MATHEMATICS AND THE IMAGINATION MATHEMATICS and the IMAGINATION Edward Kasner and James Newman With Drawings and Diagrams by Rufus Isaacs LONDON; G. BELL AND SONS, LTD ALL RIGHTS RESERVED INCLUDING THE RIGHT OF REPRODUCTION IN WHOLE OR IN PART IN ANY FORM 33247 British Edition first published 194^ ReprinUd /p 5 o, 1952, 1956, 19S9 PRINTED IN GREAT BRITAIN BY JARROLD AND SONS LIMITED, NORWICH Xo R.G. without whose selfless help and understanding there would have been no book. ACKNOWLEDGMENT We are indebted to many books, too many to enumerate. Some of them are listed in the selected bibliography. And we wish to acknowledge particularly the services of Mr. Don Mittleman of Columbia University, whose help in prepa¬ ration of the manuscript has been generous and invaluable. Table of Contents Introduction I. NEW NAMES FOR OLD Easy words Jot hard ideas . . . Transcendental . . . Non- simple curve . . . Simple curve . . . Simple group . . . Bolshe¬ viks and giraffes . . . Turbines. . . Turns and slides . . . Circles and cycles . . . Patho-circles . . . Clocks . . . Hexagons and parhexagons . . . Radicals, hyperradicals, and ultrarad¬ icals {nonpoliticaf) . . . New numbers for the nursery . . . Googol and googolplex . . . Miracle of the rising book . . . The mathescope. II. BEYOND THE GOOGOL Counting—the language of number . . . Counting, matching, and '‘"‘Going to Jerusalem'" . . . Cardinal numbers . . . Cosmic chess and googols . . . The sand reckoner . . . Mathematical induction . . . The infinite and its progeny . . . ^eno . . . Puz¬ zles and quarrels . . . Bolzano . . . Galileo''s puzzle . . . Cantor . . . Measuring the measuring rod. . . The whole is no greater than some of its parts ... The first transfinite— Alepho. . . Arithmetic Jot morons . . . Common sense hits a snag . . . Cardinality oj the continuum , . . Extravagances oj a mathematical madman . . . The tortoise unmasked . . . Motionless motion ... Private life oJ a number ... The house that Cantor built. III. TT, i, e (pie) Chinamen and chandeliers . .. Twilight oJ common sense . . . TT, i, e . . . Squaring the circle and its cousins . . . Mathe¬ matical impossibility . . . Silk purse, sow''s ear, ruler and compass . . , Rigor mortis . . . Algebraic equations and tran- vii Vlll Contents scendental numbers . . . Galois and Greek epidemics . . . Cube duplicators and angle trisectors . . . Biography of tt . . . Infancy: Archimedes^ the Bible^ the Egyptians . . . Adoles¬ cence: Vieta, Van Ceulen . . . Maturity: Wallis^ Newton, Leibniz • • • Old Age: Base, Richter, Shanks . . . Victim of schizophrenia . . . Boon to insurance companies . . . (e) . . . Logarithms or tricks of the trade . . . Mr. Briggs is surprised . . . Mr. Napier explains . . . Biography oft; or e, the bank¬ er's boon . . . Pituitary gland of mathematics: the exponential function . . . (i) . . . Humpty Dumpty, Doctor of Semantics . . . Imaginary numbers . . . The y/- 1, or Where am /?” • • • Biography of i, the self-made amphibian . . . Omar Khayyam, Cardan, Bombelli, and Gauss . . . i and Soviet Russia . . . Program music of mathematics . . . Breakfast in bed; or, How to become a great mathematician . . . Analytic geometry . . . Geometric representation of \. . . Complex plane . . . A famous formula, faith, and humility. IV. ASSORTED GEOMETRIES—PLANE AND FANCY The talking fish and St. Augustine ... A new alphabet . . . High priests and mumbo jumbo . . . Pure and applied mathe¬ matics . . . Euclid and Texas . . . Mathematical tailors . . . Geometry a game . . . Ghosts, table-tipping, and the land of the dead . . . Fourth-dimension flounders . . . Henry More to the rescue . . . Fourth dimension—a new gusher . . . A cure for arthritis . . . Syntax suffers a setback . . . The physicisCs delight ... Dimensions and manifolds . . . Distance formulae ■ - . Scaling blank walls . . . Four-dimensional geometry de¬ fined , . . Moles and tesseracts . . . A four-dimensional fancy . Romance of flatland . . . Three-dimensional cats and two- dimensional kings . . . Gallant Gulliver and the gloves . . . Beguiling voices and strange footprints . . . Non-Euclidean geometry . . . Space credos and millinery . . . Private and public space . . . Rewriting our textbooks ...The prince and the Bofthiam ...The flexible flfth ... The mathematicians unite nothing to lose but their chains . . . Lobachevsky breaks u link . . . Riemann breaks another . . . Checks and double checks in nrathematics ...The tractnx and the psendosphere ■ . . Great circles and bears ...The skeptic persists-and ts I Contents IX stepped on . . . Geodesics. . . Seventh Day Adventists . . . Curvature . . . Lobachevskian Eiffel Towers and Riemannian Holland Tunnels. V. PASTIMES OF PAST AND PRESENT TIMES 1 56 Puzzle acorns and mathematical oaks . . . Charlemagne and crossword puzzles . . . Mark Twain and the “farmer's daugh¬ ter" . . . The syntax of puzzles . . . Carolyn Flaubert and the cabin boy . . . A wolf^ a goat, and a head of cabbage . . . Brides and cuckolds . . . Til be switched . . . Poisson, the mis¬ fit.. . High finance; or. The international beer wolf . . . Lions and poker players . . . The decimal system . . . Casting out nines . . . Buddha, God, and the binary scale . . . The march of culture; or, Russia, the home of the binary system . . . The Chinese rings . . . The tower of Hanoi. . . The ritual of Benares: or, Charley horse in the Orient. . . Nim, Sissa Ben Dahir, and Josephus . . . Bismarck plays the boss . . . The 15 puzzle plague . . . The spider and the fly ... A nightmare of relatives . . . The magic square . . . Take a number from 1 to 10 . . . Fermat's last theorem . . . Mathematics' lost legacy. VI. PARADOX LOST AND PARADOX REGAINED 193 Great paradoxes and distant relatives . . . Three species of par¬ adox . . . Paradoxes strange but true . . . Wheels that move faster on top than on bottom . . . The cycloid family . . . The curse of transportation; or. How locomotives can't make up their minds . . . Reformation of geometry . . . Ensuing troubles . . . Point sets—the Arabian Nights of mathematics . . . Hausdorff spins a tall tale . . . Messrs. Banach and Tarski rub the magic lamp . . . Baron Munchhausen is stymied by a pea . . . Mathematical fallacies . . . Trouble from a bubble; or, Dividing by zero . . . The infinite—troublemaker par ex¬ cellence . . . Geometrical fallacies . . . Logical paradoxes—the folk tales of mathematics . . . Deluding dialectics of the poacher and the prince; of the introspective barber; of the number 111777; of this book and Confucius; of the Hon. Bertrand Russell. . . Scylla and Charybdis; or, What shall poor mathe- mathics do? X Contents VII. CHANCE AND CHANCEABILITY 223 The clue of the billiard cue ... A little chalk, a lot of talk . . . Watson gets his leg pulled by probable inference . . . Finds it all absurdly simple . . . Passionate oysters, waltzing ducks, and the syllogism . . . The twilight of probability . . . Inter^ esting behavior of a modest coin . . . Biological necessity and a pair of dice . . . What is probability? ... A poll of views: a meteorologist, a bootlegger, a bridge player . . . The subjective view—based on insufficient reason, contains an element of truth . . . The jackasses on Afars . . . The statistical view . . . What happens will probably happen . . . Experimental euryth- mics; or, Pitching pennies . . . Relative frequencies . . . The adventure of the dancing men . . . Scheherezode and John Wilkes Booth—a challenge to statistics . . . The red and the black . . . Charles Peirce predicts the weather . . . How far is ^''away'’'’? . . . Herodotus explains . . . The calculus of chance ... The benefits of gambling . . . De Mere and Pascal . . . Mr. Jevons omits an acknowledgment . . . The study of craps the very guide of life . . . Dice, pennies, permutations, and combinations . . . Measuring probabilities . . . D'Alembert drops the ball . . . Count Buffon plays with a needle . . . The point ... A black ball and a white ball ... The binomial theorem . . . The calculus of probability re-examined . . . Found to rest on hypothesis . . . Laplace needs no hypothesis . . . Twits Napoleon, who does . . . The Marquis de Condorcet has high hopes . . . M.le Marquis omits a factor and loses his head . . . Fourier of the Old Guard ... Dr. Darwin of the New . . . The syllogism scraps a standby . . . Mr. Socrates may not die .. . Ring out the old logic, ring in the new. VIII. RUBBER-SHEET GEOMETRY 265 Seven bridges over a stein of beer .. . Euler shivers . . . Is warmed by news from home . . . Invents topology . . . Dis¬ solves the dilemma of Sunday strollers . . . Babies' cribs and Pythagoreans . . . Taltsmen and queer figures . . . Position is everything m topology ...Da Vinci' and Dali . . . Invariants . ..Transformations ...The immutable derby . . . Com¬ petition for the cahpk's cup; or. Sifting out the suitors by Contents XI science . . . Mr. Jordan's theorem . . . Only seems idiotic . . . Dejormed circles . . . Odd facts concerning Times Square and a balloonist's head . . . Eccentric deportment of several dis¬ tinguished gentlemen at Princeton . . . Their passion for pret¬ zels . . . Their delving in doughnuts . . . Enforced modesty of readers and authors . . . The ring . . . Lachrymose recital around a Pans pissoir . . . “ Who staggered how many times around the walls of what?" . . . In and out the doughnut . . . Gastric surgery—from doughnut to sausage in a single cut . . . ^/-dimensional pretzels . . . The Mobius strip . . . Just as black as it is painted . . . Foments industrial discontent . . . Never takes sides . . . Bane of painter and paintpot alike . . . The iron rings . . . Mathematical cotillion; or., How on earth do I get rid of my partner? . . . Topology—the pinnacle of perversity; or, Removingyour vest without your coat . . . Down to earth—map coloring . . . Four-color problem . . . Euler's theorem . . . The simplest universal law . . . Brouwer's puz¬ zle .. . The search for invariants. IX. CHANGE AND CHANGEABILITY 299 The calculus and cement . . . Meaning of change and rate of change . . . ^eno and the movies . . . '‘‘‘Flying Arrow" local — stops at all points . . . Geometry and genetics . . . The arith¬ metic men dig pits . . . Lamentable analogue of the boomerang . . . History of the calculus . . . Kepler . . . Fermat . . . Story of the great rectangle . . . Newton and Leibniz • • • Archimedes and the limit . . . Shrinking and swelling; or, “ Will the circle go the limit?" . . . Brief dictionary of mathe¬ matics and physics . . . Military idyll; or. The speed of the falling bomb . . . The calculus at work . . . The derivative . . . Higher derivatives and radius of curvature . . . Laudable scholarship of automobile engineers . . . The third derivative as a shock absorber . . . The derivative finds its male . . . Integration . . . Kepler and the bungholes . . . Measuring lengths; or. The yawning regress . . . Methods of approx¬ imation . . . Measuring areas under curves . . . Method of rectangular strips . . . The definite S • • • Indefinite J' . . . One the inverse of the other . . . The outline of history and the descent of man: or, y = e^ . . . Sickly curves and orchidaceous XU Contents ones . . . The snowflake . . . Inflnite perimeters and postage stamps . . . Anti-snowflake . . . Super-colossal pathological specimen—the curve that fills space . . . The unbelieveable crisscross. EPILOGUE. MATHEMATICS AND THE IMAGINATION 357 Introduction The fashion in books in the last decade or so has turned increas¬ ingly to popular science. Even newspapers^ Sunday supplements and magazit^s have given space to relativity^ atomic physics^ and the newest marvels of astronomy and chemistry. Symptomatic as this is of the increasing desire to know what happens in laboratories and observatories, as well as in the awe-inspiring conclaves of scientists and mathematicians, a large part of modern science remains obscured by an apparently impenetrable veil of mystery. The feeling is widely prevalent that science, like magic and alchemy in the Aiiddle Ages, is practiced and can be understood only by a small esoteric group. The mathema¬ tician is still regarded as the hermit who knows little of the ways of life outside his cell, who spends his time compounding incredible and incomprehensible theories in a strange, clipped, unintelligible jargon. Nevertheless, intelligent people, weary of the nervous pace of their own existence—the sharp impact of the happenings of the day—are hungry to learn of the accomplishments of more leisurely, contemplative lives, timed by a slower, more deliberate clock than their own. Science, particularly mathematics, though it seems less practical and less real than the news contained in the latest radio dispatches, appears to be building the one permanent and stable edifice in an age where all others are either crumbling or being blown to bits. This is not to say that science has not also undergone revolutionary changes. But it has happened quietly and honorably. That which is no longer useful has been rejected only after mature deliberation, and the building has been reared steadily on the creative achievements of the past. • • • xm xiv Introduction' ThuSy in a certain sense, the popularization oj science is a duty to be performed, a duty to give courage and comfort to the men and women of good will everywhere who are gradually losing their faith in the life of reason. For most of the sciences the veil of mystery is gradually being torn asunder. Mathematics, in large measure, remains unrevealed. What most popular books on mathematics have tried to do is either to discuss it philosophically, or to make clear the stuff once learned and already forgotten. In this respect our purpose in writing has been somewhat different. ^‘Haute vulgarisation*' is the term applied by the French to that happy result which neither offends by its condescen¬ sion nor leaves obscure in a mass of technical verbiage. It has been our aim to extend the process of '■'■haute vulgarisation" to those outposts of mathematics which ate mentioned, if at all, only in a whisper; which are referred to, if at all, ordy by name; to show by its very diversity something of the character of mathematics, of its bold, untrammeled spirit, of how, as both an art and a science, it has continued to lead the creative faculties beyond even imagination and intuition. In the compass of so brief a volume there can only be snapshots, not portraits. Yet, it is hoped that even in this kaleidoscope there may be a stimulus to further interest in and greater recognition of the proudest queen of the intellectual world. MATHEMATICS AND THE IMAGINATION I tuill not go so Jar as to say that to construct a history of thought without profound study of the mathematical ideas of successive epochs is like omitting Hamlet from the play which is named after him. That would be claiming too much. But it is certainly analogous to cutting out the part of Ophelia. This simile is singularly exact. For Ophelia is quite essential to the play, she is very charm¬ ing,—and a little mad. Let us grant that the pursuit of mathematics is a divine madness of the hutnan spirit, a refuge from the goading urgency of contingent happenings. -ALFRED NORTH WHITEHEAD, Science and the Modern World. New Names for Old For out of aide feldeSy as men seithy Cometh al this newe corn fro yeer to yere; And out of olde bakes, in good feith, Cometh al this newe science that men lere. —CHAUCER Every once in a while there is house cleaning in mathe¬ matics. Some old names are discarded, some dusted off and refurbished; new theories, new additions to the household are assigned a place and name. So what our title really means is new words in mathematics; not new names, but new words, new terms which have in part come to represent new concepts and a reappraisal of old ones in more or less recent mathematics. There are surely plenty of words already in mathematics as well as in other subjects. Indeed, there are so many words that it is even easier than it used to be to speak a great deal and say nothing. It is mostly through words strung together like beads in a necklace that half the population of the world has been induced to believe mad things and to sanctify mad deeds. Frank Vizetelly, the great lexicographer, estimated that there are 800,000 words in use in the English language. But mathematicians, generally quite modest, are not satisfied with these 800,000; let us give them a few more. We can get along without new names until, as we ad¬ vance in science, we acquire new ideas and new forms. 3 2 4 Mathematics and the Imagination A peculiar thing about mathematics is that it does not use so many long and hard names as the other sciences. Besides, it is more conservative than the other sciences in that it clings tenaciously to old words. The terms used by Euclid in his Elements are current in geometry today. But an Ionian physicist would find the terminol¬ ogy of modern physics, to put it colloquially, pure Greek. In chemistry, substances no more complicated than sugar, starch, or alcohol have names like these: Meth- ylpropenylenedihydroxycinnamenylacrylic acid, or, 0- anhydrosulfaminobenzoine, or, protocatcchuicaldehyde- methylene. It would be inconvenient if we had to use such terms in every'day conversation. Who could imagine even the aristocrat of science at the breakfast table asking, Please pass the O-anhydrosulfaminobenzoic acid,” when all he wanted was sugar for his coffee? Biology also has some tantalizing tongue twisters. The purpose of these long words is not to frighten the exoteric, but to describe with scientific curtness what the literary man would take half a page to express. In mathematics there are many easy words like group, “family,” “ring,” “simple curve,” “limit,” etc. But these ordinary words are sometimes given a very peculiar and technical meaning. In fact, here is a booby- prize definition of mathematics: Mathematics is the science which uses easy words for hard ideas. In this it differs from any other science. There are 500,000 known species of insect and every one has a long Latin name. In math¬ ematics we are more modest. We talk about “fields,” “groups,” “families,” “spaces,” although much more meaning is attached to these words than ordinary con¬ versation implies. As its use becomes more and more technical, nobody can guess the mathematical meaning New Names for Old ^ of a word any more than one could guess that a “drug store ’ is a place where they sell ice-cream sodas and umbrellas. No one could guess the meaning of the word “group” as it is used in mathematics. Yet it is so impor¬ tant that whole courses are given on the theory of ‘ groups,” and hundreds of books are written about it. Because mathematicians get along with common words, many amusing ambiguities arise. For instance, the word “function” probably expresses the most important idea in the whole history of mathematics. Yet, most people hearing it would think of a “function” as meaning an evening social affair, while others, less socially minded, would think of their livers. The word “function” has at least a dozen meanings, but few people suspect the mathematical one. The mathematical meaning (which we shall elaborate upon later) is expressed most simply by a table. Such a table gives the relation between two variable quantities when the value of one variable quan¬ tity is determined by the value of the other. Thus, one variable quantity may express the years from 1800 to 1938, and the other, the number of men in the United States wearing handle-bar mustaches; or one variable may express in decibels the amount of noise made by a political speaker, and the other, the blood pressure units of his listeners. You could probably never guess the mean¬ ing of the word “ring” as it has been used in mathematics. It was introduced into the newer algebra within the last twenty years. The theory of rings is much more recent than the theory of groups. It is now found in most of the new books on algebra, and has nothing to do with cither matrimony or bells. Other ordinary words used in mathematics in a pe¬ culiar sense are “domain,” “integration,” “differentia- 6 Mathematics and the Imagination tion.” The uninitiated would not be able to guess what they represent; only mathematicians would know about them. The word “transcendental” in mathematics has not the meaning it has in philosophy. A mathemati¬ cian would say: The number tt, equal to 3.14159 . . . , is transcendental, because it is not the root of any alge¬ braic equation with integer coefficients. Transcendental is a very exalted name for a small number, but it was coined when it was thought that transcendental numbers were as rare as quintuplets. The work of Georg Cantor in the realm of the infinite has since proved that of all the numbers in mathematics, the transcendental ones are the most common, or, to use the word in a slighdy different sense, the least tran¬ scendental. We shall talk of this later when we speak of another famous transcendental number, e, the base of the natural logarithms. Immanuel Kant’s “transcen¬ dental epistemology” is what most educated people might think of when the word transcendental is used, but in that sense it has nothing to do with mathematics. Again, take the word “evolution,” used in mathematics to denote the process most of us learned in elementary school, and promptly forgot, of extracting square roots, cube roots, etc. Spencer, in his philosophy, defines evolution as ‘ an integration of matter, and a dissipation of motion from an indefinite, incoherent homogeneity to a definite, coherent heterogeneity,” etc. But that, formnately, has nothing to do with mathematical evo¬ lution either. Even in Tennessee, one may extract square roots without running afoul of the law. As \'. e see, mathematics uses simple words for com¬ plicated ideas. An example of a simple word used in a complicated way is the word “simple.” “Simple curve” New Names jor Old 7 and “simple group” represent important ideas in higher mathematics. The above is not a simple curve. A simple curve is a closed curve which does not cross itself and may look like Fig. 2. There are many important theorems about such figures that make the word worth while. Later, we are FIG. 2 going to talk about a queer kind of mathematics called “ntbber-sheet geometry,” and will have much more to say about simple curves and nonsimple ones. A French mathematician^ Tordan. gave the fundamental theorem: every simple curve has one inside and one outside. That is, every simple curve divides the plane into two regions, one inside the curve, and one outside. There are some groups in mathematics that arc “simple” groups. The definition of “simple group” is really so hard that it cannot be given here. If we wanted to get a clear idea of what a simple group was, we should 8 Mathematics and the Imagination probably have to spend a long time looking into a great many books, and then, without an extensive mathemat¬ ical background, we should probably miss the point. First of all, we should have to define the concept “group.” Then we should have to give a definition of subgroups, and then of self-conjugate subgroups, and then we should be able to tell what a simple group is. A simple group is simply a group without any self-conjugate subgroups— simple, is it not? Mathematics is often erroneously referred to as the science of common sense. Actually, it may transcend common sense and go beyond either imagination or intuition. It has become a very strange and perhaps frightening subject from the ordinary point of view, but anyone who penetrates into it will find a veritable fairy¬ land, a fairyland which is strange, but makes sense, if not common sense. From the ordinary point of view mathematics deals with strange things. We shall show you that occasionally it does deal with strange things, but mostly it deals with familiar things in a strange way. If you look at yourself in an ordinary mirror, regardless of your physical attributes, you may find yourself amus- ing, but not strange; a subway ride to Coney Island, and a glance at yourself in one of the distorting mirrors will convince you that from another point of view you may be strange as well as amusing. It is largely a matter of what you arc accustomed to. A Russian peasant came to Mos¬ cow for the first time and went to see the sights. He went to the zoo and saw the giraffes. You may find a moral in his reaction as plainly as in the fables of La Fontaine. “Look/’ he said, “at what the Bolsheviks have done to our horses.” That is what modern mathematics has done to simple geometry and to simple arithmetic.^ New Names for Old 9 There are other words and expressions, not so familiar, which have been invented even more recently. Take, for instance, the word “turbine.’* Of course, that is already used in engineering, but it is an entirely new word in geometry. The mathematical name applies to a certain diagram. (Geometry, whatever others may think, is the study of different shapes, many of them very beautiful, having harmony, grace and symmetry. Of course, there are also fat books written on abstract geom¬ etry, and abstract space in which neither a diagram nor a shape appears. This is a very important branch of mathematics, but it is not the geometry studied by the Egyptians and the Greeks. Most of us, if we can play chess at all, are content to play it on a board with wooden \ U / / / / M \ FIG. 3.—Turbines. chess pieces; but there are some who play the game blindfolded and without touching the board. It might be a fair analogy to say that abstract geometry is like blindfold chess—it is a game played without concrete objects.) Above you see a picture of a turbine, in fact, two of them. A turbine consists of an infinite number of “elements” filled in continuously. An element is not merely a point; I o Mathematics and the Imagination it is a point with an associated direction—like an iron filing. A turbine is composed of an infinite number of these elements, arranged in a peculiar way: the points must be arranged on a perfect circle, and the inclination of the iron filings must be at the same angle to the circle throughout. There are thus, an infinite number of ele¬ ments of equal inclination to the various tangents of the circle. In the special case where the angle between the direction of the element and the direction of the circle is zero, what would happen? The turbine would be a circle. In other words, the theory of turbines is a generalization of the theory of the circle. If the angle is ninety degrees, the elements point toward the center of the circle. In that special case we hav^ a normal turbine (see left-hand diagram). There is a geometry of turbines, instead of a geometry of circles. It is a rather technical branch of mathematics which concerns itself with working out continuous groups of transformations connected with differential equations and differential geometry. The group connected with the turbine bears the rather odd name of “turns and slides.” ♦ The circle is one of the oldest figures in mathematics. The straight line is the simplest line, but the circle is the simplest nonstraight line. It is often regarded as the limit of a polygon with an infinite number of sides. You can see for yourself that as a series of polygons is inscribed in a circle with each polygon having more sides than its predecessor, each polygon gets to look more and more like a circle.^ The Greeks were already familiar with the idea that as a regular polygon increases in the number of its sides, New Names for Old 11 it differs less and less from the circle in which it is in¬ scribed. Indeed, it may well be that in the eyes of an omniscient creature, the circle would look like a polygon with an infinite number of straight sides. ^ However, in the absence oi complete omniscience, we shall continue FIG. 4.—The circle as the limit of inscribed polygons. to regard a circle as being a nonstraight line. There are some interesting generalizations of the circle when it is viewed in this way. There is, for example, the concept denoted by the word “cycle,” which was introduced by a French mathematician, Laguerre. A cycle is a circle with an arrow on it, like this: If you took the same circle and put an arrow on it in the opposite direction, it would become a different cycle. The Greeks were specialists in the art of posing prob- 12 Mathematics and the Imagination lems which neither they nor succeeding generations of mathematicians have ever been able to solve. The three most famous of these problems—the squaring of the circle, the duplication of the cube, and the trisection of an angle—we shall discuss later. Many well-meaning, self-appointed, and self-anointed mathematicians, and a motley assortment of lunatics and cranks, knowing neither history nor mathematics, supply an abundant crop of “solutions” of these insoluble problems each year. However, some of the classical problems of antiquity have been solved. For example, the theory of cycles was used by Laguerre in solving the problem of Apollonius: given three fixed circles, to find a circle that touches them all. It turns out to be a matter of elementary high New Names Jor Old 13 school geometry, although it involves ingenuity, and any brilliant high school student could work it out. It has eight answers, as shown in Fig. 6(a). They can all be constructed with ruler and compass, and many methods of solution have been found. Given three circles^ there will be eight circles touching all of them. Given three cycles^ however, there will be only one cycle that touches them all. (Two cycles are said to touch each other only if their arrows agree in direction at the point of contact.) Thus, by using the idea of cycles, we have one definite answer instead of eight. Laguerre made the idea of cycles the basis of an elegant theory. FIG. 6(b). —The eight solutions of Appolonius merged into one diagram. Another variation of the circle introduced by the emi¬ nent American mathematician, C. J. Keyser, is obtained by taking a circle and removing one point.^ This creates a serious change in conception. Keyser calls it “a patho- circle,” (from pathological circle). He has used it in discussing the logic of axioms. 14 Mathematics and the Imagination We have made yet another change in the concept of circle, which introduces another word and a new di¬ agram. Take a circle and instead of leaving one point out, simply emphasize one point as the initial point. This is to be called a “clock.’’ It has been used in the theory of polygenic functions. “Pplygeiiic” is a word recently introduced into the theory of complex functions —about 1927. There was an important word, “mono¬ genic,” introduced in the nineteenth century by the famous French mathematician, Augustin Cauchy, and used in the classical theory of functions. It is used to denote functions that have a single derivative at a point, as in the differential calculus. But most functions, in the complex domain, have an infinite number of derivatives at a point. If a function is not monogenic, it can never be bigenic, or trigenic. Either the derivative has one value or an infinite number of values—either monogenic or polygenic, nothing intermediate. Monogenic means one rate of growth. Polygenic means many rates of growth. The complete derivative of a polygenic function is represented by a congruence (a double infinity) of clocks, all with different starting points, but with the same uniform rate of rotation. It would be useless to attempt to give a simplified explanation of these con¬ cepts. (The neophyte will have to bear with us over a few intervals like this for the sake of the more experienced mathematical reader.) New Names jor Old The going has been rather hard in the last paragraph, and if a few of the polygenic seas have swept you over¬ board, we shall throw you a hexagonal life preserver. We may consider a very simple word that has been intro¬ duced in elementary geometry to indicate a certain kind of hexagon. The word on which to fix your attention is ‘"parhexagon.” An ordinary hexagon has six arbitrary sides. A parhexagon is that kind of hexagon in which any side is both equal and parallel to the side opposite to it (as in Fig, 7). If the opposite sides of a quadrilateral are equal and parallel, it is called a parallelogram. By the same rea¬ soning that we use for the word parhexagon, a parallelo¬ gram might have been called a parquadrilateral. Here is an example of a theorem about the parhex¬ agon: take any irregular hexagon, not necessarily a parhexagon, ABCDEF. Draw the diagonals AC, BD, CE, DF, EA, and FB, forming the six triangles, ABC, BCD, CDE, DEF, EFA, and FAB. Find the six centers of gravity, A', B', C', D', E', and F' of these triangles. (The center of gravity of a triangle is the point at which the triangle would balance if it were cut out of cardboard and supported only at that point; it coincides with the D FIG. -ABCDEF \s an irregular hexagon. A'B' C'D'E'F' is a parhexagon. 16 Mathematics and the Imagination point of intersection of the medians.) Draw A'B', B'C', C'D', D'E', E'F', and F'A'. Then the new inner hex¬ agon A'B'C'D'ET' will always be a parhexagon. The word radical, favorite call to arms among Repub¬ licans, Democrats, Communists, Socialists, Nazis, Fas¬ cists, Trotskyites, etc., has a less hortatory and bellicose character in mathematics. For one thing, everybody knows its meaning: i.e., square root, cube root, fourth root, fifth root, etc. Combining a word previously de¬ fined with this one, we might say that the extraction of a root is the evolution of a radical. The square root of 9 is 3; the square root of 10 is greater than 3, and the most famous and the simplest of all square roots, the first in¬ commensurable number discovered by the Greeks, the square root of 2, is 1.414. . . There are also composite radicals—expressions like \/7 + "V^IO. The symbol for a radical is not the hammer and sickle, but a sign three or four centuries old, and the idea of the mathematical radical is even older than that. The concept of the ‘‘hypcrradical,” or “ultraradical,” which means some¬ thing liigher than a radical, but lower than a transcen¬ dental, is of recent origin. It has a symbol which we shall see in a moment. First, we must say a few words about radicals in general. There are certain numbers and functions in mathematics which are not expressible in the language of radicals and which are generally not well understood. Many ideas for which there are no concrete or diagrammatic representations are difficult to explain. Most people find it impossible to think without words; it is necessary to give them a word and a symbol to pin their attention. Hyperradical or ultraradical, for which hitherto there have been neither words, nor sym¬ bols, fall into this category. New Names Jor Old We first meet these ultraradicals, not in Mexico City, but in trying to solve equations of the fifth degree. The Egyptians solved equations of the first degree perhaps 4000 years ago. That is, they found that the solution of the equation ax b = 0^ which is represented in geometry by a straight line, is x = —. The quadratic equation ax^ + ix- + c = 0 was solved by the Hindus and V the Arabs with the formula x — ^ ± \/b^ — 4ac 2a The various conic sections, the circle, the ellipse, the parabola, and the hyperbola, are the geometric pictures of quadratic equations in two variables. Then in the sixteenth century the Italians solved the equations of third and fourth degree, obtaining long formulas involving cube roots and square roots. So that by the year 1550, a few years before Shakespeare was born, the equation of the first, second, third, and fourth degrees had been solved. Then there was a delay of 250 years, because mathematicians were struggling with the equation of the fifth degree—the general quintic. Finally, at the beginning of the nineteenth century, Ruffi ni and Abel showed that equations of the fifth degree couTd not be solved with radicals. The general quintic is thus not like the general quadratic, cubic or biquadratic. Never¬ theless, it presents a problem in algebra which theoret¬ ically can be solved by algebraic functions. Only, these functions are so hard that they cannot be expressed by the symbols for radicals. TThese new higher things are FIG. 9.—A portrait of two radicals. 18 Mathematics and the Imagination named “ultraradicals,” and they too have their special symbols (shown in Fig. 9). With such symbols combined with radicals, we can solve equations of the fifth degree. For example, the solution of -j- X = a may be written x = X = Jo". The usefulness of the special symbol and name is apparent. Without them the solution of the quintic equation could not be compactly expressed. + We may now give a few ideas somewhat easier than those with which we have thus far occupied ourselves. These ideas were presented some time ago to a number of children in kindergarten. It was amazing how well they understood everything that was said to them. In¬ deed, it is a fair inference that kindergarten children can enjoy lectures on graduate mathematics as long as the mathematical concepts are clearly presented. It was raining and the children were asked how many raindrops would fall on New York. The highest answer was 100. They had never counted higher than 100 and what they meant to imply when they used that number was merely something very, very big—as big as they could imagine. They were asked how many raindrops hit the roof, and how many hit New York, and how many single raindrops hit all of New York in 24 hours. They soon got a notion of the bigness of these numbers even though they did not know the symbols for them. They were certain in a little while that the number of rciindrops was a great deal bigger than a hundred. They were asked to think of the number of grains of sand on the beach at Coney Island and decided that the number of grains of sand and the number of raindrops were about the same. But the important thing is that they realized that the New Names for Old i g number \wa.s finite^ not infinite. In this respect they showed their distinct superiority over many scientists who to this day use the word infinite when they mean some big number, like a billion billion. Counting, something such scientists evidently do not realize, is a precise operation.* It may be wonderful but there is nothing vague or mysterious about it. If you count something, the answer you get is either per¬ fect or all wrong; there is no half way. It is very much like catching a train. You either catch it or vou miss it, and if you miss it by a split second you might as well have come a week late. There is a famous quotation which illustrates this: “Oh, the little more, and how much it is! And the little less, and what worlds away!” A big number is big, but it is definite and it is finite. Of course in poetry, the finite ends with about three thousand; any greater number is infinite. In many poems, the poet will talk to you about the infinite number of stars. But, if ever there was a hyperbole, this is it, for nobody, not even the poet, has ever seen more than three thousand stars on a clear night, without the aid of a telescope. With the Hottentots, infinity begins at three.t Ask a Hottentot how many cows he owns, and if he has more than three he’ll say “many.” The number of raindrops ♦ No one would say that 1 + 1 is "about equal to 2.” It is just as silly to say that a billion billion is not a finite number, simply because It IS big. Any number which may be named, or conceived of in icrins ol (he integers is finite. Infinite means something quite different, as we shall see in ihe chapter on the ^ in all fairness, it must be pointed out that some of the tribes of the Belgian Congo can count to a million and beyond. 3 20 Mathematics and the Imagination falling on New York is also “many.” It is a large finite number, but nowhere near infinity. Now here is the name of a very large number; “Goo- gol.”* Most people would say, “A googol is so large that you cannot name it or talk about it; it is so large that it is infinite.” Therefore, we shall talk about it, explain exactly what it is, and show that it belongs to the very same family as the number 1. A googol is this number which one of the children in the kindergarten wrote on the blackboard: 100000000000000000000000000000000000000000000000 000000000000000000000000000000000000000000000000 00000 The definition of a googol is: 1 followed by a hundred zeros. It was decided, after careful mathematical re¬ searches in the kindergarten, that the number of rain¬ drops falling on New York in 24 hours, or even in a year - or in a century, is much less than a googol. Indeed, the googol is a number just larger than the largest numbers that are used in physics or astronomy. All those numbers require less than a hundred zeros. This information is, of course, available to everyone, but seems to be a great secret in many scientific quarters. A very distinguished scientific publication recently came forth with the revelation that the number of snow crystals necessary to form the ice age wais a billion to the billionth power. This is very startling and also very silly. A billion to the billionth power looks like this: lOOOOOOOOO^ooo^ooooo^ A more reasonable estimate and a somewhat smaller number would be 10^®. As a matter of fact, it has been estimated that if the entire universe, which you will con- • Not even approximately a Russian author. ft Lihtmn 21 New Names for Old cede is a trifle larger than the earth, were filled with protons and electrons, so that no vacant space remained, the total number of protons and electrons would be 10 (i*c*) 10 with 110 zeros after it). Unfortunately, as soon as people talk about large numbers, they run amuck. They seem to be under the impression that since zero equals nothing, they can add as many zeros to a number as they please with practically no serious con¬ sequences. We shall have to be a little more careful than that in talking about big numbers. To return to Coney Island, the number of grains of sand on the beach is about lO^o, or more descriptively, 100000000000000000000. That is a large number, but not as large as the number mentioned by the divorcee in a recent divorce suit who had telephoned that she loved the man “a million billion billion times and eight times around the world.” It was the largest number that she could conceive of, and shows the kind of thing that may be hatched in a love nest. Though people do a great deal of talking, the total output since the beginning of gabble to the present day, including all baby talk, love songs, and Congressional debates, totals about 10*®. This is ten million billion. Con¬ trary to popular belief, this is a larger number of words than is spoken at the average afternoon bridge. A great deal of the veneration for the authority of the printed word would vanish if one were to calculate the number of words which have been printed since the Gutenberg Bible appeared. It is a number somewhat larger than 10*®. A recent popular historical novel alone accounts for the printing of several hundred billion words. The largest number seen in finance (though new records are in the making) represents the amount of 22 Mathematics and the Imagination money in circulation in Germany at the peak of the inflation. It was less than a googol—merely 496,585,346,000,000,000,000. A distinguished economist vouches for the accuracy of this figure. The number of marks in circulation was very nearly equal to the number of grains of sand on Coney Island beach. The number of atoms of oxygen in the average thimble is a good deal larger. It would be represented by perhaps 1000000000000000000000000000. The number of elec¬ trons, in size exceedingly smaller than the atoms, is much more enormous. The number of electrons which pass through the filament of an ordinary fifty-watt electric lamp in a minute equals the number of drops of water that flow over Niagara Falls in a century. One may also calculate the number of electrons, not only in the average room, but over the whole earth, and out through the stars, the Milky Way, and all the neb¬ ulae. The reason for giving all these examples of very large numbers is to emphasize the fact that no matter how large the collection to be counted, a finite number will do the trick. We will have occasion later on to speak of infinite collections, but those encountered in nature, though sometimes very large, are all definitely finite. A celebrated scientist recently stated in all seriousness that he believed that the number of pores (through which lca\ es breathe) of all the leaves, of all the trees in all the woild, would certainly be infinite. Needless to say, he was not a niathematician. The number of electrons in a single leaf is much bigger than the number of pores of all the leaves of all the trees of all the world. And still the num er of all the electrons in the entiie universe can be found by means of the physics of Einstein. It is a good New Names for Old 23 deal less than a googol—perhaps ten with seventy-nine zeros, 10^®, as estimated by Eddington. Words of wisdom are spoken by children at least as often as by scientists. The name “googol” was invented by a child (Dr. Kasner’s nine-year-old nephew) who was asked to think up a name for a very big number, namely, 1 with a hundred zeros after it. He was very certain that this number was not infinite, and therefore equally certain that it had to have a name. At the same time that he suggested “googol” he gave a name for a still larger number: “Googolplex.” A googolplex is much larger than a googol, but is still finite, as the inventor of the name was quick to point out. It was first suggested that a googolplex should be 1, followed by writing zeros until you got tired. This is a description of what would happen if one actually tried to write a googolplex, but different people get tired at different times and it would never do to have Camera a better mathematician than Dr. Ein¬ stein, simply because he had more endurance. The goo¬ golplex then, is a specific finite number, with so many zeros after the 1 that the number of zeros is a googol. A googolplex is much bigger than a googol, much bigger even than a googol limes a googol. A googol times a googol would be 1 with 200 zeros, whereas a googolplex is 1 with a googol of zeros. You will get some idea of the size of this very large but finite number from the fact that there would not be enough room to write it, if you went to the farthest star, touring all the nebulae and put¬ ting down zeros every inch of the way. One might not believe that such a large number would ever really have any application; but one who felt that way would not be a mathematician. A number as large as the googolplex might be of real use in problems of 24 Mathematics and the Imagination combination. This would be the type of problem in which it might come up scientifically: Consider this book which is made up of carbon and nitrogen and of other elements. The answer to the ques¬ tion, “How many atoms are there in this book?” would certainly be a finite number, even less than a googol. Now imagine that the book is held suspended by a string, the end of which you are holding. How long will it be necessary to wait before the book will jump up into your hand? Could it conceivably ever happen? One answer might be “No, it will never happen without some external force causing it to do so.” But that is not correct. The right answer is that it will almost certainly happen sometime in less than a googolplex of years—per¬ haps tomorrow. The explanation of this answer can be found in physical chemistry, statistical mechanics, the kinetic theory of gases, and the theory of probability. We cannot dispose of all these subjects in a few lines, but we will try. Molecules are always moving. Absolute rest of molecules | would mean absolute zero degrees of temperature, and I absolute zero degrees of temperature is not only non- I existent, but impossible to obtain. All the molecules of 1 the surrounding air bombard the book. At present the bombardment from above and below is nearly the same and gravity keeps the book down. It is necessary to wait or the favorable moment when there happens to be an enormous numb- • of molecules bombarding the book from below and x. ry few from above. Then gravity will be overcome and the book will rise. It would be some¬ what hke the effect known in physics as the Brownian movement, which describes the behavior of small par- tic es m a liquid as they dance about under the impact New Names Jot Old 25 of molecules. It would be analogous to the Brownian movement on a vast scale. But the probability that this will happen in the near future or, for that matter, on any specific occasion that we might mention, is between ^—- and -5-. googol googolplex To be reasonably sure that the book will rise, we should have to wait between a googol and a googolplex of years. When working with electrons or with problems of combination like the one of the book, we need larger numbers than are usually talked about. It is for that reason that names like googol and googolplex, though they may appear to be mere jokes, have a real value. The names help to fix in our minds the fact that we are still dealing with finite numbers. To repeat, a googol is a googolplex is 10 to the googol power, which may be written We have seen that the number of years that one would have to wait to see the miracle of the rising book would be less than a googolplex. In that number of years the earth may well have become a frozen planet as dead as the moon, or perhaps splintered to a number of meteors and comets. The real miracle is not that the book will rise, but that with the aid of mathematics, we can project ourselves into the future and predict with accu¬ racy when it will probably rise, i.c., some time between today and the year googolplex. ♦ We have mentioned quite a few new names in mathe¬ matics new names for old and new ideas. There is one more new name which it is proper to mention in con¬ clusion. Watson Davis, the popular science reporter, has given us the name “mathescope.” With the aid of the 26 Mathematics and the Imagination magnificent new microscopes and telescopes, man, mid¬ way between the stars and the atoms, has come a little closer to both. The mathoscope is not a physical instru¬ ment; it is a purely intellectual instrument, the ever- increasing insight which mathematics gives into the fairy¬ land which lies beyond intuition and beyond imagina¬ tion. Mathematicians, unlike philosophers, say nothing about ultimate truth, but patiently, like the makers of the great microscopes, and the great telescopes, they grind their lenses. In this book, w'e shall let you see through the newer and greater lenses which the mathe¬ maticians have ground. Be prepared for strange sights through the mathescope! FOOTNOTES 1. Sec the Chapter on pie. —P. 10. 2. See the Chapter on Change and Changeability—Section on Path¬ ological Curves.—P.ll. 3. NMi. iliis is a diagram which the reader will have to imagine, lor It IS beyond the capacity of any printer to make a circle with one point omitted. A point, having no dimensions, will, like many of the persons on the Lord High Executioner’s list, never c inissc . o the circle with one point missing is purely con- cepiual, not an idea wliich can be pictured.—P.13. Beyond the Googol If you do not expect the unexpected^ you will not find it; for It IS hard to be sought out, and difficult. —HERACLITUS Mathematics may well be a science of austere logical propositions in precise eanonical form, but in its count¬ less applications it serves as a tool and a language, the language of description, of number and size. It describes with economy and elegance the elliptic orbits of the plan¬ ets as readily as the shape and dimensions of this page or a corn field. The whirling dance of the electron can be seen by no one; the most powerful telescopes can re¬ veal only a meager bit of the distant stars and nebulae and the cold far corners of space. But with th(' aid of mathematics and the imagination the very small, the very large—all things may be brought within man’s domain. To count is to talk the language of number. To count to a googol, or to count to ten is part of the same ijroeess; the googol is simply harder to pronounce. The essential thing to realize is that the googol and ten arc kin, like the giant stars and the electron. Arithmetic—this count¬ ing language—makes the whole v/orld kin, both in space and in time. To grasp the meaning and importance of mathematics, to appreciate its beauty and its value, arithmetic must first be understood, for mostly, since its beginning, mathc- 27 28 Mathematics and the Imagination matics has been arithmetic in simple or elaborate attire. Arithmetic has been the queen and the handmaiden of the sciences from the days of the astrologers of Chaldea and the high priests of Egypt to the present days of relativity, quanta, and the adding machine. Historians may dispute the meaning of ancient papyri, theologians may wrangle over the exegesis of Scripture, philosophers may debate over Pythagorean doctrine, but all will con¬ cede that the numbers in the papyri, in the Scriptures and in the writings of Pythagoras are the same as the num- bers of today. As arithmetic, mathematics has helped man to cast horoscopes, to make calendars, to predict the risings of the Nile, to measure fields and the height of the Pyramids, to measure the speed of a stone as it fell from a tower in Pisa, the speed of an apple as it fell from a tree in Woolsthorpe, to weigh the stars and the atoms, to mark the passage of time, to find the curvature of space. And although mathematics is also the calculus, the theory of probability, the matrix algebra, the science of the infinite, it is still the art of counting. * Everyone who will read this book can count, and yet, what is counting? The dictionary definitions are about as helpful as Johnson’s definition of a net: “A series of reticulated interstices.’* Learning to compare is learning to count. Numbers come much later; they are an artificiality, an abstraction. Counting, matching, comparing are al¬ most as indigenous to man as his fingers. Without the faculty of comparing, and without his fingers, it is un¬ likely that he would have arrived at numbers. One who knows nothing of the formal processes of counting is still able to compare two classes of objects, to determine which is the greater, which the less. With- Beyond the Googol 29 out knowing anything about numbers, one may ascertain whether two classes have the same number of elements; for example, barring prior mishaps, it is easy to show that we have the same number of fingers on both hands by simply matching finger with finger on each hand. To describe the process of matching, which underlies counting, mathematicians use a picturesque name. They call it putting classes into a “one-to-one reciprocal cor¬ respondence” with each other. Indeed, that is all there is to the art of counting as practiced by primitive peoples, by us, or by Einstein. A few examples may serve to make this clear. In a monogamous country it is unnecessary to count both the husbands and the wives in order to ascertain the number of married people. If allowances are made for the few gay Lotharios who do not conform to either custom or statute, it is sufficient to count either the husbands or the wives. There are just as many in one class as in the other. The correspondence between the two classes is one-to-one. There are more useful illustrations. Many people are gathered in a large hall where seats are to be provided. The question is, are there enough chairs to go around? It would be quite a job to count both the people and the chairs, and in this case unnecessary. In kindergarten children play a game called “Going to Jerusalem”; in a room full of children and chairs there is alwa>a one less chair than the number of children. At a signal, each child runs for a chair. The child left standing is “out.” A chair is removed and the game continues. Here is the solution to our problem. It is only necessary to ask everyone in the hall to be seated. If everyone sits down and no chairs are left vacant, it is evident that there 30 Alatfumatics and the Imagination are as many chairs as people. In other words, without actually knowing the number of chairs or people, one does know that the number is the same. The two classes— chairs and people—have been shown to be equal in number by a one-to-one correspondence. To each person corresponds a chair, to each chair, a person. In counting any class of objects, it is this method alone which is employed. One class contains the things to be counted; the other class is always at hand. It is the class of integers, or “natural numbers,” which for convenience we regard as being given in serial order: 1, 2, 3, 4, 5, 6, 7 . . . Matching in one-to-one correspondence the ele¬ ments of the first class with the integers, we experience a common, but none the less wonderful phenomenon—the last integer necessary to complete the pairings denotes how many elements there are. * In clarifying the idea of counting, we made the un¬ warranted assumption that the concept of number was everyone. The number concept may seem intuitively clear, but a precise definition is required. While the definition may seem worse than the disease, it is not as difficult as appears at first glance. Read it t you will find that it is both explicit and economical. Given a class C containing certain elements, it is possible to find other classes, such that the elements of each may be matched one to one with the elements of C. (Each of these classes is thus called “equivalent to C.”) All such classes, including C, whatever the character of their elements, share one property in common: all of them have the same cardinal number^ which is called the cardinal number of the class Cd Beyond the Googol 21 The cardinal number of the class C is thus seen to be the symbol representing the set of all classes that can be put into one-to-one correspondence with C. For example, the number 5 is simply the name, or symbol, attached to the set of all the classes, each of which can be put into one-to-one correspondence with the fingers of one hand. Hereafter we may refer without ambiguity to the number of elements in a class as the cardinal number of that class or, briefly, as “its cardinality.’* The question, “How many letters are there in the word mathematics?'^ is the same as the question, “What is the cardinality of the class whose elements are the letters in the word mathematics?" Employing the method of one-to-one cor¬ respondence, the following graphic device answers the question, and illustrates the method: M A T H E M A T I C S 1 I t t i t t k t t t 4- 1 X X X X X X X 1 2 3 4 5 6 1 8 9 10 11 It must now be evident that this method is neither strange nor esoteric; it was not invented by mathema¬ ticians to make something natural and easy seem un¬ natural and hard. It is the method employed when we count our change or our chickens; it is the proper method for counting any class, no matter how large, from ten to a googolplex—and beyond. Soon we shall speak of the “beyond” when we turn to classes which are not finite. Indeed, we shall try to measure our measuring class —the integers. One-to-one correspond¬ ence should, therefore, be thoroughly understood, for an amazing revelation awaits us: Infinite classes can also be counted, and by the very same means. But before 32 Mathematics and the Imagination we try to count them, let us practice on some very big numbers—big, but not infinite. * “Googol” is already in our vocabulary: It is a big number one, with a hundred zeros after it. Even bigger is the googolplex: 1 with a googol zeros after it. Most numbers encountered in the description of nature are much smaller, though a few are larger. Enormous numbers occur frequently in modern sci¬ ence. Sir Arthur Eddington claims that there are, not approximately, but exactly 136-2256 protons,* and an equal number of electrons, in the universe. Though not easy to visualize, this number, as a symbol on paper, takes up little room. Not quite as large as the googol, it is completely dwarfed by the googolplex. None the less, Eddington’s number, the googol, and the googolplex are finite. A veritable giant is Skewes* number, even bigger than ^ googolplex. It gives information about the distribution of primes^ and looks like this: 10 10 ID Or, for example, the total possible number of moves in a game of chess is: 10 10 oO And speaking of chess, as the eminent English mathe¬ matician, G. H. Hardy, pointed out—if we imagine the • Let no one suppose that Sir Arthur has counted them. But he if'n ^ ^ justify his claim. Anyone with a better theory may challenge Sir Arthur, for who can be referee? Here is his number 653,951,181,555,468,- he says,’ to t’he" Hst ,425,076,185,631,031,276-accurate, Beyond the Googol 33 entire universe as a chessboard, and the protons in it as chessmen, and if we agree to call any interchange in the position of two protons a “move” in this cosmic game, then the total number of possible moves, of all odd coin¬ cidences, would be Skewes’ number: No doubt most people believe that such numbers are part of the marvelous advance of science, and that a few generations ago, to say nothing of centuries back, no one in dream or fancy could have conceived of them. There is some truth in that idea. For one thing, the ancient cumbersome methods of mathematical notation made the writing of big numbers difhcult, if not actually impossible. For another, the average citizen of today en¬ counters such huge sums, representing armament ex¬ penditures and stellar distances, that he is quite conver¬ sant with, and immune to, big numbers. But there were clever people in ancient times. Poets in every age may have sung of the stars as infinite in number, when all they saw was, perhaps, three thousand. But to Archimedes, a number as large as a googol, or even larger, was not disconcerting. He says as much in an introductory passage in The Sand Reckoner, realizing that a number is not infinite merely because it is enor¬ mous. There are some, King Gelon, who think that the number of the sand is infinite in multitude; and I mean by the sand, not only that which exists about Syracuse and the rest of Sicily, but also that which is found in every region whether inhabited or uninhabited. Again there are some who, without regarding 34 Mathematics and the Imagination it as infinite, yet think that no number has been named which is great enough to exceed its multitude. And it is clear that they who hold this view, if they imagined a mass made up of sand in other respects as large as the mass of the earth, in¬ cluding in it all the seas and the hollows of the earth filled up to a height equal to that of the highest of the mountains, would be many times further still from recognizing that any number could be expressed which exceeded the multitude of the sand so taken. But I will try to show you by means of geometrical proofs, which you will be able to follow, that, of the numbers named by me and given in the work which I sent to Zeuxippus, some exceed not only the number of the mass of sand equal in magnitude to the earth filled up in the way described, but also that .of a mass equal in magnitude to the universe. The Greeks had verv definite ideas about the infinite. Just as we are indebted to them for much of our wit and our learning, so are we indebted to them for much of our sophistication about the infinite. Indeed, had we always retained their clear-sightedness, many of the prob¬ lems and paradoxes connected with the infinite would never have arisen. Above everything, we must realize that “very big” and “infinite’' arc entirely difTejrent.* By using the method of one-to-one correspondence, the protons and electrons in the universe may theoretically be counted as easily as the buttons on a vest. Sufficient and more than sufficient lor that task, or for the task of counting any finite collection, are the integers. But measuring the * There is no point where the very big starts to merge into the infinite. \ou may write a number as big as you please; it will be no nearer the infinite than the number 1 or the number 7. Make sure that you keep this distinction very clear and you will have mastered many of the subtleties of the transfinitc. Beyond the Googol 3 ^ totality of integers is another problem. To measure such a class demands a lofty viewpoint. Besides being, as the German mathematician Kronecker thought, the work of God, which requires courage to appraise, the class of integers is infinite—which is a great deal more in¬ convenient. It is worse than heresy to measure our own endless measuring rod! ♦ . The problems of the infinite have challenged man’s I mind and have fired his imagination as no other single problem in the history of thought. The infinite appears both strange and familiar, at times beyond our grasp, at times natural and easy to understand. In conquering it, man broke the fetters that bound him to earth. All his faculties were required for this conquest—his rea.son- ing powers, his poetic fancy, his desire to know. To establish the science of the infinite involves the principle of mathematical imita tion. This principle affirms the power of reasoning by recurrence. It typifies almost all mathematical thinking, all that we do when we construct complex aggregates out of simple elements. It is, as Poincare remarked, “at once necessary to the mathematician and irreducible to logic.” His statement of the principle is: “If a property be true of the number one, and if we establish that it is true of h -f 1,* provided it be of «, it will be true of all the whole numbers.” Mathematical induction is not derived from experience, rather is it an inherent, intuitive, almost instinctive property of the mind, “ What we have once done we can do again.’’’* If we can construct numbers to ten, to a million, to a googol, we are led to believe that there is no stopping, * Where n is any integer. 4 36 Mathematics and the Imagination no end. Convinced of this, we need not go on forever; the mind grasps that which it has never experienced— the infinite itself. Without any sense of discontinuity, without transgressing the canons of logic, the mathema¬ tician and philosopher have bridged in one stroke the gulf between the finite and the infinite. The mathematics ol the infinite is a sheer affirmation of the inherent power of reasoning by recurrence. In the sense that “infinite” means “without end, with¬ out bound,” simply “not finite,” probably everyone un¬ derstands its meaning. No difficulty arises where no precise definition is required. Nevertheless, in spite of the famous epigram that mathematics is the science in which we do not know what we are talking about, at least we shall have to agree to talk about the same thing. Apparently, even those of scientific temper can argue bitterly to the point of mutual vilification on subjects ranging from Marxism and dialectical materialism to group theory and the uncertainty principle, only to find, on the verge of exhaustion and collapse, that they are on the same side of the fence. Such arguments are generally the results of vague terminology; to assume that everyone is familiar with the precise mathematical definition of infinite is to build a new Tower of Babel. Before undertaking a definition, we might do well to glance backwards to see how mathematicians and philos¬ ophers of other times dealt with the problem. The infinite has a double aspect—the infinitely large, and the infinitely small. Repeated arguments and demon- stiations, of apparently apodictic force, were advanced, overwhelmed, and once more resuscitated to prove or disprove its existence. Few of the arguments were ever Beyond the Googol 37 refuted—each was buried under an avalanche of others. The happy result was that the problem never became any clearer. * The warfare began in antiquity with the paradoxes of Zeno; it has never ceased. Fine points were debated with a fervor worthy of the earliest Christian martyrs, but without a tenth part of the acumen of medieval theologians. Today, some mathematicians think the infinite has been reduced to a state of vassalage. Others are still wondering what it is. Zeno’s puzzles may help to bring the problem into sharper focus. Zeno of Elea, it will be recalled, said some disquieting things about motion, with reference to an arrow, Achilles, and a tortoise. This strange company was employed on behalf of the tenet of Eleatic philosophy —that all motion is an illusion. It has been suggested, probably by “baffled critics,” that “Zeno had his longue in cheek when he made his puzzles.” Regardless of mo¬ tive, they arc immeasurably subtle, and perhaps still defy solution.! One paradox—the Dichotomy—states that it is im¬ possible to cover any given distance. The argument: First, half the distance must be trav'ersed, then half of the remaining distance, then again half of what remains, * No one has written more brilliantly or more wittily on this subject than Bertrand Russell. Sec particularly his essays in the volume Mys¬ ticism and Lof^ic. tTo be sure, a variety of explanations have been t^iven for the paradoxes. In the last analysis, the explanations for the riddles rest upon the interpretation of the foundations of mathematics. Mathe¬ maticians like Brouwer, who reject the infinite, would probably not accept any of the solutions given. 38 Mathematics and the Imagination and so on. It follows that some portion of the distance to be covered always remains, and therefore motion is impossible! A solution of this paradox reads: The successive distances to be covered form an infinite geometric series: i + i + i + l + _L+ 3 2 ^ 4 ^ 816 ^ 32 ^ *" each term of which is half of the one before. Although this series has an infinite number of terms, its sum is finite and equals 1. Herein, it is said, lies the flaw of the Dichotomy. Zeno assumed that any totality composed of an infinite number of parts must, itself, be infinite, whereas we have just seen an infinite number of elements which make up the finite totality—1. The paradox of the tortoise states that Achilles, running to overtake the tortoise, must first reach the place where it started: but the tortoise has already departed. This comedy, however, is repeated indefinitely. As Achilles arrives at each new point in the race, the tortoise having been there, has already left. Achilles is as unlikely to catch him as a rider on a carrousel the rider ahead. Finally. the arrow in flight must be moving every instant of time. But at ever)’ instant it must be somewhere in space. However, if the arrow must always be in some Beyond the Googol 3g one place, it cannot at every instant also be in transit, for to be in transit is to be nowhere. Aristotle and lesser saints in almost every age tried to demolish these paradoxes, but not very creditably. Three German professors succeeded where the saints had failed. At the end of the nineteenth century, it seemed that Bolzano, Weierstrass and Cantor had laid the infinite to rest, and Zeno’s paradoxes as well. The modern method of disposing of the paradoxes is not to dismiss them as mere sophisms unworthy of serious attention. The history of mathematics, in fact, recounts a poetic vindication of Zeno’s stand. Zeno was, at one lime, as Bertrand Russell has said, “A notable victim of pos¬ terity’s lack of judgement.” That wrong has been righted. In disposing of the infinitely small, Weierstrass showed that the moving arrow is really always at rest, and that we live in Zeno’s changeless world. The work of Georg Cantor, which we shall soon encounter, showed that if we are to believe that Achilles can catch the tortoise, we shall have to be prepared to swallow a bigger paradox than any Zeno ever conceived of: the whole is no GREATER THAN MANY OF ITS PARTS* The infinitely small had been a nuisance for more than two thousand years. At best, the innumerable opinions it evoked deserved the laconic verdict of Scotch Juries: “Not proven.” Until Weierstrass appeared, the total advance was a confirmation of Zeno’s argument against motion. Even the jokes were better. Leibniz, according to Carlyle, made the mistake of trying to explain the infinitesimal to a Queen—Sophie Charlotte of Prussia. She informed him that the behavior of her courtiers made her so familiar with the infinitely small, that she needed no mathematical tutor to explain it. But philos- 40 Mathematics and the Imagination ophers and mathematicians, according to Russell, ‘‘hav¬ ing less acquaintance with the courts, continued to dis¬ cuss this topic, though without making any advance.’’ Berkeley, with the subtlety and humor necessary for an Irish bishop, made some pointed attacks on the infini¬ tesimal, during the adolescent period of the calculus, that had the very best, sharp-witted, scholastic sting. One could perhaps speak, if only with poetic fervor, of the infinitely large, but what, pray, was the infinitely small? The Greeks, with less than their customary sagacity, introduced it in regarding a circle as differing infini¬ tesimally from a polygon with a large number of equal sides. Leibniz used it as the bricks for the infinitesimal calculus. Still, no one knew what it was. The infinitesimal had wondrous properties. It was not zero, yet smaller than any quantity. It could be assigned no quantity or size, yet a sizable number of infinitesimals made a very definite quantity. Unable to discover its nature, happily able to dispense with it, Weierstrass interred it alongside of the phlogiston and other once-cherished errors. * The infinitely large offered more stubborn resistance. Whatever it is, it is a doughty weed. The subject of reams of nonsense, sacred and profane, it was first discussed fully, logically, and without benefit of clergy-like prej¬ udices by Bernhard Bolzano. Die Paradoxien des ZJnendlichen, a remarkable little volume, appeared posthumously in 1851. Like the work of another Austrian priest, Gregor Mendel, whose distinguished treatise on the principles of heredity escaped oblivion only by chajr^e, this im¬ portant book, charmingly written, made no great im¬ pression on Bolzano’s contemporaries. It is the creation of a clear, forceful, penetrating intelligence. For the Beyond the Googol 41 first time in twenty centuries the infinite was treated as a problem in science, and not as a problem in theology. Both Cantor and Dedekind are indebted to Bolzano for the foundations of the mathematical treatment of the infinite. Among the many paradoxes he gathered and explained, one, dating from Galileo, illustrates a typical source of confusion: Construct a square— ABCD. About the point .1 as cen¬ ter, with one side as radius, describe a quarter-circle, in¬ tersecting the square at B and Z). Draw PR parallel to dZ), cutting AB at P, CD at /?, the diagonal AC at N, and the quarter-circle at A/. A FIG. 11.—Extract triangle APM from the figure. It is not hard to see that its three sides equal respectively the radii of the three circles. Thus RP - R^i = R,2 or, wRi^ - = irRs^ or, the two shaded areas are equal. 42 Mathematics and the Imagination By a well-known geometrical theorem, it can be shown that if PjV, PM and PR are radii, the following relation¬ ship exists: irPN = itTR - wPM^ (1) Permit PR to approach AD. Then circle with PH as radius becomes smaller, and the ring between the circles with PiM and PR as radii becomes correspondingly smaller. Finally, when PR becomes identical with AD, the radius PH vanishes, leaving the point A, while the ring between the two circles PM and PR contracts into one periphery with AD as radius. From equation (1) it may be concluded that the point A takes up as much area as the circumference of the circle with AD as radius. Bolzano realized that there is only an appearance of a paradox. The two classes of points, one composed of a single member, the point A^ the other of the points in the circumference of the circle with AB as radius, take up exactly the same amount of area. The area of both is zero! The paradox springs from the erroneous conception that the number of points in a given configuration is an indication of the area which it occupies. Points, finite or infinite in number, have no dimensions and can therefore occupy no area. Through the centuries such paradoxes had piled up. Born of the union of vague ideas and vague philosophical reflections, they were nurtured on sloppy thinking. Bol¬ zano cleared away most of the muddle, preparing the way for Cantor. It is to Cantor that the mathematics of the in¬ finitely large owes its coming of age. * Georg Cantor was born in St. Petersburg in 1845, SIX years before Bolzano’s book appeared. Though born m Russia, he lived the greater part of his life in Germany, Beyond the Googol 43 where he taught at the University of Halle. While Weier- strass was busy disposing of the infinitesimal, Cantor set himself the apparently more formidable task at the other pole. The infinitely small might be laughed out of existence, but who dared laugh at the infinitely large? Certainly not Cantor! Theological curiosity prompted his task, but the mathematical interest came to subsume every other. In dealing with the science of the infinite, Cantor realized that the first requisite was to define terms. His definition of “infinite class” which we shall paraphrase, rests upon a paradox, an infinite class has the unique PROPERTY THAT THE WHOLE IS NO GREATER THAN SOME OF ITS PARTS. That statement is as essential for the mathe¬ matics of the infinite as the whole is greater than any OF ITS PARTS is for finite arithmetic. When we recall that two classes are equal if their elements can be put into one-to-one correspondence, the latter statement be¬ comes obvious. Zeno would not have challenged it, in spite of his scepticism about the obvious. But what is obvious for the finite is false for the infinite; our extensive experience with finite classes is misleading. Since, for example, the class of men and the class of mathemati¬ cians are both finite, anyone realizing that some men arc not mathematicians would correctly conclude that the class of men is the larger of the two. He might also conclude that the number of integers, even and odd, is greater than the number of even integers. But we see from the following pairing that he would be mistaken: 1 1 2 2 T i 4 3 t 6 4 5 6 1 ... t t t t i 4 ' i f 8 10 12 14 ... 44 Mathematics and the Imagination Under every integer, odd or even, we may write its double—an even integer. That is, we place each of the elements of the class of all the integers, odd and even, into a one-to-one correspondence with the elements of the class composed solely of even integers. This process may be continued to the googolplex and beyond. Now, the class of integers is infinite. No integer, no matter how great, can describe its cardinality (or numer- osity). Yet, since it is possible to establish a one-to-one correspondence between the class of even numbers and the class of integers, we have succeeded in counting the class of even numbers just as we count a finite collection. The two classes being perfectly matched, we must con¬ clude that they have the same cardinality. That their cardinality is the same we know, just as we knew that the chairs and the people in the hall were equal in number when every chair was occupied and no one was left standing. Thus, we arrive at the fundamental paradox of all infinite classes:—There exist component parts of an infinite class which are just as great as the class itself. THE WHOLE IS NO GREATER THAN SOME OF ITS PARTS! The class composed of the even integers is thinned out as compared with the class of all integers, but evidently thinning out” has not the slightest effect on its .cardi¬ nality. Moreover, there is almost no limit to the number • of times this process can be repeated. For instance, there are as many square numbers and cube numbers as there are integers. The appropriate pairings are: ] 23456... 1 2 3 4 5 6... I I I I I I I I I I I I 1 4 9 16 25 36... 1 g 27 64 125 216.. . U 2“ 3^ 4^ 5^ 6^ p 2^ 3^ 4^ 53 Beyond the Googol 45 Indeed, from any denumerable class there can always be removed a denumerably infinite number of denumer- ably infinite classes without affecting the cardinality of the original class. * Infinite classes which can be put into one-to-one cor¬ respondence with the integers, and thus “counted,’’ Cantor called countable^ or denumerably infinite. Since all finite sets are countable, and we can assign to each one a number, it is natural to try to extend this notion and assign to the class of all integers a number representing its cardinality. Yet, it is obvious from our description of “infinite class” that no ordinary integer would be ade¬ quate to describe the cardinality of the whole class of in¬ tegers. In effect, it would be asking a snake to swallow itself entirely. Thus, the first of the transfinite numbers was created to describe the cardinality of countable infinite classes. Etymologically old, mathematically new, (aleph), the first letter of the Hebrew alphabet, was suggested. However, Cantor finally decided to use the compound symbol (Aleph-Null). If asked, “How many integers are there?” it would be correct to reply, “There are integers.” Because he suspected that there were other transfinite numbers, in fact an infinite number of transfinites, and the cardinality of the integers the smallest. Cantor affixed to the first N a small zero as subscript. The cardinality of a denumerably infinite class is therefore referred to as Xo (Aleph-Null). The anticipated transfinite numbers form a hierarchy of alephs; ^<3 . . . All this may seem very strange, and it is quite excus¬ able for the reader by now to be thoroughly bewildered. Yet, if you have followed the previous reasoning step 4 ^ Mathematics and the Imagination by step, and will go to the trouble of rereading, you will see that nothing which has been said is repugnant to straight thinking. Having established what is meant by counting in the finite domain, and what is meant by number, we decided to extend the counting process to infinite classes. As for our right to follow such a pro¬ cedure, we have the same right, for example, as those who decided that man had crawled on the surface of the earth long enough and that it was about time for him to fly. It is our right to venture forth in the world of ideas as it is our right to extend our horizons in the physical univer.se. One restraint alone is laid upon us in these adventures of ideas: that we abide by the rules of logic. Upon extending the counting process it was evident at once that no finite number could adequately describe an infinite class. If any number of ordinary arithmetic describes the cardinality of a class, that class must be finite, even though there were not enough ink or enough space or enough time to write the number out. We shall then require an entirely new kind of number, nowhere to be found in finite arithmetic, to describe the cardi¬ nality of an infinite class. Accordingly, the totality of inte¬ gers was assigned the cardinality “aleph.” Suspecting that there were other infinite classes with a cardinality greater than that of the totality of integers, we supposed a whole hierarchy of alephs, of which the cardinal number of the totality of integers was named Aleph-Null to indicate it was the smallest of the transfinites. Having had an interlude in the form of a summary, let us turn once more to scrutinize the alephs, to find if, upon closer acquaintance, they may not become easier to understand. The arithmetic of the alephs bears little resemblance Beyond the Googol 47 to that of the finite integers. The immodest behavior of No is typical. A simple problem in addition looks like this: No + 1 = No No + googol = No No + No = No The multiplication table would be easy to teach, easier to learn: .1 X No = No 2 X No = No 3 X No = No « X No = No where n represents any finite number. Also, (No) 2 = No X No = No And thus, (No)" = No when is a finite integer. There seems to be no variation of the theme; the monotony appears inescapable. But it is all very deceptive and treacherous. We go along obtaining the same result, no matter what we do to No, when suddenly we try: (No)''^ This operation, at last, creates a new transfinite. But before considering it, there is more to be said about countable classes. * Common sense says that there are many more fractions than integers, for between any two integers there is an in- 48 Mathematics and the Imagination finite number of fractions. Alas—common sense is amidst alien corn in the land of the infinite. Cantor discovered a simple but elegant proof that the rational fractions form a denumerably infinite sequence equivalent to the class of integers. Whence, this sequence must have the same car¬ dinality.* The set of all rationed fractions is arranged, not in order of increasing magnitude, but in order of ascending numerators and denominators in an array: FIG. 12.—Cantor's diagonal array. Since each fraction may be written as a pair of integers, i.e., f as (3,4). the familiar one-to-one correspondence It has been suggested that at this point the tired reader puts the book down with a sigh—and goes to the movies. VVe can only offer Beyond the Googol with the integers may be effected. This is illustrated in the above array by the arrows. 1 2 3 4 5 6 7 8 I I I I I I I I (1,1) (2,1) (1,2) (1,3) (2,2) (3,1) (4.1) (3,2) (2,3) Cantor also found, by means of a proof (too technical to concern us here) based on the “height" of algebraic equations, that the class of all algebraic numbers, num¬ bers which are the solutions of algebraic equations with integer coefficients, of the form: ^ + . . . + = 0 is denumerablv infinite. But Cantor felt that there were other transfinites, that there were classes which were not countable, which could not be put into one-to-one correspondence with the integers. And one of his greatest triumphs came when he succeeded in showing that there are classes with a cardinality greater than No. The class of real numbers composed of the rational and irrational numbers! is such a class. It contains those irrationals which are algebraic as well as those which are not. The latter are called transcendental numbers^ in mitigation that this proof, like the one which follows on the non¬ countability of the real numbers, is tough and no bones about it. You may grit your teeth and try to get what you can out of them, or conveniently omit them. The essential thing to come away with is that Cantor found that the rational fractions are countable but that the set of real numbers is not. Thus, in spite of what common sense; tells you, there are no more fractions than there are integers and there are more real numbers between 0 and 1 than there are elements in the whole class of integers. ^ Irrational numbers are numbers which cannot be expressed as rational fractions. For example, \/2, VX e, tt. The class of real numbers is made up of rationals like 1, 2, 3, i, ‘ .J, and irrationals as above. Mathematics and the Imagination Two important transcendental numbers were known to exist in Cantor’s time: tt, the ratio of the circumference of a circle to its diameter, and the base of the natural logarithms. Little more was known about the class of transcendcntals: it was an enigma. What Cantor had to prove, in order to show that the class of real numbers was nondenumerable (i.e., too big to be counted by the class of integers), was the unlikely fact that the class of transcendcntals was nondenumerable. Since the rational and the algebraic numbers were known to be denumer¬ able, and the sum of any denumerable number of de¬ numerable classes is also a denumerable class, the sole remaining class which could make the totality of real numbers nondenumerable was the class of transcendcntals. He was able to devise such a proof. If it can be shown that the class of real numbers between 0 and 1 is non- dcnumerable, it will follow a Jortiori that all the real numbers arc nondenumerable. Employing a device often used in advanced mathematics, the reductio ad absurdum^ Cantor assumed that to be true which he suspected was false, and then showed that this assumption led to a contradiction. He assumed that the real numbers be¬ tween 0 and 1 were countable and could, therefore, be paired with the integers. Having proved that this as¬ sumption led to a contradiction, it followed that its opposite, namely, that the real numbers could not be paired with the integers (and were therefore not count¬ able), was true. To count the real numbers between 0 and 1, it is required that they all be expressed in a uniform way and a method of writing them down in order be devised so that they can be paired one to one with the integers. The first requirement can be fulfilled, for it is possible Beyond the Googol to express every real number as a nonterminating dec¬ imal. Thus, for example; ® 1 3 1 9 . 3333 ... . 1111111 ... . 2142857121428571 ... 1 . 414 . . . - = .707 Now, the second requirement confronts us. How shail we make the pairings? What system ma>- be devised to en¬ sure the appearance of every decimal? We did find a method for ensuring the appearance of every rational fraction. Of course, we could not actually write them all, any more than we could actually write all the integers; but the method of increasing numerators and denomina¬ tors was so explicit that, if we had had an infinite time in which to do it, we could actually have set down all the fractions and have been certain that we had not omitted any. Or, to put it another way: It was always certain and determinate after a fraction had been paired with an integer, what the next fraction would be, and the next, and the next, and so on. On the other hand, when a real number, expressed as a nonterminating decimal, is paired with an integer, what method is there for determining what the next decimal in order should be? You have only to ask your¬ self, which shall be the first of the nonterminating dec¬ imals to pair with the integer 1, and you have an inkling of the difficulty of the problem. Cantor however assumed that such a pairing does exist, without attempting to give its explicit form. His scheme was: With the integer 1 pair the decimal .aia 2 a 3 . . . , with the integer 2, .bib 2 b 3 . . . , etc. Each of the letters represents a digit of the nonterminating decimal in which it appears. The 5 52 Mathematics and the Imagination determinate array of pairing between the decimals and the integers would then be: —^ 0 . a\ a2 63 ^4 ^5 . 3 <—> 0. iTi C2 Cz Cs . 4 ^ * 0 . (/2 di d\ di, . 5 <—► 0. ^2 ^3 ^4 ^6 . The new decimal may be written:— 0. ai 02 03 04 as . . .; where oi differs from oi, 03 differs from /> 2 , 0.3 from ^3 04 from d^y 03 from ^ 5 , etc. Accordingly, it will differ from each decimal in at least one place, from the «th decimal in at least its nth place. This proves conclusively that there is no way of including all the decimals in any possible array, no way of pairing them off with the inte¬ gers. Therefore, as Cantor set out to prove: 1. The class of transcendental numbers is not only infinite, but also not countable, i.e., nondcnumerably infinite. 2. The real numbers between 0 and 1 are infinite and not countable. 3. A fortiori, the class of all real numbers is nondenumerable. ♦ To the noncountable class of real numbers, Cantor as¬ signed a new transfinite cardinal. It was one of the alephs, but which one remains unsolved to this day. It is sus¬ pected that this transfinite, called the “cardinal of the continuum,” which is represented by c or C, is identical with Ki. But a proof acceptable to most mathematicians has yet to be devised. 54 Mathematics and the Imagination The arithmetic of C is much the same as that of Xo* The multiplication table has the same dependable mon¬ otone quality. But when C is combined with Xo> it swal¬ lows it completely. Thus: C + Xo = C <7- Xo = C C X X 0 = C and even C X C = C Again, we hope for a variation of the theme when we come to the process of involution. Yet, for the moment, we are disappointed, for C^o = C. But just as (Xo)^® does not equal Xo, so does not equal C. We are now in a position to solve our earlier problem in involution, for actually Cantor found that (Xo)^° = C. Likewise gives rise to a new transfinite, greater than C. This transfinite represents the cardinality of the class of all one-valued functions. It is also one of the X’s, but again, which one is unknown. It is often designated by the letter F.® In general, the process of involution, when re¬ peated, continues to generate higher transfinites. Just as the integers served as a measuring rod for classes with the cardinality Xo, the class of real numbers serves as a measuring rod for classes with the cardinality C. Indeed, there are classes of geometric elements which can be measured in no other way except by the class of real numbers. From the geometric notion of a point, the idea is evolved that on any given line segment there are an infinite number of points. The points on a line segment are also, as mathematicians say, “everywhere dense.” This means that between any two points there is an infinitude of others. The concept of two immediately adjoining points is, therefore, meaningless. This property of being “everywhere dense,” constitutes one of the es- Beyond the Googol 55 sential characteristics of a continuum. Cantor, in referring to the “cardinality of the continuum,” recognized that it applies alike to the class of real numbers and the class of points on a line segment. Both are everywhere dense, and both have the same cardinality, C. In other words, it is possible to pair the points on a line segment with the real numbers. Classes with the cardinality C possess a property similar to classes with the cardinadity No: they may be thinned out without in any way affecting their cardinality. In this connection, we see in very striking fashion another illustration of the principle of transfinite arithmetic, that the whole is no greater than many of its parts. For instance, it can be proved that there are as many points on a line one foot long as there are on a line one yard long. The line segment AB in Fig. 13 is three times as long as the line A*B‘. Nevertheless, it is possible to put the class of all points on the segment AB into a one- to-one correspondence with the class of points on the segment A'B\ L FIG. 13. Let L be the intersection of the lines AA' and BB'. If then to any point Ad of AB^ there corresponds a point 5 ^ Mathematics and the Imagination M' oi A'B\ which is on the line LM, we have established the desired correspondence between the class of points on A'B' and those on AB. It is easy to see intuitively and to prove geometrically that this is always possible, and that, therefore, the cardinality of the two classes of points is the same. Thus, since A*B^ is smaller than AB^ it may be considered a proper part of AB^ and we have again established that an infinite class may contain as proper parts, subclasses equivalent to it. There are more startling examples in geometry which illustrate the power of the continuum. Although the statement that a line one inch in length contains as many points as a line stretching around the equator, or as a line stretching from the earth to the most distant stars, is startling enough, it is fantastic to think that a line segment one-millionth of an inch long has as many points as there are in all three-dimensional space in the entire universe. Nevertheless, this is true. Once the principles of Cantor s theory of transfinites is understood, such state¬ ments cease to sound like the extravagances of a mathe¬ matical madman. The oddities, as Russell has said, “then become no odder than the people at the antipodes who used to be thought impossible because they would find it so inconvenient to stand on their heads.*’ Even conceding that the treatment of the infinite is a form of mathemati¬ cal madness, one is forced to admit, as does the Duke in Measure for Measure: If she be mad,—as I believe no other,— Her madness hath the oddest frame of sense, Such a dependency of thing on thing. As e’er I heard in madness.” ♦ Until now we have deliberately avoided a definition Beyond the Googol 57 of “infinite class.” But at last our equipment makes it possible to do so. We have seen that an infinite class, whether its cardinality is C, or greater, may be thinned out in a countless variety of ways, without affecting its cardinality. In short, the whole is no greater than many of its parts. Now, this property does not belong to finite classes at all; it belongs only to infinite classes. Hence, it is a unique method of determining whether a class is finite or infinite. Thus, our definition reads: An infinite class is one which can be put into one-to-one reciprocal correspondence with a proper subset of itself. Equipped with this definition and the few ideas we have gleaned we may re-examine some of the paradoxes of Zeno. That of Achilles and the tortoise may be ex¬ pressed as follows: Achilles and the tortoise, running the same course, must each occupy the same number of distinct positions during their race. However, if Achilles is to catch his more leisurely and determined opponent, he will have to occupy more positions than the tortoise, in the same elapsed period of time. Since this is man¬ ifestly impossible, you may put your money on the tortoise. But don’t be too hasty. There are better ways of saving money than merely counting change. In fact, you had best bet on Achilles after all, for he is likely to win the race. Even though we may not have realized it, we have just finished proving that he could overtake the tortoise by showing that a line a millionth of an inch long has just as many points as a line stretching from the earth to the furthest star. In other words, the points on the tiny line segment can be placed into one-to-one corre¬ spondence with the points on the great line, for there is no relation between the number of points on a line 58 Mathematics and the Imagination and its length. But this reveals the error in thinking that Achilles cannot catch the tortoise. The statement that Achilles must occupy as many distinct positions as the tortoise is correct. So is the statement that he must travel a greater distance than the tortoise in the same time. The only incorrect statement is the inference that since he must occupy the same number of positions as the tortoise he cannot travel further while doing so. Even though the classes of points on each line, which cor¬ respond to the several positions of both Achilles and the tortoise are equivalent, the line representing the path of Achilles is much longer than that representing the path of the tortoise. Achilles may travel much further than the tortoise without successively touching more points. The solution of the paradox involving the arrow in flight requires a word about another type of continuum. It is convenient and certainly familiar to regard time as a continuum. The time continuum has the same properties as the space continuum: the successive instants in any elapsed portion of time, just as the points on a line, may be put into one-to-one correspondence with the class of real numbers; between any two instants of time an infinity of others may be interpolated; time also has the mathematical property mentioned before—it is every¬ where dense. Zeno’s argument stated that at every instant of time the arrow was somewhere, in some place or position, and therefore, could not at any instant be in motion. Although the statement that the arrow had at every moment to be in some place is true, the conclusion that, therefore, it could not be moving is absurd. Our natural tendency to accept this absurdity as true springs from our firm conviction that motion is entirely different from rest. Beyond the Googol We are not confused about the position of a body when it is at rest we feel there is no mystery about the state of rest. We should feel the same when we consider a body in motion. When a body is at rest, it is in one position at one instant of time and at a later instant it is still in the same position. When a body is in motion, there is a one-to-one correspondence between every instant of time and every new position. To make this clear we may construct two tables: One will describe a body at rest, the other, a body in motion. The “rest” table will tell the life history MOTION On Bedloe’s Island 9 a.m. In the city On Bedloe’s Island 11 a.m. Over the river On Bedloe’s Island 3 p.m. In the mountains. FIG. 14.—At the times shown, the Statue of Liberty is at the point shown, while the taxi’s passengers see the different scenes shown at the right. 6o Mathematics and the Imagination and the life geography of the Statue of Liberty, while the “motion” table will describe the Odyssey of an auto¬ mobile. The tables indicate that to every instant of time there corresponds a position of the Statue of Liberty and of the taxi. There is a one-to-one space-time correspondence for rest as well as for motion. No paradox is concealed in the puzzle of the arrow when we look at our table. Indeed, it would be strange if there were gaps in the table; if it were impossible, at any instant, to determine exactly what the position of the arrow is. Most of us would swear by the existence of motion, but we are not accustomed to think of it as something which makes an object occupy different positions at different instants of time. We are apt to think that motion endows an object with the strange property of being continually nowhere. Impeded by the limitations of our senses which prevent us from perceiving that an object in motion simply occupies one position after another and does so rather quickly, we foster an illusion about the nature of motion and weave it into a fairy tale. Mathe¬ matics helps us to analyze and clarify what we perceive, to a point where we are forced to acknowledge, if we no longer wish to be guided by fairy tales, that we live either in Mr. Russell’s changeless world or in a world where motion is but a form of rest. The story of motion is the same as the story of rest. It is the same story told at a quicker tempo. The story of rest is: “It is here.” The story of motion is; “It is here, it is there.” Because, in this re¬ spect, it resembles Hamlet’s father’s ghost is no reason to doubt its existence. Most of our beliefs are chained to less substantial phantoms. Motion is perhaps not easy for our Beyond the Googol 61 senses to grasp, but with the aid of mathematics, its essence may first be properly understood. * At the beginning of the twentieth century it was generally conceded that Cantor’s work had clarified the concept of the infinite so that it could be talked of and treated like any other respectable mathematical concept. The controversy which arises wherever mathematical philosophers meet, on paper, or in person, shows that this was a mistaken view. In its simplest terms this con¬ troversy, so far as it concerns the infinite, centers about the questions: Does the infinite exist? Is there such a thing as an infinite class? Such questions can have little meaning unless the term mathematical “existence” is first explained. In his famous “Agony in Eight Fits,” Lewis Carroll hunted the snark. Nobody was acquainted with the snark or knew much about it except that it existed and that it was best to keep away from a boojum. The infinite may be a boojum, too, but its existence in any form is a matter of considerable doubt. Boojum or garden variety, the infinite certainly does not exist in the same sense that we say, “There are fish in the sea.” For that matter, the statement “There is a number called 7” refers to something which has a different existence from the fish in the sea. “Existence” in the mathematical sense is wholly different from the existence of objects in the physical world. A billiard ball may have as one of its properties, in addition to whiteness, roundness, hardness, etc., a relation of circumference to diameter involving the number tt. We may agree that the billiard ball and TT both exist; we must also agree that the billiard ball and TT lead different kinds of lives. 62 Mathematics and the Imagination There have been as many views on the problem of existence since Euclid and Aristode as there have been philosophers. In modern times, the various schools of mathematical philosophy, the Logistic school. Formalists, and Intuitionists, have all disputed the somewhat less than glassy essence of mathematical being. All these dis¬ putes are beyond our ken, our scope, or our intention. A stranger company even than the tortoise, Achilles, and the arrow, have defended the existence of infinite classes —defended it in the same sense that they would defend the existence of the number 7. The Formalists, who think mathematics is a meaningless game, but play it with no less gusto, and the Logistic school, which considers that mathematics is a branch of logic—both have taken Cantor’s part and have defended the alephs. The defense rests on the notion of self-consistency. “Existence” is a metaphysical expression tied up with notions of being and other bugaboos worse even than boojums. But the ex¬ pression, “self-consistent proposition” sounds like the language of logic and has its odor of sanctity. A propo¬ sition which is not self-contradictory is, according to the Logistic school, a true existence statement. From this standpoint the greater part of Cantor’s mathematics of the infinite is unassailable. New problems and new paradoxes, however, have been discovered arising out of parts of Cantor’s structure because of certain difficulties already inherent in class¬ ical logic. They center about the use of the word “all.” The paradoxes encountered in ordinary parlance, such as “All generalities are false including this one,” con¬ stitute a real problem in the foundations of logic, just as did the Epimenides paradox whence they sprang. In the Epimenides, a Cretan is made to say that all Cretans are Beyond the Googol 63 liars, which, if true, makes the speaker a liar for telling the truth. To dispose of this type of paradox the Logistic school invented a “Theory of Types.” The theory of types and the axiom of reducibility on which it is based must be accepted as axioms to avoid paradoxes of this kind. In order to accomplish this a reform of classical logic is required which has already been undertaken. Like most reforms it is not wholly satisfactory—even to the reformers—but by means of their theory of types the last vestige of inconsistency has been driven out of the house that Cantor built. The theory of transfinites may still be so much nonsense to many mathematicians, but it is certainly consistent. The serious charge Henri Poin¬ care expressed in his aphorism, “La logistique n’est plus sterile; elle engendre la contradiction,” has been success¬ fully rebutted by the logistic doctrine so far as the infinite is concerned. To Cantor’s alephs then, we may ascribe the same existence as to the number 7. An existence statement free from self-contradiction may be made relative to either. For that matter, there is no valid reason to trust in the finite any more than in the infinite. It is as permissible to discard the infinite as it is to reject the impressions of one’s senses. It is neither more, nor less scientific to do so. In the final analysis, this is a matter of faith and taste, but not on a par with rejecting the belief in Santa Claus. Infinite classes, judged by finite standards, generate para¬ doxes much more absurd and a great deal less pleasing than the belief in Santa Claus; but when they are judged by the appropriate standards, they lose their odd appear¬ ance, behave as predictably as any finite integer. At last in its proper setting, the infinite has assumed a respectable place next to the finite, just as real and just as Mathematics and the Imagination dependable, even though wholly different in character. Whatever the infinite may be, it is no longer a purple cow. FOOTNOTES 1. We distinguish cardinal from ordinal numberSy which denote the re¬ lation of an element in a class to the others, with reference to some system of order. Thus, we speak of the first Pharaoh of Egypt, or of the fourth integer, in their customary order, or of the third day of the week, etc. These are examples of ordinals. P. 30. 2. For the definition of primes, see the Chapter on pie. —P. 32. 3. This series is said to converge to a limit —1. Discussion of this concept must be postponed to the chapters on pie and the cal¬ culus.—P. 38. 4. A transcendental number is one which is not the root of an algebraic equation with integer coefficients. See pie. —P. 49. 5. Any terminating decimal, such as .4, has a nonterminating form .3999. . . —P. 51. 6. A simple geometric interpretation of the class of all one-valued functions F is the following: With each point of a line segment, associate a color of the spectrum. The class F is then composed of all possible combinations of colors and points that can be conceived.—P. 54. PIE (TrJ.e) Transcendental and Imaginary In order to reach the Truth, it is necessary, once in one's life, to put everything in doubt—so far as possible. —DESCARTES Perhaps pure science begins where common sense ends; perhaps, as Bergson says, “Intelligence is characterized by a natural lack of comprehension of life.’' ' But we have no paradoxes to preach, no epigrams to sell. It is only that the study of science, particularly mathematics, often leads to the conclusion that one need only say that a thing is unbelievable, impossible, and science will prove him wrong. Good common sense makes it plain that the earth is flat and stands still, that the Chinese and the Antipo- deans walk about suspended by their feet like chandeliers, that parallel lines never meet, that space is infinite, that negative numbers are as real as negative cows, that -1 has no square root, that an infinite series must have an infinite sum, or that it must be possible with ruler and compass alone to construct a square exactly equal in area to a given circle. Just how far have we been carried by common sense in arriving at these conclusions? Not very far! Yet some of the statements seem quite plausible, even inescapable. It would be wrong to say that science has proved that all are false. We may still cling to the Euclidean hypothesis that parallel lines never meet and remain always equi- 66 Mathematics and the Imagination distant, as long as we remember it is merely a hypothesis, but the statements about the squaring of the circle, the square root of —1, and about infinite series belong in a different category. The circle can not be squared with ruler and compass, — 1 has a square root. An infinite series can have a finite sum. Three symbols, tt, e, have enabled mathemati¬ cians to prove these statements, three symbols which rep¬ resent the fruits of centuries of mathematical research. How do they stand up to common sense? * The most famous problem in the entire history of math¬ ematics is the “squaring of the circle.’^ Two other prob¬ lems which challenged Greek geometers, the “duplication of the cube” and the “trisection of an angle,” may, as a matter of interest, be briefly considered with the first, even though squaring the circle alone involves tt. In the infancy of geometry, it was discovered that it was possible to measure the area of a figure bounded by straight lines. Indeed, geometry was devised for that very purpose—to measure the fields in the valley of the Nile, where each year the floods from the rising river obliter¬ ated every mark made by the farmer to indicate which fields were his and which his neighbors. Measuring areas bounded by curved lines presented greater difficulties, and an effort was made to reduce every problem of this type to one of measuring areas with straight boundaries. Clearly, if a square can be constructed with the area of a given circle, by measuring the area of the square, that of the circle is determined. The expression “squar¬ ing the circle” derives its name from this approach. The number tt is the ratio of the circumference of a circle to its diameter. The area of a circle of radius r is PIE (tt, i, e)—Transcendental and Imaginary 67 given by the formula Now the area of a square with side of length A is A^. Thus, the algebraic statement: A^ = irr^ expresses the equivalence in area between a given square and a circle. Taking square roots of both sides of this equation yields A = As r is a known quantity, the problem of squaring the circle is, in effect, the computation ^ of the value of tt. Since mathematicians have succeeded in computing TT with extraordinary exactitude, what then is meant by the statement, “It is impossible to square the circle?’* Unfortunately, this question is still shrouded in many misapprehensions. But these would vanish if the problem were understood. * Squaring the circle is proclaimed impossibUy but what does “imfxjssible” mean in mathematics? The first steam vessel to cross the Atlantic carried, as part of its cargo, a book that “proved” it was impossible for a steam vessel to cross anything, much less the Atlantic. Most of the savants of two generations ago “proved” that it would be forever impossible to invent a practical heavier-than-air flying machine. The French philosopher, Auguste Comte, demonstrated that it would always be impossible for the human mind to discover the chemical constitution of the stars. Yet, not long after this statement was made the spectroscope was applied to the light of the stars, and we now know more about their chemical constitution, in¬ cluding those of the distant nebulae, than we know about the contents of our medicine chest. As just one illustra¬ tion, helium was discovered in the sun before it was found in the earth. Museums and patent offices are filled with cannons, clocks, and cotton gins, already obsolete, each of which 6 68 Mathematics and the Imagination confounded predictions that their invention would be im¬ possible. A scientist who says that a machine or a project is impossible only reveals the limitations of his day. What¬ ever the intentions of the prophet, the prediction has none of the qualities of prophecy. “It is impossible to fly to the moon” is meaningless, whereas “We have not yet devised a means of flying to the moon” is not. Statements about impossibility in mathematics are of a wholly different character. A problem in mathematics which may not be solved for centuries to come is not always impossible. “Impossible” in mathematics means theoretically impossible, and has nothing to do with the present state of our knowledge. “Impossible” in mathe¬ matics does not characterize the process of making a silk purse out of a sow’s ear, or a sow’s ear out of a silk purse; it does characterize an attempt to prove that 7 dmes 6 is 43 (in spite of the fact that people not good at arithmetic often achieve the impossible). By the rules of arithmetic 7 times 6 is 42, just as by the rules of chess, a pawn must make at least 5 moves before it can be queened. Where theoretical proof that a problem cannot be solved is lacking, it is legitimate to attempt a solution, no matter how improbable the hope of success. For centuries the construction of a regular polygon of 17 sides was rightly considered difficult, but falsely considered im¬ possible, for the nineteen-year-old Gauss in 1796 suc¬ ceeded in finding an elementary construction.® On the other hand, many famous problems, such as Fermat’s Last Theorem,^ have defied solution to this day in spite of heroic researches. To determine whether we have the right to say that squaring the circle, trisecting the angle, or duplicating the cube is impossible^ we must find logical proofs, involving purely mathematical reasoning. Once PIE (tt, z, e)—Transcendental and Imaginary 69 such proofs have been adduced, to continue the search for a solution is to hunt for a three-legged biped.^ * Having determined what mathematicians mean by impossible, the bare statement, “It is impossible to square the circle” still remains meaningless. To give it meaning we must specify how the circle is to be squared. When Archimedes said, “Give me a place to stand and I will move the earth,” he was not boasting of his physical powers but was extolling the principle of the lever. When it is said that the circle cannot be squared, all that is meant is that this cannot be done with ruler and compass alone^ although with the aid of the integraph or higher curves the operation does become possible. Let us repeat the problem; It is required to construct a square equal in area to a given circle, by means of an exact theoretical plan, using only two instruments: the ruler and compass. By a ruler is meant a straightedge, that is, an instrument for drawing a straight line, not for measuring lengths. By a compass is meant an instru¬ ment with which a circle with any center and any radius can be drawn. These instruments are to be used a finite number of times, so that limits or converging processes with an infinite number of steps may not be employed.® The construction, by purely logical reasoning, depending only on Euclid’s axioms and theorems, is to be absolutely exact. The concepts of “limit” and “convergence” are more fully explained elsewhere,’ but a word about them here is in place. Consider the familiar series 1 + ^ + l + 3 T "t" • - • The sum of the first 5 terms of this series is 1.9375; the sum of the first 10 terms is 1.9980 . . . ; the 70 Mathematics and the Imagination sum of the first 15 is 1.999781 . . . What is readily appar¬ ent is that this series tends to choke off, i.e., the additional terms which are added become so small that even a vast number will not cause the series to grow beyond a finite bound. In this instance the bound, or limit, is 2. Such a series which chokes off is said to ^^convergd"^^ to a ^^limitJ*^ FIG. 15,—An infinite number of terms with a finite sum. If the width of the first block is one foot, the width of the second foot, of the third \ foot, of the fourth J foot, and so one, then an infinite number of blocks rests on the 2-foot bar, that is: The geometric analogues of the concepts of limit and convergence are equally fruitful. A circle may be re¬ garded as the limit of the polygons with increasing num¬ ber of sides which may be successively inscribed in it, or circumscribed about it, and its area as the common limit of both of these sets of polygons. This is not a rigorous definition of limit and conver¬ gence, but too often mathematical rigor serves only to bring about another kind of rigor —rigor mortis of math¬ ematical creativencss. PIE (tt, i, e)—Transcendental and Imaginary 71 To return to squaring the circle: the Greeks, and later mathematicians, sought an exact construction with ruler and compass, but always failed. As we shall see later, all ruler and compass constructions are geometric equiva¬ lents of first- and second-degree algebraic equations and combinations of such equations. But the German mathe¬ matician Lindemann, in 1882, published a proof that tt is a transcendental number and thus any equation which satisfies it cannot be algebraic and surely not algebraic of first or second degree. It follows that the statement, “The squaring of the circle is impossible with ruler and compass alone,” is meaningful. So far as the other two problems are concerned, thanks in part to the work of “the marvelous boy . . . who perished in his prime,” the sixteen-year-old Galois, it was established about one hundred years ago that the dupli¬ cation of the cube and the trisection of an angle are also impossible with ruler and compass. We may allude to them briefly. There is a story among the Greeks that the problem of duplicating the cube originated in a visit to the Delphic oracle. There was an epidemic raging at the time, and the oracle said the epidemic would cease only if a cubical altar to Apollo were doubled in size. The masons and architects made the mistake of doubling the side of the cube, but that made the volume eight times as great. Of course the oracle was not satisfied, and the Greek math¬ ematicians, on re-examining the problem began to see that the right answer involved, not doubling the side, but multiplying it by the cube root of 2. This could not be done geometrically with ruler and compass. They finally succeeded by using other instruments and higher curves. The oracle was appeased and the epidemic 72 Mathematics and the Imagination ceased. You may believe the story or not, much as you choose, but you cannot “duplicate the cube.”® The trisection of an angle has received a good deal of attention in the newspapers during the past few years because monographs continue to crop up which claim to solve the problem completely. The fallacies contained in these “solutions” are of four kinds: they are sometimes merely approximate and not exact; instruments other than the ruler and compass are occasionally used, either wittingly or unwittingly; at times there is a logical fallacy in the intended proof; and often only special and not general angles are considered. An angle can be bisected but not trisected by elementary geometry, since the first problem involves merely square roots, while the second involves cube roots, which, as we have stated, cannot be constructed with ruler and compass. ★ The difficulty in squaring the circle, as stated at the outset, lies in the nature of the number tt. This remark¬ able number, as Lindemann proved, cannot be the root of an algebraic equation with integer coefficients.^® It is therefore not expressible by rational operations, or by the extraction of square roots, and as only such operations can be translated into an equivalent ruler and compass construction, it is impossible to square the circle. The pa¬ rabola is a more complicated curve than a circle, but nevertheless, as Archimedes knew, any area bounded by a parabola and a straight- line can be determined by ra¬ tional operations, and hence the “parabola can be squared.” Lindemann’s proof is too technical to concern us here. If, however, we consider the history and development of TT, we shall be in a better position to understand its PIE (x, z, e)—Transcendental and Imaginary 73 purpose without being compelled to master its difficulties. If a triangle is inscribed in a circle (fig. 16), the area of the inscribed triangle will be less than the area of the circle: no. 16.—^The circle as the limit of inscribed and circumscribed polygons. The difference between the area of the circle and the triangle are the three shaded portions of the circle. Now consider the same circle with a triangle circumscribed about it (Fig. 16). The area of the circumscribed triangle will be greater than the area of the circle. The three shaded portions of the triangle again represent the difference in area. It may readily be seen that if the number of sides of the inscribed figure is doubled, the area of the resulting hexagon will be less than the area of the circle, but closer to it than the area of the inscribed triangle. Similarly, if the number of sides of the circumscribed triangle is doubled, the area of the circumscribed hexagon will still be greater than the area of the circle but, again, closer to *74 Mathematics and the Imagination it than the area of the circumscribed triangle. By well- known, simple, geometric methods, employing only ruler and compass, the number of sides of the inscribed and circumscribed polygons may be doubled as many times as desired. The area of the successively inscribed polygons will approach that of the circle, but will always remain slightly less; the area of the circumscribed polygons will also approach that of the circle but their area will always remain slightly greater. The common value approached by both is the area of the circle. In other words, the circle is the limit of these two series of polygons. If the radius of the circle is equal to 1, its area, which equals ttt^, is simply TT. This method of increasing and decreasing polygons for computing the value of tt was known to Archimedes, who, employing polygons of 96 sides, showed that tt is less than and greater than • Somewhere in between lies the area of the circle. Archimedes’ approximation for tt is considerably closer than that given in the Bible. In the Book of Kings, and in Chronicles, tt is given as 3. Egyptian mathematicians gave a somewhat more accurate value—3.16. The fa¬ miliar decimal—3.1416, used in our schoolbooks, was already known at the time of Ptolemy in 150 a.d. Theoretically, Archimedes’ method for computing tt by increasing the number of sides of the polygons may be extended indefinitely, but the requisite calculations soon become very cumbersome. None the less, during the Middle Ages such calculations were zealously carried out. Francisco Vieta, the most eminent mathematician of the sixteenth century, though not a professional, made a great advance in the calculation of tt in determining its PIE (tt, 2 , e)—Transcendental and Imaginary 75 value to ten decimal places. In addition to giving the formula: a nonterminating product, and making many other im¬ portant mathematical discoveries, Vieta rendered service to King Henry IV of France, in the war against Spain, by deciphering intercepted letters addressed by the Span¬ ish Crown to its governors of the Netherlands, The Span¬ iards were so impressed that they attributed his discovery of the cipher key to magic. It was neither the first nor the last time that the efforts of mathematicians were branded as necromancy. In 1596 Ludolph van Ceulen, the German mathe¬ matician, long a resident in Holland, calculated 35 decimal places for tt. Instead of the epitaph, “died at 40, buried at 60,” appropriate where cerebration ceases just when life is supposed to begin, van Ceulen, who worked on tt almost to the day of his death at the age of 70, requested that the 35 digits of tt which he had com¬ puted be inscribed as a fitting epitaph on his tombstone. This was actually done. The value he gave for tt is, in part, 3.14159 26535 89793 23846... In memory of his achievement the Germans still call this number the Lu- dolphian number. We propose to call tt the Archimedean number. * The number tt reached maturity with the invention of the calculus by Newton and Leibniz. The Greek method was abandoned and the purely algebraic device of convergent infinite series, products, and continued fractions came into vogue. John Wallis (1616-1703), the 76 Mathematics and the Imagination Englishman, contributed one of the most famous prod ucts: Leibniz’ infinite series, unlike Wallis’ product for tt, is a sum: The successive products and sums of the terms of these series yield values of tt as accurate as desired. These proc¬ esses, typical of the powerful methods of approximation used not only in mathematics but in the other sciences, although much less cumbersome than the method em¬ ployed by the Greeks, still entail a great deal of calcula¬ tion. The products of Wallis’ series are: 1 ’13 3’1^3^3 9’1^3 3 5 45’ etc. 2 FIG. 17.—Wallis’ product. PIE (tt, i, e)—Transcendental and Imaginary 77 Taking the successive sums of Leibniz’ series, we obtain; 1 , = 1.1 1 76 ’ 3 3 ’ ^ 3 ^ 5 “ 15 ’ 1 — i -|- i _ 1 = 3 ^ 5 7 105 ’ etc. 1 •’A ♦ Vs -Vt y I _ I + ...)- FIG. 18. —Leibniz’ series. ^ = 0.795 . . . 5 = i -1 + 1_1 + !_± + After taking the first 50 terms of these series, the next 50 will not yield an appreciably more accurate value of TT, for the series converge rather slowly. The rapidly convergent series = J- +J_^ 4 \5 3-53 ^5-5^ 7-5^ U _^ + _^ \239 3-2393 ^ 5.2393 7.2397 -r . . -y is much more useful, and is frequently employed in mod¬ ern mathematics. Its relation to tt was established by Machin (1680-1752). Using even more rapidly converg¬ ing series, Abraham Sharp, in 1699, calculated tt to 71 decimal places. Dase, a lightning calculator employed by Gauss, worked out 200 places in 1824. In 1854, Richter computed 500 places, and finally, in 1873, Shanks, an English mathematician, achieved a curious kind of im¬ mortality by determining tt to 707 decimal places. Even 78 Mathematics and the Imagination today it would require 10 years of calculation to deter¬ mine TT to 1000 places. Yet that does not seem like a waste of time as compared with the billions of hours spent by millions of people on crossword puzzles and contract bridge, to say nothing of political debates. Of course Shank’s result has no conceivable use in ap¬ plied science. No more than 10 decimal places for tt are ever needed in the most precise work. The famous Amer¬ ican astonomer and mathematician, Simon Newcomb, once remarked, “Ten decimal places are sufficient to give the circumference of the earth to the fraction of an inch, and thirty decimals would give the circumference of the whole visible universe to a quantity imperceptible with the most powerful telescope.” Why, then, has so much time zind effort been devoted to the calculation of tt? The reason is twofold. First, by studying infinite series mathematicians hoped they might find some clue to its transcendental nature. Second, the fact that tt, a purely geometric ratio, could be evolved out of so many arithmetic relationships—out of infinite series, with apparendy little or no relation to geometry—was a never-ending source of wonder and a never-ending stimulus to mathematical acdvity. Who would imagine—that is, who but a mathemati¬ cian—that the number expressing a fundamental relation between a circle and its diameter could grow out of the curious fraction communicated by Lord Brouncker (1620-1684) to John Wallis? 4 1 + P 2 + 32 2 + 52 2 + 72 . . . TT PIE (tt, z, e)—Transcendental and Imaginary 79 But just such relations between infinite series and tt illustrate the profound connection between most mathe¬ matical forms, geometric or algebraic. It is mere coinci¬ dence, a mere accident that tt is defined as the ratio of the circumference of a circle to its diameter. No matter how mathematics is approached, tt forms an integral part.*^ In his Budget of Paradoxesy Augustus De Morgan illustrated how little the usual definition of tt suggests its origin. He was explaining to an actuary what the chances were that, at the end of a given time a certain proportion of a group of people would be alive, and quoted the for¬ mula employed by actuaries which involves tt. On ex¬ plaining the geometric meaning of tt, the actuary, who had been listening with interest, interrupted and ex¬ claimed, “My dear friend, that must be a delusion. What can a circle have to do with the number of people alive at the end of a given time?” To recapitulate briefly, the problem of squaring the circle turns out to be an impossible construction with ruler and compass alone. The only constructions possible with these instruments correspond to first- and second- degree algebraic equations. Lindcmann proved that tt is not only not the root of a first- or second-degree algebraic equation, but is not the root of any algebraic equation (with integer coefficients), no matter how great the de¬ gree; therefore tt is transcendental. Here, then, is the end of every hope of proving this classical problem in the in¬ tended way. Here is mathematical impossibility. * When the Greek philosophers found that the square root of 2 is not a rational number, t^ey celebrated the discovery by sacrificing 100 oxen. The much more pro¬ found discovery that tt is a transcendental number de- 8o Mathematics and the Imagination serves a greater sacrifice. Again mathematics triumphed over common sense, tt, a finite number the ratio of the circumference of a circle to its diameter is accurately expressible only as the sum or product of an infinite series of wholly different and apparently unrelated numbers. The area of the simplest of all geometric figures, the circle, Ccinnot be determined by finite (Euclidean) means. e In the seventeenth century, perhaps the greatest of all for the development of mathematics, there appeared a work which in the history of British science can be placed second only to Sir Isaac Newton’s monumental Principia. In 1614, John Napier of Merchiston issued his Mirifici Logariihmorum Canonis Descriptio, (“A Description of the Admirable Table of Logarithms”), the first treatise on logarithms.*® To Napier, who also invented the decimal point, we are indebted for an invention which is as im¬ portant to mathematics as Arabic numerals, the concept of zero, and the principle of positional notation.*'* With¬ out these, mathematics would probably not have ad¬ vanced much beyond the stage to which it had been brought 2000 years ago. Without logarithms the com¬ putations accomplished daily with ease by every math¬ ematical tyro would tax the energies of the greatest math¬ ematicians. Since e and logarithms have the Scime genealogical tree and were brought up together, we may for the moment turn our attention to logarithms to ascertain something of the nature of the number e. Stupendous calculations being required to construct trigonometric tables for navigation and astronomy, Na¬ pier was prompted to invent some device to facilitate these computations. Although contemporaries like Vieta PIE (tt, z, e)—Transcendental and Imaginary 8i and Ceulen vied with each other in performing almost unbelievably difficult feats of arithmetic, it was at best a labor of love, an exalted drudgery and self-immolation, with love’s labor often lost as the result of one small slip. Napier succeeded in achieving his purpose, in abbre¬ viating the operations of multiplication and division, op¬ erations “so fundamental in their nature that to shorten them seems impossible.” Nevertheless, by means of loga¬ rithms, every problem in multiplication and division, no matter how elaborate, reduces to a relatively easy one in addition and subtraction. Multiplying and dividing googols and googolplexes becomes as easy as adding a simple column of figures. Like many another of the profound and fecund inven¬ tions of mathematics, the underlying idea was so simple that one wonders why it had not been thought of earlier. Cajori recounts that Henry Briggs (1556-1631), professor of geometry at Oxford, “was so struck with admiration of Napier’s book, that he left his studies in London to do homage to the Scottish philosopher. Briggs was de¬ layed in his journey, and Napier complained to a com¬ mon friend, ‘Ah, John, Mr. Briggs will not come.’ At that very moment knocks were heard at the gate, and Briggs was brought into the lord’s chamber. Almost one quarter of an hour was spent, each beholding the other without speaking a word. At last Briggs began: ‘My lord, I have undertaken this long journey purposely to see your person, and to know by what engine of wit or ingenuity you came first to think of this most excellent help in as¬ tronomy, viz. the logarithms; but, my lord, being by you found out, I wonder nobody found it out before, when now known it is so easy.* ” 32 Mathematics and the Imagination Napier’s conception of logarithms was based on an ingenious and well-known idea: a comparison between 2 moving points, one of which generates an arithmetical, the other a geometric progression. The two progressions: Arithmetical— 0 1 23 4567 8... Geometric— 1 2 4 8 16 32 64 128 256 ... bear to each other this interesting relationship: If the terms of the arithmetical progression are regarded as exponents (powers) of 2, the corresponding terms of the geometric progression represent the quantity resulting from the indicated operation. Thus,'^ 2° = 1, 2' == 2, 22 ^ 4 23 = 8, 2^ = 16, 2^ = 32, etc. Furthermore, to determine the value of the product 2^ X 2^, it is necessary to add the exponents, obtaining 2^^ = 2®, which is the desired product. Calling 2 the base, each term in the arithmetical progression is the logarithm oJ the corre¬ sponding term of the geometric progression. Napier explained this notion geometrically as follows: A point S moves along a straight line, AB, with a velocity at each point Si proportional to the remaining distance SiB. Another point R moves along an unlimited line, CD, with a uniform velocity equal to the initial velocity of S. If both points start from A and C at the same time, then the logarithm of the number measured by the distance S\B is measured by the distance CR\. FIG. 19.—Napier’s d*;Tiamic interpretation of logarithms. PIE (ir, iy e)—Transcendental and Imaginary 83 By this method, as S^B decreases, its logarithm CR^ increases. But it soon became apparent that it was advan¬ tageous to define the logarithm of 1 as zero, and to have the logarithm grow with the number. Napier changed his system accordingly. One of the fruits of the higher education is the illumi¬ nating view that a logcu-ithm is merely a number that is found in a table. We shall have to widen the curric¬ ulum. If a, by and c are three numbers related by the equation ~ c, then by the exponent of a, is the loga¬ rithm of c to the base a. in other words, the logarithm of a number to the base a is the power to which a must be raised to obtain that number. In the example, 2^ = 8, the logarithm of 8 to the base 2 is 3. Or 10^ = 100, and the logarithm of 100 to the base 10 equals 2. The concise way of expressing this is: 3 = log 2 8, 2 = log 10 100. The simple table below gives all the essential properties of logarithms: (1) loga {b X c) (2) log. 0 = = loga b -f loga loga b — loga C, (3) loga b‘ = C X loga b. (4) logaV^ b = (7) Equations (1) and (2) indicate how to multiply or divide two numbers; nothing more is required than to add or subtract their respective logarithms. The result obtained is the logarithm of the product, or quotient. Equations (3) and (4) show that with the aid of loga¬ rithms the operations of raising to powers and extracting roots may be replaced by the much simpler ones of multiplication and division. 7 84 Mathematics and the Imagination Extensive tables of logarithms were soon constructed to the base 10 and to the Napierian or natural base e. So widely were these tables distributed that mathematicians all over Europe were able to avail themselves of the use of logarithms within a very short time of their invention. Kepler was one who not only saw the tables of Napier but himself advanced their development; he was thus one of the first of the legion of scientists whose contributions to knowledge were greatly facilitated by logarithms. The two systems of logs to the two bases, 10 and e (the Briggs and the natural base respectively), are the principal ones still in use, with e predominating.^® Like TT, the number e is transcendental and like tt it is what P. W. Bridgman names a “progrcim of procedure,” rather than a number, since it can never be completely expressed ( 1 ) in a finite number of digits, ( 2 ) as the root of an algebraic equation with integer coefficients, (3) ^ a nonterminating but repeating decimal. It can only be expressed with accuracy as the limit of a convergent infinite series or of a continued fraction. The simplest and most familiar infinite series giving the value of e is: Accordingly, its value may be approximated as closely as we please by taking additional terms of the series. To the tenth decimal places = 2.718281285. A glance at the table below will indicate how an infinite convergent series behaves as more and more of its terms are summed. ( 1 ) 1 + 1 ; = 2 - ( 2 ) 1 + ^ + ^ = 2.5 PIE (tt, i, e)—Transcendental and Imaginary 85 ( 3 ) 1 + ^ ^ ^ = 2.6666666 ... (4) l+^ + ^ + l + l = 2.7083334 ... (5) l+i + ^ + i^ + l + ± = 2.7166666 ... W 1 + ^ + . . . + i = 2.7180555 ... (■ 7 ) 1 + + • . . + = 2.7182539 ... (8) 1 + ^ + . . . + ^ = 2.7182787 ... (9) 1 + ^ + . . . + ^ = 2.7182818 . .. Upon taking a few more terms, e looks like this: 2.7182818284590452353602874 . . . Euler, who undoubtedly had the Midas touch in math¬ ematics, not only invented the symbol e and calculated its value to 23 places, but gave several very interesting ex¬ pressions for it, of which these two are the most important: (1) e = 2+ 1 1 -hi _ 2 + 2 3 + 3 4 + 4 5 + 5... 86 Mathematics and the Imagination ( 2 ) = 1+1 T +j_ 1 + 1 _ 1 + 1 _ 5 + ^ 1 1 +J_ 1 +2 9... The need for navigational tables was not alone re¬ sponsible for the development of logarithms. Big business, particularly banking, played its part as well. A remark¬ able series, the limiting value of which is e, arises in the preparation of tables of compound interest. This series is obtainable from the expansion of ^1 comes infinite. The origin of this important expression is interesting. Suppose your bank pays 3 per cent interest yearly on deposits. If this interest is added at the end of each year, for a period ot three years, the total amount to your credit, assuming an original capital of $1.00, is given by the formula; (1 + .03)®. If the interest is compounded semiannually, after the three-year period the total of principal plus interest would be .03\ 2X3 t) ■ Imagine however that you are fortunate enough to find a philanthropic bank which decides to pay 100 per cent interest a year. Then the amount to your credit at the end of the year will be(l + 1)^ = $2.00. If the inter¬ est is compounded semiannually, the amount will be PIE (tt, i, e)—Transcendental and Imaginary 87 ~1” 2 )^^ “ or $2.25. If it is compounded quarterly, it will be (1 + = $2.43. It seems clear that the more often the interest is compounded, the more money you will have in the bank. By a further stretch of the imagina¬ tion, you may conceive of the possibility that the philan¬ thropic bank decides to compound the interest continuously, that is to say at every instant throughout the year. How much money will you then have at the end of the year? No doubt a fortune. At least, that is what you would sus¬ pect, even allowing for what you know about banks. In¬ deed you might become, not a millionaire, not a billion¬ aire, but more nearly what could be described as an “in- finitaire.” Alas, banish all delusions of grandeur, for the process of compounding interest continuously, at every instant, generates an infinite series which converges to the limit e. The sum on deposit after this hectic year, with its apparent promise of untold riches, would be not quite $2.72. For, if one takes the trouble to expand as n becomes very large, the successive values thus ob¬ tained approximate to the value of e, and where n be¬ comes infinite, actually yields the infinite series for e: Besides serving as the base for the natural logarithms, f is a number useful everywhere in mathematics and applied science. No other mathematiccil constant, not even tt, is more closely connected with human affairs. In economics, in statistics, in the theory of probability, and in the exponential function, e has helped to do one 88 Mathematics and the Imagination thing and to do that better than any number yet dis¬ covered. It has played an integral part in helping mathe¬ maticians describe and predict what is for man the most important of all natural phenomena—that of growth. The exponential function, y = is the instrument used, in one form or another, to describe the behavior of growing things. For this it is uniquely suited: it is the only function of x with a rate of change with respect to x equal to the function itself.'^ A function, it will be remembered, is a table giving the relation between two variable quan¬ tities, where a change in one implies some change in the other. The cost of a quantity of meat is a function of its weight; the speed of a train, a function of the quantity of coal consumed; the amount of perspiration given off, a function of the temperature. In each of these illustrations, a change in the second variable; weight, quantity of coal consumed, and temperature, is correlated with a change in the first variable: cost, speed, and volume of perspira¬ tion. The symbolism of mathematics permits functional relationships to be simply and concisely expressed. Thus, y = x^ y = x"^, y — sin x, y = csch x^ y = e^ are ex¬ amples of functions. A function is not only adequate to describe the behav¬ ior- of a projectile in flight, a volume of gas under changes of pressure, an electric current flowing through a wire, but also of other processes which entail change, such as growth of population, growth of a tree, growth of an amoeba, or as we have just seen, growth of capital and interest. What is peculiar to every organic process is that the rate of growth is proportional to the state of growth. The bigger something is, the faster it grows. Under ideal conditions, the larger the population of a country be¬ comes, the faster it increases. The rate of speed of many PIE (ttj 2, e)—Transcendental and Imaginary 89 chemical reactions is proportional to the quantity of the reacting substances which are present. Or, the amount of heat given off by a hot body to the surrounding me¬ dium is proportional to the temperature. The rate at which the total quantity of a radioactive substance dimin¬ ishes at any instant, owing to emanations, is proportional to the total quantity present at that instant. All these phe¬ nomena, which either are, or resemble, organic processes, may be accurately described by a form of the exponential function (the simplest beings = for this has the prop¬ erty that its rate of change is proportional to the rate of change of its variable. ♦ A universe in which e and tt were lacking, would not, as some anthropomorphic soul has said, be inconceivable. One could hardly imagine that the sun would fail to rise, or the tides cease to flow for lack of tt and But without these mathematical artifacts, what we know about the sun and the tides, indeed our ability to describe all natural phenomena, physical, biological, chemical or statistical, would be reduced to primitive dimensions. Alice was criticizing Humpty Dumpty for the liberties he took with words: “When I use a word,” Humpty replied, in a scornful tone, “it means just what I choose it to mean—neither more nor less.” “The question is,” said Alice, whether you can make a word mean so mai\y different things.” “The question is,” said Humpty, “which is to be master, that’s all.” Those who are troubled (and there are many) by the word imaginary” as it is used in mathematics, should hearken unto the words of H. Dumpty. At most, of course, it is a small matter. In mathematics familiar go Mathematics and the Imagination words are repeatedly given technical meanings. But as Whitehead has so aptly said, this is confusing only to minor intellects. When a word is precisely defined, and signifies only one thing, there is no more reason to criticize its use than to criticize the use of a proper name. Our Christian names may not suit us, may not suit our friends, but they occasion litde misunderstanding. Con¬ fusion arises only when the same word packs several meanings and is what Humpty D. calls a “portmanteau.** Semantics, a rather fashionable science nowadays, is devoted to the study of the proper use of words. Yet there is much more need for semantics in other branches of knowledge than in mathematics. Indeed, the larger part of the world’s troubles today arise from the fact that some of its more voluble magnificoes are definitely anti- semantic. An imaginary number is a precise mathematical idea. It forced itself into algebra much in the same way as did the negative numbers. We shall see more clearly how imaginary numbers came into use if we consider the development of their progenitors—the negatives. Negative numbers appeared as roots of equations as soon as there were equations, or rather, as soon as mathe¬ maticians busied themselves with algebra. Every equa¬ tion of the form ax -\- b — 0, where a and b are greater than zero, has a negative root. The Greeks, for whom geometry was a joy and algebra a necessary evil, rejected negative numbers. Unable to fit them into their geometry, unable to represent them by pictures, the Greeks considered negative numbers no numbers at all. But algebra needed them if it were to grow up. Wiser than the Greeks, wiser than Omar Khayyam,the Chinese and the Hindus recognized negative numbers even before the Christian era. Not as PIE (ttj 2, e) ’Transcendental and Imaginary gi learned in geometry, they had no qualms about numbers of which they could draw no pictures. There is a repeti¬ tion of that indifference to the desire for concrete repre¬ sentation of abstract ideas in the contemporary theories of mathematical physics, (relativity, the mechanics of quanta, etc.) which, although understandable as symbols on paper, defy diagrams, pictures, or adequate metaphors to explain them in terms of common experience. Cardan, eminent mathematician of the sixteenth cen¬ tury, gambler, and occasional scoundrel, to whom al¬ gebra is vasdy indebted, first recognized the true im¬ portance of negative roots. But his scientific conscience twitted him to the point of calling them “fictitious.” Raphael Bombelli of Bologna carried on from where Cardan left off. Cardan had talked about the square roots of negative numbers, but he failed to understand the con¬ cept of imaginaries. In a work published in 1572, Bom¬ belli pointed out that imaginary quantities were essential to the solution of many algebraic equations. He saw that equations of the form -\- a = Oy where a is any num- ber g^reater than 0, could not be solved except with the aid of imaginaries. In trying to solve a simple equation + 1 = 0, there are two alternatives. Either the equa¬ tion is meaningless, which is absurd, or is the square root of —1, which is equally absurd. But mathematics thrives on absurdities, and Bombelli helped it along by accepting the second alternative. ♦ Three hundred and fifty years have gone by since Bombelli made his choice. Philosophers, sciendsts, and those with that minor-key quality of mind known as plain common sense have criticized, in ever-increasing diminuendo, the concept of the imaginary. All of these worthies are dead, most of them forgotten, while imagi- 92 Mathematics and the Imagination nary numbers flourish wickedly and wantonly over the whole field of mathematics. Occasionally, even the. masters snickered. Leibniz thought: “Imaginary numbers are a fine and wonderful refuge of the Holy Spirit, a sort of amphibian between being and not being.’* Even the mighty Euler said that numbers like the square root of minus one “are neither nothing, nor less than nothing, which necessarily con¬ stitutes them imaginary, or impossible.” He was quite right, but what he omitted to say was that unaginanes were useful and essential to the development of mathe¬ matics. And so they were allotted a place in the number domain with all the rights, privileges, and immunities thereunto appertaining. In time, the fears and queasiness about their essence all but vanished, so that the judgment of Gauss is the judgment of today: Our general arithmetic, so far surpassing in extent the geometry of the ancients, is entirely the creation of modem times. Starting originally from the notion of absolute integers, it has gradually enlarged its domain. To integers have been added fractions, to rational quantities, the irrational, to pos¬ itive, the negative, and to the real, the imaginary. This ad¬ vance, however, had always been made at first with timorous and hesitating steps. The early algebraists called the negative roots of equations false roots, and this is indeed the case when the problem to which they relate has been stated in such a form that the character of the quantity sought allows of no opposite. But just as in general arithmetic no one would hesitate to admit fractions, although there are so many count¬ able things where a fraction has no meaning, so we would not deny to negative numbers the rights accorded to positives, simply because innumerable things admit of no opposite. The reality of negative numbers is sufficiently justified since m innumerable other cases they find an adequate interpretauon. PIE (tt, z, e)—Transcendental and Imaginary 93 This has long been admitted, but the imaginary quantities, formerly, and occasionally now, improperly called impossible, as opposed to real quantities—are still rather tolerated than fully naturalized; they appear more like an empty play upon symbols, to which a thinkable substratum is unhesitatingly denied, even by those who would not depreciate the rich contribution which this play upon symbols has made to the treasure of the relations of real quantities. * Imaginary numbers, like four-dimensional geometry, developed from the logical extension of certain proc¬ esses. The process of extracting roots is called evolution. It is an apt name, for imaginary numbers were literally evolved out of the extension of the process of extracting ro ots. I f 's/a, y/l, 's/ll had meaning, why not \/ — 4 , a/ ~ 7, \/ — 11? If — 1 =0 had a solution, why not + I =0? The recognition of imaginaries was much like the United States recognizing Soviet Russia—the existence was undeniable, all that was required was for¬ mal sanction and approval. is the best-known imaginary. Euler represented It by the symbol “z"” which is still in use.^^ It is idle to be concerned with the question, “What number when multiplied by itself equals —1?’’ Like all other numbers, i is a symbol which represents an abstract but very precise idea. It obeys all the rules of arithmetic with the added convention that iXi = —1. Its obedience to these rules and its manifold uses and applications justify its existence regardless of the fact that it may be an anomaly. The formal laws of operation for i are easy: Since the rule of signs provides: ( + 1) X (+1) = + 1 I /(-I) X ( + 1) = -1 ( + 1) X (-1) = - 1 / \(-l) X (-1) = +1 Mathematics and the Imagination Accordingly: i X ( + 1 ) = i X ( - 1 ) = - V - 1 - t X ( - 1 ) = + ^ = i X ! = = - 1 i X i y. i = _ = (V“) = - 1 ) = - ixiyiyi = ,_ = ( - 1 ) X ( - 1) = + 1 !' X z' X i X i X I = ^ = (V^)* (V -1 )Mv' - 1) = ( - 1) X ( - 1) X V~^ = ( + 1) X V - 1 = — 1, etc.* • From which we may construct a convenient table: 11 < 11 . = V -1V -1 = -1 i* = (V~1)’-(V -D’ = +1 = +iV -1 = >■ = +i(V^)* = -1 p = -i (V- i)'= +1 _ The table shows that odd powers of i are equal to — *, or + *, and even powers of i are equal to — 1 or +1. PIE (tt, iy e') TTanscendental and Imaginary Extension of the use of imaginaries has led to complex numbers of the form a + iby where a and b are real numbers (as distinguished from imaginaries). Thus 3 + 4i, 1 — 2 4- 3/ are examples of complex numbers. The enormously fruitful field of function theory is a direct consequence of the development of complex num¬ bers. While this is a subject too technical and specialized, we shall have occasion to mention complex numbers again when we explain the geometric representation of imaginaries. To that end, we must turn for a moment to that mathematical idea which, as Boltzmann once said seems almost cleverer than the man who invented it— the science of Analytical Geometry. ♦ Program music is distinguished from absolute music, which owes its coherence to structure, in that the purposes of the former is to teU a story. In a certain sense, analyti¬ cal geometry can be distinguished from the geometry of the Greeks as program music from absolute music. Geom- etry, practical in its origin, was cultivated and developed for its own sake both as a logical discipline and as a study of form. Geometry was a manifestation of a striving for the ideal. Shapes and forms that were beautiful, har¬ monious, and symmetric were appreciated and eagerly studied. But the Greeks cultivated the practical only as long as it had a beautiful side; beyond that, their math¬ ematics was hampered by their aesthetics. There was left to Descartes the task of writing the program music of mathematics, of devising a geometry which tells a story. When it is said that every algebraic equation has a picture, we are describing the relation between analytical geometry and algebra. And just as program music is as important and significant in itself as g6 Mathematics and the Imagination the stories it illustrates, so analytical geometry has its own dignity and importance—is an autonomous math¬ ematical discipline. * The Jesuit Fathers were often very wise: at their school at La Fleche, young Rene Descartes was permitted, be¬ cause of his delicate health, to remain in bed each day until noon. What McGufFey would have prophesied about the future of such a child is not difficult to imagine. But Descartes did not turn out a complete profligate. Indeed, his delightful habit of staying in bed until noon bore at least one remarkable fruit. Analytical geometry came to him one morning as he lay pleasantly in bed. It is powerful, this idea of a co-ordinate geometry, yet easy to understand. Consider two lines (axes) in a plane: intersecting at right angles at a point R: Y FIG. 20.—The point p has the co-ordinates {rriy in'). Any point in the entire plane may then be unique y determined by its perpendicular distance from the lines xx' Sind yy'. The point P, for example, by the distances m and Thus, a pair of numbers representing scalar PIE (tt, 2 , e) Transcendental and Imaginary gy distances along jirV and^' will determine every point in the plane, and conversely, every point in the plane determines a pair of numbers. These numbers are called the co-ordinates of the point. All distances on ata:' measured to the right of R are called positive, to the left of /?, negative. Similarly, all distances measured on yy' above R are positive, all distances below, negative. The point of intersection, the origin, is designated by the co-ordinates (0, 0). The con- y' FIG. 21.—The co-ordinate axes in the real plane. vention for writing co-ordinates is to put down the distance/rom they/ axis (i.e. the distance along the A.r' axis) first, the distance/rom the xx' axis, along the yy' axis second; thus: (0, 0), (4, 3), ( - 1, 5), (6, 0),' (0, 6), Mathematics and the Imagination ( _ 6, - 5), (3, - 3), (- 8, 0), (0, - 8) are the co¬ ordinates of the points in Fig. 21. Y' Y FIG. 22(a).—Graphic represenution of the equa¬ tion y = Jc®. FIG. 22(b).—Graphic representation of the equation y — sin X. This is the famous wave curve used to represent many regular and periodic phenomena, i.e., electrical current, the motion of a pendulum, radio trans^ssion, sound and light waves, etc. (For the meaning of sin x, see note in the chapter on the calculus.) Coupling this notion with that of a function, it is not difficult to see how an equation may be pictured in the PIE (tt, e, e)—Transcendental and Imaginary 99 plane of analytic geometry. When a- and^ are functionally related, to each value of x there corresponds a value of which two values determine a point in the plane. The totality of such number pairs, that is, all the values of y corresponding to all the values of ;r, when joined by a smooth curve as in Figs. 22(a,b,c), make up the geometri¬ cal portrait of an equation. FIG. 22(c).—Graphic representation of the equa¬ tion^ = e*. This curve illustrates the property com¬ mon to all phenomena of growth: rate of growth is O proportional to state of growth. Employing co-ordinate geometry, how shall we rep¬ resent an imaginary number like A theorem in A 8 100 Mathematics and the Imagination elementary geometry, relating to the geometric mean, furnishes the clue (see Fig.- 23). In the right triangle ABC, the perpendicular AD divides BC into two portions: BD^ DC. The length of the perpen¬ dicular AD equcils \/BD X DC, and is called the geo¬ metric mean of BD and DC. (Fig. 23.) A Norwegian surveyor, Wessel, and a Parisian book¬ keeper, Argand, at the close of the eighteenth and begin¬ ning of the nineteenth centuries, independendy found that imaginary numbers could be represented by the application of this theorem. In Fig. 24: Y' Y FIG. 24.—Geometric interpretation of i. the distance S, from the origin to +1, is the geometric mean of the triangle, bounded by the sides L and L , and the base formed by that portion of the xx' axis from — 1 to 1. Then 6' = -\/—1+1 — — \ — i PIE (tt, 2 , e)—Transcendental and Imaginary \ o i Here, then, is a geometric representation of an imaginary number. Extending this idea, Gauss built up the entire complex plane. In the complex plane every point represented by a complex number of the form x iy corresponds to the point in the plane fixed by the co-ordinates a: and y. In other words, a complex number may be regarded as a pair of real numbers with the addition of the number i. The use of i appears only on performing the operations of multiplication and division. Conceive of a line joining the point {a -j- ib) to the origin R. Then the operation of multiplying by — 1 is equivalent to rotating that line about the origin through 180° and shifting the point from (+(2 -\-ib) to ( a — ib). The effect of multiplying a number by i is such that when performed twice, P is obtained, which is equivalent to multiplication by —1. FIG. 25.—Multiplication by i is a rotation throueh 90°.- Let /* = (a -j- ib'). Then, P X i = (a ib) X i = (a X i) + (A X I X i) = ta + 6 • -1 = —b ia = d- 102 Mathematics and the Imagination Therefore, multiplication by i is a rotation through only 90°. Complex numbers may be added, subtracted, multi¬ plied, and divided, just as though they were real numbers. The formal rules of these operations (the most interesting being the substitution of — 1 for i^) are illustrated in the examples below. (1) X iy = x' iy' if, and only \S x = x' znd y = y' (2) {x + iy) + ix' + iy') = {x + x') + i(y + /) (3) (x -I- iy) - (x' + ty') = (x - x') -h i{y - /) (4) (x -h iy) (x' -h iy') = (xx' - y/) -h Hxy' -f yx') FIG. 26.—The complex plane. Figure 26 shows the same points in the plane given in Fig. 21, except that for the co-ordinates of x and ^ of each point we have substituted the corresponding com¬ plex number x -|- iy. By virtue of the peculiar properties of i, complex num- PIE (tt, 2 , e)—Transcendental and Imaginary 103 bers may be used to represent both magnitude and direc¬ tion. With their aid some of the most essential notions in physics such as velocity, force, acceleration, etc., are con¬ veniently represented. Enough has now been said to indicate the general nature of i, its purpose and importance in mathematics, its challenge to and final victory over the cherished tenets of common sense. Undaunted by its paradoxical appear¬ ance, mathematicians used it as they used tt and e. The result has been to make possible almost the entire edifice of modern physical science.* * One thing remains. There is a famous formula—per¬ haps the most compact and famous of all formulas—de¬ veloped by Euler from a discovery of the French mathe¬ matician, De Moivre: + 1 =0. Elegant, concise and full of meaning, we can only reproduce it and not stop to inquire into its implications. It appeals equally to the mystic, the scientist, the philosopher, the mathematician. For each it has its own meaning. Though known for over a century, De Moivre’s formula came to Benjamin Peirce, one of Harvard's leading mathematicians in the nine¬ teenth century, as something of a revelation. Having discovered it one day, he turned to his students and made a remark which supplies in dramatic quality and ap¬ preciation what it may lack in learning and sophistica- * Let us have this much balm for the reader who has bravely gone through the pages on analytical geometry and complex numbers. The average college course on analytic geometry (not including complex numbers) takes six months. It is, therefore, a little too much to expect that it can be learned in about five pages. On the other hand, if the basic idea has been put over, that every number, every equation of algebra, can be graphically represented, the harrowing details may be left to more intrepid adventurers. 104 Mathematics and the Imagination tion: “Gentlemen,” he said, “that is surely true, it is absolutely paradoxical; we cannot understand it, and we don’t know what it means, but we have proved it, and therefore, we know it must be the truth.” When there is so much humility and so much vision everywhere, society will be governed by science and not by its clever people. APPENDIX BIRTH OF A CURVE (1) Let US consider the equation^ = x^. Take a few sample values of x and find the corresponding values of arranging the results in a table: X y 0 0 1 1 2 4 3 9 4 16 That is 2^ = 4, 3^ = 9, etc. Plotting these points on the co-ordinate plane, we obtain Fig. A. (2) Now, what about the negative values of xl We see, for example, ( — 2)^ = —2X —2 = 4. This is evidently true for all values of x; thus there corresponds to every point plotted in Fig. A another point which is its mirror image, the axis OT being the mirror. Adding these gives the second figure (Fig. B). (3) The arrangement of the points suggests that we draw a smooth curve through them. (Fig. C.) PIE (tt, iy e) 'Transcendental and Imaginary 105 Y FIO. B. But does this curve embrace other points which arise m our functional table. Let us test this, tabulating some fractional values of x. io6 Mathematics and the Imagination FIG. C, If we plot these new points, it may be seen that they all lie on the curve (Fig. D). Indeed, if we continue further, we would find that every point which might arise in the table will lie on the curve; the totality of such points will form the curve known as the parabola. PIE (ttj e)—Transcendental and Imaginary 107 The parabola is formed by the section of a cone cut by a plane parallel to the opposite edge. You can make a parabola for A jet of water forms a parabola, yourself with the help of a flash- So does the path of a projectile, light, holding it so that the upper But the curve formed by a loop of boundary of the beam will be string held at the ends, hangin parallel to the floor. freely, is not a parabola, but catenary. io8 Mathematics and the Imagination FOOTNOTES 1. Henri Bergson, Creative Evolution. —P. 65. 2. It is a simple matter geometrically to determine the square root of a given length.—P. 67. FIG. 27.—Let AB be the given length. Extend it to C so that BC = 1. Draw a semicircle having AC as diameter. Erect a perpendicular at B meeting semicircle at D. BD is the required square root ofL. 3. Gauss made an exhaustive study to determine what other poly¬ gons could be constructed with ruler and compass. The Greeks had been able to construct regular polygons of 3 and 5 sides, but not those with 7, 11, or 13 sides. Gauss, with marvelous precocity, gave the formula which showed what polygons were constructible in the classical way. It had been thought that only regular poly¬ gons, the number of whose sides could be expressed by the forms: 2", 2” X 3, 2" X 5, 2" X 15 (where n is an integer), could be so constructed. Gauss’ formula proves that polygons with a prime number of sides may be constructed as follows; Let P be the num¬ ber of sides and n any integer up to 4, then P =* 2*" -|- 1. If « = 0, 1, 2, 3, 4, P = 3, 5, 17, 257, 65537. Where n is greater than 4, there are no known primes of the form 2^" + 1. (A prime number is one which is not evenly divisible by any number other than 1 or itself. Thus, 2, 3, 5, 7, 11, 13, 17 are examples of primes. A famous proof of Euclid, which appears in his Elements^ shows that the number of primes is infinite. See p. 192.) It is an amazing fact that of all the possible polygons, the number of whose sides is prime, only the five given above are known to be constructible with ruler and compass.—P. 68. 4. See Chap. 5.—P. 68. 5. As long ago as 1775, the Paris Academy was so overwhelmed PIE (ttj z, e )— Transcendental and Imaginary 109 with pretended solutions from circle squarers, angle trisectors, ■ and cube duplicators, that a resolution was passed that no more would be accepted. But at that time the impossibility of these solutions was only suspected and not yet mathematically demon¬ strated; thus the arbitrary action of the academy can only be explained on the grounds of self-preser\'ation.—P. 69. 6 . Limits and converging processes with an infinite number of steps, as we shall soon see, were used in computing tt.—P. 69. 7. See the chapter on the calculus.—P. 69. 8 . Most infinite series are divergenty that is, the sum of the series exceeds any assignable integer. A typical divergent series is 1 + ^ + ^+ ^ + !+ . . . This series seems to differ very little from the convergent series given in the text, and only the most subtle mathematical operations reveal whether a series is convergent or divergent. —P. 70. 9. A square can be duplicated by dra^ving a square on the diagonal of the given square, but a cube cannot be duplicated because the cube root of 2 is involved, and this, like tt, is not the root of an algebraic equation of first or second degree, and therefore cannot be constructed with ruler and compass. In four-dimensional space, the figure which corresponds to the cube, called a “tes- saract” (see the chapter on assorted geometries) can be duplicated by ruler and compass, because the fourth root of 2 , which is what is required, can be written as the square root of the square root of 2.—P. 70. 10. What is meant by “the root of an algebraic equation with integer coefficients”? A word may suffice to jog the memory of those who have had a course in elementary algebra. The root of an equation is the value that must be substituted for the unknown quantity in the equation in order to satisfy it. Thus, in the equation a: — 9 = 0, 9 is the root, since if you substitute 9 for X, the equation is satisfied. Similarly —4 and 4 are the roots of the equation x* — 16 = 0 , because when either value is substituted for x, the equation balances. “Algebraic” equations arc the kind of equations we have just been talking about. There are also trigonometric equations, differential equations and others, and the term “algebraic” is intended to distinguish equations of the form flox" + <2ix"~‘ + + . . . + an-\x + = 0. The coefficients of an equation are the numbers which appear before the unknown quantity or quantities. In the equation 3x* + 17x3 -f >/2 x3 - /x + ttx = 0 I 10 Mathematics and the Imaginatibn 3, 17, \/2, z, and x are the coefficients. This is an example of an algebraic equation with queer coefficients. In defining an algebraic number (see page 49), we demand that n be a positive integer and that the a’s be integers.—P. 72. 11. See Buffon’s Needle Problem in the chapter on Chance and Chanceability.—P. 79. 12. The y/l when written as a decimal is just as complicated as x, for it never repeats, never ends, and there is no known law giving the succession of its digits; yet this complicated decimal is easily obtained with exactitude by a ruler and compass con¬ struction. It is the diagonal of a square whose side is equal to 1. —P. 79. 13. Jobst Biirgi of Prague had prepared tables of logarithms before Napier’s Descriptio appeared. Biirgi however failed to publish his tables until 1620 because, as he explained, he was busy on some other problem.—P. 80. 14. According to the principle of positional notation, the value of a digit depends on its position in relation to the other digits in the number in which they appear.—P. 80. 15. The rules for operating with exponents in multiplication and division are: A) Multiplication a”* 'K — 42 "*+"; thus, X = d®; or, d® X d* = (d d d) X (d d) B) DiiAsion = d® ^8 = a d®-* = d But, if wj b equal to n, d"* „ ^ d" d® — = /, 3-3 = ^0 = ? = 1 d® 4X^X4 TTierefore we agree upon a® = 1 _p. 82. 16 17 18 19 20 , 21 PIE (tt, iy e) Ttanscendental and Imaginary 111 22 . 23. Because e possesses certain unique properties valuable in many branches of mathematics, particularly the calculus, because of the relation between logarithmic and exponential functions, e is the natural” base for the logarithmic system.—P. 84. The first proof that e is transcendental (i.c., not the root of an algebraic equation with integer coefficients), was given by Her- mite, the distinguished French mathematician, in 1873, nine years before Lindemann’s proof of the transcendental character of TT a^eared. Since that time several others succeeded in simplify¬ ing Hermite’s proof. The general method is to “assume e to be the root of an algebraic equation,/(e) = 0, and show that a multiplier iW can be chosen such that when each side of the equation is mul- Uphed by Af, (the value of) Mj{e) is reduced to the sum of an integer ny zero and a number between 1 and 0, showing that the assumption that e can be the root of an algebraic equation is un¬ tenable.” See U. G. Mitchell and M. Strain, in Osiris, Studies in History of Science, Vol. I.—P. 84. The symbol ! as used in mathematics does not indicate surprise or excitement, although in this case it might not be amiss, since the simplicity and beauty of this series is amazing. ! means take the factorial of the number after which ! appears.” The factorial of a number is the product of its components; thus 11 = 1,2! = 1X2, 3! = IX2X3, 41 = 1X2X3X4 5! = 1X2X3X4X5.-P.84. X ^ X 3 X 4, Actually « need only be equal to 1000 (i.e., the interest computed thnee daily) to give S2.72.—P. 87. The derivative of _>- = e* h equal to the function itself. For a further discussion of the derivative and of problems involving rate of change, see the chapter on the calculus.—P. 88. being the author of the well-worn Rubaiyat, ’ was also a mathematician of distinction, but one whose prophetic vision failed for negative numbers.— P 90 Translated in Dantzig, Mumb.r, the Language ojScience (New York; Macmillan), 1933, p. 190.—P. 93. It was once suggested that appropriate symbols for the two constants, r and should be 0 for r. and (p for f in order to avoid confusion. But printers balked at making new type and the old •symbol^s remained. More often than is realized, such considera¬ tions determined the character of mathematical notation.- Assorted Geometries—Plane and Fancy They say that habit is second nature, nature is only first habit? Who knows but —PASCAL Among our most cherished convictions, none is more precious than our beliefs about space and time, yet none is more difficult to explain. The talking fish of Grimm’s fairy tale would have had great difficulty in explaining how it felt to be always wet, never having tasted the pleasure of being dry. We have similar difficulties in talking about space, knowing neither what it is, nor what it would be like not to be in it. Space and time are “too much with us late and soon” for us to detach ourselves and describe them objectively. “For what is time?” asked Saint Augustine. “Who can easily and briefly explain it? Who even in thought can comprehend it, even to the pronouncing of a word con¬ cerning it? But what in speaking do we refer to more familiarly and knowingly than time? And certainly we understand when we speak of it; we understand also when we hear it spoken of by another. What, then, is time? If no one ask of me, I know; if I wish to explain to him who asks, I know not.”^ And this could as well be said of space. Though space cannot be defined, there is little difficulty in measuring distances and areas, in moving about, in charting vast I 12 Assorted Geometries—Plane and Fancy \ 13 courses, or in seeing through millions of light years. Everywhere there is overwhelming evidence that space is our natural medium and confronts us with no in¬ superable problems. But this professes to be no philosophical treatise and no German Handbook on an Introduction to the Theory of Space in 14 volumes. Our intention is to explain in the simplest, most general manner, not the physical space of sense perception, but the space of the mathematician. To that end, all preconceived notions must be cast aside and the alphabet learned anew. In this chapter we propose to discuss two kinds of geometry—four-dimensional and non-Euclidean. Neither of these subjects is beyond the comprehension of the non¬ mathematician prepared to do a little straight thinking. To be sure, they have both been described, like the theory of relativity (to which they are in some ways related) in high and mighty mumbo jumbo. High priests in every profession devise elaborate rituals and obscure language as much to conceal their own ineptness as to awe the uninitiate. But the corruptness of the clergy should not deter us. The basic ideas underlying four-dimensional and non-Euclidean geometry are simple, and this we aim to prove. ♦ Euclid, in writing the Elements^ recognized no great obstacles. Starting with certain fundamental ideas (pre¬ sumably understood by everyone) expressed as postu¬ lates and axioms, he built upon these as foundations. This ideal method for developing a logical system has never been improved upon, although occasionally it has been neglected or forgotten with sad results. Although Euclid’s Elements constitute an imposing in- 114 Mathematics and the Imagination tellectual achievement, they fail to make an important distinction between two types of mathematics —fure and applied —a distinction which has only come to light in modern theoretical developments in mathematics, logic and physics. A geometry which treats of the space of experience, is applied mathematics. If it says nothing about that space —if, in other words, it is a system composed of abstract notions, elements, and classes, with rules of combination obeying the laws of formed logic, it is pure mathematics. Its propositions are of the form: If A is true, then B is true, regardless of what A and B may possibly be.^ Should a system of pure mathematics be applicable to the physi¬ cal world, its fruitfulness may be regarded either as mere chance, or as further evidence of the profound connection between the forms of nature and those of mathematics. Yet, in either case, this essential fact must be borne in mind—the fruitfulness of a logical system neither dimin¬ ishes nor augments its validity. As applied mathematics, Euclid’s geometry is a good approximation within a restricted field. Good enough to help draw a map of Rhode Island, it is not good enough for a map of Texas or the United States, or for the meas¬ urement of either atomic or stellar distances. As a system of pure mathematics, its propositions are true in a most general way. That is to say, they have validity only as propositions of logic, only if they have been correctly deduced from the axioms. Other geometries with dif¬ ferent postulates are therefore possible—indeed, as many others as the mathematician chooses to devise. All that is necessary is to assemble certain fundamental ideas (classes, elements, rules of combination), declare these to be undefinable, make certain that they are not self-con- Assorted Geometries—Plane and Fancy 115 tradictory, and the groundwork has been laid for a new edifice, a new geometry. Whether this new geometry will be fruitful, whether it will prove as useful in survey¬ ing or navigation as Euclidean geometry, whether its fundamental ideas measure up to any standard of truth other than self-consistency, doesn’t concern the mathe¬ matician a jot. The mathematician is the tailor to the pntry of science. He makes the suits, anyone who fits into them can wear them. To put it another way, the mathematician makes the rules of the game; anyone who wishes may play, so long as he observes them. There is no sense in complaining afterwards that the game was without profit. * If we wish to pay a mathematical system the highest possible compliment, to indicate that it partakes of the same generality and has the same validity as logic, we may call it a game. A four-dimensional geometry is a game: so is the geometry of Euclid. To object to four¬ dimensional geometry on the grounds that there are only three dimensions is absurd. Chess can be played as well by those who believe in comrades or dictators as by those who cling to the vanishing glory of kings and queens. What sense is there in objecting to chess on the grounds that kings and queens belong to a past age, and that, in any case, they never did behave like chess pieces —no, not even bishops. What merit is there to the con¬ tention that chess is an illogical game because it is im¬ possible to conceive that a private citizen may be crowned queen merely by moving forward five steps. Perhaps these are ridiculous examples, but they are no more so than the complaints of the faint of heart who say that three dimensions make space and space makes 9 116 Mathematics and the Imagination three dimensions, “that is all ye know on earth and all ye need to know.” If we can rake the doubters fore, we can rake them aft—indeed, from stem to stern. For there is no proof, in the scientific sense, that space is three- dimensional, or for that matter, that it is four-, five-, six-, or anything but rz-dimensional. Space cannot be proved three-dimensional by geometry considered as pure mathematics, because pure mathematics is concerned only with its own logical consistency and not with space or anything else. Nor is this the province of applied math¬ ematics, which does not generally inquire into the nature of space, but assumes its existence. All that we have learned from applied mathematics is that it is convenient, but not obligatory, to consider the space of our sense per¬ ceptions as three-dimensional. To the objection that a fourth dimension is beyond imagination we may reply chat what is common sense today was abstruse reasoning—even wild speculation— yesterday. For primitive man to imagine the wheel, or a pane of glass, must have required even higher powers than for us to conceive of a fourth dimension. Someone may still object: “You tell me that four¬ dimensional geometry is a game. I will believe you. But it seems to be a game that doesn’t concern itself with anything real, with anything I have ever experienced.” We may answer in the Socratic way with another ques¬ tion. “If a four-dimensional geometry treats of nothing real, what does the plane geometry of Euclid consider? Anything more real? Certainly not! It doesn’t describe the space accessible to our senses which we explain in terms of sight and touch. It talks about points that have no dimensions, lines that have no breadth, and planes that have no thickness—all abstractions and idealiza- Assorted Geometries—Plane and Fancy 11 7 tions resembling nothing we have ever experienced or encountered.’* The notion of a fourth dimension, although precise, is very abstract, and for the greatest majority beyond imagi¬ nation and in the purest realm of conception. The de¬ velopment of this idea is as much due to our rather child¬ ish desire for consistency as to anything more profound. In this same striving after consistency and generality, mathematicians developed negative numbers, imaginar- ies, and the transcendentals. Since no one had ever seen minus three cows, or the square root of minus one trees It was not without a struggle that these now rather com¬ monplace ideas were introduced into mathematics. The same struggle was repeated to introduce a fourth dimen¬ sion, and there are still skeptics in the camp of the opposi¬ tion. Every possible allegory and fiction was proposed to coax and cajole the doubters, to make the idea of a fourth dimension more palatable. There were the ro¬ mances which described how impossible a three-dimen¬ sional world would seem to creatures in a two-dimen¬ sional one, there were stories of ghosts, table-tipping, and the land of the dead. It required illustrations from the land of the living, which were still less comprehen¬ sible than the fourth dimension, to win even a partial victory. From this, it should not be inferred that a greater absurdity was enlisted in support of a lesser one. Beginning as usual with Aristotle, it was proved again and again that a fourth dimension was unthinkable and impossible. Ptolemy pointed out that three mutually per¬ pendicular lines could be drawn in space, but a fourth, perpendicular to these, would be without measure or 118 Mathematics and the Imagination depth. Other mathematicians, unwilling to risk a heresy greater even than going contrary to the Bible—that is, contradicting Euclid—advised that to go beyond three dimensions was to go “against nature.” And the English mathematician, John Wallis, of whom one might prop¬ erly have expected better things, referred to that “fansie,” a fourth dimension, as a “Monster in Nature, less possible than a Chimera or a Centaure.” Unwittingly, a philosopher, Henry More, came to the rescue, although mathematicians today would hardly acknowledge his support. His suggestion was not an unmixed blessing. Ghosdy spirits, said More, surely have four dimensions. But Kant delivered an earthly blow by laying down his intuitive notions of space which were hardly compatible with either a four-dimensional or a non-Euclidean geometry. In the nineteenth century several leading mathemati¬ cians espoused the apparently hopeless cause, and be¬ hold—a new mathematical gusher. The great paper of Riemann On the Hypotheses Which Underlie the Foundations of Geometry^ followed by the works of Cayley, Veronese, Mobius, Pliicker, Sylvester, Bolyai, Grassmann, Lobach¬ evsky, created a revolution in geometry. The geometry of four and even higher dimensions became an indis¬ pensable part of mathematics, related to many other branches. When finally, there came, as for some mysterious rea¬ son they always come, direct uses and applications of four-dimensional geometry to mathematical physics, to the physical world, when the unwanted child was sud¬ denly recognized and rechristened “Time, the fourth di¬ mension!” the rejoicing made the cup flow over. Curious Assorted Geometries—Plane and Fancy 11 g and marvelous things were said. The fourth dimension would solve all the awful mysteries of the universe, and ultimately might prove a cure for arthritis. So far in the general jubilation did the mathematicians forget them¬ selves that some of them began to refer to it as *‘’‘the fourth dimension/* as though, instead of being merely an idea shaken loose from the ends of their pencils, only the fourth in a class of infinite possibilities, it was a physical reality, like a new element. Thus the lamentable confusion spread from mathematics to grammar, from the principles of the 2 -}- 2 to the science of the proper uses of the definite and indefinite article. * Physicists may consider time to be a fourth dimension, but not the mathematician. The physicist, like other sci¬ entists, may find that his latest machine has just the right place for some new mathematical gadget; that does not concern the mathematician. The physicist can borrow new parts for his changing machine every day for all the mathematician cares. If they fit, the physicist says they are useful, they are true, because there is a place for them in the model of his world in the making. When they no longer fit, he may discard them or “destroy the whole ma¬ chine and build a new one as we are ready to buy a new car when the old one doesn’t run well.** ^ The practice of calling time a dimension points to the necessity of explaining what is meant by that troublesome word. In this way, too, we shall arrive at a clearer image of four-dimensional geometry. Instead of referring to “a space,” or to “spaces,” we shall use the more fashionable and more general term— manijold* A manifold bears a rough resemblance to a 120 Mathematics and the Imagination class. A plane is a class composed of all those points uniquely determined by two co-ordinates. It is therefore a two-dimensional manifold. FIG. 28(a).—A two-dimensional manifold. Each point requires a pair of numbers to individualize it. A = (3, 2) B = (-5>-. 4) C = {x,y) D = (0, -3) E - (0, 0) FIG. 28(b).—The same idea can be extended to a three- dimensional manifold (space). Each point requires 3 numbers to individualize it. Thus, P = {x,y, z) The space studied in three-dimensional analytical geometry may be regarded as a three-dimensional mani¬ fold, because exactly three co-ordinates are required to I2I Assorted Geometries—Plane and Fancy fix every point in it. Generally, if n numbers are necessary to specify, to individualize, each of the members of a manifold, whether it be a space, or any other class, it is called an n-dimensional manifold. Thus, for the word dimension, with its many mysterious connotations and linguistic encrustations, there has been substituted a simple idea—that of a co-ordinate. And in place of the physical word space, the mathematician intro¬ duces the more general and more accurate concept of class, or manifold. ♦ It is now possible, as a consequence of these refine¬ ments, to introduce an idea already familiar from our discussion of analytical geometry, which shall serve to uniquely characterize space manifolds. We shall use some geometrical reasoning. The Pythagorean theorem states that,'in a right-angle 122 Mathematics and the Imagination triangle, the length of the hypothenuse equals the square root of the sums of the squares of the other two sides. FIG. 30.—The Pythagorean theorem in three di¬ mensions. ^2 = ^2 + ^2 -I- f2 For ^ c^-\- {eY and {ey = ^ When this is carried over into analytical geometry of two dimensions, the result is the well-known distance for¬ mula, according to which the distance between any two points in the plane, having the co-ordinates and (x',y') respectively, is^{x — x')^ -f- {jf yV- (1) Distance AB = V(x — x'y -h — yy (2) Distance AB = \/(x — x'y -h (j — /y -h U — ■c')* Assorted Geometries—Plane and Fancy 123 Similarly, in three-dimensional analytical geometry the distance between any two points having the co-ordinates z)y {x\ y, respectively, is _ y{x-x'y + {y-yy+ u - z')\ Now, in either two or three dimensions the concept of distance, as both the mathematician and the layman understand it, is the same. The layman is satisfied with an intuitive grasp; the mathematician demands an exact formulation. However, in the higher dimensions, while the layman is halted by a blank wall—the natural limitations of his senses—the mathematician scales the wall using his extended formula as a ladder. Distance in four dimensions means nothing to the layman. Indeed, why should it? For even a four-dimensional space is wholly beyond ordinary imagination. But the mathematician, who rests the con¬ cept upon an entirely different base, is not called upon to struggle with the bounds of imagination, but only with the limitations of his logical faculties. Accordingly, there is no reason for not extending the above formula to 4, 5, 6, ... or rz dimensions. Thus, in a four-dimensional Euclidean manifold, the distance of an element, i.e., point, having the co-ordinates (x^y, Zy u) from an element with co-ordinates {x\ y\ z\ n') is V{x — x'Y {y —y'y {z — zV + (« — uy. This method enables us to define in terms of analytical geometry a 2, 3, 4, ... or n-dimensional Euclidean manifold. An analogous definition can be given for the manifolds of other geometries, in which case some other distance formula would apply. We have chosen analytical geometry and taken the Pythagorean distance formula to distinguish the Euclidean manifolds. A condensed definition of a three- and four-dimen¬ sional Euclidean manifold in terms of analytical geom¬ etry reads: ^ 124 Mathematics and the Imagination 1. A three-dimensional Euclidean manifold is the class of all number triples: z), (x\y, z'), z"), etc., to any two of which there may uniquely be assigned a measure (called the distance between them) defined by the formula - x'Y + {y - y'Y + ( 4 : - z'Y. Cer- tain subclasses of this class are called points, lines, and planes, etc. The theorems derived from these definitions constitute a mathematical system called “Analytical Geometry of Three Dimensions.” 2. A four-dimensional Euclidean manifold is the class of all number quadruples: (x, y, z> u), (x', y', z', u'), etc., to any two of which there may uniquely be assigned a measure^ (called the distance be¬ tween them) defined by the formula ~ + (y ~y'Y {z — z'Y -h (« — u'y. Certain subclasses of this class are called points, lines, planes, and hyperplanes. Analytical four-dimensional Eu¬ clidean geometry is the system formed by theorems de¬ rived from these definitions. Note that nothing has been said in either of these defini¬ tions about space; neither the space of our sense percep¬ tions, nor the space of the physicist, nor that of the philos¬ opher. All that we have done is to define two systems of mathematics which are logical and self-consistent, which may be played like checkers, or charades, according to stated rules. Anyone who finds a resemblance between his game of checkers or charades and the physical reality of his experience is privileged to point morals and to make capital of his suggestion. * But having established that we are in the realm of pure conception, beyond the most elastic bounds of imagina¬ tion, who is satisfied? Even the mathematician would like Assorted Geometries—Plane and Fancy 125 to nibble the forbidden fruit, to glimpse what it would be like if he could slip for a moment into a fourth dimension. It’s hard to grub along like moles down here below, to hear someone tell of a fourth dimension, to make careful note of it, and then to plow along, giving it no further thought. To make matters worse, books on popular sci¬ ence have made everything so ridiculously simple—rela¬ tivity, quanta, and what not—that we are shamed by our inability to picture a fourth dimension as something more concrete than time. Graphic representations of four-dimensional figures have been attempted: it cannot be said these efforts have been crowned with any great success. Fig. 31(a) illus¬ trates the four-dimensional analogue of the three-dimen¬ sional cube, a hypercube or tesseract: Our difficulties in drawing this figure are in no way diminished by the fact that a three-dimensional figure can only be drawn in perspective on a two-dimensional surface—such as this page—, while the four-dimensional object on a two di¬ mensional page is only a perspective of a “perspective.” Yet since equals the area of a square, the volume FIG. 31 (a).—Two views of the tesseract. 126 Mathematics and the Imagination of a cube, we feel certain that describes something, whatever that something may be. Only by analogy can we reason that that “something” is the hypervolume (or content) of a tesseract. Reasoning further, we infer that the tesseract is bounded by 8 cubes (or cells), has 16 vertices, 24 faces and 32 edges. But visualization of the tesseract is another story. Fortunately, without having to rely on distorted dia¬ grams, we may use other means, using familiar objects to help our limping imagination to depict a fourth dimen¬ sion. The two triangles A and B in Fig. 32 are exacdy alike. Geometrically, it is said they are congruent, * meaning that by a suitable motion, one may be perfecdy super¬ posed on the other. Evidendy, that motion can be carried out in a plane, i.e., in two dimensions, simply by sliding triangle A on top of triangle 5.1 But what about the two triangles C and D in Fig. 33? One is the mirror image of the other. There seems to be no reason why by sliding or turning in the plane, C * See the chapter on paradoxes for an exact definition. ^ Actually, “sliding on top oP’ would be impossible in a physical two-dimensional world. Assorted Geometries—Plane and Fancy 127 cannot be superimposed on D. Strangely enough, this cannot be done. C or D must be lifted out of the plane, from two dimensions into a third, to effect superposition. Lift C up, turn it over, put it back in the plane, and then it can be slid over D. Now, if a third dimension is essential for the solution of certain two-dimensional problems, a fourth dimension would make possible the solution of otherwise unsolvable problems of three dimensions. To be sure, we are in the realm of fancy, and it need hardly be pointed out that a fourth dimension is not at hand to make Houdinis of us all. Yet, in theoretical inquiries, a fourth dimension FIG. 34.—This is no blueprint but an actual house in Fladand. 128 Mathematics and the Imagination is of signal importance, and part of the warp and woof of modem theoretical physics and mathematics. Ex¬ amples chosen from these subjects are quite difficult and would be out of place, but some simpler ones in the lower dimensions may prove amusing. If we lived in a two-dimensional world, so graphically described by Abbott in his famous romance, Flatland, our house would be a plane figure, as in Fig. 34. Entering through the door at we would be safe from our friends and enemies once the door was closed, even though there were no roof over our head, and the walls and windows were merely lines. To climb over these lines would mean getting out of the plane into a third dimension, and of course, no one in the two-dimensional world would have any better idea of how to do that than we know how to escape from a locked safe-deposit vault by means of a fourth dimension. A three-dimensional cat might peek at a two-dimensional king, but he would never be the wiser. When winter comes to Flatland, its inhabitants wear gloves. Three-dimensional hands look like this; FIG. 35. 130 Mathematics and the Imagination Modem science has as yet devised no relief for the man who finds himself with two left gloves instead of a right and a left. In Flatiand, the same problem would exist. But there, Gulliver, looking down at its inhabitants from the eminence of a third dimension, would see at once that, just as in the case of the two triangles on page 127, all that is necessary to turn a right glove into a left one is to lift it up and turn it over. Of course, no one in Fladand would or could lift a finger to do that, since it involves an extra dimension. If then, we could be transported into a fourth dimen¬ sion, there is no end to the miracles we could perform— starting with the rehabilitation of all ill-assorted pairs of gloves. Lift the right glove from three-dimensiond space into a fourth dimension, turn it around, bring it back and it becomes a left glove. No prison cell could hold the four-dimensional Gulliver—far more of a men¬ ace than a mere invisible man. Gulliver could take a knot and untie it without touching the ends or breaking it, merely by transporting it into a fourth dimension and slipping the solid cord through the extra loophole. Or he might take two links of a chain apart without breaking them. All this and much more would seem absurdly simple to him, and he would regard our he p- lessness with the same amusement and pity as we 00 upon the miserable creatures of Flatiand. ♦ Our romance must end. If it has aided some in making a fourth dimension more real and has satis e a common anthropomorphic thirst, it has served its pur pose. For our own part, we confess that the fables ave never made the facts any clearer. . . An idea originally associated with ghosts and spirits Assorted Geometries—Plane and Fancy 131 needs, if it is to serve science, to be as far removed as possible from fuzzy thinking. It must be clearly and courageously faced if its true essence is to be discovered. But it is even more stupid to reject and deride than to glorify and enshrine it. No concept that has come out of our heads or pens marked a greater forward step in our thinking, no idea of religion, philosophy, or science broke more sharply with tradition and commonly accepted knowledge, than the idea of a fourth dimension. Eddington has put it very well: ® However successful the theory of a four-dimensional world may be, it is difficult to ignore a voice inside us which whispers: “At the back of your mind, you know that a fourth dimension is all nonsense.” I fancy that voice must often have had a busy time in the past history of physics. What nonsense to say that this solid table on which I am writing is a collection of electrons moving with prodigious speed in empty spaces, which relatively to electronic dimensions are as wide as the spaces between the planets in the solar system! What nonsense to say that the thin air is trying to crush my body with a load of 14 lbs. to the square inch! What nonsense that the star cluster which I see through the telescope, obviously there nowy is a glimpse into a past age 50,000 years ago! Let us not be beguiled by this voice. It is discredited. . . . We have found a strange footprint on the shores of the un¬ known. We have devised profound theories, one after another to account for its origin. At last, we have succeeded in recon¬ structing the creature that made the footprint. And lo! It is our own. * We have emphasized the fact that pure geometry is divorced from the physical space which we perceive about us, and we are now prepared to tackle an idea which is slightly tougher. There is no harm, however, 132 Mathematics and the Imagination in first distinguishing somewhat differently than before between space as it is ordinarily conceived and the space manifolds of mathematics. Perhaps this distinction will help to make our new concept—the non-Euclidean geometries—seem less strange. We are quite used to thinking of space as infinite, not in the technical mathematical sense of infinite classes, but simply meaning that space is boundless without end. To be sure, experience teaches us nothing of the kind. The boundaries of a private citizen rarely go much further than the end of his right arm. The boundaries of a nation, as bootleggers once learned, do not go beyond the twelve-mile limit. Most of what we believe about the infinitude of space comes to us by hearsay, and another part comes from what we think we see. Thus, the stars look as if they were millions of miles away, although on a dark night a candle half a mile off would give the same impression. Moreover, if we imagined ourselves the size of atoms, a pea at a distance of one inch would appear mightier and far more distant than the sun. The distinction between the space of the individual and “public space” soon becomes apparent. Our personal knowledge of space does not show it to be either infinite, homogeneous, or isotropic. We do not know it to be infinite because we crawl, hop, and fly around in only tiny portions. We do not know it to be homogeneous because a skyscraper in the distance seems much smaller than the end of our nose; and the feather on the hat o the lady in front of us shuts off our vision of the cinema screen. And we know it is not isotropic, that is, it does not possess the same properties in every direction, because there are blind spots in our vision and our sense Assorted Geometries—Plane and Fancy 133 of sight is never equally good in all directions. The notion of physical or “public space” which we abstract from our individual experience is intended to free us from our personal limitations. We say physical space is infinite, homogeneous, isotropic, and Euclidean. These compliments are readily paid to an ideal entity about which very little is actually known. If we were to ask the physicist or astronomer, “What do you think about space?” he might reply: “In order to carry out experimental measurements and describe them with the greatest convenience, the physical scientist decides upon certain conventions with respect to his measuring appa¬ ratus and operations performed with it. These are, strictly speaking, conventions with regard to physical objects and physical operations. However, for practical purposes, it is convenient to assume for them a generality beyond any particular set of objects or operations. They then become, as we say, properties of space. That is what is meant by physical space, which we may define, in brief, as the abstract construct possessing those properties of rigid bodies that are independent of their material content. Physical space is that on which almost the whole of physics is based, and it is, of course, the space of everyday affairs.” ® On the other hand, the spaces, or more generally the manifolds, which the mathematician considers are con¬ structed without any reference to physical operations, such as measurement. They possess only those properties expressed in the postulates and axioms of the particular geometry in question, as well as those properties dcducible from them. It may well be that the postulates are themselves suggested, in part or in whole, by the physical space of 134 Mathematics and the Imagination our experience, but they are to be regarded as full-grown and independent. If experiments were to show that some, or all, of our ideas about physical space are wrong (as the theory of relativity has, in fact, done) we would have to rewrite our texts on physics, but not our geometries. ♦ But this approach to the concept of space, as well as to geometry, is comparatively recent. There has been no more sweeping movement in the entire history of science than the development of non-Euclidean geometry, a movement which shook to the foundations the age-old belief that Euclid had dispensed eternal truths. Compe¬ tent and accurate as a measuring tool since Egyptian times, intuitively appealing and full of common sense, sanctified and cherished as one of the richest of intel¬ lectual legacies from Greece, the geometry of Euclid stood for more than twenty centuries in lone, resplendent, and irreproachable majesty. It was truly hedged by divinity, and if God, as Plato said, ever geometrized, he surely looked to Euclid for the rules. The mathemadcians who occasionally had doubts soon expiated their heresy by vodve offerings in the form of further proofs in corroboration of Euclid. Even Gauss, the “Prince of Mathematicians,” dared not offer his criticisms for fear of the vulgar abuse of the “Boethians.” Whence came the doubts? Whence the inspiration of those who dared profane the temple? Were not the postu¬ lates of Euclid self-evident, plain as the light of day? And the theorems as unassailable as that two plus two equals four? The center of the ever-increasing storm, which finally broke in the nineteenth century was the famous fifth pos¬ tulate about parallel lines. This postulate may be restated as follows: “Through Assorted Geometries—Plane and Fancy 135 any point in the plane, there is one, and only one, line parallel to a given line.” There is some evidence to show that Euclid, himself, did not regard this postulate as “quite so self-evident” as his others.® Philosophers and mathematicians, intent on vindicating him, attempted to show that it was really a theorem and thus deducible from his premises. All of these attempts failed for the very good reason which Eu¬ clid, much wiser than those who followed him, had al¬ ready recognized, namely, that the fifth postulate was merely an assumption and hence could not be mathe¬ matically proved. * More than two thousand years after Euclid, a German, a Russian, and a Hungarian came to shatter two in¬ disputable “facts.” The first, that space obeyed Euclid; the second, that Euclid obeyed space. Gauss we credit on faith. Not knowing the extent of his investigations, in deference to his greatness as well as to his integrity, we are hospitable to his statement that he had independently arrived at conclusions resembling those of the Hungarian, Bolyai, some years before Bolyai’s father informed Gauss of his son’s work. Lobachevsky, the Russian, and Bolyai, both in the 1830 ’s, presented to the very apathetic scientific world their remarkable theories. They argued that the trouble¬ making postulate could not be proved, could not be deduced from the other axioms, because it was only a postulate. Any other hypothesis about parallels could be substituted in its place, and a different geometry—just as consistent and just as “true”—would follow. All the other postulates of Euclid were to be retained, only, in place of the fifth, a substitution was to be made: 136 Mathematics and the Imagination “Through any point in the plane, there are two lines parallel to any given line.” Overnight, mathematics had thrown off its chains, and a new line of richly fruitful theoretic and practical inquiry was born. * In the figure are two parallel lines: B A c FIG. 39. How is it possible, you may ask, that another line different from BC, yet parallel to DE may be drawn through A? The answer is that the reader is talking about the physical plane and lines drawn with a pencil. He is haunted by the ghosts of common sense instead of reasoning in teims of pure geometry. Tou might go further and say that in your system, in Euclidean geometry, any line different from BC will meet DE if sufficiently extended. We would reply that that rule holds in your game, not in ours—Lobachevskian geome¬ try. Neither of us, if we are mathematicians, are talking about physical space, but even if we were, there is better ground to believe that we are speaking the truth than you. Lobachevsky’s geometry may be introduced in this way: In Fig. 40 line AB is perpendicular to CD. If we per¬ mit it to rotate about A counterclockwise, it will intersect CD at various points to the right of B until it reaches a limiting position EF^ when it becomes parallel to CD. Assorted Geometries—Plane and Fancy 137 Continuing the rotation, it will start to intersect CD to the left of B. Euclid assumed that there is only one position for the line, namely £*F, when it would be parallel to CD. Lobachevsky assumed that there were two such positions, represented by A*B’ and C'D\ and further, that all lines falling within the angle 0, while not parallel to CD, would never meet it, no matter how far extended. A'. _ r — c D B FIG- 40 . Now this is an assumption, and there is no sense in arguing from the diagram that it is evident that if A'D\ or C'D' were extended sufficiently far, they would eventu¬ ally intersect CD. If, as Professor Cohen has pointed out, we rely wholly on our intuition of space, which is finite, there will always be an angle 0 which grows smaller as our space is extended, but which never vanishes, and all lines falling within 0 will fail to intersect the given line.*® + What happens to the geometry of Euclid when its parallel postulate is replaced by that of Lobachevsky? Many of its important theorems, those which in no way depend upon the fifth postulate, are carried over. Thus, in both geometries: 1. If two straight lines intersect, the vertical angles are equal: 138 Mathematics and the Imagination FIG. 41.—Angle 1 = Angle 2 (because each one = 180° — Angle 3). 2. In an isosceles triangle, the base angles are equal: FIG. 42.—If AB — BC, then Angle 1 = Angle 2. 3. Through a point, only one perpendicular can be drawn to a straight line: A 9 c n _? B FIG. 43.—Through the point A one and only one perpendicular can be drawn to CD. On the other hand, some very important theorems of Euclidean geometry are changed when another postu¬ late is substituted for the fifth, with startling results. Thus, in Euclidean geometry, the sum of the angles of every tri¬ angle equals 180 degrees, whereas in Lobachevsky’s geomr etry, the sum of the angles of every triangle is less than 180 Assorted Geometries—Plane and Fancy 139 degrees. Parallel lines in Euclidean geometry never inter¬ sect and remain, no matter how far extended, a constant distance apart. Parallel lines in Lobachevsky’s geometry never meet, but approach each other asymptotically —that is, the distance between them becomes less as they are further extended. To cite one more interesting theorem, two triangles in Euclidean geometry may have the same angles but different areas; i.e., one may be a magnification of the other. But in Lobachevsky’s geometry, as a triangle in¬ creases in areay the sum of its angles decreases; thus, only tri¬ angles equal in area can have the same angles. (See Fig. 47b,) ♦ The brilliant Riemann, in his famous inaugural lecture On the Hypotheses Which Underlie the Foundations of Geometry, proposed still another substitute for Euclid’s fifth postu¬ late differing from that of Lobachevsky and Bolyai, This assumption holds: “Through a point in the plane, no line can be drawn parallel to a given line.” In other words, every pair of lines in the plane must intersect. It should be noted that this contradicts Euclid’s tacit supposition that a straight line may be infinitely ex¬ tended. In this connection, Riemann pointed out the important distinction between infinite and unbounded: Thus, space may be finite though unbounded. Moving in any given direction, like the hands of a clock, we can keep going forever, forever retracing our steps. As might be expected, Riemann’s hypothesis also affects those theorems of Euclid dependent on the fifth postulate. Both Euclid’s and Lobachevsky’s geometries state that only one perpendicular can be drawn to a straight line from a given point. But in Riemann’s any number of 140 Mathematics and the Imagination perpendiculars can be drawn from an appropriate point to a given straight line. Again, the sum of the angles of any triangle is greater than 180 degrees in Riemann’s geometry, and the angles increase as the triangle grows larger. (See Fig. 47a, page 144.) ♦ We thus have three postulate systems; Euclid’s, Loba¬ chevsky’s, and Riemann’s. From these, three geometries have been developed: the first, Euclidean, the other two, non-Euclidean. The non-Euclidean geometries are, of course, vastly indebted to the postulates and the methods of Euclid. So far as the postulates are concerned, they dif¬ fer only with respect to the parallel postulate. The theo¬ rems differ greatly in many respects. A little earlier we laid down the criterion for every mathematical system—its postulates must be consistent, must lead to no contradictions. But how are we to discover whether the non-Euclidean geometries of Lobachevsky and Riemann are consistent? For that matter, it may well be asked, how are we to be certain that the postulates of Euclid will engender no contradictions? Evidentiy, we may pile up theorem after theorem without encountering any, but that is no proof that at some future time one may not arise. Is it that we are no better off than if we were PIG. 44.—^The pseudosphere. Assorted Geometries—Plane and Fancy 141 verifying an hypothesis in physics or any other experi¬ mental science? Fortunately mathematicians have devised a trick which satisfies their conscience on this score. It consists in show¬ ing, for example, in non-EucIidean geometry, that a set of entities which exist in Euclidean geometry would sat- 0)W FIG. 45(a).—One way of generating the tractrix. The toy locomotive L is tied to the watch the string being perpendicular to the track. When the locomotive starts pulling, the path of the watch is a tractrix. 142 Mathematics and the Imagination isfy the non-Euclidean theorems. It is assumed that these entities, themselves, are “free from contradictions, and that they in effect, fully embody the axioms,” and the latter are therefore shown to involve no inconsistencies. Let us take separate examplesfrom Lobachevsky’s and Riemann’s geometries to illustrate what is meant. Figure 44 illustrates a surface generated by revolving the curve known as the tractrix about a horizontal line. The tractrix itself may be obtained as follows: On a pair of mutually perpendicular axes, as in Cartesian geometry, imagine a chain lying along TY\ To one end of this chain there is attached a watch; the other end coincides with the point of origin 0. Keep the chain taut, and pull the free end slowly along the X axis, to the right of 0. Then repeat this procedure to the left. The path of the watch in either case generates a tractrix. If this curve is now revolved about the line XX , a “double trumpet surface,” as E. T. Bell calls it, is formed. FIG. 45(b).—The tractrix is also that curve which is perpendicular to a family of equal circles with their centers on a straight line. This surface Beltrami named a pseudosphere. We find that the geometry applicable to a pseudosphere is that of Lobachevsky. For example, on the pseudosphere, Assorted Geometries—Plane and Fancy 143 through a given point two lines may be drawn parallel to a third line, which will approach them asymptotically without ever intersecting. Thus, Lobachevsky’s geom¬ etry is satisfied by an entity from Euclid’s geometry, and this complies with the mathematician’s criterion of con¬ sistency. FIG. 45(c).—If perpendiculars are drawn to the curve (called the catenary) formed by a chain held at both ends, the curve which just touches all the perpendiculars is again the tractrix. The geometry of Riemann is applicable to a very famil¬ iar object—the sphere. It may be seen from Fig. 46 that a plane which passes through the center of a sphere cuts the surface in a great circle. Although the earth is somewhat oblate, for the purpose of this discussion we may consider it spherical. Every circle passing through the North and the South Poles on the earth’s surface is a great circle (longitude), but with the exception of the equator, the circles of latitude are not. Straight lines drawn on the surface of the earth are always parts of great circles, and even if two suck lines are perpendicular to a third line {which, in Euclidean geometry, would imply they are parallel), they will always 144 Mathematics and the Imagination intersect at a pair of poles. Thus, the elements for a geometry which will satisfy the surface of the earth are identical FIG. 46. with those of Riemannian geometry. For example, a triangle drawn on the surface of the earth will have FIG. 47(a).—Triangle A is small compared ynth the sphere; thus it is nearly a plane triangle and its angle sum is near 180®. . . • i. r But let it grow into triangle B, the sides of which he on three perpendicular great circles, and the angle sum = 90® + 90® + 90® = 270®. In the still larger triangle C, the angles of which arc all obtuse, the sum is greater than 270®. Assorted Geometries—Plane and Fancy 145 angles totaling more than 180 degrees, and the larger the triangle, the greater the sum of the angles. FIG. 47(b).— This is the reverse of what happens on a sphere, Fig. 47(a). On a pscudosphere, the larger the triangle, the smaller the sum of the angles. Furthermore, two straight lines drawn on the earth’s surface, if sufficiently extended, will always enclose an area. It is convenient at this point to recall the im{>ortant distinction noted by Riemann that a surface may be finite but unbounded, so that straight lines drawn upon the surface of the earth can be infinitely extended, al¬ though the surface is evidently not infinite, but merely unbounded. The Riemannian properties of the sphere are amusingly set out by the following riddle: A group of sportsmen, having pitched camp, set forth to go bear hunting. They walk 15 miles due south, then 15 miles due east, where they sight a bear. Bagging their game, they return to camp and find that altogether they have traveled 45 miles. What was the color of the bear? ♦ Our brief discussion of non-Euclidean geometry is bound to raise in the mind of the reader many questions 146 Mathematics and the Imagination outside our province, but the literature, even the popular literature, is so extensive that anyone sufficiendy inter¬ ested and curious need not go begging for answers. Yet it is perhaps proper that we should consider one very natural question which might take this form. On a sphere, two straight lines, even though parallel at one place, are certain (if sufficiently extended) to intersect, and may enclose an area. Why, then, call such lines ‘straight’? Are they not really curved?” At the outset it is obvious that whether a line is straight or not depends on the definition of “straight. In mathematics, it has been found convenient to formu¬ late such a definition only with reference to the particu¬ lar surface under consideration. One way of defining a straight line is to say that it is the shortest distance be¬ tween two points. On the other hand, everyone knows, from the many references in recent times to aeronautical exploits, that the shortest route between two points on the surface of the earth can be covered by following the arc of the great circle lying between them. Conveniently enough, through each two points on the surface of a sphere there does pass a great circle. The great circle, then, on the sphere, corresponds to the straight line in the plane—it is the shortest distance between two points. Suitable curves may be found for other types of surfaces, for instance, the pseudosphere, or a saddle-shaped surface which will fulfill the same role. Generalizing this notion, a curve which is the shortest distance between two points (analogue of the straight line in the plane) on any kind of a surface is called a geodesic of that surface. When we sought entities that would satisfy the geometry of Lobachevsky, and that of Riemann, we were really looking for surfaces, the geo- Assorted Geometries—Plane and Fancy 147 desics of which would obey the parallel postulates of these geometries. In the plane, if we adopt Euclid’s hypothesis, a pair of geodesics meet in one point, unless they are parallel, in which case they do not meet at all. On a sphere, a pair of geodesics (arcs of great circles), even if parallel, always meet in two points, and therefore the sphere obeys the geometry of Riemann. On a pseudosphere, obeying Lobachevsky’s geometry, parallel geodesics may ap¬ proach one another asymptotically, but never intersect. POSITIVE NEGATIVE ZERO FIG. 48.—Curvature. The geodesics of a surface are determined by its curvature. Curvature is not easy to explain, although we all have an intuitive notion of what it means. A plane has a curvature of 0. A surface like that of a sphere or an ellipsoid is one of positive curvature, whereas the saddle-shaped surface or the pseudospherc is said to be of negative curvature. We can imagine more complicated Mathematics and the Imagination surfaces, parts of which may have a positive, parts a negative, and parts a 0 curvature. The geodesics of a surface, as well as its most appropriate geometry, depend upon such curvature—positive, negative, or 0. Whence the geometry of a surface of constant negative curvature is Lobachevskian, that of a surface of constant positive curvature Riemannian, and that of a surface of 0 curva¬ ture Euclidean. All that has been said about non-Euclidean geometry, while evident enough when we talk of geometry^ tends to become obscure when applied to everyday surroundings. We are inclined to pity the inhabitants of a two-dimen¬ sional world, as much for their ignorance as for their physical limitations. They cannot even dream of doing things which to us are perfectly commonplace. Yet we tend to show the same intellectual limitations in picturing our world to ourselves. Indeed, we go further, for we deliberately reject our own experience. Our experience is that space is finite but unbounded, and that the straight lines we are able to draw on the surface on which we live can never recdly be straight, but must be curved. (Of course the earth’s curvature differs from 0.) But we continue to confuse infinity and unboundedness, to reject the latter which constitutes our actual spatial knowledge and to embrace the former for religious and aesthetic rea¬ sons. And, although every intelligent person knows the earth’s surface is curved, and every navigator practices great-circle sailing, most of us behave like Seventh-Day Adventists in reasoning that our straight lines are drawn in a plane of 0 curvature—or, in effect, in a world that is flat. From this it is only a step to the belief that Euclid s fifth postulate is sacred and any substitute is “against Assorted Geometries—Plane and Fancy 149 nature.” A little curvature, even more than a little learn¬ ing, has its disadvantages. Although we know a good deal more about the surface we inhabit than about the physical space in which we live, there is hardly any choice between the absurdities of our beliefs about either one. The geometry of Euclid, which considers surfaces of 0 curvature, in the strictest sense (disregarding convenience in computation) does not suit the surface on which we live as well as that of Rie- mann. Unmistakably, our geometries, though suggested by our sense perceptions, are not dependent upon them. The geometries we have discussed are only three of an infinite number of possible ones. Any geometry, whatever its postulates (provided they lead to no contradictions), will be just as “true” as the geometry of Euclid. For every surface, however complex its curvature, there is a peculiarly suited geometry. It is true we start our geom¬ etries as purely logical structures, but, as in other branches of mathematics, we find that Nature has anticipated us, and that a surface often waits upon our inventiveness. For that reason, the non-Euclidean mathematics has found enormously important fields of application in the weird topsy-turvy of modern physics. While we have considered the applications of two- dimensional non-Euclidean geometries to familiar sur¬ faces, the mathematical physicist studies the application of higher-dimensional non-Euclidean geometries to higher-dimensional space manifolds. In attempting to discover experimentally what space we actually live in, scientists have obtained results which lead them to believe that space is curved rather than straight. Having emanci¬ pated ourselves from the primitive idea that we live on a plane surface, curved space should not be so hard to take. 150 Mathematics and the Imagination There is a final point: If we consider the geometries of Euclid, Lobachevsky and Riemann as applied, and not as pure, mathematics, if we ask which one is most suitable to the space immediately surrounding us and the surface on which we live, what shall our answer be? Ex¬ periment and measurement alone can answer that ques¬ tion. It turns out that Euclid’s geometry is the most con¬ venient, and the one, in consequence, which we shall continue to use to build our bridges, tunnels, skyscrapers, and highways. The geometries of Lobachevsky, or Rie¬ mann, properly handled, would do just as well.^® Our skyscrapers would stand it, and so would our bridges, tunnels, and highways; our engineers might not. The geometry of Euclid is easier to teach, fits in more readily with misguided common sense, but above everything, is easier to use. And we are concerned, after all, in such matters with living, and not with logic. Yet our vistas have widened and our vision is clearer. Mathematics has helped us to transcend those sense impressions which we now say “deceive us never, while lying ever.” FOOTNOTES 1. St. Augustine, Confessions. —P. 12. 2. An illustration of pure mathematics: * Consider the following propositions, which are the axioms for a special kind of geometry. Axiom 7. If A and B are distinct points on a plane, there is at least one line containing both A and B. Axiom 2. If A and B cire distinct points on a plane, there is not more than one line containing both A and B. * Morris Raphael Cohen and Ernest Nagel, An Introduction to Logic and Scientific Method {Nc^Yot]l\ Harcourt Brace, 1936), pp- 133- 139. Assorted Geometries—Plane and Fancy 151 Axiom 3. Any two lines on a plane have at least one point of the plane in common. Axiom 4. There is at least one line on a plane. Axiom 5. Every line contains at legist three points of the plane. Axiom 6. Ail the points of a plane do not belong to the same line. Axiom 7. No line contains more than three points of the plane. These axioms seem clearly to be about points and lines on a plane. In fact, if we omit the seventh one, they are the assump¬ tions made by Veblen and Young for “projective geometry” on a plane in their standard treatise on that subject. It is un¬ necessary for the reader to know anything about projective geometry in order to understand the discussion that follows. But what are points, lines and planes? The reader may think he “knows” what they are. He may “draw” points and lines with pencil and ruler, and perhaps convince himself that the axioms state truly the properties and reladons of these geometric things. This is extremely doubtful, for the properdes of marks on paper may diverge noticeably from those postulated. But in any case the quesdon whether these actual marks do or do not conform is one of applied and not of pure mathematics. The axioms them¬ selves, it should be noted, do not indicate what points, lines, and so on “really” are. For the purpose of discovering the implicadons of these axioms, it is unessential to know what we shall understand by points, lines, and planes. These axioms imply several theorenns, not in virtue of the visual representation which the reader may give them, but in virtue of their logical form. Points, lines, and planes may be any entities whatsoever, undetermined in every way except by the relations stated in the axioms. Let us, therefore, suppress every explicit reference to points, lines, and planes, and thereby eliminate all appeal to spatial intuition in deriving several theorems from the axioms. Suppose, then, that instead of the word “plane,” we employ the letter S; and instead of the word “point,” we use the phrase '^element of •S’.” Obviously, if the plane (5) is viewed as a collection of points (elements of 5), a line may be viewed as a class of points (ele¬ ments) which is a subclass of the points of the plane (i*). We shall therefore substitute for the word “line” the expression ^'L-class'* Our original set of axioms then reads as follows: Axiom V. If A and B are disdnet elements of S, there is at least one L-class containing both A and B. Axiom 2'. If A and B are distinct elements of S, there is not more than one L-class containing both A and B. Mathematics and the Imagination Axiom 3'. Any two L’classes have at least one element of S' in common. Axiom 4'. There exists at least one L-class in .S'. Axiom 5'. Every L-