) and (4), respectively, become y = Ao + Bi cos X -\- B2 cos 2x + B3 cos 3a; + • • • (7) and 2 f' Bn = ~ \ y COS nx ' dx (8) 204 CALCULUS. 112. The vibration of a plucked string completely formulated. — In Art. 110 the problem of the plucked string was reduced to the problem of adding together a number of terms like A sin nx cos nvt so as to get an expression for y which is the equation of the curve formed by the string when t = 0. This is evidently the same thing as expanding the given function y '\n o. series of sines as explained in Art. HI. That is, we must find the coefficients in the series: 2/ = ill sin a; + Ai sin 2x + ^la sin 3a; + • • • (1) and the general expression for the coefficient An is: An = - \ 2/ sin nx • dx (2) according to equation (5) of Art. Ill, the use of which means that we have assumed the length of our string as tt units and middle of string -string Fig. 97o. have completed the curve as indicated by the heavy dotted line in Fig. 97a. Now the equation of the curve formed by the string at the beginning {t = 0), as shown in Fig. 97a, is: y 2c from X = to ^ = -^ 2c y = 2c X from x = ^ to x = IT Z (3) (4) DIFFERENTIAL EQUATION OF WAVE MOTION. 205 Therefore the integral (2) must be evaluated in two parts, namely: An = - t — xsmnx ' ax + - I [2c x ] smnx-dx (5) which gives: . Scl . nx ir^w- 2 Therefore the equation of the curve formed by the string at the beginning is: y = -^ (sin x — ^ sin 3x + ^V sin 5a; — • • • ) (6) and the equation of the curve formed by the string at an instant t seconds after the beginning is: y = —^ (sin X cos vt — ^ sin Sx cos 3vt ■}- • • • ) (7) Description of the motion of a plucked string. — It would seem from the elaborate difficulties of the above problem [and perhaps >qTTTTTTTpr TTTTTlXTTrt\^ ^ "T^iXnTTTTTTTTTTpf Fig. 976. 206 CALCULUS. the complexity of the result as given in equation (6) would suggest] that the motion of a plucked string is very complicated. As a matter of fact, however, the motion of a plucked string is extremely simple and easy to describe. Thus Fig. 976 shows six successive snapshots of a string which has been plucked in the middle and released. The moving part of the string is straight and parallel to the equilibrium position of the string and the motion is indicated by the short arrows. PROBLEMS. The most important practical problem involving Fourier's theorem is, perhaps, the finding of the coefficients in equation (1) of Art. Ill when the function y is given not as an algebraic function of x but as an experimentally determined curve. A very good set of directions for making the necessary calculations is given in Bedell and Pierce's Direct and Alternating Current Manual, 2nd edition, pages 331-338, D. Van Nostrand, New York, 1913; and a discussion of the origin and proof of the method is given on pages 339-344. The harmonic analyzer is a machine for finding the coefficients in equation (1) of Art. Ill when the function y is given as an experimentally determined curve. The eariiest harmonic analyzer is that of Lord Kelvin. A brief description of this machine is given in Franklin's Electric Waves, pages 340- 342, The Macmillan Co., New York, 1909. See also section 37 of the article on Tides in the ninth edition of the Encyclopedia Britannica. See also articles by James Thomson and by Sir Wm. Thomson (Lord Kelvin) in Proceedings of the Royal Society, Vol. XXIV, 1876, page 262 and pages 269 and 271. The harmonic analyzer of G. U. Yule is described in an article by J. N. Le Conte, Physical Review, Vol. VII, pages 27-34, 1898. An especially interesting de- scription of a harmonic analyzer is given on pages 68-74 of A. A. Michelson's Light waves and their uses, University of Chicago Press, Chicago, 1903. 1. Determine the coefficients in Fourier's series to give the curve cc which is shown in Fig. p\. Fig. pi. DIFFERENTIAL EQUATION OF WAVE MOTION. 207 Ans: y = (sin x + J sin 2a; + J sin 3a; + • • • ) TT 2. Determine the coefficients in Fourier's series to give the curve cc which is shown in Fig. p2. Ans; y = -^ (sin a; + J sin 3a; + J sin 5a; + • • •) TT y-axis c T r 1 c \a c 1 1 * ! x-ojcw 1 1 1 c \a c ' i c h K -^ -J(— ^- — ^ Fig. p2. 3. Determine the coefficients in Fourier's series to give the curve cc which is shown in Fig. p3. Ans: y = —2 (cos a; — | cos 3a; + i cos 5a; — | cos 7a; + • • •) y-axis e c n r — f X c c x-curis 1 I 1 1 L J c 1 la c 1 i 7( *-^ , Fig. p3. 4. Determine the coefficients in Fourier's series to give the curve cc which is shown in Fig. p4. Ans: y = — (sin a; — J sin 3a; + -jV sin 5a; — :jV sin 7a; + • • •) 208 CALCULUS. 5. Determine the coefficients in Fourier's series to give the curve cc which is shown in Fig. p5. y-€ixi8 Sa Ads: y = -^ (cos a; + i cos 3a; + ^ cos 5a; + • • • ) y-axis ./ " ^^ »-axw ^ 1 \. ■■--.--'• ^^\i>^ 1 "^^-^^- < — n ^—-n i Fig. p5. 6. Determine the coefficients in Fourier's series to give the curve cc which is shown in Fig. p6. Ans: y = — (sin a; — J sin 2a; + J sin 3a; — i sin 4a; + • • • ) t DIFFERENTIAL EQUATION OF WAVE MOTION. 209 7. Derive the answer to problem 4 by integrating the answer to problem 3. Note. — Any Fourier series gives a convergent series when integrated term by term. 8. Plot the curve which represents the integral of the curve given in Fig. p2, the constant of integration being such as to make the integral curve pass through the origin; and determine the coefficients in a Fourier's series to give the integral curve. Note. — The constant of integration is zero in problem 7 and therefore easily overlooked. 9. Derive the answer to problem 3 from the answer to problem 2 by shifting the origin ^ to the right. 10. Differentiate the answer to problem 4 and interpret the result. Note- — A Fourier series gives a divergent series by differentiation unless the function ij is continuous. Thus y in Fig. p4 is continuous, whereas y in Figs, pi, p2, p3 and p6 is discontinuous. 15 CHAPTER IX VECTOR ANALYSIS. SCALAR AND VECTOR FIELDS. 113. Space analysis. — A system of space analysis, commonly called vector analysis, is developed in this chapter, and it is ex- tremely useful and important in every branch of physics where variations in space are involved. Thus the theory of electricity and magnetism (aside from the electron theory) is a simple appli- cation of vector analysis. Also the theory of heat flow, the theory of fluid motion, and a great part of the theory of elasticity and wave motion are applications of vector analysis. It is almost impossible for a student to make a beginning in any branch of theoretical physics without some understanding of vector analysis. Vector analysis as outlined in this chapter is a very different thing from the familiar use of vectors and of complex quantity in the theory of alternating currents. In the one case we have a comprehensive system of space analysis, and in the other case we have a narrow scheme for representing the solution of a linear partial differential equation, as explained in chapter VII. Indeed the kind of "vector analysis'' which is used in the theory of alternating currents cannot possibly be extended to three dimen- sions as a consistent system of space analysis. Vector analysis originated in Sir William Hamilton's theory of quaternions.* The theory of quaternions, however, contains much that is of doubtful value in theoretical physics. Indeed Maxwell in his great treatise on Electricity and Magnetism (Oxford, 1873) used only a few of Hamilton's ideas. The study of Max- * See Sir William Rowan Hamilton's Lectures on Quaternions, and Elements of Quaternions. These books are now out of print but they may be found in most good libraries. An Elementa ry Treatise on OuMernions by P. G. Tait, second edition, Oxford, 1873, is the standard treatise on the subject. 210 VECTOR ANALYSIS. 211 well's treatise has kept the subject of space analysis before the serious student of physics for more than a generation, and two significant attempts have been made to formulate the more useful of Hamilton's ideas in what is now called vector analysis. The first of these was that of Willard Gibbs, whose view of vector analysis was outlined in a very condensed form in a pamphlet printed (for private circulation, only) in New Haven in 1883.* The second attempt to formulate the more useful of Hamilton's ideas was that of Oliver Heaviside.f The first attempt to place vector analysis before the student of physics in simple form was that of E. L. Nichols and W. S. Frank- lin.J The best discussion of vector analysis for the student is that of Abraham and Foppl in their Theorie der Electridtdt, Vol. I, pages 1-125, Leipsig, 1907. 114. Scalar and vector quantities. — A scalar quantity is a quantity which has magnitude only. Thus every one recog- nizes at once that to specify 10 cubic yards of sand, 25 pounds of sugar, 5 hours of time is in each case to make a complete speci- fication. Such quantities as volume, mass, time and energy are scalar quantities. A vector quantity is a quantity which has both magnitude and direction, and to completely specify a vector quantity one must give both its magnitude and its direction. This necessity of specifying both the magnitude and the direction of a vector is especially evident when one is concerned with the relationship of two or more vectors. Thus if a man travels a stretch of 10 miles and then a stretch of 5 miles more, he is by no means necessarily 15 miles from home. If T)ne man pulls on a post with a force of * Professor Gibbs' point of view is set forth in Vector Analysis by E. B. "V^ilson, New Haven, Yale University Press, 1901. t See Heaviside 's Electromagneti c Theory ^ Vol. I, pages 132-305, The Electrician Publishing Co., London, 1893. This discussion of Heaviside'a is unusually interesting, but we cannot agree with Heaviside in his statement that vector analysis is independent of the quaternion. X See Nichols and F rankUn's Elements of Physics, Vol. II, first edition, The Macmillan Co., New York, 1895. 212 CALCULUS. 200 units and another man with a force of 100 units, the total force acting on the post is by no means necessarily equal to 300 units. A New Yorker traveling steadily at a speed of 60 miles an hour would by no means necessarily reach Boston in 5 hours, because he might be traveling in some other direction. Many cases arise in physics where it is necessary to consider the single force which is equivalent to the combined action of several given forces; where it is necessary to consider the actual velocity of a body due to the combined action of several causes, each of which alone would produce a certain amount of velocity in a given direction; and so on. The single force or single velocity is in each case called the vector sum or the resultant of the given forces or given velocities. The addition of vector quantities is exemplified by the addition of forces as follows: Two given forces are represented by the lines a and 6, in Fig. 98, and the sum or resultant of the forces is represented by the diagonal r of the parallelogram constructed on a and h as sides. It is evident that the geometrical relation between a, b and r is completely shown by the triangle in Fig. 99, in which the line which represents force h is drawn from the extremity of the line which represents force a, and r is the closing line of the triangle. The geometrical construction of Fig. 99 gives a method of adding any number of forces with the least amount of compli- cation, as shown in Fig. 100. Thus to add four given forces a, h, c and d: draw a Hne representing force a; from the end of this line draw a line representing force b; from the end of this VECTOR ANALYSIS. 213 line draw a line representing force c; and so on. Then the closing line r of the polygon abed represents the sum of the forces in magnitude and in direction. To represent a vector in an alge- braic discussion it is usually most convenient to represent the vec- tor in terms of its components in the directions of three chosen rectangular axes of reference. Thus if X, F, Z are the compo- nents of a given force F (or any p. .^ vector whatever) then the force may be represented as F = X+Y + Z. In this expression vector addition is of course understood because X, Y and Z are vectors. ■ 115. Vector products. Case I. Parallel vectors. — The product of two parallel vectors is a scalar. This fact is exemphfied in every branch of physics. Thus to multiply a force by the distance a body has moved in the direction of the force gives the work done by the force, and work is a scalar quantity. To multiply a force by the velocity with which a body moves in the direction of the force gives the power developed by the force, and power is a scalar quantity. The square of the velocity of a body determines its kinetic energy, and energy is a scalar quantity. A plane area is a vector quantity and its vector direction is the direction of the normal to the area. To multiply the area of the base of a prism by the altitude of a prism gives the volume of the prism, and volume is a scalar quantity. The quotient of two parallel vectors is a scalar quantity. Thus let F be the force exerted by a fluid on a flat surface of area a. F Then the quotient - is the hydrostatic pressure of the fluid, and a hydrostatic pressure is a scalar quantity. 214 CALCULUS. Case II. Orthogonal vectors. — Vectors which are at right angles to each other are said to be orthogonal. The product of two orthogonal vectors is a third vector at right angles to both of the given vectors. Thus to multiply the length by the breadth of a rectangle gives the area of the rectangle, and area is a vector as above explained. To multiply a force F by the perpendicular distance I from the Une of action of the force to a chosen axis gives the torque action IF of the force about the axis, and torque action is a vector quantity. The quotient of two orthogonal vectors is a third vector at right angles to both of the given vectors. Thus to divide the area of a rectangle by the length of one side gives the length of the other side. Case III. Oblique vectors. — The product of two oblique vectors consists of two parts; one part is a scalar and the other part is a vector. This proposition is a necessary result of the above statements concerning the product of parallel vectors and the product of orthogonal vectors. Thus Fig. 101 shows two obUque vectors U and V. The vector V can be resolved into the ">^ components 7' and V" parallel to U vy\ j and perpendicular to U respectively, y^ ^\ 1 ^ and then 7' + V" can be substi- v ^ ^^ ^ ^1 tuted for V in the product UV, giv- — ^— ' ing UV + VV". But UV is the pj jQj product of two parallel vectors and it is a scalar, and UV" is the product of two orthogonal vectors and it is a vector. From Fig. 101 it is evident that V = V cos 6 and that 7" = 7 sin 6. Therefore we have the following important propositions: (a) The scalar part of the product of two oblique vectors is equal to the product of the numerical values of the respective vectors and the cosine of the angle between them. (6) The numerical value of the vector part of the product of two oblique vectors is equal to the product of the numerical values VECTOR ANALYSIS. 215 of the respective vectors and the sine of the angle between them; and of course the direction of the vector part of the product is at right angles to the plane of the given oblique vectors. There are but few cases in physics where the scalar part and the vector part of the product of two oblique vectors are both important, although there are many cases where the scalar part of a vector product is important, and many other cases where the vector part of a vector product is important. Thus a force F acts on a car as shown in Fig. 102. Imagine the car to move a distance I in Fig. 102. the direction of the track, then the work done by the force is the scalar part of the product of the two obhque vectors F and l^ or: the work done is equal to numerical value of I X numerical value of F X cos 6. The vector part of the product IF in this case has no very important meaning. A force F acts on a crank as shown in Fig. 103. The torque action of the force about the axis of the crankshaft is the vector part of the product of the two oblique vectors F and I; the numerical value of the torque is equal to numerical value of I X numerical value of F X sin 6, and the direction of the torque, considered as a vector, is parallel to the axis of the shaft. Exact direction of the vector part of a vector product. — ^The product UV in Fig. 101 is a vector at right angles to the plane of the paper, but is it towards the reader or away from the reader? Consistency* requires us to admit that if the product UV" is * This is shown by the following discussion of Fig. 104 and by the inter- pretation of equation (1). Sir William R. Hamilton gives a very full discus- sion of this in his Lectures and in his Elements of Quaternions. 216 CALCULUS. towards the reader in Fig. 101, then V'U must be away from the reader. That is, we must admit that UV" = — V"U where U and V" are orthogonal vectors. This matter is exemphfied as follows: A force F acts upon a crank-arm I as shown in Fig. 104. The vector part of the product IF is the torque action of the force about the axis (perpendicular to the plane of the paper). Now the torque action of F in Fig. 104 can be ex- pressed in terms of the components of F, namely, X and F, and the com- ponents of I, namely, x and y. Thus xF is a counter-clockwise torque about 0, and yX is a clockwise torque about in Fig. 104. as positive, we have*. Therefore, taking counter-clockwise torques Total torque action of F about in Fig. 104, or the vector part of the product IF I =a;y- yx (1) The vector direction of a torque may be thought of as the direc- tion (along the axis of the torque) in which a right-handed screw would travel if turned by the torque. Adopting this convention and choosing the positive direction of the 2-axis of reference towards the reader in Fig. 104 we have from equation (1) : (a) xY is an x-vector multiplied by a ^/-vector, and it is a vector in the positive direction of the 2-axis. (6) yX is a ?/-vector multiplied by an a;-vector, and it is a vector in the negative direction of the 2-axis. Expressing F and I in terms of their components as shown in Fig. 104, we have: F = X + y (2) and l = x-\-y (3) Multiplying equations (2) and (3) member by member, using the VECTOR ANALYSIS. 217 ordinary rules of algebra, we get : IF = xX -\- yY -{- xY -{- yX (4) But it is shown above that xY and yX are opposite in sign. Now the opposite signs oi xY and yX are shown in equation (4), according to the above discussion, hy the fact that in one case we have an x-vector multiplied by a y-vector and in the other case we have a y-vector multiplied hy an x-vector. But it is very inconvenient to have algebraic signs so indirectly indicated; it is better to indicate the sign explicitly and take xY — yX as the vector part of the product IF. When this is done we pay no attention to the order of the factors, because the inverse order in the two terms is taken account of in the negative sign of the second term. With this understanding, therefore, we have the following important propo- sitions: (a) The scalar part of the product IF in Fig. 104 is equal to xX -h yY. This is evident when we consider that xX and yY are the only scalar terms in equation (4). (&) The vector part of the product of IF in Fig. 104 is equal to xY — yX, and it is in the direction towards the reader in Fig. 104. 116. Scalar fields. — It is sometimes important to consider the temperature at various points in a body, the pressure at various points in a fluid, the density at various points of a substance or the electric charge per unit volume at various points in a region. Such a distribution of temperature, hydrostatic pressure, or density is called a scalar field because the quantity under consideration is a scalar and it has a definite value at each point in a region or field of space. The distribution is said to be homogeneous or uniform when the scalar quantity has the same value at every point in the field; otherwise the distribution is said to be non- homogeneous. An example of a non-homogeneous scalar field is the temperature of an iron rod one end of which is red hot and the other end of which is cold. Atmospheric pressure is also non- homogeneously distributed because it is different at different places and different at different altitudes. 218 CALCULUS. A scalar field is sometimes called a distrihvied scalar. Thus when the scalar nature of temperature is to be emphasized it is helpful to call temperature (throughout a hot iron rod, for example) a distributed scalar. A scalar field is sometimes called a scalar function, A distri- buted scalar has a definite value at each point of space; that is, the value of a distributed scalar at a point is a function of the coordinates of the point. When this fact is to be emphasized it is helpful to call temperature (or any distributed scalar) a scalar function. 117. Vector fields. — It is sometimes important to consider the velocity at different points of a fluid, the direction and intensity of heat flow at different points of a substance, or the direction and intensity of an electric or magnetic field at different points in space. Such a distribution of fluid velocity, or other vector, is called a vector field because the quantity imder consideration is a vector, and it has a definite value and direction at each point in a region or field of space. The distribution is said to be homo- geneoiLs or uniform when the vector has the same value and is in the same direction at every point in the field; otherwise the distri- bution is said to be non-homogeneous. The water in a rotating bowl is an example of a non-homogeneous vector field because the water moves in different directions and at different velocities at different points in the bowl. The magnetic field around an electric wire is a non-homogeneous vector field because the mag- netic field does not have the same intensity and the same direction everywhere. A vector field is sometimes called a distributed vector. Thus when the vector nature of fluid velocity is to be emphasized it is helpful to call fluid velocity a distributed vector. A vector field is sometimes called a vector function. Each component (the a;-component, the ^/-component and the 2-com- ponent) of the velocity of a fluid has a definite value at each point of space; that is, the values of the components at a point p are VECTOR ANALYSIS. 219 functions of the coordinates of p. When this fact is to be empha- sized it is helpful to call fluid velocity (or any distributed vector) a vector function. 118. Volume integral of a distributed scalar. — For the sake of simplicity let us consider a special case, namely, the density of a substance, and let us assume that the density varies from point to point. Let ^ be the density at a given point, then \l/ ' dr is the mass of material in the volume element dr at the point, and the total mass M of the body is: M = frP ' dr (1) This integral is called the volume integral of the distributed scalar ^. The physical significance of volume integral is not in every case so simple as in the case of density. If \l/ is the volume density of electric ch?irge, then equation (1) gives the total electric charge in the region throughout which the integration is extended. If ^ is the energy density in an electric or magnetic field, or in a strained solid, or in a moving fluid, then equation (1) gives the total energy in the region throughout which the integration is extended. 119. Gradient of a distributed scalar. — Any distributed scalar like temperature, or density, or hydrostatic pressure has a definite value at each point of a field and therefore the value of any distri- buted scalar at a point may be thought of as a function of the coordinates x, y and z of the point. Indeed, this matter has already been discussed in Arts. 62 to 66, where it is explained that a distributed scalar has a definite gradient at each point. Thus if ^ is the temperature at a point in a body, then (1) (2) (3) X ■ dx Y dyp dy Z drP dz 220 CALCULUS. where X, Y and Z are the component gradients of \f/. The gradient of a distributed scalar has a definite value and a definite direction at each point and it is, therefore a distributed vector. PROBLEMS. 1. A vessel one meter wide, one meter long, and one meter deep, contains a fluid of which the density is one gram per cubic centimeter at the top, increasing uniformly to two grams per cubic centimeter at the bottom. Find the volume integral of the density of the fluid. Ans. 1,500,000 grams. 2 . The gradient of the density in problem 1 is uniform through- out the vessel and equal to one gram per cubic centimeter per meter, and it is directed vertically downwards. Suppose the downward gradient of the density to be a linear function of the distance from the top of the vessel, changing from zero at the top to two grams per cubic centimeter per meter at the bottom. Find the volume integral of the density of the fluid, the density being one gram per cubic centimeter at the top. Ans. 1,333,333 grams. 3. The value of a distributed scalar at any point p is \p = — where Q is a constant and r the distance of p from the origin of coordinates. Find the a:, y and z components of the gradient Qx of \l/. Ans. The x-component is — -r-r-; — t"; — 5Ti' {x^ + 2/2 H- z^)* 4. Find the gradient of ^ ( = ~ ) in the direction of r. Ans. 5. The a;-component of the gradient of any distributed scalar may be thought of as a distributed scalar. Find the ^/-component of the gradient of the a;-component of the gradient of -• Ans. SQxy (3^-\-y^ + 2^)1 VECTOR ANALYSIS. 221 120. Permanent and varying states of scalar distribution. — When the temperature at each point of a body remains unchanged, the distribution of temperature throughout the body is said to be permanent, although, of course, the temperature of the body may not be everywhere the same. When the temperature at each point of a body is changing we have what is called a varying state of distribution. Thus the density in a gas which is being com- pressed, and the temperature throughout a body which is being heated or cooled are examples of varying scalar distributions. 121. Stream lines of a distributed vector. — A line drawn through a moving fluid so as to be at each point in the direction in which the fluid at the point is moving is called a stream line. The geometrical idea of a stream line applies to any vector field what- ever, and the manner of distribution of a vector is clearly repre- sented by the use in imagination of such lines. In electric and magnetic fields these lines are called lines of force, but the term stream line will be used in general statements. 122. Permanent and varying states of vector distribution. — When the velocity of a moving fluid remains unchanged in magni- tude and direction at every point, we have what is called a perma- nent state of fluid motion. When the velocity of a fluid is changing at each point, we have what is called a varying state of fluid motion. Thus, when an orifice in a large tank of water is suddenly opened, a perceptible time elapses before the jet of water becomes estab- lished. During this time the velocity of the water is changing rapidly at each point in the jet. After the jet becomes steady, however, the velocity of the water at each point remains constant in magnitude and in direction. The magnetic field in the neigh- borhood of a moving magnet or in the neighborhood of a moving or changing electric current is an example of a varying vector distribution. Rate of change of a distributed vector at a point. — Let the line a, Fig. 105, represent the value at a given instant, of the velocity of a fluid at the point p, and let the line a + Aa represent the 222 CALCULUS. velocity of the fluid at the same point after a time-interval A^ has elapsed. The Hmiting value oi -r- as At approaches zero is the rate of change of a at the point p.* This rate of change of a has a definite value and is in a definite direction at each point of space and it is therefore a distributed vector. 123. Line integral of a distributed vector. — Consider a line or path pp' in a vector field as shown in Fig. 106. Let As be an element of this line or path, let R be the value of the distributed ^direction o/j^atp N\ ""\^. ^^^ >P' Fig. 105. Fig. 106. vector at the element As, and let e be the angle between U and As. Then R cos € is the resolved part of R parallel to As, and R cos € • As is the scalar part of the product of R and As; and: E = f Rcose- ds (1) is called the line integral of the distributed vector R along the line or path over which this summation is extended. The angle e is reckoned between R and the positive direction of As, the positive * In this illustration the velocity under consideration is the changing velocity of the successive particles of the fluid as they pass the point p, not the changing velocity of a given particle while it is traveling along near p. The latter velocity may be changing from instant to instant even though the former is invariable. VECTOR ANALYSIS. 223 direction of As being the direction in which As would be passed over in traveling along the line pp' in a chosen direction. If the chosen direction be changed, cos e will change sign at each ele- ment. Therefore the line integral from p to p' is equal to the line integral from p' to p in Fig. 106, but opposite in sign. Examples. — The line integral of electric field along a path is called the electromotive force along the path. The Une integral of a magnetic field along a path is called the magnetomotive force along the path. The line integral of fluid velocity along a path is called the circulation of the fluid along the path. Cartesian expression for line integral. — Let X, Y and Z be the components of the vector R, and let dx, dy and dz be the components of the line element ds. Then we have: R = X +Y + Z (a vector equation) (2) and ds = dx + dy + dz (a vector equation) (3) The product of the two vectors R and ds is part scalar and part vector as explained in Arts. 114 and 115, and the scalar part of the product is R cos e - ds or X - dx + Y - dy + Z • dz. Therefore equation (1) may be written: E = J{X 'dx+Y 'dyi-Z -dz) (4) Line integral of the gradient of a distributed scalar.— Consider a distributed scalar rf/. Let us call it temperature for the sake of inteUigibility. Let R in Fig. 106 be the gradient of \p, that is R is the temperature gradient. Then the line integral of R along the path pp' is equal to \J/' — ^, where ^' is the tempera- ture at p' and \f/ is the temperature at p. This is evident from the following considerations. The product R cos € in Fig. 106 is the resolved part of the temperature gradient in the direction of the line element As, and i^ cos € • As is the change of tempera- ture along As. Therefore 2/2 cos e • As is the sum of the changes of temperature along all parts of the path pp' or the total change of temperature from p to p'. 224 CALCULUS. Line integral of a gradient along co-terminous paths. — The line integral of R along any path pp' is the temperature difference between p and p'. Therefore the line integral of R is the same for all paths from p to p\ There is an important exception to this proposition as follows: Imagine an ordinary auger to be placed with its axis at right angles to the plane of the paper in Fig. 107, the plane of the Fig. 107. Slope-lines on an auger-hill. Fig. 108. Magnetic lines of force around an electric wire. paper being the base plane and the winding-stair-like surface of the auger being looked upon as the surface of a hill raised above the base plane. This hill we will call an auger-hill, and its slope lines are shown by the fine-line circles in Fig. 107. The height of this auger-hill above any point p has many values differing from each other by, say, one inch, where one inch is the ** pitch" of the auger. Now the line integral of the slope of a hill along any path is the difference of level of the ends of the path. A path which returns to its starting point (a closed path) comes back to its initial level on an ordinary hill, but a path from p back to p in Fig. 107 does not come back to its initial level if the path circles round the axis of the auger. Therefore the line integral of slope along the path 1 from p to p' in Fig. 107 is not the same as the line integral of slope along path 2 from p to p'. VECTOR ANALYSIS. 225 The fine-line circles in Fig. 108 are the magnetic lines of force (slope lines of the magnetic-potential hill) in the neighborhood of an electric wire perpendicular to the plane of the paper. The line integral of the magnetic-potential slope along path 1 from p to p' in Fig. 108 is not the same as the line integral of the magnetic-potential slope along path 2 from p to p'. 124. Potential of a vector field. — Consider any distributed vec- tor R, Then the '^height" at each point of space of an imagined "hill'' whose slope or gradient is everywhere equal to R is called the potential of R. Thus the temperature at each point of a body is the "height" at that point of a hill whose slope is every- where equal to the temperature gradient, therefore temperature is the potential of temperature gradient. The "height" in volts at each point of space of an imagined "hill" whose slope is every- where equal to the electric field intensity (in volts per centimeter) is called electric potential. The "height" at each point of space of an imagined "hill" whose slope is everywhere equal to the velocity of a moving fluid is called the velocity potential of the fluid. The temperature at a point of a body can of course be deter- mined by placing a thermometer at that point. That is, tempera- ture is an actual physical condition. The electric potential at a point, however, is not an actual measurable physical condition at that point; indeed electric potential exists only in the imagi- nation as a sort of mathematical fiction whose usefulness grows out of the fact that it is often very helpful to think of a given vector field as a gradient. An example showing the usefulness of the idea of potential is given in Art. 133. Theorem as to the existence of potential. — Let ^ be a distri- buted scalar, like temperature. Then the component gradients of yp are given by equation (1), (2) and (3) of Art. 119. Differ- entiating the first of these equations with respect to y and the second with respect to x we have: d^yp dX dy ' dx dy 16 226 CALCULUS, and ^ dY (2^ dx ' dy dx But J , and , , are always identical as stated in dy ' dx dx ' dy Art. 59 and as demonstrated in Art. 134. Therefore from equa- tions (1) and (2) we have: ^ ^dY dy dx Similarly from equations (2) and (3) of Art. 119, we obtain: f-^ = (4) dz dy and frona equations (1) and (3), we obtain: dZ dX , V In these equations (3), (4) and (5) X, Y and Z are the com- ponent gradients of any scalar function \J/, and any distributed vector whose components satisfy equations (3), (4) and (5) can have a potential, or, in other words, any distributed vector whose com- ponents satisfy (3), (4) and (5) can be looked upon as the gradient of an imagined "hilV Multivalued potential. — The auger-hill which is represented in Fig. 107 has a multiplicity of heights above any point p. Similarly the magnetic-potential in Fig. 108 has a multiplicity of values at any point p. Of course the "height" of the potential "hill" in Fig. 108 may be thought of as an actual height measured upward from the plane of the paper; but one must not forget that the magnetic field under consideration fills all space, and that there is a multiplicity of values of magnetic potential at each point in space. See Art. 66. VECTOR ANALYSIS. 227 125. Surface integral of a distributed vector. Flux. — Let AOOB, Fig. 109, be a diaphragm stretching across a closed loop of wire, the dots A and B being where the wire passes through the plane of the paper. Let AS be the area of an ele- diaphram Fig. 109. Fig. 110. ment of the diaphragm, let R be the value at AS of a distributed vector, and let e be the angle between R and the normal to AS this normal being always drawn outward from the same side 00 of the diaphragm. Then R cos e is the resolved part of R normal to A>S, and R cos e ' AS is the scalar part of R - AS; and: ^ = f Rcose ' dS (1) is called the surface integral of R over the portion of the diaphragm over which the integration is extended. ^ If the normal to AS in Fig. 109 is reversed, as shown in Fig 110, then cos e will be reversed in sign, and the surface integral, retaining its numerical value, will be reversed in sign. In the integration over a closed surface like a box or sphere, the normal is understood, throughout the following discussion, to be drawn outwards. Examples. — If R in Fig. 109 is the velocity of a fluid at AS, then R cos e is the resolved part of the velocity normal to AS, 228 CALCULUS. R cos c • A*S is the volume of fluid per second flowing across AS, and JR cos e • dS is the total volume of fluid per second flowing through the loop of wire AB. The volume of fluid per second passing through a loop is called the fliix of the fluid through the loop, and the surface integral of any distributed vector over a surface is called the fiux of the vector across the surface. Thus magnetic flux is the surface integral of magnetic field and electric flux is the surface integral of electric fleld. The surface integral of a distributed vector over a closed surface like a box or sphere is called the flux into or oiU of the region enclosed by the box or sphere. Cartesian expression for surface integral. — Let X, Y and Z be the componenfs of the vector R in Fig. 109, and let da, db and dc be the areas of the projections of dS on the yz, on the xz, and on the xy planes, respectively. Then: R = X + Y + Z (a vector equation) (2) and dS = da -{- dh + dc (a vector equation) (3) The scalar part of the product R - dS is X ' da + Y ' dh -\- Z ' dc i= R cos € ' dS), But the shapes of the surface elements da, db and dc may be anything whatever. Therefore we may write dy • dz for da, dx • dz for dh, and dx • dy for dc. Hence equation (1) becomes: ^ = ff (X ' dy ' dz -{- Y ' dx ' dz + Z ' dx ' dy) (4) 126. Divergence of a distributed vector. — In some cases a dis- tributed vector "flows" outwards or emanates from a region of space. Thus the liquid in a tank flows outwards from the end of a supply pipe. When a gas is expanding each portion of the gas is growing less dense and there is an outward flow from every small part of the region occupied by the expanding gas. What is called electric field emanates from electrically charged bodies, VECTOR ANALYSIS. 229 and when electric charge is spread throughout a region, electric field emanates from every small part of the region occupied by the charge. Consider a small volume element Ar in the neighborhood of a point p in a vector field. Let A$ be the flux of the distributed vector R out of the volume element Ar. It can be shown,* when i2 is a continuous function of the coordinates x, y and 2;, A$ that the ratio 7- approaches a definite limiting value as the At volume element Ar grows smaller and smaller. This limiting value, -J- is called the divergence of the vector field at the point p. Therefore, representing the divergence of a vector field at a point by p, we have: d^ = P' dr (1) in which d^ is the flux of R coming out of the volume element dr. When a vector field flows into each small part of a region, the divergence is negative. Negative divergence is sometimes called convergence. The flux $ of a vector field across a surface is a scalar quantity, as is evident from the discussion of equation (1) of Art. 125. (A\ — 1 is a scalar quantity; and since a vector field has a definite diver- gence at each point (of course the divergence may be zero) of space, it is evident that the divergence of a vector function (a distributed vector) is a scalar function (a distributed scalar). Cartesian expression for divergence. — Consider a small cube of which the edges are dx, dy and dz as shown in Fig. 111. Let Xj Y and Z be the components of the given distributed vector R A€» * The actual proof that — has a definite limiting value may be estab- lished without great difficulty by considering the portion v" of the linear vector field in Art. 133. 230 CALCULUS. at the point p. Then the flux of R into the cube across the face a is X * dy ' dzj because X is the component of R normal to the face a and dy • dz is the area of face a. If X is the ic-component of R at any point of the face a, then f X + -7- * dx\ z-axis ^y^r- p^ JL. ^^#^ 1 1 \' 1 X-axis yy-axi B X 1 / dx Fig. 111. is the a;-component of R at the corresponding point of the face 6, and if X - dy - dz is the flux into the cube across face a then X + ^ ' dx) ' dy ' dz is the flux out of the cube across face &.* Therefore the flux of R out of the cube across face h exceeds the jy flux of R m^o the cube across face a by the amount -7- -dx'dy'dz. In the same manner it may be shown that the net flux of R out of the cube across the two faces which are perpendicular to the y-axis is -T~ ' dx • dy ' dz; and that the net flux oui of the cube across the two faces which are perpendicular to the 2-axis is -r 'dx-dy- dz. Therefore the total flux out of the cube is: ,^ /dX,dY^dZ\ , , , ^^=Kdx-^dy^-dz)' <^^'dy'dz (2) *The argument here given is not entirely rigorous, and the peculiar wording of this sentence is intended to suggest a form of argument which is rigorous. VECTOR ANALYSIS. 231 But dx • dy ' dz is the volume dr oi the cube. Therefore, using equation (1) we have: . dX.dY.dZ .^. PROBLEMS. 1. All space is filled with a fluid moving parallel to the a;-axis of reference at a uniform velocity of 10 centimeters per second. Find an expression for the velocity potential of the fluid. Ans. xp = 10a; + any constant. 2. The velocity components of a moving fluid parallel to the axes of reference are everywhere equal to a, h and c respectively. Find the velocity potential. Ans. rj/ = ax + hy -\- cz + any con- stant. 3. The velocity components of a fluid parallel to the axes of reference are ax, hy and cz, respectively. Find the velocity potential. Ans. \J/ = \ax^ + \hy'^ + \cz^ + any constant. 4. A viscous fluid flowing over a plane has a velocity which is everywhere given by the equation X = ay as shown in Fig. p4. Show that no velocity potential exists. y-axi8 *■ i ,> X a* gj/ Jub:. x-axi8 Fig. p4. 5. A vessel of water rotates uniformly at a speed of two revo- lutions per second about a vertical axis (the 2-axis). Derive expressions for the velocity components of the moving water and show that the velocity has no potential. 6. Liquid flows over an infinite plane towards a circular spot 232 CALCULUS. 20 centimeters in radius where the liquid leaks through the plane at the rate of two cubic centimeters per second for each square centimeter of the leaky portion of the plane. The depth of the liquid is everywhere 10 centimeters. Choose the x and 2/-axes of reference in the plane with the origin at the center of the leaky portion, and derive expressions for the x and 2/-components of the fluid velocity in the region over the leaky portion. Ans. x- component = r^:. Note. — In this and the following problems consider only the horizontal part of the motion and assume the velocity to be the same from top to bottom of the liquid. 7. Find an expression for the velocity potential in the region x^ -I— ifi over the leaky portion. Ans. — ^ — h any constant. 8. Derive expressions for the x and y components of the fluid velocity in the region outside of the leaky portion. Ans. x- 40x component = -^x~2- 9. Find an expression for the velocity potential in the region outside of the leaky portion. Ans. 20 logc (x^ -f y"^) + any constant. 10. Determine the constants of integration involved in the answers to problems 7 and 9 on the assumption that the velocity potential is zero at the edge of the leaky spot. Ans. (a) — 20, (6) - 20 log. 400. 11. A straight "fence" 20 centimeters long and 10 centimeters high is placed in the layer of moving liquid on the plane in problem 6. Find the surface integral of the fluid velocity over the fence (a) when the ends of the fence are at a distance of Vl25 centimeters from the ce nter of the leaky spot, and (6) when the ends of the fence are V 1,000 centimeters from the center of the leaky spot. Ans. (a) 100 cubic centimeters per second, (6) 800 tan~i 0.333 cubic centimeters per second. 12. Find the divergence of the fluid velocity in problem 6 (a) over the leaky spot and (6) in the region outside of the leaky spot. VECTOR ANALYSIS. 233 Ans. (a) 0.2 cubic centimeter per second per cubic centimeter, (h) zero. 13. Find the divergence of the fluid velocity specified in prob- lem 3. Ans. a -f- 6 + c; 14. One end of a rubber band is fixed, and the band is stretched by moving the other end of the band at a velocity of 2 centimeters per second. The band is 24 centimeters long at a given instant. Find the divergence of v where v is the distributed velocity of the band. Assume that the band does not contract laterally. Ans. 0.0833 cubic centimeters per second per cubic centimeter. 15. Find the distribution of pressure in a tank of water rotating at a speed of one revolution per second (w = 27r radians per second), the density of water being 62.5 pounds per cubic foot. Ans. p = 123.3 r^ where p is the pressure in poundals per square foot at a point r feet from the axis of rotation. Note. — Consider an element of the rotating liquid as indicated by the shaded area in Fig. pl5, the dimen- sion perpendicular to the plane of the paper being I. The area of side a is Ir . do and the outward push on a is plr . do. The area of the side h is Z(r 4- dr) . do, and the pressure at this face is (p + dp) ; therefore, dropping infinitesimals of the third order, the force pushing inwards on face h is plr . do -{- pl-dr .do +lr .pd.dd. The area of faces c and e is I . dr and the pressure over these faces may be considered as equal to p; therefore the normal forces on faces c and e are pl.dr as shown. The outward compo- nent of these two forces, according to Fig. pl5. Art. 51, is pi . dr Xr (omitting infinitesimals of the third order). Therefore the net inward force due to pressure acting on the element of liquid is Ir . dp . de, and this must be equal to the product of the mass Dlr . dd . dr of the element of liquid and its radial acceleration coV. 16. Find the expressions for the pressure gradients in the 234 CALCULUS. rotating liquid of problem 15 in the directions of x and y axes lying in a plane at right angles to the axis of rotation. Ans. X = 246.6X, Y = 246.6y. 17. Find the divergence of the pressure gradient in the rotating fluid of problem 15. Ans. 493.2. 127. Reduction of a surface integral over a closed surface to a volume integral extended throughout the enclosed region. — A very important transformation in the theory of electricity and magnetism is the reduction of a surface integral to a volume integral or vice versa. Let J? be a distributed vector, and let us consider its surface integral over a closed surface (normal directed outwards). Imagine the entire enclosed region to be broken up into small cells Uke the individual bubbles in a mass of foam. Then the integral of R over the hounding surface of the region is equal to the sum of the integrals of R over the hounding surfaces of the individuxil cells (normal directed outwards in each case). This proposition is evident when we consider that every wall which separates two cells is integrated over twice with directions of normal opposite (see Art. 125), so that the surface integrals over dividing walls are thus cancelled. The only surface integrals which are not thus cancelled are the integrals over the parts of the surface which bounds the region. Let d$ be the surface integral over one of the cells (the flux of R out of the cell), and let Jr cos e - dS be the surface mtegral of R over the bounding surface of the enclosed region. Then from the above proposition we have; fR cos €' dS = fd^ (1) But d^ = p ' dr according to equation (1) of Art. 126, so that equation (1) becomes: Jr cos € • dS = fp ' dr (2) This equation may be expressed in words as follows : The integral of a distributed vector R over a closed surface is equal to the integral VECTOR ANALYSIS. 235 of the divergence of R throughout the enclosed region. The first member of (2) is of course a surface integral and the second member is a volume integral. 128. Solenoidal vector fields. — A vector field is said to be solenoidal when its divergence is everywhere zero. The surface integral of such a vector field over any closed surface is zero according to equation (2) of Art. 127. Tube of flow. — Imagine stream lines to be drawn in a vector field from each point of the periphery of a closed curve or loop. These stream lines form a tubular surface which is called a tube of flow of the given distributed vector. Consider a number of diaphragms which stretch across a tube of flow. The flux is the same across each diaphragm. This is evident when we consider that any two diaphragms, together with the walls of the tube, constitute a closed surface out of which the total flux must be zero because the distributed vector under consideration is assumed to be solenoidal; but the flux across the walls of the tube is zero because the vector field is everywhere parallel to the walls. There- fore the flux into the enclosed space across one diaphragm must be equal to the flux out of the enclosed space across the other diaphragm. Unit tube. — A tube of flow is called a unit tube when the flux through the tube is unity. Thus, for example, a unit tube has a sectional area of 0.1 of a square inch at a point in a moving fluid where the velocity of the fluid is 10 inches per second. A unit tube has a sectional area of 0.1 of a square centimeter at a point in a magnetic fleld where the intensity of the field is 10 gausses. Imagine the solenoidal region of a distributed vector to be divided up into unit tubes. Then the flux across any surface anywhere in the region will be equal to the number of these unit tubes which pass through the surface. Each unit tube may be conveniently represented in imagination by the single stream line along the axis of the tube. When stream lines are drawn in this way, so that each stream line represents a unit tube, then the flux across any 236 CALCULUS. surface in the vector field is equal to the number of stream Unes which pass through the surface. The lines of force in a magnetic or electric field are always thought of as being drawn so that each line represents a unit tube, and the quantity of magnetic flux or electric flux through a surface is expressed by the number of lines of force which pass through the surface. 129. Curl of a distributed vector. — In Art. 126 examples were given of vector fields which emanate or "flow out" from a region or from every small part of a region. There are important cases in which a vector field curls round a region or round every small part of a region. Thus the stream lines in a rotating bowl of water curl round the central portions of the bowl. The magnetic field due to an electric wire curls round the wire. The electric field induced by an iron rod while the rod is being magnetized curls round the rod. Consider a small plane area AS at a point p in a. vector field. Let AL be the fine integral of the distributed vector R round the boundary of this element of area. It can be shown,* when R is a continuous function of the coordinates Xy y and z, that the ratio -r-s approaches a definite limiting value as the element of area AS grows smaller and smaller. This limiting value, -r^ is the component of a new vectorf C in the direction of the normal to dSf and this new vector C is called the curl of the given vector R at the point p. From this definition we have: dL = C cose- dS (1) where dL is the line integral of R round the boundary of a * The actual proof that — „ approaches a definite limiting value may be established without great difficulty by considering the part v' of the linear vector field in Art. 133; see Art. 135. t See Art. 135. VECTOR ANALYSIS. 237 small plane element of area dS, C is the curl of R at the element of area, and e is the angle between C and the normal to dS. Example of curl. — Consider a uniformly rotating bowl of water, the speed of rotation being oi radians per second. Consider the character of the fluid motion in the immediate neighborhood of the point p distant D from the axis of rotation of the bowl as shown in Fig. 112. The motion of a small portion of the water near p may be thought of as a combination of (1) A motion of translation at velocity Deo, and (h) A simple motion of rotation at a speed of w radians per second about an axis through the point p.* Now in considering the line integral of the fluid velocity around the circle cc, the motion of translation evidently need not be considered. Let r be the radius of the circle cc. The velocity of the fluid at every point of the circumference of cc (rotatory motion only being considered) is cor and it is everywhere in the direction of the circumference. There- fore the angle € is zero and the expression for the line integral around the circle reduces to circumference X (>ir. Therefore AL = 27rr2aj, where AL is the line integral of the fluid velocity around the circle cc. But the area of the circle is dS = ttt^. Therefore from equation (1) we get: C = 2aj. That is, the curl of the fluid velocity in the rotating bowl is Yig. 112. equal to two times the speed of the bowl in radi- ans per second; and since the expression for C does not contain D it is evident that C has the same value everywhere in the bowl. Now the angular velocity co of the bowl is a vector and its vector direction is parallel to the axis of rotation and towards the reader in Fig. 112. Therefore C is a vector, and it is towards the reader at every point in Fig. 112. The "stream lines" of the vector C are straight lines parallel to the axis of rotation of the bowl. * See Franklin and MacNutt's Mechanics and Heat, pages 174-175, The Macmillan Co., New York, 1910. 238 CALCULUS. Cartesian expression for curl. — Consider a small rectangular plane area of which the edges are dy and dz, as shown in Figs. 113 and 114. Let X, Y and Z be the components at p of the y-axia z-axU y-axi$ Q Z•^^'f z ^ z-axia Fig. 113. Fig. 114. given distributed vector R. To find the line integral of R around the small rectangle in the direction of the small curved arrow, let us consider first th^ two edges a and 6 in Fig. 113. The line integral along the edge a is — Y • dy because the component Y is counter to the direction of the small curved arrow. If Y is the i/-component of R at any point of the edge a, then the 2/-component of R at the corresponding point of the edge h dz; and if the line being -3- • dz greater, is equal to F + , az dz integral of R along edge a is — Y ' dy, then the line integral of R along the edge h is + (y -{- -^-dzj-dy* Therefore the net value of the line integral along the two edges a and b is -f ' dy » dz. In the same way it may be shown from Fig. 114 that the net line integral (in the direction of the small curved arrow) along the edges e and /is — -r- ' dy - dz. Therefore dZ , -Ty'^y * The argument here given is not entirely rigorous, and the peculiar wording of this sentence is intended to suggest a slightly modified argument which is rigorous. VECTOR ANALYSIS. 239 the total line integral dL of R around the small rectangle is: But dy ' dz is the area dS of the rectangle. Therefore using equation (1) and remembering that the x-component of the curl is at right angles to the rectangle in Figs. 113 and 114, we have: ^^- dz~ dy ^^^ In a similar manner we may get: ^'~ dx dz ^^^ and ^-g-S («) where d, Cy and C z are respectively the x^ y and z com- ponents of the curl of the distributed vector R whose x, y and z components are X, Y and Z respectively. The curl of a gradient is necessarily equal to zero. — This is evident when we consider that equations (3), (4) and (5) of Art. 124 are satisfied in a vector field when the vector is a gradient. 130. Reduction of a line integral around a closed loop to a surface integral over a diaphragm bounded by the loop. — A very important transformation in the theory of electricity and mag- netism is the transformation of a line integral to a surface integral or vice versa. Let E be a distributed vector, and let us consider its fine integral around a closed curve or loop AB, Fig. 115, the heavy arrow showing the direction in which the line integral is taken. Imagine a diaphragm of any shape whatever stretched across the loop AB, and imagine this diaphragm to be divided up into small meshes as if the diaphragm were made of wire 240 CALCULUS. gauze. Then the integral of R round the loop AB is equal to the sum of the integrals of R round the individuul meshes of the diaphragm, the integrals round the meshes being taken in the same direction as the integral round the loop as indicated by the small curled arrows. This proposition is evident when we consider that every line dividing two meshes is inte- grated over twice in opposite direc- tions (see Art. 123), and when we consider that in integrating round the various meshes we eventually integrate along every portion of the loop AB once in the direction of the heavy arrow in Fig. 115. Let dL be the line integral of R round one of the meshes, and let jR cos e - ds be the Hne integral of R roujid the loop AB, Then from the above proposition we have: Fig. 115. J R cos € ' ds = f dL (1) But dL = C cos € ' dS, according to equation (1) of Art. 129, so | that equation (1) becomes: ■ J R cos e ' ds = J C cos e • dS (2) This equation may be expressed in words as follows: The integral of a distributed vector R round a closed loop is eqvxil to the integral of the curl of R over any diaphragm stretched across the loop. The first member of (2) is of course a line integral and the second member is a surface integral. The divergence of the curl of a distributed vector is always and everjnvhere equal to zero. — Consider two diaphragms stretched across a closed loop. The integral of C is the same over both of these diaphragms because each surface integral is equal to the line integral of R round the loop. If the direction of the normal VECTOR ANALYSIS. 241 be reversed in the integral over one of the diaphragms the integral will be reversed in sign (see Art. 125). Therefore the surface integral of C over a closed surface, namely, the two diaphragms, with normal directed outwards, is equal to zero. That is C is a vector whose flux out of or into a closed surface is always equal to zero. Therefore C is a solenoidal vector and its divergence is everywhere zero. See Art. 127. 131. Vector potential. — Let us imagine a distributed vector of which a given distributed vector is the curl. Then the imagined distributed vector is called the vector potential of the given distri- buted vector. The divergence of a curl is always and everywhere necessarily equal to zero, as explained in the previous article. Therefore, to have a vector potential, a given distributed vector r. . dX , dY , dZ . _^, ,. must have no divergence. But ^ — \- -^ — \- -f~ is the divergence of a distributed vector whose components at a point are X, Y and Z, according to equation (3) of Art. 126. Therefore the existence of a vector potential (of a given distributed vector) is determined by the condition; dx^ dy^ dz " ^ ^ where X, Y and Z are the components at a point of the given distributed vector. Example. — In certain cases of fluid motion each particle of the fluid is rotating at a definite speed about a definite axis. This spinning motion of the particles of a fluid is a distributed vector* and it is called vortex motion. Now it was shown in Art. 129, that the curl of the velocity of the water at a point p in a rotating bowl is equal to two times the spin velocity w of the particle of water at p. That is, ignoring the factor 2, the curl of fluid velocity * Spin is a vector quantity, its vector direction being along the axis of spin in the direction in which a right-handed screw would travel if turned in the direction of the spin. 17 242 CALCULUS. IS the vortex motion; therefore the vortex motion of a fluid is a distributed vector whose vector potential is the fluid velocity. In this case the vector potential (of vortex motion) is physically existent; in general, however, the vector potential of a given distributed vector is merely imagined to exist. The idea of vector potential is useful because it is sometimes helpful to think of a given distributed vector as a curl. 132. Rotational and irrotational vector fields. Vector potential and scalar potential. — ^When the velocity of a moving fluid has curl the particles of the fluid are in rotation. In consequence of this fact any vector field which has curl is called a rotational vector ( ^ field. A vector field which has no curl is called an irrotational vector fiM. In an irrotational vector field equations (3), (4) and (5) of Art. 129 reduce to equations (3), (4) and (5) of Art. 124. There- fore an irrotational vector field can be thought of as a gradient, or, in other words, an irrotational vector field has a potential (a scalar potential). In a rotational vector field equations (3), (4) and (5) of Art. 124 are not satisfied, and therefore a rotational vector field cannot be thought of as a gradient, or, in other words, a rotational vector field has no potential (no scalar potential). A vector field which has no divergence is called a solenoidal vector fiM, and a solenoidal vector field can be thought of as a curl, or, in other words, a solenoidal vector field has a vector potential. A vector field which has divergence is called a non-solenoidal vector field, and such a vector field cannot be thought of as a curl, or, in other words, a non-solenoidal vector field has no vector potential. PROBLEMS. 1 . A viscous liquid moves over the xy plane so that X = az. Find the curl. Ans: — a. 2. Find the vector potential of fluid velocity when the fluid is everywhere moving in the direction of the 2-axis at a velocity VECTOR ANALYSIS. 243 of + 100 centimeters per second. Ans. IOO2! centimeters squared per second squared. Note. — A distributed vector which has neither divergence nor curl in a given region may have a scalar potential and it may have a vector potential. 133. Character of a very small portion of any vector field. — Let X, Y and Z be the components of any distributed vector v at a point p whose coordinates are x, y and z. Then X, Y and Z are functions of x, y and 2, and indeed we will assume them to be continuous functions. If one is to reach a clear understanding of divergence and curl it is necessary to examine carefully into the manner of distribution of v in a very small region near a chosen point. It is most convenient to locate the origin of coordinates at the chosen point. Then the coordinates X, y and z of any point in the very small region are infinitesimals, and the squares and products and higher powers of x, y and z are negligible. Now the components X, Y and Z being continuous functions of X, y and z can be expanded by Maclaurin's theorem as ex- plained in Art. 90, and all terms in these expansions which contain squares and products and higher powers of x, y and z may be discarded, giving: X = Xo + aix + a2y + a^z (1) F = Fo + 61X + h,y + hz (2) and Z = Zo + CiX + C22/ + Cziz (3) in which the constant coefficients are the values of the derivatives at the origin, as shown in the following schedule : (I) dX dx dX "'-dy dX "'- dz ^'- dx ^' = dy dY ^'= dz dZ dZ '^-dy dZ 244 CALCULUS. The " linear " vector field. — The simplest type of vector field is the homogeneous field; in such a field X has the same value throughout the field, Y has the same value throughout the field, and Z has the same value throughout the field. When X, Y and Z are linear functions of x, y and z we have what is called a linear vector field. Thus equations (1), (2) and (3) represent a linear vector field. Resolution of a linear vector field into simple parts. — The dis- cussion of divergence and curl in Arts. 126 and 129 is not rigorous, and, as is always the case in a merely plausible discussion, the harder one tries to understand it the more vague and unintelligible it becomes. It is now proposed, therefore, to establish rigorously the ideas of divergence and curl by considering a linear vector field as expressed by equations (1), (2) and (3). For this purpose it is necessary to resolve the given linear vector field into three parts, and to be able to speak of these parts intelligibly let us think of the given linear vector field as the velocity t; of a moving fluid, the components of v being X, Y and Z as given in equa- tions (1), (2) and (3). The three parts into which v is to be resolved are: 1. A uniform translatory motion of the entire body of fluid. This part of v will be represented by Vo and its components are Xo, Yq and Zq. 2. A simple motion of rotation of the entire body of fluid. This part of v will be represented by v' and its components will be represented by X', Y' and Z', 3. A continuous stretching of the entire body of fluid in three mutually perpendicular directions. This part of v will be repre- sented by v" and its components will be represented by X", Y" and Z", Discussion of v'. — Consider a rotating body of which the axis of rotation passes through the origin of coordinates, and let cox, 03 y and 0)2 be the components of the angular velocity of the body around the x, y and z axes of reference, respectively. VECTOR ANALYSIS. 245 Consider a point p of the body of which the coordinates are X, y and z as indicated in Figs. 116, 117 and 118. The velocity Fig. 116. Fig. 117. Fig. 118. of p due to CO a; consists of two components yo)x and — zccx as indicated in Fig. 116, and similar statements may be made con- cerning Figs. 117 and 118. Therefore, picking out the x-compo- nents of the velocity of p in Figs. 117 and 118, we get the value of X', and in a similar manner we get expressions for Y' and Z' as follows: X' = " ' — oi^y + ojyZ (4) y = + 0)eX • • • — OiixZ (5) Z' = — 03yX + £0x2/ (6) The values of cox, coj, and co^, expressed in terms of the coefficients in equations (1), (2) and (3), can best be determined after the expressions for v" have been formulated. Discussion of v''. — As in case of v' it is necessary to find the forms of expressions which give X'', F'' and Z'\ For this purpose consider three mutually perpendicular axes of reference Xi, yi and Zi. Then a continuous stretch of the portion of fluid parallel to the Xi-axis is represented by the equation Xi = axi, where Xi is a velocity parallel to the rci-axis and a is a constant. Analogous expressions using j8 and y as constants represent continuous stretches parallel to the 2/1 and Zi axes. Therefore 246 CALCULUS. continuous stretches in three mutually perpendicular directions (which directions are for the moment chosen as the axes of refer- ence) are expressed by the equations: Xi = axi (7) 1^1 = Pyi (8) and Zi = 721 (9) It is desired to transform these equations so as to express exactly the same fluid motion, but to express it by giving the com- ponents of v" parallel to new axes a;, y and z, and as functions of the new coordinates x,y and z. This transformation is enorm- ously simplified by expressing the given fluid motion [equations (7), (8) and (9)] in terms of its velocity potential P, which is a distributed scalar whose Xi-gradient is 3— = axu whose 2/1-gradi- axi ent is ^ = Pyi and whose 2i-gradient is -p = yzi. Integrating these three differential equations and ignoring the constant* of integration we get: P = IciXr'^ + \W + \W (10) Now to transform equations (7), (8) and (9) as stated above we need only to substitute in equation (10) the following values for Xif 2/1 and 2i [see equations (11), (12) and (13)], and then find the gradients of P in the directions of the new axes. In this way we get equations (14), (15) and (16). Xi = hx + m\y -\- niZ (11) 2/1 = Ux + rn^y -f n2.z (12) Zi = I2PC -{- mzy + n^ (13) where U mi ni, U W2 ^2 and h rriz riz are the direction cosines of * The three equations are partial differential equations, but the functions of integration reduce to a single imdetennined constant. I VECTOR ANALYSIS. 247 the new x, y and z axes referred to the old Xi, 2/1 and Zi axes. X" =fx + py + qz (14) F' = px + gy + rz (15) Z" = qx + ry-\-hz (16) The coefficients in these equations are of course expressions involv- ing a, jS and 7 and the direction cosines in equations (11), (12) and (13), but they are represented by the single letters /, g, h, p, q and r for the sake of clearness. Anyway, the only im- portant feature about equations (14), (15), and (16) is the sym- metry of the coefficien s which means that X'\ Y" and Z" are the components of a fluid motion consisting of three continuous mutually perpendicular stretches, because this degree of symmetry is the only necessary consequence of equations (7), (8) and (9). Determination of coefficients in equations (4), (5) and (6) and in equations (14), (15) and (16) in terms of the coefficients in equations (1), (2) and (3). — Equations (4), (5) and (6) and (14), (15) and (16) show only the degree of symmetry that the co- efficients must have in order that equations (4), (5) and (6) may represent a simple rotation and in order that equations (14), (15) and (16) may represent three continuous mutually perpendicu- lar stretches, and the coefficients in equations (4), (5) and (6) and (14), (15) and (16) may be thought of as undetermined. It remains to determine these coefficients cox, co^, co^, /, g, h, p, q and r, so that Vo, v' and v" may be component parts of the given fluid velocity v which is represented by equations (1), (2) and (3). To do this add equations (4) and (14) and place the coefficients of x, y and z in the resulting equations equal to the coefficients of x, y and 2, respectively, in equation (1); and proceed in a similar manner with the other pairs of equations. In this way we get the following schedule of equations: f = ai p — coz = az q + o)y = as^ P + 03z = hi g = ^2 r - COx = 63 p (II) q — 03y = Ci r + Wx = C2 h = C3J 248 CALCULUS. from which we get the following values: w, = i(c2 — hz) o)y — \{az — ci) Or, using the values of ai, a^, 03, 61, 62, etc., from the schedule I of equations, we have: dZ dY 2-^ = dy dz (17) dX dZ 2co« = -3 3- ^ dz dx (18) _ dY dX 2"^ - dx dy (19) and from equations II we get also: dX ■' dx (20) dY ^~ dy (21) dz (22) P = ^{dx + dy) (23) JdX , dZ\ ^ = Kdz+dx) (24) ,/dZ , dY\ (25) PROBLEMS. 1. Find the flux of v" out of a rectangular parallelopiped which is I feet long in the direction of the a:i-axis, w feet wide in the direction of the 2/1-axis, and t feet thick in the direction of the 2i-axis, the constants in equations (7), (8) and (9) being expressed in reciprocal seconds. Ans. lwt(a + 18 + 7) cubic feet per second. VECTOR ANALYSIS. 249 2. Find the divergence of v" , Ans. a + /S + 7- "Note. — It is here intended that the student find the flux $ which comes out of a given volume t, and then determine the limit of - as t approaches zero. It is simplest to take for r the rectangular parallelopiped of problem 1. 3. Show that the integral of v" around any closed curve what- ever is zero. ^ote. — Choose the axes of reference as in equations (7), (8) and (9). Let ds be an element of the closed path or curve. Then the scalar part of the product v" . ds can easily be found, and it is easy to show that the integral of this scalar product aroimd the closed curve is zero. 4. It is evident that the velocity v\ which is a simple motion of rotation, has zero flux into or out of any closed region whatever. Therefore the divergence of v' is everywhere equal to zero. Find the line integral of v' around the ellipse of Fig. 122, the radius of the cylinder being r, the axis of the cylinder being parallel to the axis of rotation of the fluid, and the angular velocity of rotation being co. Ans. l-KV^iti. 134. Proof that ^ -j- and -i, -r- are identical when z is dy • dx dx • dy a continuous function of x and y. — This proposition may be stated so as to appeal to one's geometric sense as follows: Let z be a function of x and y and let this function be represented by a hill built upon the xy plane. Then z is the height of this hill above the point p' in the base plane, x and y being the coordinates of p' as shown in Figs. 41a and 416 in Art. 62. Let x(^-p\ be the a;-component of the slope of the hill at the point p (see Figs. 41a and 416), and let ^( = ^) be the ^/-component of the slope of the hill at p. Figs. 119 and 120 represent a top view of the hill. Let dz be the difference of level of the points p and q on the hill in Figs. 250 CALCULUS. 119 and 120. Let us travel from p to g along the sides a and h x^axU dx />>. <^-»e r X e 9 A X-^^'dy IHveu ^X-axis AS dx^ < y-axl "a c b Y 8 ^^gM Fig. 119. Fisc. 120. of the infinitesimal square, and the rise will be; dz = X • rfa; + F + dY y ' dx dx] ' dy Let us travel from p to q along the sides c and e of the in- finitesimal square, and the rise will be: dz = Y ' dy ■{- (x -\-^ ' d^ ' dx Now if z is a continuous function of x and y these two expressions for dz must be identical, that is, one must rise by the same amount in going along a and 6 as in going along c and e in Figs. 119 { and 120. Therefore dz dY ^ , 1 . ^^ -J- must be equal to -y- . dx ^ dy But X = — dx and 7 = dy' Therefore ^z ., , ^ d^z -3 -J- must be equal to -; dy ' dx dx dy The above argument is not entirely rigorous but it brings out very cleariy the geometrical significance of equations (3), (4) and (5) of Art. 124. To make the argument rigorous a slightly altered point of view suffices, as follows: The gradient of the hill at any point in the side e of the infinitesimal square exceeds the gradient at the corresponding point in the side a by the amount dX dy dy, so that the rise along e exceeds the rise along a by the amount VECTOR ANALYSIS. 251 ( 7P • % ) • ^^- Similarly the rise along side b exceeds the rise along side c by the amount (-j- » dx) ' dy. Therefore the rise along sides c and e exceeds the rise along a and b by the amoimt l^ ^ J • dx • dy. But the rise along c and d must be the same as the rise along a and 6. There- fore dX dY -r- must be equal to zero. dL dS 135. To show that -^ is the resolved part of a vector (the curl of R) in a direction normal to dS, where dL is the line integral of R around the boundary of a surface element dS. — It can be shown without difficulty that the two parts Vo and v" of the Hnear vector field v of Art. 133 have zero Hne integrals around any closed curve. Furthermore any distributed vector R may- be looked upon as having a linear distribution throughout a very small region. Therefore to establish the above proposition it is sufficient to consider the portion v' of a linear vector field. Now v' is a simple motion of rotation about a definite axis. Let CO be the angular velocity of this rotation. The line integral of v' around the circle cc in Fig. 121 is equal to 27rr^aj, so that cylinder Fig. 121. normal to ^-^ ^ >^ Fig. 122. the line integral divided by the area of the circle is equal to 2a>, as explained in Art. 129. Consider a cylindrical surface of which the end view is the circle cc in Fig. 121. A side view of this 262 CALCULUS. cylinder is shown in Fig. 122. Let dS be the area of an oblique plane section of the cylinder. Let dL be the Une integral of v' around the boundary of dS. It can be shown without difl&culty that dL is equal to 2Trr^o). But the area of dS is ;:. There- cos 6 fore -75 is equal to 2co cos 6. That is, 2w is a vector whose component in the direction of the normal to dS is equal to -75-. 136. The distortion of a small portion a body which has been twisted or bent in any manner whatever. — Consider a particle p of an unbent beam. Imagine the beam to be bent or twisted in any manner whatever, and let 1; be a line drawn from p to p', where p' is the position of the particle after the beam is bent. Then the line v is a vector, it is called the displacement vector of the particle p, and it may be thought of as being at p because it refers to the point p. Now v has a definite value and a definite direction at each point of the unbent beam, therefore it is a distri- buted vector. That is the unbent beam may be thought of as a vector field in its relation to the bent or twisted beam. Consider a very small portion of the unbent beam. The displacement vector v may be thought of as having a linear distri- bution throughout this small portion of the beam as explained in Art. 133. Therefore the displacement-vector field v may be resolved into three parts as explained in Art. 133. One part Vq is a simple translatory displacement of the entire small portion of the beam; another part v' is a simple rotation of the small portion of the beam about a definite axis; and a third part v" consists of three mutually perpendicular stretches.* This is a fundamental proposition in the theory of elasticity.! * The expression cordinuom stretches as used in Art. 133 referred to what takes place in a rubber band while it is being stretched, and the word stretch as here used refers to what has taken place in a rubber band after it has been stretched. t See the chapter on elasticity in Franklin and MacNutt's Mechanics and Heat, The Macmillan Co., New York, 1910. I I VECTOR ANALYSIS. 253 PROBLEM. 1. One end of a rubber band is fixed and the band is stretched by moving the other end of the band. The initial length of the band is 10 inches, the increase of length is 0.5 inch. Express the displacement of each particle of the stretched band in terms of the initial distance x of the particle from the fixed end of the band, ignoring the lateral contraction of the band. Ans. Displace- ment = O.OSic. APPENDIX A. PROBLEMS GROUP 1. (After Art. 34.) Differentiate the following: 1. 2/ = n+ 1' dx g = 20.3(. + l). 3. y = 6a;8 - 5x^ + 4^^ - Sx^ dy dx 4,y = x^x + 7), 5. t/ = 3a;'(a;2 - 5a;), 6. e = ar*(6r3 - cr + d), 7^b<*-^ ^ = ar3(76r3 - 5cr + 4d). ar 7. = (r + 3)(2r + 5), ^^ ^ I 11 -=- = 4r + 11. ar 8. e = (2r2 + 3r)(3r2 + 2r), ^ = 3r(8r2 + 13r + 4). CALCULUS. dr ^ = (r2 - r + l)(5r2 - 3r + 1). ar 37 = 2act -{- ae + be, at 9. = (2 - r)(3 - r), dr '' 10. = (r^ - r2 + r)(r2 - r + 1), de ^ dr 11. s = iat + h)(ct-\-e), ds ^ dt 12. « = (a - 60(c - eO, ds _ dt ~ 13. s = (3^ - 2< - 1)(1 - 3^2 4- 5^3)^ 14. 8 = (2 H- 0^ 15. 5 = (1 + 2x)^ 16. 8 = ^(2< 4- 3)«, 5? =26e<- (ae + bc). at 5^ = 75<* - 76^3 + 3^ + 12^ - 2. at ^ = 3(2 + t)\ ^ = 6(1 + 2x)\ ax I =2«(7< + 3)(2< + 3)^ 17. s = a«H6^ + c« + e), ^ = a^(56^ + 4d + 3e). 18. 2/ = (1+ 2x)H2x2 + 1)\ ^ = 2(1 + 2a;)2(2x2 + 1)3(3 + Sx + 22x2). ax 19. 2/ = (ax + by{a + 6x)3, ^ = (ax + 6) (a + 6x)2(5a6x + 2a2 + 36^). ax APPENDIX A L. 2°-^ = x + 2' A^: dy 2 v- da; {X + 2Y' ^t CI — X dy a — h dx~ (b-xy ^^- ^" (2x + 5)4' dy 2(a; + 3)(2a; + 7) dx {2x + 5)» 23 .-(^^ + ^)' t z' 1 + a; \ d^ _ {ax + bYiSae - 76c - 4aca;) dx ~ (ex + ey 25. 26. /2x- 1\ /ax^ -hx + cY dy _ dx 3(1 + xy {2x + 3)* • dy 12(2x - -D' dx- (X- 2)' • dy __ 5{ax^ — c)(ax'^ — hx -\- c) dx e^x^ 27. Find the slope of the curve y = x^ — 2x^ + Z at (a) x = 0, (6) X = 2, and (c) x = - 2. Ans. (a) 0, (6) 4, (c) - 20. 28. At what angles does the curve y = x{x — 2){x — 3) cut the a;-axis? Ans. (a) At a; = 0, slope is 6, (6) at x = 2, slope is — 2, (c) at a; = 3, slope is 3. 18 4 CALCULUS. 2a; 4- 3 29. Is -. — r-^ an increasing or a decreasing function of x? 4x + o Ans. Decreasing. 30. Is o _t A ^^ increasing or a decreasing function of «? oX -p 4 Ans. Increasing. 31. When y = 2x^ — Qx + 5, for what values of x is t/ (o) an increasing function, and (6) a decreasing function? Ans. (a) All values of x less than — 1 or greater than + 1, (6) all values of X between — 1 and + 1. PROBLEMS GROUP 2. (After Art. 89.) Differentiate the following functions Note. — The forms referred to in the answers to the following problems are found in the table of integrals, see Appendix B. Thus the answer to problem 4 is given by form 45, as dy 1 dx V^ + hx the answer to problem 5 is found to be dy _ x =, etc. l.V = ^-3x + l-l„ dy _ dx 2x-3-%-^. 2. J/ = 3V?-2V?, dy dx 2Vx' 5 dy _ dx 3 , , 20 , b Form 45. S. y= Va;2 + a^, Form 61. 6. 2/ = - Va^ — x^y Form 53. APPENDIX A. 5 7. e = (r^- a^y^, ^ = 5r%r' - a')K 8. y = ib V(aT6i)^ Form 42. 9. y = i V(rc2 + a2)3, Form 58. 10. 2/ = - J V(a2 - a;2)3, Form 49, 11. 1/ = (a. - x.)t, ^ = ^. d^ r3(7 - 8r) 12. d == r^-ylr - r^, rfr 2 -^7- _ 7.2 2(2a - hx) ^a + hx ^ 14. 2/ = — 25^ , Form 46. _ 2(2a - Zhx) V(a + M^ ^ 15. 2/ = -^ ^^ '-y Form 43. 2(8a2 - 4:ahx + 36V) Va + 6rc 16. 2/ = ~- — ^553 > Form 47. 17. 2/ = -^ 1Q553 ^> Form 44. 18. 2/ = 7,/ ■ I. x > Form 34. o{a + oaj) 19 ^ = _ ^^l-~^\ Form 64. 20. y --= , ,1 , Form 66. aNa;2 + a2' 22. „ = - ^'^^ - ^, Form 73. ^ ax - (3^' - Q')^ ^ = (a;^ - a^)K7a;^ + 9a') ^^' ^ ~ a;5 ' dx . 6a;§ |a2"- aj2 <^^ - 4a2a; " NoH^T^' S ^ 3(a2 - a;2)§(a2"+^t CALCULUS. 25., = log„.n, 1 = ^ 26. u = log (3t;2 + 4t;3)», du ^ 18(1 + 2t;) dv v{3 + 4t;) ' 27. 2/ = ^ log (a + 6a;), Form 33. 28. 2/ = ^ log ^a;2 + ^V Form 39. 29. 2/ = log (x + Vl + x^)y Form 10. 30. y = log (a; + V^~^), Form 11. 31. 2/ = log (x + Va;2 + a^), Form 60. 32. 2/ = - log(^ +iJJ+l), From 16. 33. 2/ = - log (i + ij^2 - l)» Form 15. 35. s - ti log (a^ - 60* I = %faT-^ W "^ ^* ^""^ ^""^ " ^^^' ^^•^ = ^^^S^' ^^^^^• 37. 2/ = i log i-±^, Form 13. 1 X 38. 2/ = - log — 1 — P==, Form 63. 39. 2/ = - log "^^' + ^' - ^ Form 63. "a X __J___, 2cx + 6 - Ate2 _ ^ac ^ .. 40. 2/ = h, . log X r-i—F — ■ f Form 41. ^ V62 - 4ac *^ 2cx + 6 + a/^^ - 4ac 41. 2/ = p a + 6x — a log (a + 6a;) , Form 35. 42. 2/ = J [log (a + 6a;) + ^^^], Form 36. APPENDIX A. 7 43. 2/ = ^a^ - x^ — a log — ' , Form 51. X 44. 1/ - I -VSM^ + |- log (a; + Va;2 + a^), Form 57. 45. 2/ = I Va;2 + a^ - ^ log (a; + V^^ + ^2)^ Form 62. 46. 2/ = |(2a;2 + Sa^) Va:^ + a^ + ~ log {x + Vx^ + a^), Form 65. 47. y = |(2a;2 + a^) ^2 _|_ ^2 _ |' i^g (^^ ^^2-^:^2)^ Form 59. 48. 2/ = log. (30. + 2), J = 3^-2 49. ^ = ^ % = _ L±J2g^ a; log x* dx {x log xY ' 50. 2/ = a:i°«=", 1^ = loga-x*°««-i 51. 2/ = a^, ^ = 2a;a-Moga. ^"^^ ^ "^ » da: ~ ic • 53. ^ = c=^e^ 3- = c^e^(l + log c). 54. 2/ = e^'x^ -^ = e=^a;^-K^ + ^)' 55. 2/ = ^^x\ ^ = 4-a:3(4 + x log 4). _ ?!±? dy /- 2\ ?!±2 56.2/ = e«, ^=(^l-_je-^. 58„-5£_+3^^ # 24X-W 5'- ^=(^''-2)'. , %-W^r 8 CALCULUS. 11 1 dy 1 + log X 60. y = loglog^ - j^, ^ = -^(bi^- 61. y = ^{e^ - e-2-), ^ = c{e'^^ + 6-2-). 62. s = (2e2* - 1)(1 - 3e-30, ^^ = 4e2< + Ge"* - Qe-^'. /:o -.. i^« log ^ ^2/ I MO, y -^^^l + logx' 64. y = glog„x _ a;log««^ 65. y e' + x- dx a; log X (1 + log x) ^= loga_e/ ,,^,_^,,^A dx iC \ / ^ _ 2e'xf'-\a — x) dx " (e* - x")2 66. 2/ = x[(log xY - 2 log X + 2], ^ = (log x)2. 2/ - log gnx _^ 1 _^ g-nx » ^a- - g2«x + 1 + e'^"* * riogi(e- + l) __ log (6"- -1) 1 = log (e"'' - 1) %^ L e"^- 1 e"^ + 1 J ^^* ^ log (€"- +1)' da; [log (e-- + 1)]2 PROBLEMS GROUP 3. {After Art. 42.) Differentiate the following: 1. 2/ = sin hx, 2. 2/ = cos nXy 3. 2/ = sin 3a;2, 4. 2/ = 2 + 4 Sin 2a;, a; 1 5. 2/ = 2 "" 4 sin 2a;, d2/ 5 cos 5a;. dx~ — n sin na;. dy_ dx 6a; cos Sa;^. Form 21. Form 20. APPENDIX A. y dii 6. y = sin Zx cos 2x + cos 3a; sin 2x, 3^ = 5 cos 5x. sin (m - n)x sin (m + n)a; ^- ^ 2(m - n) ^ 2(m + n) ' '''' ^ «m (m - n)x _ sin (m + n)a; ^^^^ 24. ^ 2(m - n) 2(m + n) ' ^ cos (m - n)x cos(m + n)a: ^^^^ 25. ^ 2(m - ri) ^ 2(m + n) ' 10. y = log sin x, 11. 2/ = — log cos a;, 12. s = sin(<2 -st + 2), 13. s = log cos (1 - f), 14. y =log(a sin2 x dy ^ 2(a — h) tan x + & cos^ x), dx a tan^ x -{- h ' 15. w = 2x2 sin 2a; dy . ^ ^ , o o . ^ / = 4a:2 cos 2a;. + 2a; cos 2a; — sm 2a;, dx 16. 2/ = 6=^ sin (2 - a;2), ^ = e^ [sin (2-a;2) -2a; cos (2-a;2)]. ^o re -«^ ^^ 0[2 cos (e« - 6"^) - ^e^ + 6-«) 17. r - 02 cos (e^ — e "), t^ = • / « _flM ^ ^' dd sin (e^ — e '')]. , cos a; dy sin c 18. i/ = log Form 8. Form 7. J^ = (2^ - 3)cos(i2 _ 3« + 2). 1 = 2aan (1 - i2). cos (a; + c) ' da; cos x cos (a; + c) 19. V = sin' 2x cos^ 3a;, i^ = 6 sin^ 2a; cos 3a; cos 5a; . ^ aa; 20. v = e^^'^^^ -^ = 2a;cosa;2.e^^^^ ^ dx * ^ ^ sin a; + cos a; ' da; cos 2a; * , sin J(a; — a) d?/ sin a 22 1/ = log — - — ^ = ' ^ sin |(a; + a)' dx 2 sin J(a; — a) sin |(a; + a) 10 CALCULUS. 23. The lengths of crank radius and connecting rod of a steam engine are 2 feet and 10 feet respectively, as shown in Fig. p23, and the crank revolves at the uniform speed of 2 revolutions per second. Find the velocity of the piston for the following values of the angle 6. (a) 6 = 0, (h) 6 = 45°, (c) 6 = 90° and (d) 6 = 135°. Ans. (a) zero, (h) 20.32 feet per second, (c) 25.13 feet per second, (d) 15.24 feet per second. Fig. p2Z is 1 J" high X 4" long. Fia. p23. PROBLEMS GROUP 4. {After Art. 43.) Differentiate the following: 1. 2/ = cot X, 2. y = sec x, 3. 2/ = cosec Xf 4. 2/ = vers x, 5. 2/ = tan {x^ — 2a:), 6. 2/ = sec" nx, 7. 2/ = log tan-, J = — cosec^ X. ax dy . -J- = sec X tan x, ax dy . :r — — cosec x cot ». aa; = smx. dy dx ^ = 2(x - 1) sec2 (a;2 - 2x). dy „ ■f^ = rv- sec" na; tan nx. ox Form la APPENDIX A. 11 S, y = log tan ^ + I j, 9. y = log (sec x + tan x), 10. y = log (cosec x — cot x), 11. 2/ = e^* cosec 2x, Form 18. Form 18. Form 19. 4y _ 12. 1/ = (tan w— Scot w) Vtan^.-^p = r— — v— . ' dw 2 tan5 w da; dy = 2e2' cosec 2x(l - cot 2x), 13. s = <" cosec"* nt, 2taiid 15. 1/ = log [tan (e^ + S)*]^, 16. e = log tan ^ — 2 17. r = 2 tan ^ - 1 a sin ^ + 6 vers ^ -T. =ni"~^ cosec*" n<(l — mi cosnQ. di/ _ 2 sec^ e Te~{l - tan2 ^)2* dx sin [2(e^ + 3)i]* dg^ 3 dr 4 — 5 sin r dr 2ab vers 6 a sin — 6 vers d^ (a sin — 6 vers ^)2' PROBLEMS GROUP 5. (After Art. U-) Differentiate the following: dy 1 1.1/ = COS~^ X, 2. y = tan~^ a;, 3. 2/ = cot~^ x, 4. y = sec~^ a;, 5. 2/ = cosec"^ a;, 6. 2/ = vers~^ a;, da; Vl- a;2 Form 12. dy dx 1 l^x^' Form 14. dy _ 1 dx x^lx'' - 1 Form 17. 12 CALCULUS. X 7. 2/ = sin-i -, Form 52. 8. y = -tan-i-, Form 37. c c 2 , 2cx + h Form 40. 1 X 10. J/ = - sec-i-, Form 67. o a 11. y = vers-i -, Form 70. 12. 2/ = x sin-i X + Vl - a^, Form 27. 13. 2/ = x cos-i a: — V 1 — x^ Form 28. 14. 2/ = I Va2 - x2 + ^sin-i-, Form 48. 15. 2/ = - I Vo^"^^ + ^sin-i -, Form 54. 16. 2/ = 1(2x2 - a^) Va2-x2 + |^sin-i -, Form 50. 17. 2/ = J(5a2 - 2a;2) Va^-rc^ + ^' sin-^-, Form 55. o o O 18. to = COS"* Vvers t -jj= — h Vl + sec i. ^ ,3^ - 2 , ^ , 3(9 - 12 dr ^ 19. , = tan-^- + cot- -^^^, ^ = 0. 20. r = cos-i-3- - 2^-^, rf", = aJ^ - r 21. s = t^ sec-i I - 2 Vi2-4, ^=2^ sec-i|. 22. 0) = V2[sec-* (sec i + tan 01, dt V (sec t + tan tan t «. ^ ,4 + 5 tan x dy S 23. 2/ = tan 3 d^ = 5 + 4sin2x OA - -l!i±£l' ^ - - 2 Z4. s - tan ^, _ ^_,, ^^ - ^2, ^ g-2f APPENDIX A 25. y = tan-' (|tan x) - tan-'^, g = - 26. y = tan 13 ab sin^ X + ¥ cos^ x' J g^ tan a; - 6^ dy a6 a?)(l + tan xY 27. 1/ = V2ax — x^ + a vers" 28. 2/ = — V 2aa; — x^+ a vers"* — . X — a ITT 5 , a^ , X 29. 2/ = — ^ — V 2ax — a;^ + — vers~i -, 30, y = ' — ^ V 2ax - a;2 + — vers-^ -, dx o? sin^ X ■\-\P- cos^ x Form 72. Form 71. Form 68. Form 69. PROBLEMS GROUP 6 {Ajter Art. ^.) Differentiate the following: 1. 2/ = ^'^'j 2. = (1 - ty-', 3. s = Zt'\ 4. 2/ = x^^''^*^", 5. 2/= (iogxy°«- 6. 2/ = (3x2 + 1)2^^-2), ^ = nx"*(l + log x). ax I = - (1 - 0^-'[i + log(i - 01 | = 3<^'ni + 31og<]. ^= (n+l)(logx)«x^'"«*>"-i ^'[n-log(logx)]. dy ^ (log x) dx dy dx (3x= + !)«-«' {2 log(3a;» + 1) + ^3^^^ -^} • 14 APPENDIX A. PROBLEMS GROUP 7. {After rule III of Art. 67.) Integrate the following differential equations. 1. dy = (3x'-7^+ 1 - I'jdx, y = ^-1^+ log^ +^ + C. 2. dy = (2 - ^)V ■ dx, j,=^_^ + |'+C. 3. dy = 3(2j? + S)3? • dx, y = ^2?f-±-^+C. dx - , dx ^ , X ' dx 7. dy = Va + hx' • dx, Form 42. dx 8. di/ = ■ , , Form 45. VaH- hx 9. dy = X Va2 - a;^ • c?x Form 49. 3/ * dx 10. di/ = ■ , Form 53. Sa^ — x^ dx 11. dy =—-==, Form 73. ^<2ax-x^ 12. dy = a' ' dx, Form 4, 13.%=(e- + a- + 36-)dx, , = ^^+^_^^+C. 15. dy — sin X cos x ' dx y = — ^ — + C 16. dy = tan a; • dx, Form 7. JVote. — ^Take tan x = . C06 X APPENDIX A. 15 17. dy = cot X • dXj Form 8. -o , . . , , 2qo^^x cos^x , ^ 18. a?/ = sm^ xdx, y = — cos x H ^ ^ f- C. iVote. — Take sin" x = iX — cos^ x)" sin x. ,-, J -1 K J sin^a; 2siD^a; , sin® a; , ^ 19. dy = sm* a; cos^ x - dXy y = —z = 1 ^ h C. 20. dy = sin^ a; • c^a;, Form 20. 1 — cos 2a; iVote. — Put sin^ x = 2 21. c?2/ = cos^ X ' dx, Form 21. - - , fi 7 5a; , sin 2a; sin^ 2a; , 3 sin 4a; , ^ 22. dy = cos' X'dx, 2/ = jg + -j is" "*" "~64~ "^ ^• 23 . (??/ = sin mx sin na; • da;, Form 24. iVote. — Put sin mx sin nx = }^ cos(m — n)a; — 3^ cos (m + n)a;. - . , , J. , tan^ a; tan^ a; ' , ^ 24. ai/ = tan^ x - dx, y = — -. ^r log cos x -^ C. 4: ^ Note. — Put tan" x = tan' x (sec^ x — I). cot^ X 25. dy = cot^ X ' dx, y = x -\- cot x ^ h C. o^ J fi , ^ X 2 cot^ a; cot^ x 26. ai/ = cosec^ x - dx, y = C— cot x i^— ^ — . o 5 iVofe. — Put cosec* x = (1 + cot^ x)''. ofT J J. 1 fi J tan'' x , tan^ a; , tan^ x , ^ 27. dy = tan' a; sec^ x - dx, y = — 1 1 ^ f-C 4 o o iVo^e.—Put sec* x = (1 + tan^ a:)2. v,o J , . , J sec^ X 2 sec^ a; , sec' x , ^ 28. dy = tan^ a; sec' x - dx, y = —^ = 1 ^r — |- C. 7 o o Note. — Put tan^ x = (sec^ x — 1)^ and write tan" x sec' x = (sec* x — 1) sec^x-sec xtanx-dx. 29. dy = -p=^=, 2/ = i sin-i ~ + C, ^- , dx 1 , * " VliP'^l' 2^ = 5 log (5x+V25x2-4) + C. 16 APPENDIX A. ^^' '^'' = 4T25:^' J/ = 2 tan-' 2- + C. ,~ J dx 1 , 2 + 6a; , - 35. ., = ^^ . dx, , = I log (4x-6)+^-?^ '°^ £|+^- 36. Find the length of the arc of the parabola in Fig. 13 on page 24 from x = to x = 10 inches, the value of A; being 2 when X and y are both expressed in inches. Ans. 201 square inches. PROBLEMS GROUP 8. (After rule IV of Art. 67.) Verify the following: 1. Jx smx • dx = — a; cos a; + sin a; + C. Note. — Use rule IV. Try u = em x, dv = x-dx] also try u = x, dv = mnx-dx. 2. fx COS a; • da; = a; sin a; + COS a; + C /x^ x^logx ' dx = -J (log X — I) + C. 4. fx{^' + e-'')dx = I (e^x _ g-sx) _ i (^sx + g-3x) + q. 5. J* log (ax + 6) • cte = - (ax + h) log (ax + &) — a; + C. 6. fa^smx'dx= — x^ cos x + Sx^ sin x + 6x cos x — 6 sin x + C. Note. — Applying rule IV, taking u = x^ and dv = smx'dx, we get an expression involving Jx^ cos x-dx. This expression can again be reduced by APPENDIX A 17 applying rule IV (to the part to be integrated), making w = a;^ and dv » cosx'dx] and so on. 7. fx^ sin 2x ' dx = §(2a;2 - 1) sin 2x - iC^s^ - 3x)cos 2x + C, „ /• „^ . , J e°^(a sin 6x — 6 cos 6a;) , ^ 8. I e«^ sm hx • (ix = — ^^ . , ,, ^ + C. •^ a^ -jr 0^ Note. — Integrate by parts, taking u = e<**, then: /, • I. J e"" C08 hx , a r ^ , , ,^ . e'^'^embx-dx = ^^ ^ hj ^ ^°^ oa;*aa; (1) Integrate by parts, taking u = sin bx, then: /e"* sin bx'dx = / e"* cos bx'dx (2) a a«/ Take the sum of equation (1) multiplied by ¥ and equation (2) multiplied by a^ in order to eliminate the term containing J e°* cos 6a: • dx and the result given above is obtained. By subtracting equation (2) from equation (1), J e"* sin bx'dx can be eliminated and the value of I e"* cos bx-dx obtained. 9 . J e^^ cos 5x ' dx = -^{5 sin 5x + 3 cos 5x) + C. e~2* sin a; • (ia; = ^ (2 sin x + cos ic) + C X 11. fe~^ COS ^' dx = -TY" (2 sin r- — 3 cos ^j + C. PROBLEMS GROUP 9. (Special methods a of Art. 67.) Verify the following: r 5a; + 4 _ /" 3 - c?a; /•2j^ ^•Ja;2 + x-2*'^~Ja;-l"*"Ja; + 2 = 3 log (a; - 1) + 2 log (a; + 2) + C. Note. — A full discussion of the integration of rational fractions is given in Byerly's Integral Calculus. See note on page 86. In this particular case the denominator is equal to (x — 1) (x + 2) and 18 APPENDIX A. 5x4-4 «« + a;-2 may be broken up into what are called partial fractions as follows: &C + 4 _ A B a:* + x-2 X - 1 x-{-2 Clear of fractions and equate coefficients of like powers of x and we find A = 3 and 5=2. 2. f ^^^^l^ -dx = \og(x + i) + \og(x+l) + C. + \ log (x - 2) + C. = ^' + 15x - 6 log (x - 1) + 41 log (x - 2) + C. Note. — \Mienever the degree of the numerator of a rational fraction is gre^ater than or equal to the degree of the denominator, the fraction should be reduced to a mixed quantity. 5-/^^^-'i- = - + glog(--3)-|log(. + 2) + \ log (x + 1) + C. '^ • / (X - IRx' + 1 ) • "^ ° ^°g ^^ ~ ^^ " W^) - I log (x^ + 1) + C. X* Note. — __ ^ - ^ is broken into partial fractions by putting it equal .^ A B , Cx-\-D ^ x-l'^ ix-1)*'^ x« + 1 * «• J i(^^-:^4? • ^^ = - i^^4 + ^^S^^^ + ^- APPENDIX A. 19 PROBLEMS GROUP 10. (Special methods b of Art. 67.) Integrate the following and for answers see specined forms in the table of integrals of Appendix B. 1. /Vo^-^^ • dx Form 48. Note. — Let a: = a sin ^, then dx — a cos d-dd; and we get: (^CL^ — x^-dx = C55irr^. d:c = V2ax-x2 + a vers-^?. J X o, 73. f ^ = - V2aa; - 7^ J X'^2ax — a? ox APPENDIX C A SELECTED LIST OF TREATISES ON MATHEMATICS AND ON THE VARIOUS BRANCHES OF MATHEMATICAL PHYSICS. HISTORY. See A History of Mathemati cs by Florian Cajori. The Macmillan Co., New York, 1894 A good sketch of the history of calculus is given in the article Infinitedmal Calculus in the 9th edition of the Encyclopedia Britannica. Sir David Brewster's Life of Newton j Thomas Constable & Co., Edinburgh, 1855, is well worth reading by anyone who is inter- ested in the history of mathematics. The theory of infinitesimals and the theory of limits have occasioned a very great deal of discussion in the past. These matters are discussed somewhat at length in the Britannica article above mentioned, but the most edifying discussion of this subject is that which is given in the introductory chapter to Augustus De Morgan's Differential and Integral Calculus, London, 1842. GENERAL TREATISES. An extremely interesting and facinating book for the student of mathematics is W. K. Cliff ord's* Common Sense _oi the Ejcact Sciences, The International Scientific Series, D. Appleton & Co., New York, 1888. One of the best general treatises is Edouard_Gou.rsat's Cot^rs d^ Analyse M athematique , two volumes, Gauthier-Villars, Paris, 1905. This valuable work has been translated into English by *W. K. Clifford's philosophical essays (two volumes, Macmillan & Co., London, 1879) and especially his small book entitled Seeing and Thinking (Macmillan & Co., London, 1880) are extremely readable and interesting. 29 30 APPENDIX C. E. R. Hedrick and it is published as Goursat-Hedrick Mathe- matical Analysis, Ginn & Co., Boston. The various mathematical articles in the ninth and eleventh editions of the Encyclopedia Britannica are a great help to the student of mathematics. The great reference work in pure and applied mathematics is the Encyclopddie der mathematischen Wissenschaften, B. G. Teub- ner, Leipsig. The publication of this encyclopedia was begun in 1905 and it is not yet finished. In the meantime the work is being republished in a revised and enlarged form in a French translation. The most extensive general treatise on mathematical physics is Winkelman^s Handbuch der Physik in six volumes, second edition, J. A. Earth, Leipsig, 1909. THEORY OF FUNCTIONS. The theory of functions is one of the most facinating branches of pure mathematics. A good discussion of this subject is given in the second volume of Goursat's Cours d^ Analyse Mathematique (Goursat-Hedrick Mathematical Analysis). One of the best books on the subject is H. Durdge's Elements of the Theory of Functions (English trans- lation by Fisher and Schwatt), Fisher and Schwatt, Philadelphia, 1896. Two very extensive treatises on the function theory are A. R. Forsythe's Theory of Functions of a Complex Variable, Cambridge University Press, 1893; and Harkness and Morley^s Theory of Functions^ Macmillan & Co., London, 1893. DIFFERENTIAL EQUATIONS A very good elementary treatise is D. A. Mur ray ^s IntrodmAory Cours&JtL^DiJlerential Equ ations , Longmans Green & Co., New York, 1897. See also W. W. Johnson ' s Differential Eq uations^ John Wiley & Sons, New York, 1890. The second volume of APPENDIX C. 31 Goursat-Hedrick's Mathematical Analysis contains a good dis- cussion of differential equations. A very good treatise on partial differential equations is W. E. Byerly^s treatise entitled Fou rier's Series and Spherical, C ylindrical and Ellipsoidal Harmonics, Ginn & Co., Boston, 1893. An important general method of solving linear differential equations is given by P. A. Lambert in the Annals of Mathematics. First paper in Vol. XI, pages 185-192, July, 1910; second paper, Vol. XIII, pages 1-10, September, 1911. QUATERNIONS AND VECTOR ANALYSIS. (See page 211.) HYPER-GEOMETRY AND RELATIVITY. A very interesting development of mathematics is what is called non-Euclidean geometry or hyper-geometry. Perhaps the most interesting thing foi a beginner to read on this subject is Helmholtz's popular lecture entitled **Ueber den Ursprung und die Bedeutung der geometrischen Axiome" Vortrdge und Reden, Vol. II, F. Vieweg & Son, Braunschweig, 1884. These very inter- esting lectures of Helmholtz have been translated into English, first series by Dr. Pye-Smith, second series by E. Atkinson, and both series are published by Longmans, Green & Co. See section VI of article Geometry in the 11th edition of the Encyclopedia Britannica. A good discussion of non-Euclidean geometry is given in A. N. Whitehe ad's Universal Alg ebra, Cam- bridge University Press, 1898. See also Stackel und Engel. TMm:ie_jderParallellinim von Eu klid his auf Gauss, Leipsig. 1895; F. Klein, Nicht Euklidi sche Geometric, Gottingen, 1893; and P. Barbarin, La Geometric non-Euclidienne, Paris, 1902. The subject of non-Euclidean geometry is closely related to a very recent development in theoretical physics called the theory or relativity. A very simple discussion of this subject is given by W. S. Franklin in an article entitled The Principle of Relativity, Journal of the Franklin Institute, July, 1911. A very interesting article on this subject is the eighth lecture in Max Planck's Acht 32 APPENDIX C. Volemngen uher theoretische Physik, S. Hirzel, Leipsig, 1910. The following articles by Q, N. Lewis^ R. C. Tolm an and E. B. Wilson^ are interesting, and all but the last one are easy to read. Philo- sophical Magazine, Vol. 16, pages 705-717; American Academy Pro- ceedings, Vol. 44, pages 711-724; American Academy Proceedings, Vol. 48, pages 389-507. A general treatise is M. Laue's Das Relativi- tdtsprimip, Braunschweig, F. Vieweg und Sohn, second edition, 1913. ASTRONOMY.* Most of the recent researches in astronomy have been in the line of what is called astro-physics, and but few additions have been made to the older theoretical astronomy with its beautifully complete mathematical theory. Good books on theoretical astronomy for the beginner are C. A . Young's General Astronomy, Ginn & Co., Boston, 1898; and W. W » Camp bei rs Elements of Practical Astronomy , The Macmillan Co., New York, 1899. Perhaps the best treatise on astronomical measurements is Wm.. Qhauvenet's Syherical and Prac tical Astronomy , two volumes, J. B. Lippincott, Philadelphia, 1863. For a treatise on astro- nomical calculations see Jas. C. W atson's The oretical Astronomy , J. B. Lippincott, Philadelphia, 1867. See also ^ahnbes timmung der_ Komet en und Pla neten, T. R. v. Oppolzer, two volumes, Wm. Engelmann, Leipsig, 1882. Perhaps the greatest treatise on theoretical astronomy is Traiti. d&.JLechanique Cileste, P. S. Laplace, five volumes, Paris, 1799. English translation by Nathaniel Bowditch, Boston, 1829. Theoretical astronomy goes far beyond every other branch of * Although we are here concerned chiefly with books on mathematical theory it is worth while, perhaps, to mention some of the extremely interesting descriptive books on astronomy: ^imon New co mb^s Popular Astronomy f Harper Bros., New York, 1882. J. Norm an Lockyer' s Star gazin g, Macmillan & Co., London, 1878. Richard A. Proctor's Other Worlds tha n Ours , Longmans, Green & Co., London, 1878. Also C. A. Young's General Astronomy is ar- ranged so that the general reader can easily omit the more difficult parts and the book then becomes a very good descriptive treatise. APPENDIX C. 33 applied mathematics in the legitimate use of elaborate formulas and in the tremendously laborious numerical calculations that are involved. The student can get an idea of the former and a faint suggestion of the latter by looking over the four large volumes of The Collected Mathematical Works of George W. Hill which have been recently pubhshed (1905) by The Carnegie Institution of Washington and placed in every important library in the United States. Dr. Hill's greatest contribution to theoretical astronomy was a new method for calculating the motion of the moon. PROBABILITY. In many respects the theory of probability is the most important branch of mathematics for the experimental scientist. The article Prdbahility in the eleventh edition of the Encyclopedia Britannica is a very good outline of the theory of probability. See also De Morgan's treatise on probabiHty in the Cabinet Cyclopedia, London, 1838. Laplace's Theorie analytiq ue des Prohahilites^Vsins, 1820; is one of the great treatises on the subject. The application of the theory of probability to measurement is treated in Mansfi eld Mer riman's Least Squares, John Wiley & Sons, New York, eighth edition, 1903; in R. S. Woodward's Prdbahility and Theory of Errors, John Wiley & Sons, New York, 1906; and in A. de F. Palmer's Theory of Measurements, McGraw- Hill Book Co., New York, 1912. The application of the theory of probability to the study of statistics of all kinds and especially to biometrics (the quantitative study of variation of plants and animals) is of great importance. A simple discussion of this subject is given in the article Variation and Selection in the eleventh edition of the Encyclopedia Britannica. The kinetic theory of gases, one of the most important branches of theoretical physics is a phase of the theory of probability. See Watson's Kinetic Theory of Gases, The Clarendon Press, Oxford, 1893. See also the epoch-making work, T^dwig "Rn1t7mfl.Tm\q Ymksungm^uh^.GaMMQrie^iT^ two parts, J. A. Barth, Leipsig, 1895 and 1898. The kinetic theory of gases is treated in a very 34 APPENDIX C. general way by J. Willard Gi bbs in his Stat istical MechanicSf Charles Scribner's Sons, New York, 1902. A very good discussion of the kinetic theory of gases is given in Nernst's Theoretical Chemistry, Chapter II, translated by Lehfeldt, Macmillan & Co., London, 1904. The ideas of the kinetic theory of gases are extensively used in the recent theories of the discharge of electricity through gases and in the theories of radioactivity. See especially J. J. Thomson's Conduction of Electricity through Gases, Cambridge University Press, 1906. See also E. Rutherford's Radioactive Transforma- tions, Charles Scribner's Sons, New York, 1906. The modem statistical theory of radiant heat is also a branch of the theory of probability. Some idea of this subject can be obtained from lecture 6 of Acht Vorlesungen iiher theoretische Physik, Max Planck, S. Hirgel, Leipsig, 1910. THERMODYNAMICS. There is a widespread notion that theoretical thermodynamics makes a very severe demand upon the methods of higher mathe- matics, whereas, as a matter of fact its demands are less perhaps than any other branch of mathematical physics. An extremely simple development of the mathematical theory of thermodynamics is given in Franklin and MacNutt's Mechanics and Heat, pages 350-397, The Macmillan Co., 1910. A very good advanced treatise is Max Planck's Treatise on Ther modynamics (English translation by Alexander Ogg), Long- mans Green & Co., London, 1903. Another important treatise is Edgar Buckin gham's Theo ry o f Th ermodynamics ^ The Macmillan Co., New York, 1900. R. Clau sius' Mec hanische Wdrmetheorie , third edition, F. Vieweg & Sons, Braunschweig, 1887, is a very important work. The student interested in theoretical thermodynamics will find it well worth while to read the celebrated paper of J. Van't Hoff entitled "The function of osmotic pressure in the analogy between solutions and gases," Philosophical Magazine, Vol. XXVI, pages 81-105, August, 1888. APPENDIX C. 35 The best treatment of thermodynamics for the experimental chemist is that which is given in a more or less disconnected manner in W. Nernst's Theoretical Chemistry , English translation by R. A. Lehfeldt, Macmillan & Co., London, 1904. A good treatment of thermodynamics for the steam engineer is that which is given in chapters II and III (pages 37-159) of JL_A,_-Ewing' s The Steam Engin e _ and othe r Heat Engines ^ third edition, Cambridge University Press, 1910. THEORETICAL MECHANICS. . There are many good beginner's treatises on theoretical me- chanics. A good advanced treatise is Alexander Ziwet's Ele- mentary Treatise on Theoretical Mechanics, in two parts, The Mac- millan Co., New York, 1893. Every student of mathematical physics should read a portion, at least, of Sir Is aac Newton's Pri ndp ia. An extremely interesting and readable book is Poinso t's Theorie Nouvelle de la Rotation des Corps , second edition, Paris, 1851. Harold Crabtree's Spinning Tops, Longmans Green & Co., London, 1909 (about), is a good mathematical treatment of the theory of the rotation of a rigid body. The most exhaustive treatises on theoretical mechanics are; Geo. M. Minchin's Treatise on Statics, third edition, two volumes. The Clarendon Press, Oxford, 1886; Ed ward J. Rout h's Ele- mentary Rigid Dynamics, Macmillan & Co., London, 1860; and Edward J. Routh's Advanced Rigid Dynamics, Macmillan & Co., London, 1860. See lecture 7 in Acht Vorlesungen uber theoretische Physik by Max Planck, S. Hirzel, Leipsig, 1910. This lecture deals with the most fundamental principle of physics, the so-called principle of least action. THEORY OF SOUND. ♦ One of the best books on this subject for the beginner is J. H. Poynting and J. J. Thomson's Sound, Charles Griffin & Co., London, 1899. See also the article Acoustics in the 9th edition 36 APPENDIX C. of the Encyclopedia Britannica. Helmholtz^s great work The Sensations of Tone (English translation by Alexander J. Ellis), contains a series of appendices on mathematical theory. The most comprehensive treatise is L ord R ayleigh^s Theory of Sound in two volumes, Macmillan & Co., London, 1877. Second edition revised and enlarged 1894. THEORY OF ELECTRICITY AND MAGNETISM. An understanding of vector analysis, as developed in chapter IX of this text, is absolutely necessary before one can begin the study of Maxwell's theory of electricity and magnetism. A very simple discussion of Maxwell's theory is given in Frank - lin's Electric Waves, pages 186-196, The Macmillan Co., New York, 1909. The beginner will be greatly helped by E. Atkinson's translation of Mascart and Joubert's Treatise on Electricity and Magneti sm, Vol. I, Thos. de la Rue & Co., London, 1883. The great treatise on this subject is Maxwell' s original treatise in two volumes entitled Electricity and Magnetism. The first edition of this work appeared in 1873. Third, edition very slightly altered, The Clarendon Press, 1891. The study of Maxwell's treatise is greatly facilitated by Mascart and Joubert's treatise above mentioned ; by the study of Heinrich Hertz's Ele ctric Waves , translated by D. E. Jones, Macmillan & Co., London, 1893; and by the study of A. G. Webster's Electricity and Magnetism , Macmillan & Co., London, 1897. A most excellent treatise for the student is Abr aham and Fop pl's Theor ie de r Ehctrizitdt; Vol. I, Einfuhrung in die Max- wellsche Theorie, 3d edition, Leipsig, 1907; Vol. II, Electromag- neschie Theorie der Strahlung (Electronentheorie) , Leipsig, 1905. Recen t Researches Electricity and Magnetism by J. J. Thomson, The Clarendon Press, Oxford, 1893, contains a great deal of interest especially on the subject of electric waves. See also J. J . Thomson's _ Cowc^t^c^ion qf_ Electricity thr ough Gases, Cambridge University Press, 1906. This important book deals principally with the electron theory, although, of course, the book deals almost entirely with experimental researches. APPENDIX C. 37 THE THEORY OF LIGHT. One of the best books for the student is Thomas P rest on's Theory of Lig hts third edition, Macmillan & Co., London, 1901. An extremely interesting book, partly theoretical and partly descriptive, is A. A. Michelson's Ligh t Waves and th eir U seSf Uni- versity of Chicago Press, 1903. R. W . Wood's PhydcaLQpiic&f second edition. The Macmillan Co., New York, 1910, contains a great deal of interesting and important theory. The theory of lenses and optical instruments is treated in a very exhaustive manner by Czapski, von Rohr and Eppfenstein in Winkelmann's Handbuch der Phydk. One of the most interesting series of original memoirs is that of Augustin Fresnel, see Fresnel's Oevres Completes^ Vol. I, pages 1-382, Paris, 1866. A very complete treatise on the theory of light (electro- magnetic) is that of P. Prude ; Enghsh translation by C. R. Mann and R. A. MiUikan, Longmans Green & Co., New York, 1902. HYDROMECHANICS. One of the best treatises on this subject for the student is the article Hydromechanics in the ninth edition of the Encyclopedia Britannica. Part I of this article is devoted to Hydrostatics. In this part the problem of the figure of the earth is discussed and also the important practical problem of the stability of floating bodies. Part II of this article is devoted to Hydrodynamics, the highly mathematical theory of the motion of a frictionless fluid. Part III of this article is devoted to Hydraulics from the point of view of the experimental physicist and the engineer. Parts I and II were written by A. G. Greenhill and Part III was written by W. C. Unwin. Professor Unwin's article has been pubUshed as a separate treatise, Macmillan & Co., London, 1907. See also G. M. Minchin's Treatise on Hydrostatics, two volumes, second edition, Clarendon Press, Oxford, 1912; Horace Lamb's Hydrodynamics I Cambridge University Press, 1879; third edition 1906; and A. B. Bassett's Hydrodynamics, two volumes, Deighton Bell & Co., Cambridge, 1888. 20 38 APPENDIX C. THEORY OF ELASTICITY. An understanding of vector analysis as developed in chapter IX of this text is very helpful in the study of the theory of elasticity. Thus the ideas which are established in Art. 136 are the foundation of the elementary theory of elasticity as given ii; Franklin and MacNutt's Mechanics and Heat, pages 182-218, The Macmillan Co., New York, 1910. This is perhaps the simplest existing elementary treatise on the theory of elasticity. The matter which is discussed in Art. 136, namely, the theory of elastic strains, is discussed in W. K. Clifford's Kinematic, part I, Macmillan & Co., London, 1878, pages 158-221. One of the best treatises on the theory of elasticity is W. J. Ibbetson's Mathematical Theory of Elasticity, Macmillan & Co., London, 1887; see also A. E. H. L ove's Mathematical Theory of Elasticity, two volumes, Cambridge University Press, 1892. The greatest reference work in the theory of elasticity is Karl Pearson's History. CRYSTALLOGRAPHY. It is not generally known that one of the most interesting and remarkable branches of mathematical physics is the theory of crystallography. Thus the purely mathematical theory of regu- lar-point-systems in space is in complete accord with experi- mental studies of crystal forms. A regular-point-system in space is called a space lattice. The purely mathematical theory of the space lattice is treated in Sohncke's Theorie der Krystallstruktur, B. G. Tuebner, Leipsig, 1879. Very complete treatises on crystallography are Groth's Physikalische Krystallographie, Wm. Englemann, Leipsig, 1895; and Liebisch's Grundriss der Physi- kalischen Krystallographie, Veit & Co., Leipsig, 1896. In both of these books the geometrical theory of crystallography is fully treated. INDEX Acceleration and velocity, 55 in circular motion, 70 Algebraic integration, 82 Angular acceleration and torque, 146 Arbitrary variations, 9 Area under a curve, 28 Artificial functions and natural func- tions, 15 Average, discussion of, 121 value of a function, 121 Barrel hoop, discussion of, 72 Beam, the problem of the bent, 116 Bent beam, problem of, 115 Boat, starting of a, 177 stopping of a, 178 Center of gravity, 123 of mass. See center of gravity, of pressure, 134 Change, rate of, 2 Circle, the osculating, 66 Complex quantity, definition of, 163 geometrical representation of, 163 use of, 153 Component slopes and resultant slopes, 96 Constant of integration, 33 Constant quantities and variable quantities, 1 Continuous variables and discon- tinuous variables, 4 Convergence of a vector field, 229 Convergent series, 43 Cosines, hyperbolic, 166 Curl, cartesian expression for, 238 divergence of, 240 example of, 237 of a gradient, 239 of a vector field, 236 Curvature, 61 radius of, 65 Cyclometer, the, 74 Dam, force exerted upon, 136 Damped oscillations, 185 Decrements and increments, 2 Demoivre's theorem, 162 Dependent variables, 13 Derivative, Graphical representation of a, 18 of a function, 17 Derivatives, successive, 54 Definite integrals and indefinite in- tegrals, 34 Differential equation, 31 and 168 Degree and order of, 168 general solutions and par- ticular solutions of, 172 linear, 172 of wave motion, 190 ordinary and partial, 168 partial, 89 pure and mixed, 169 Differentiation and integration, 32 formulas for, 38 by development, 38 by rule, 38 geometric, 70 graphical, 38 of an exponential function, 48 of a function of a function, 42 of an implicit function, 91 of a logarithm, 46 of a product, 40 of a quotient, 41 of a sum, 39 of trigonometric functions, 50 partial, 87 successive, 88 physical, 38 rules for, 38-54 successive, 54 Discontinuous variables and con- tinuous variables, 4 Distributed scalar, 218 vector, 218 Divergence, Cartesian expression for, 229 of a curl, 240 of a vector field, 228 40 INDEX. Equation, differential, partial, 89 Errors of observation, influence upon a result, 109 Euler's expression for sine and cosine, 164 Expansions in series, 153 Fields, scalar and vector, 217, 218 Forced oscillations, 188 Fourier's theorem, 199 application of, 204 Function, definition of, 13 derivative of a, 17 graphical representation of a, 16 implicit differentiation of, 91 law of growth of, 31 tabulation of, 15 Functions, natural and artificial, 15 which have the same derivative, 25 Gravity, center of, 123 Geometric differentiation, 70 Grade, definition of, 6 Gradient, curl of, 239 definition of, 6 line integral of, 223 of a distributed scalar, 219 of a hill, 93 temperature, discussion of, 7 Gradients, examples of, 98 Graphical representation of a deriva- tive, 18 of a function, 16 Gyration, radius of, 142 Harmonic motion, 58 Hill, gradient of a, 93 slope of a, 93 surface, area of, 105 volume of a, lOO Hooke's law, 115 Hoop, discussion of, 72 HyperboUc sines and cosines, 166 Implicit function, differentiation of, 91 Increments and decrements, 2 Indefinite integrals and definite inte- grals, 34 Independent variables, 13 Inertia, moment of. 140 moments of, about parallel axes, 147 Inertia, moments of, table of, 151 Infinitesimals, method of, 22 Integrals, definite and indefinite, 34 table of. See Appendix B. Integrating machine, 74 Integration, 74 algebraic, 82 and differentiation, 32 formulas for, 38 by rules. See footnote, page 83. by steps, 79 constant of, 33 mechanical, 74 multiple. See Integration, Par- tial. partial, 87 and 99 rules for, 84 Irrotational vector fields, 242 Kelvm's law, 113 Line integral, Cartesian expression for, 223 of a gradient, 223 of a vector field, 222 reduction of to a surface in- tegral, 239 Linear vector field, the, 243 Maclaurin's theorem, 153 appHed to the function of two variables, 159 Mass, center of. See Center of gravity. Maximum and minimum values, 111 Mechanical integration, 74 Median line of beam, definition of, 117 Method of infinitesimal, the, 22 Minimum and maximum values. 111 Moment of inertia, 140 of section of a beam, 148 of inertia about parallel axes, 147 Table of, 151 Motion, harmonic, 58 in a circle, 70 Multiple integration. See Partial Integration. Natural functions and artificial func- tions, 15 Oscillations, damped, 185 forced, 188 undamped, 181 Osculating circle, 66 INDEX. 41 Partial differential equation, 89 differentiation, 87 successive, 88 integration, 87 and 99 Planimeter, the, 74 Plucked string, the vibration of, 197 Potential, existence theorem of, 225 multivalued, 226 of a vector field, 225 scalar, 225 and vector, 242 Pressure, center of, 134 Principle of superposition, 174 Prismoid formula, the, 102 Quantities, constant and variable, 1 Radius of curvature, 65 of gyration, 142 Rate of change, 2 Resultant slopes and component slopes, 96 Rotational vector fields, 242 Scalar and vector quantities, 211 field, 217 gradient of, 219 volume integral of, 219 function, 218 potential, 225 and vector potential, 242 Series, convergent, 43 Sines, hyperbolic, 166 Slope of a hill, 93 Slopes, component and resultant, 96 Solenoidal vector fields, 235 Space analysis, 210 Speed-time curve, the, 79 Spring, work required to stretch a, 26 Starting of a boat, 177 Stopping of a boat, 178 Stream lines, 221 Stretched string, equation of motion of, 193 String, equation of motion of, 193 plucked, vibration of, 197 Successive derivatives, 54 differentiation, 54 partial differentiation, 88 Superposition, the principle of, 174 Surface integral, Cartesian expression for, 228 of distributed vector, 227 Surface integral, reduction of to line mtegral, 239 of to volume integral, 234 of a hill, area of, 105 Table of moments of inertia, 151 Tabulation of a function, 15 Tangent plane, equation of, 94 Taylor's theorem, 157 Temperature gradient, discussion of, 7 Torque and angular acceleration, 146 Tube of flow in a vector field, 235 Undamped oscillations, 181 Unit tube in a vector field, 235 Variable and constant quantities, 1 Variables, continuous and discon- tinuous, 4 dependent and independent, 13 Variations, arbitrary, 9 Vibrations, damped, 185 forced, 188 undamped, 181 Vector analysis, 210 and scalar quantities, 211 field, 218 convergence of, 229 curl of, 236 divergence of, 228 line integral of, 222 potential of, 225 rotational and irrotational, 242 solenoidal, 235 surface integral of, 227 the linear, 243 function, 218 potential, 241 and scalar potential, 242 products, 213 Velocity and acceleration, 55 Volume integral of scalar field, 219 reduction of to surface in- tegral, 234 of a hill, 100 Vortex motion, 241 Water gate, force exerted upon, 134 Watt-hour-meter, the, 74 Wave motion, differential equation of, 190 14 DAY USE RSTUKN TO DESK FROM WHICH BORROWED LOAN DEPT. This book is due on the last date stamped below, or on the date to which renewed. Renewed books are subject to immediate recall. REC'D L-u UCT 2^1960 m \i »y*>^ l71uP62lE REC'D LD JUN G 4961 JUL 3 1962 ^ 1, 24Se)r'6?«!.': it^ ' ^ >0^ QU LD <^ ^ RECP LD 4^kii!i REC'D LP mmk V mi OCT 2 7 '63 -3 PM LIBRARY USE ^«Wl*l LD 21A-50m-4.'60 (A95628l0)476B LD 21-100m-6,'56 (B931l8l0)476 General Library University of California Berkeley General Library University of California Berkeley •VS! 4r - . ^■y^z^'f Q/9 3^3 .. :^^y'- UNIVERSITY OF CALIFORNIA UBRARY * w Kt