Zero and Infinity in the History of Mathematics “The Sultan Abdul Hamid the Second, perpetrator of the terrible nineteenth-century Armenian massacres, had his censors, they say, remove any reference to H20 from chemistry books entering his empire, [as he was] convinced that the symbol stood for ‘Hamid the Second is Nothing!’” “You flatter me – flatten, I mean,” said the Tortoise; “for you are a heavy weight, and no mistake! Well now, would you like to hear of a race-course, that most people fancy they can get to the end of in two or three steps, while it really consists of an infinite number of distances…?” As children – and perhaps as adults – each of us presumably wondered about infinity and nothingness: in the seeming infinity of a Möbius Band, the thought of what it would be like had we not been born, or the endlessness of the night sky: “twinkle, twinkle, little star, how I wonder what you are….” Perhaps in grammar school we told our classmates, “ I know how to count higher than infinity… Infinity plus one!”. Maybe someone read us Can you count to a Googol? by Robert Wells, and we wondered when they told us numbers did not stop at googol. Throughout the ages, the Infinite and the Nothing have been a perennial motif for all thinkers, woven in and out of the fabric of culture and time, politics, religions, and peoples; but they have been a particular hobbyhorse of theologians, mathematicians, philosophers, cosmologists, and mystics. Modern studies of infinity and zero still leave us spinning. As G.K. Chesterton penned, “ The poet only asks to get his head into the heavens. It is the logician who seeks to get the heavens into his head. And it is his head that splits.” Perhaps advances toward the understanding of nothingness and infinity have simply qualified the void and not quantified it. Nonetheless, our limited understanding of zero and infinity has proven to be useful and significant for the development of mathematics and society. Zero’s exact birthday into the world of written numbers is not known. It made “…shadowy appearances only to vanish again almost as if mathematicians were searching for it yet did not recognize its fundamental significance even when they saw it.” Zero has two distinct functions, namely a symbol for null (the way we use a “0” by itself) and a placeholder in a place-value number system. In modern science, zero is also used to indicate the degree of accuracy. The last digit of the number is the only number that involved subjectivity, e.g. in the case of 3.000, 3.00 is absolutely accurate and the last zero approximated. These functions of zero developed independently. Especially for those who only calculated so-called “real” problems (e.g. A farmer owned five horses and two died; how many are left?), zero was“…far from an intuitive concept.” One of those so-called “real” problems was astronomy, a developing science whose calculations and inductions depended on the ability to keep lists. While the ancient Greek mathematicians, such as Euclid, were working on deductive proofs using magnitudes and not numerical quantities, astronomers were working on inductive proofs based on empirical data. Although ancient Greek mathematicians had little need for the number zero, the Greek astronomers used the null to record what they saw (perhaps it was used more often on cloudy nights). Here, in ancient Greece, was the first appearance of our modern symbol for zero, the “O”. However, the concept of zero had not developed; the omicron “O” simply represented nothing, of which the Greeks cared to understand nothing. One of the important uses of zero is as a placeholder in a place-value number system. The ancient Babylonians were already using sexagesimal (base-60) place-value system for complex calculations at the time of Hammurabi (ca. 1800BC). In the first millennium BC (the “Seleucid” period), a separator symbol was introduced to indicate an empty space within a number. For example, the ancient reader would need to infer whether “2 6” meant 26, 206, or 2006 based on the context. To prevent confusion, Scribes put a small imprint between “2” and “6” when they wanted to indicate the number 206. This little imprint was a forerunner of one function of our modern zero. By 665 AD, the Mayans had developed a vigesimal (base-20) place-value system complete with a null-value placeholder similar to the Babylonians. The primary uses were calendrical or astrological. Unfortunately, Mayan mathematics all but died out with the European conquest and colonization. Mayan culture has no extant heirs. Although the Babylonians, Greeks, and Mayans began to use zero in their own right, its true conceptual development took place in India. Some historians have argued that the Indian mathematicians just developed near-Mediterranean concepts: “Pygmies placed on the shoulders of giants see more than the giants themselves.” However, there is evidence that Indian mathematics were already highly developed at the peak of Greek astronomy: A symbol for null was used in India as early as 200 BC, marked by the binary system described in Pingala’s Chandra-sutra. Even before this, Panini used a zero operator in his Astadhyayi. Kak states, “The development of the zero sign in India was motivated by numerical calculations…. contrasted from the manner in which the zero signs arose in Babylon and Mexico [the Mayans], where the motivation was from the areas of astronomy and calendrical calculations.” Three prominent Indian mathematicians attempted to discover the functions zero: Brahmagupta, Mahavira, and Bhaskara. In ca. 600 AD, Brahmagupta drafted several accurate laws for addition, subtraction, and multiplication operations involving zero. In the case of his statements regarding zero and division, he sometimes grappled with the truth, e.g. n divided by zero is a fraction with numerator n and zero as the denominator, or missed the truth altogether, e.g. zero divided by zero is zero. No major developments were made in the two hundred years following Brahmagupta. Mahavira clarified some of Brahmagupta’s ideas; but as indicated by his err in Ganita Sara Samgraha, division remained an enigma. Three centuries later, in ca. 1100 AD, Bhaskara outlined a number of laws concerning zero and came close to understanding division by zero. He wrote (n/0 = 8), which is close: today we know that division by zero is impossible. About one century later, Arabic culture fell heir to the Indian number system, which included zero, but the concept of zero as a number was lost or weakened at best. Fibonacci introduced the Hindu-Arabic number system to Europe at around the same time, but zero was not granted equal status with real numbers. Perhaps the dominant philosophical ideas hindered it: St. Thomas Aquinas’ philosophy of Ens, that “There is an Is”, dominated the intellectual climate of Europe at the time. Mortal Ens (“being”) is predicated of God’s infinite, perfect, and eternal Ens. Most likely, null was not seen as an entity but as a non-entity and was therefore deemed unnecessary. It was certainly not in vogue. It was not until the seventeenth century that zero had filtered down into common use. Infinity’s history is much more juicy. Hilbert’s Infinite Hotel (a common demonstration of 8 + 1) may be more like a haunted house, trapping the hapless transient with mirror rooms. Unlike zero, infinity is not a quantity we encounter in usual mathematical situations; we cannot find it at the grocery, the office, the home, or the street corner. We rarely encounter it, if ever, in scientific work, except in irrational numbers, such as p or the square root of two, but we “round” the decimal or our calculator does rounds it for us. Usually we only use the word “infinity” when speaking about non-mathematical things. We speak seldom of actual infinity but often of potential infinity:“I could look into her eyes forever” or “It seems I have an infinite number of papers on my desk”. Perhaps for these reasons, mathematicians did not see infinity as a worthy pursuit and therefore relegated it to the philosophers and mystics. Infinity’s salad days were spent amongst the dreamers. The first of these dreamers, Zeno of Elea (495-435 BC), is best known for his paradoxes, especially “Achilles and the Tortoise,” also demonstrated by his Dichotomy of the Room: Achilles can never overtake the Tortoise if he makes each step half as long as the preceding step. This illustrates “the unlimited density of points in any interval on the real number line.” The second of these dreamers, Pythagorus (569-500 BC), was the father of a philosophical movement. His fascination with infinity, irrational numbers, and numbers in general became what we now call number mysticism, or the worship of numbers. The Pythagoreans did make some accurate contributions to mathematics, namely the Pythagorean theorem, and to music, namely the understanding of intervallic ratios. Plato (428-347 BC) adopted the Pythagorean emphasis on number without the accompanying mysticism. The mathematicians trained at his school in Athens began to work with magnitudes instead of numbers. Euclid of Alexandria (c. 330-275 BC), for example, does not use actual numbers in his Elements. While he touched on potential infinity in his definition of a line or a plane, it was merely that one could extend these things to infinity if he wanted to, not that he would. Eudoxus of Cnidus (408-355 BC), also of Plato’s academy, used the idea of potential infinity in his “computation of the total area of a volume of a curved surface.” Eudoxus’ ideas were extended by Archimedes of Alexandria approximately a century later. At this point, infinity passed out of the hands of the mathematicians into the hands of the mystics and philosophers. Kabbalah, founded by Jewish Rabbi Joseph ben Akiva (50-132 AD), were a group of such mystics. By the thirteenth century, the Kabbalah believed that “Since God is Infinity and cannot be comprehended, the Sefirot [ten permutations of the letters of God’s name, YHWH] are the finite aspects… gleaned from the immensity of the Ein Sof [the infinite]…” which can be “studied and meditated upon, and prayed with.” The Kabbalah believed that Yahweh had given them a finite, bite-sized piece of His infinity. The infinite could be perceived. St. Augustine of Hippo (354-430 AD) also contemplated infinity and time. He demonstrates a complex understanding of infinity and the 8 + 1 principle in City of God, Book XII, Chapter 18. However, this understanding is used solely for the purpose of religious contemplation: “Does [God’s] knowledge extend only to a certain height in numbers, while of the rest He is ignorant…? All infinity is in some ineffable way made finite to God, for it is comprehensible to His knowledge.” Nicholas of Cusa (1401-1464) used the concept of infinity for similar meditations. Through the ingenious mathematical and astronomical cookery of Galileo Galilei (1564-1642), Infinity returned to the menu of mathematicians as “Infinity Florentine.” Ecclesiastics saw Galileo’s iconoclastic tendencies and his challenge to the earth-centered view of the universe as a recipe for disaster and eventually condemned Galileo to house arrest. Knowing that the authorities had cooked his proverbial goose, Galileo focused his attention on pure mathematics, and especially on the infinite, because he was no longer at liberty to conduct experiments. In his Dialogues on Two New Sciences (1638), he wrote that the number of squares of integers in an infinite set is not less than the number of integers (essentially 82 is not < 8 in an infinite set). In common language, this means, “an infinite set is not less than a part of itself.” Infinity lay dormant like the sleeping beauty in academic ivory towers for another two centuries, only to be awakened in Europe in the nineteenth century. Bernhard Bolzano (1781-1848) discovered what Zeno searched for in his paradoxes: aggregate infinity, or the concept that there are unlimited infinitesimally small numbers between any two points on the number line. He went on to prove that there are equally infinite numbers between zero and one as there are between zero and two. Infinity qua infinity was not articulated until the research of Georg Cantor (1845-1918). As infinity was for the Kabbalah, the Jewish number mystics, so it was for the Jewish mathematician, Cantor: to comprehend infinity was to comprehend the divine Mind. Cantor’s studies of infinity are possibly the most confusing and openly metaphysical of all mathematics. Cantor applied set theory to infinity. (Set theory was first drafted by Guiseppe Peano (1858-1932). “Zero [was defined as] the empty set. One was then defined as the set containing the empty set. Two was the set that contained the empty set and the set containing the empty set.” ) He assigned “¿”, or Aleph, the first letter of the Hebrew alphabet, to an infinite set. The notation for the infinity of integers and rational numbers, ¿0 , enabled the articulation of some laws of infinity, namely the following: ¿0 + 1 = ¿0 ; ¿0 + n = ¿0 ; ¿0 + ¿0 = ¿0 ; and ¿0 x ¿0 = ¿0. Infinity could now be expressed in equations. Cantor then proved that there are equally infinite numbers of points in one dimension (the line) as in two (the plane), three (cube), four dimensions, …ad infinitum. Thus, with infinity, it was possible for the part to be equal to the whole, contrary to Euclid’s fifth Common Notion. Cantor also sought to classify and differentiate levels of infinity. We must remember Peano’s set theory (vide supra): for every set of numbers, the power set (the set of all possible subsets) is greater. In the case of the number three, the power set would be the set of all possible subsets one could form the set from the set of the three elements. The power set of the set of three elements has eight elements: 23 = 8. Cantor then applied this logic to Infinity. Thus, the cardinality of the numerical continuum could be expressed thus: c=n¿0 , where c is the cardinality and n is any number on the numerical continuum. In the case of infinity, the power set was a “greater” infinity. It was later expanded by Felix Hausdorff (1868-1942) to the general statement 2¿a = ¿a+1. The number of ¿’s could thus go on infinitely, because ¿n always had a power set of greater cardinality. Cantor recognized the infinite number of infinites – that there was no cardinal number that bound the infinite – but he could not finish without implying finality. He named the set of all infinites z, or taf, the last letter of the Hebrew alphabet. Later, Kurt Gödel (1906-1978) developed the incompleteness theorem, part of which implies that the cardinality of z must be outside of the realm of the universe. Friedrich Nietzsche (1844-1900) once said, “If you gaze for long into the abyss, the abyss also gazes into you.” Many of those who attempt to fathom the infinite have lost their sanity (e.g. Cantor, Gödel, Rabbi Ben Zoma). At the climax of The Man Who Was Thursday, Chesterton’s captivating “nightmare,” Gogol, Syme, and Bull are all in fast pursuit of Sunday (who represents Nature distinct from God). As he is running, Sunday throws a parcel high in the air. Gogol stops to examine it, but finds it “…to consist of thirty-three pieces of paper of no value wrapped one round the other. When the last covering was torn away it reduced itself to a small slip of paper, on which was written: ‘The word, I fancy, should be ‘pink’’…” Is there anything at the end of infinity? Does it matter? It seems the infinite is a bridge between material and immaterial and the physical and metaphysical. In the end, what is the function of zero and infinity? Some mathematicians and philosophers would say that these concepts are dead ends in themselves, only useful by their application. Others, like Plato, would claim that contemplation of incommensurables is worthwhile because it sets the mind on higher things. Zero is a much more commonly used and understood than infinity, perhaps because of how common it has become in the past four centuries, but more likely because we can grasp it. Numerous new “sciences” depend on it, especially Statistics. Although we cannot contain “nothingness” in a box as with the bag of wind Aeolus gave to Odysseus in Homer’s Odyssey, we see it quite often. As a student in college, it was: I have zero money; as a new mother: I have no time of my own; as a city-dweller: I have no backyard. It only seems fitting that our numerical vocabulary should include null. Using zero as a placeholder in an empty column was not intuitive, however, but it certainly was the largest advance toward the immediate clarity and intelligibility of written numbers. It eliminated the possibility for multiple interpretations of one number. Infinity, on the other hand, has not had practical application in the world of commerce or technology. However, infinity and zero are not much ado about nothing; they have had a tremendous impact during their brief entrances on the world stage. Consider the confusion over whether it was 2000 or 2001that marked the new millennium. Infinity and zero remain as an idée fixe for many mathematicians and philosophers, a wildcard in the theologian’s hand, and as an enigma for all who dare to think about nothing. Works Cited and Selected Bibliography St. Augustine of Hippo, City of God, trans. Marcus Dods (New York: Modern Library, 2000). James Brackett, “Children’s conceptualizations of infinity,” Journal of Interdisciplinary Mathematics Vol. 1, No. 1. (1998). Lewis Carroll, “What the Tortoise Said to Achilles,” Mind 4, No. 14 (April 1895): 278- 280. Also found online at http://www.ditext.com/carroll/tortoise.html. G.K. Chesterton, Orthodoxy (Colorado Springs, Colorado: Shaw Books, 2001). G.K. Chesterton, The Annotated Thursday, ed. Martin Gardner (San Francisco: Ignatius Press, 1999), Joseph Dauben, Georg Cantor (Princeton: Princeton University Press, 1979). Stuart Hollingdale, Makers of Mathematics (London: Penguin Books, 1989). Robert Kaplan, The Nothing that Is (Oxford: Oxford University Press, 2000). Eli Maor, To Infinity and Beyond (Princeton: Princeton University Press, 1987). Richard Morris, Achilles in the Quantum Universe (New York: Henry Holt and Company, 1997). J.J. O’Connor and E.F. Robertson, “A history of Zero”, http://www-history.mcs.st-andrews.ac.uk/history/HistTopics/Zero.html